You are on page 1of 225

Modelling the Neuroanatomical Progression of

Alzheimer’s Disease and Posterior Cortical Atrophy

Supervisors:
Author: Prof. Daniel C. Alexander
Răzvan V. Marinescu Dr. Sebastian Crutch
Dr. Neil P. Oxtoby

A dissertation submitted in partial fulfillment


of the requirements for the degree of
Doctor of Philosophy
of
University College London

Centre for Medical Image Computing, University College London

April 8, 2019
I, Răzvan Valentin Marinescu, confirm that the work presented in this thesis is my
own. Where information has been derived from other sources, I confirm that this has
been indicated in the thesis.
Abstract

In order to find effective treatments for Alzheimer’s disease (AD), a devastating neurode-
generative disease affecting millions of people worldwide, we need to identify subjects
at risk of AD as early as possible. To this end, disease progression models have been
recently developed, which not only to perform early diagnosis, but also estimate a unique
disease signature that is used to predict the subjects’ disease stages and future evolution.
However, these models have not yet been applied to rare neurodegenerative diseases,
are not suitable to understand the complex dynamics of biomarkers, work only on large
multimodal datasets, and their predictive performance has not been objectively validated.
In this work I developed novel models of disease progression and applied them to
estimate the progression of Alzheimer’s disease and Posterior Cortical atrophy, a rare
neurodegenerative syndrome causing visual deficits. My first contribution is a study on
the progression of Posterior Cortical Atrophy, using models already developed: the Event-
based Model (EBM) and the Differential Equation Model (DEM). My second contribution
is the development of DIVE, a novel spatio-temporal model of disease progression that es-
timates fine-grained spatial patterns of pathology, potentially enabling us to understand
complex disease mechanisms relating to pathology propagation along brain networks.
My third contribution is the development of Disease Knowledge Transfer (DKT), a novel
disease progression model that estimates the multimodal progression of rare neurodegen-
erative diseases from limited, unimodal datasets, by transferring information from larger,
multimodal datasets of typical neurodegenerative diseases. My fourth contribution is the
development of novel extensions for the EBM and the DEM, and the development of
novel measures for performance evaluation of such models. My last contribution is the
organization of the TADPOLE challenge, a competition which aims to identify algorithms
and features that best predict the evolution of AD.
Impact Statement

The work presented in this thesis furthers our understanding of the temporal evolution
of Posterior Cortical Atrophy and Alzheimer’s disease. The disease progression models
and evaluation techniques that we developed can help towards understanding underlying
disease mechanisms, aid patient stratification and drug evaluation in clinical trials for
Alzheimer’s disease and Posterior Cortical Atrophy, and can be used in clinical practice
for predicting the future evolution of subjects that are at risk of developing Alzheimer’s
disease.
I published the work in this PhD thesis in two first-author papers (DIVE and TAD-
POLE chapters), and will soon submit another two papers (DKT and PCA chapters).
I have also communicated my results in international conferences. I have also engaged
with the broader scientific community by organising the TADPOLE Challenge, as well
as a couple of hackathons at the PyConUK conference and the CMIC Summer School.
Acknowledgements

There are many great people who have helped my PhD project become reality. First of all,
I’d like to thank my supervisor Daniel Alexander, for his great advice, ideas and research
directions. He has always encouraged me to pursue interesting ideas and supported me
in developing them. Secondly, I’d also like to thank Alexandra Young and Neil Oxtoby
for teaching me disease progression modelling, especially in the early years of my PhD.
I’d also like to thank Sebastian Crutch, Tim Shakespeare, Keir Yong, and other DRC
collaborators, for their help and advice on Posterior Cortical Atrophy and other clinical
aspects of my work. I’d like to thank Marco Lorenzi, for trying to explain mathematics
to a wanna-be mathematician like myself. Marco and Neil are also great guitar players,
which I had the opportunity to hear a few times. I’d also like to thank Sara Garbarino,
for her great spirit, for taking the time to repeatedly listen to my presentations when
rehearsing them, and for reminding me that I was probably the biggest nerd in CMIC.
I’d further like to thank the POND group, for the help they offered me throughout my
PhD, for the great coffees we had after our meetings, and for reminding me that I can’t
deal with non-working technology in hotels during our trips in the Netherlands. I’d like
to thank Gary Zhang for coaching me on how to present my work without putting half
of the audience to sleep, as well as others in MIG and CMIC, for teaching me about
diffusion MRI, machine learning and other imaging techniques. I remember coming to
those meetings in early days of my PhD and not understanding what was being discussed.
In terms of the social aspect, I had a wonderful time at UCL. I’ll miss the trips
organised by Pawel Markiewicz around Wales and Cornwall, where we had a lot of fun
surfing, playing frisbee and BBQ-ing on the beach. I’ll also miss the great camping trips
with the CMIC folks in Peak District and Lake district, when I attempted driving –
successfully! – for the first time in the UK! I’ll also miss the great time I had with Thore
Bucking, Emma Hill and Kin Quan during the MRes year. I’ll miss the dinners and
lunches such as the EuroPOND celebratory lunch, when we got so excited that we each
ordered 4 glasses of champagne, which got me tipsy. When we came back to UCL after
lunch I realised I was actually breaking the code instead of doing anything useful.
Finally, I’d like to thank my parents, Aurora and Dan Marinescu, for their love and
support, without which I wouldn’t have been able to start the PhD in the first place.
My brother Robert Marinescu, for his funny jokes and good spirit. My grandmother
Anghelută Constantina, for her funny and charismatic character. And my friends and
housemates, in particular Vibhav Mishra, Carlos Gavidia and Mikael Brudfors, for the
wonderful time spent in the Ifor residence, as well as Georgiana Ghetie, Alexandru Barbu
and Oana Lang, for their light-hearted spirit, conversations and for the fun we had in the
last few months of my PhD.
Contents

List of Figures 15

List of Tables 17

1 Introduction 19
1.1 Alzheimer’s Disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.2 Posterior Cortical Atrophy . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.3 Disease Progression Models . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.4 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.5 Justification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.5.1 Longitudinal Modelling of Posterior Cortical Atrophy . . . . . . . 21
1.5.2 Current Disease Progression Models Cannot Model Complex Dy-
namics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
1.5.3 Comparative Performance of Different Disease Progression Models 22
1.6 Thesis Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.6.1 Longitudinal Neuroanatomical Progression of Posterior Cortical At-
rophy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.6.2 DIVE: A Spatiotemporal Progression Model of Brain Pathology in
Neurodegenerative Disorders . . . . . . . . . . . . . . . . . . . . . 23
1.6.3 Disease Knowledge Transfer across Neurodegenerative Diseases . . 23
1.6.4 Novel Extensions to the Event-based Model and Differential Equa-
tion Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.6.5 TADPOLE Challenge: Prediction of Longitudinal Evolution in
Alzheimer’s Disease . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.7 Thesis Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2 Background – Alzheimer’s Disease 27


2.1 Alzheimer’s Disease . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.1.1 Symptoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.1.2 Disease Causes and Mechanisms . . . . . . . . . . . . . . . . . . . 28
2.1.3 Other Risk Factors . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.1.4 Biomarkers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.1.5 Diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.2 Progression of Alzheimer’s Disease . . . . . . . . . . . . . . . . . . . . . 38
2.2.1 Braak Staging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.2.2 Neuroimaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.3 Posterior Cortical Atrophy . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.3.1 Symptoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
10 Contents

2.3.2 Causes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.3.3 Diagnosis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.3.4 Management . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.3.5 Neuroimaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.3.6 Heterogeneity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

3 Background – Disease Progression Models 45


3.1 Hypothetical Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.2 Models of Progression using Symptomatic Groups . . . . . . . . . . . . . 47
3.3 Regression Against One Biomarker . . . . . . . . . . . . . . . . . . . . . 47
3.4 Survival Analysis Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.5 Scalar Biomarker Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.5.1 The Event-Based Model . . . . . . . . . . . . . . . . . . . . . . . 49
3.5.2 Differential Equation Model . . . . . . . . . . . . . . . . . . . . . 53
3.5.3 The Disease Progression Score Model . . . . . . . . . . . . . . . . 55
3.5.4 The Self-Modelling Regression Model . . . . . . . . . . . . . . . . 58
3.5.5 The Manifold-based Mixed Effects Model . . . . . . . . . . . . . . 59
3.6 Spatiotemporal Disease Progression Models . . . . . . . . . . . . . . . . . 60
3.6.1 The Voxelwise Disease Progression Model . . . . . . . . . . . . . 61
3.6.2 Cortical Atrophy Progression Model . . . . . . . . . . . . . . . . . 63
3.6.3 Parameter Estimation . . . . . . . . . . . . . . . . . . . . . . . . 64
3.6.4 Advantages and Limitations . . . . . . . . . . . . . . . . . . . . . 64
3.7 Mechanistic Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.7.1 The Network Diffusion Model . . . . . . . . . . . . . . . . . . . . 65
3.8 Machine Learning Methods . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.8.1 Advantages and Limitations . . . . . . . . . . . . . . . . . . . . . 67
3.9 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

4 Longitudinal Neuroanatomical Progression of PCA 71


4.1 Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.3 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.3.1 Participants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.3.2 Image Acquisition and Preprocessing . . . . . . . . . . . . . . . . 75
4.3.3 Statistical Methods . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4.4.1 Progression of PCA and Typical AD . . . . . . . . . . . . . . . . 78
4.4.2 Progression of PCA Subgroups . . . . . . . . . . . . . . . . . . . 79
4.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
4.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

5 Novel Extensions to the EBM and DEM 85


5.1 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.3 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.3.1 EBM Extensions . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
5.3.2 DEM – Optimised Trajectory Alignment . . . . . . . . . . . . . . 88
5.3.3 Performance Evaluation . . . . . . . . . . . . . . . . . . . . . . . 89
5.3.4 Data Preprocessing . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Contents 11

5.3.5 The Dementia Research Centre Cohort . . . . . . . . . . . . . . . 90


5.3.6 The Alzheimer’s Disease Neuroimaging Initiative Cohort . . . . . 91
5.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5.4.1 DRC Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5.4.2 ADNI Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
5.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
5.5.1 Model Performance on DRC cohort . . . . . . . . . . . . . . . . . 94
5.5.2 Model Performance on ADNI cohort . . . . . . . . . . . . . . . . 95
5.5.3 Staging-based Metrics . . . . . . . . . . . . . . . . . . . . . . . . 95
5.5.4 Diagnosis Prediction Metrics . . . . . . . . . . . . . . . . . . . . . 96
5.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
5.6.1 Limitations and Future Work . . . . . . . . . . . . . . . . . . . . 96
5.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

6 DIVE: A Spatiotemporal Progression Model of Brain Pathology 99


6.1 Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
6.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
6.3 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
6.3.1 DIVE Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
6.3.2 Modelling Subject-specific Parameters . . . . . . . . . . . . . . . 102
6.3.3 Modelling Biomarker Trajectory for a Single Vertex . . . . . . . . 102
6.3.4 Modelling Biomarker Trajectories for all Vertices . . . . . . . . . 103
6.3.5 Modelling Spatial Correlation . . . . . . . . . . . . . . . . . . . . 104
6.3.6 Fitting the Model using Generalised Expectation-Maximisation . 104
6.3.7 Implementation Details . . . . . . . . . . . . . . . . . . . . . . . . 107
6.3.8 Simulation Experiments . . . . . . . . . . . . . . . . . . . . . . . 108
6.3.9 Data Acquisition and Pre-processing . . . . . . . . . . . . . . . . 109
6.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
6.4.1 Results on Synthetic Data . . . . . . . . . . . . . . . . . . . . . . 111
6.4.2 Results with ADNI and DRC Datasets . . . . . . . . . . . . . . . 113
6.4.3 Model Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
6.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
6.5.1 Summary and Key Findings . . . . . . . . . . . . . . . . . . . . . 118
6.5.2 Limitations and future work . . . . . . . . . . . . . . . . . . . . . 118
6.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

7 Disease Knowledge Transfer across Neurodegenerative Diseases 121


7.1 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
7.2 Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
7.3 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
7.4 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
7.4.1 DKT Framework . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
7.4.2 Modelling Biomarker Trajectories . . . . . . . . . . . . . . . . . . 125
7.4.3 Parameter Estimation . . . . . . . . . . . . . . . . . . . . . . . . 125
7.4.4 Synthetic Experiment . . . . . . . . . . . . . . . . . . . . . . . . . 126
7.4.5 Data Acquisition and Preprocessing . . . . . . . . . . . . . . . . . 127
7.5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
7.5.1 Synthetic Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
12 Contents

7.5.2 Results on TADPOLE and DRC Datasets . . . . . . . . . . . . . 128


7.6 Validation on DTI Data in PCA . . . . . . . . . . . . . . . . . . . . . . . 133
7.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
7.8 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

8 TADPOLE Challenge: Prediction of Evolution in Alzheimer’s Disease 137


8.1 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
8.2 Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
8.3 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
8.4 Competition Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
8.5 Forecasts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
8.6 Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
8.6.1 ADNI Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
8.6.2 Image Preprocessing . . . . . . . . . . . . . . . . . . . . . . . . . 141
8.7 TADPOLE Datasets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
8.8 Submissions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
8.9 Forecast Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
8.9.1 Clinical Status Prediction . . . . . . . . . . . . . . . . . . . . . . 144
8.9.2 Continuous Feature Predictions . . . . . . . . . . . . . . . . . . . 145
8.10 Prizes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
8.11 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
8.12 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

9 Conclusions 149
9.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
9.2 Future Research Directions . . . . . . . . . . . . . . . . . . . . . . . . . . 151
9.2.1 Applications to Neurodegenerative Diseases . . . . . . . . . . . . 151
9.2.2 Applications to Clinical Trials . . . . . . . . . . . . . . . . . . . . 154
9.2.3 Methodological Developments . . . . . . . . . . . . . . . . . . . . 154
9.2.4 Model Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . 156

A Longitudinal Neuroanatomical Progression of PCA 157

B DIVE: A Spatiotemporal Progression Model of Brain Pathology 171


B.1 Simulations - Error in Estimated Trajectories and DPS . . . . . . . . . . 172
B.2 Comparison Between DIVE and Other Models . . . . . . . . . . . . . . . 172
B.2.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
B.2.2 Experiment Design . . . . . . . . . . . . . . . . . . . . . . . . . . 173
B.2.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
B.3 Derivation of the Generalised EM Algorithm . . . . . . . . . . . . . . . . 173
B.3.1 E-step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
B.3.2 M-step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
B.3.3 Optimising Trajectory Parameters . . . . . . . . . . . . . . . . . . 176
B.3.4 Estimating Subject Time Shifts - α, β . . . . . . . . . . . . . . . 178
B.3.5 Estimating MRF Clique Term - λ . . . . . . . . . . . . . . . . . . 179
B.4 Fast DIVE Implementation - Proof of Equivalence . . . . . . . . . . . . . 180
B.4.1 Trajectory Parameters - θ . . . . . . . . . . . . . . . . . . . . . . 181
B.4.2 Fast Implementation . . . . . . . . . . . . . . . . . . . . . . . . . 181
B.4.3 Slow Implementation . . . . . . . . . . . . . . . . . . . . . . . . . 181
Contents 13

B.4.4 Noise Parameter - σ . . . . . . . . . . . . . . . . . . . . . . . . . 182


B.4.5 Subjects-specific Time Shifts - α, β . . . . . . . . . . . . . . . . . 182
B.4.6 Fast Implementation . . . . . . . . . . . . . . . . . . . . . . . . . 182
B.4.7 Slow Implementation . . . . . . . . . . . . . . . . . . . . . . . . . 183

C Disease Knowledge Transfer across Neurodegenerative Diseases 185

D Novel Extensions to the EBM and DEM 187


D.1 EBM Fitting using Expectation-Maximisation . . . . . . . . . . . . . . . 187
D.1.1 M-step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
D.1.2 E-step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

E TADPOLE Challenge: Prediction of Longitudinal Evolution in AD 191


E.1 Expected Number of Subjects and Available Data for D4 . . . . . . . . . 191

F Bibliography 193
14 Contents
List of Figures

2.1 Prevalence of dementia around the world . . . . . . . . . . . . . . . . . . 27


2.2 Diagram showing the amyloid hypothesis . . . . . . . . . . . . . . . . . . 29
2.3 Different genes and associated risk for AD . . . . . . . . . . . . . . . . . 31
2.4 Different risk factors for AD related to lifestyle and the associated level of
evidence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.5 Intercalated pentagons used in the Mini-Mental State Examination (MMSE) 33
2.6 Comparison between the MRI brain scans of healthy subjects and subjects
with mild cognitive impairment . . . . . . . . . . . . . . . . . . . . . . . 34
2.7 FDG PET images of healthy and AD subjects . . . . . . . . . . . . . . . 35
2.8 Diffusion tensor image diagram . . . . . . . . . . . . . . . . . . . . . . . 36
2.9 Diagram showing the cerebro-spinal fluid (CSF). . . . . . . . . . . . . . . 37
2.10 Visual deficits and neuroimaging pathology in Posterior Cortical Atrophy 40

3.1 Hypothetical biomarker signatures in two diseases . . . . . . . . . . . . . 46


3.2 Biomarker cascade by Jack et al. [1] . . . . . . . . . . . . . . . . . . . . . 47
3.3 Event-based model diagram . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.4 MCMC perturbation rules in the event-based model . . . . . . . . . . . . 51
3.5 Event-based model - MCMC sampling diagram . . . . . . . . . . . . . . 52
3.6 Diagram of the Differential Equation Model (DEM) . . . . . . . . . . . . 54
3.7 ADNI biomarker trajectories estimated by Jedynak et al. [2] . . . . . . . 55
3.8 ADNI Biomarker trajectories estimated by Donohue et al. [3] . . . . . . . 58
3.9 Voxelwise disease progression model by Bilgel et al. [4] . . . . . . . . . . 61
3.10 Diagram of the cortical atrophy progression model by Koval et al. [5] . . 64
3.11 Diagram of the network diffusion model by Raj et al. [6]. . . . . . . . . . 65

4.1 Diagram of the Differential Equation Model . . . . . . . . . . . . . . . . 76


4.2 Atrophy progression in PCA and tAD patients according to the event-
based model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.3 PCA and tAD positional variance diagrams estimated by the EBM . . . 80
4.4 PCA and tAD trajectories estimated by the DEM . . . . . . . . . . . . . 81
4.5 PCA and tAD trajectories aligned in the same space, with samples from
the posterior distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.6 Early atrophy progression within the three cognitively-defined PCA sub-
groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

6.1 Diagram of the proposed DIVE model. . . . . . . . . . . . . . . . . . . . 101


6.2 The DIVE parameter estimation algorithm. . . . . . . . . . . . . . . . . 106
6.3 DIVE Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
6.4 DIVE Results on ADNI and DRC cohorts . . . . . . . . . . . . . . . . . 115
16 List of Figures

6.5 DIVE estimated clusters and trajectories over the 10 cross-validation folds 116
6.6 Scatter plot of DIVE-derived DPS scores vs cognitive tests . . . . . . . . 117

7.1 Diagram of the proposed framework for joint modelling of multiple diseases.123
7.2 The algorithm for estimating the DKT parameters . . . . . . . . . . . . . 126
7.3 DKT Simulation Results - Comparison between true and DKT-estimated
biomarker trajectories and subject time-shifts. . . . . . . . . . . . . . . . 129
7.4 Estimated biomarker trajectories for the ”synthetic PCA” disease, plotted
alongside true trajectories . . . . . . . . . . . . . . . . . . . . . . . . . . 130
7.5 DKT results - biomarker trajectories in the occipital unit and dysfunction-
ality scores for tAD and PCA . . . . . . . . . . . . . . . . . . . . . . . . 131
7.6 Estimated multi-modal trajectories for the PCA cohort. . . . . . . . . . . 132

8.1 Diagram showing the TADPOLE Challenge design . . . . . . . . . . . . 139


8.2 Venn diagram of the TADPOLE datasets derived from ADNI data. . . . 142

A.1 Labels of the different areas analysed in the EBM progression snapshots . 157
A.2 EBM bootstrap samples of the atrophy sequence for PCA and tAD . . . 160
A.3 Hypothesis testing of ordering of events within PCA and tAD . . . . . . 161
A.4 Positional variance diagram estimated by the event-based model, for three
PCA sugroups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
A.5 EBM bootstrap samples of the atrophy sequence, for three PCA subgroups 163
A.6 Hypothesis testing of the ordering of events within the three PCA subgroups.164
A.7 Testing for statistically significant differences in positions of each biomarker
in the EBM abnormality sequences, for both PCA and typical AD. . . . 168
A.8 Testing for statistically significant differences in biomarker positions in the
EBM sequences of PCA subgroups. . . . . . . . . . . . . . . . . . . . . . 169

B.1 DIVE: Error in DPS scores and trajectory estimation in simulations . . . 172

C.1 Estimated biomarker trajectories for the ”synthetic AD” disease, plotted
alongside true trajectories. . . . . . . . . . . . . . . . . . . . . . . . . . . 186
List of Tables

3.1 Comparison of features of various disease progression models. . . . . . . . 69

4.1 Demographic details for participants in the PCA study . . . . . . . . . . 73


4.2 Baseline population demographics for PCA subgroups . . . . . . . . . . . 74

5.1 Baseline population demographics for DRC data . . . . . . . . . . . . . . 90


5.2 Baseline population demographics for the ADNI cohort. . . . . . . . . . . 91
5.3 Model performance according to staging-based metrics on PCA subjects
from the DRC cohort . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
5.4 Model performance according to staging-based metrics on typical AD sub-
jects from the DRC cohort. . . . . . . . . . . . . . . . . . . . . . . . . . 93
5.5 Model performance at diagnosis prediction on DRC data. . . . . . . . . . 93
5.6 Model performance according to staging metrics on ADNI data. . . . . . 94
5.7 Model performance at prediction of conversion from MCI to AD on ADNI
data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

6.1 Demographics of the four cohorts from ADNI and DRC . . . . . . . . . . 110
6.2 Performance evaluation of DIVE and two simplified models on the ADNI
MRI dataset . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

7.1 Performance evaluation of DKT and other models . . . . . . . . . . . . . 134

8.1 The format of the forecasts for three example subjects. Participants have
to predict, for each subject, the probability of clinical diagnosis (CN/M-
CI/AD), the ADAS-Cog13 score and Ventricle volume, as well as the 50%
confidence range. RID - Roster ID is the unique identifier for ADNI sub-
jects, ADAS - ADAS-Cog13, CI - confidence range. Note that, even if the
CN/MCI/AD probabilities don’t sum to one, we will normalise them anyway.140
8.2 Subject statistics and available data in the TADPOLE datasets D1, D2,
D3 and D4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
8.3 Types of TADPOLE submissions that can be made by participants. . . . 144
8.4 TADPOLE prize allocation scheme using funds from AD charities . . . . 147

A.1 Statistical testing for significant differences in volumes of different brain


regions of PCA subjects at -10 years beforet0 . . . . . . . . . . . . . . . . 165
A.2 Statistical testing for significant differences in volumes of different brain
regions of PCA subjects at t0 . . . . . . . . . . . . . . . . . . . . . . . . 165
A.3 Statistical testing for significant differences in volumes of different brain
regions of PCA subjects at 10 years after t0 . . . . . . . . . . . . . . . . . 165
18 List of Tables

A.4 Statistical testing for significant differences in volumes of different brain


regions of tAD subjects at -10 years before t0 . . . . . . . . . . . . . . . . 166
A.5 Statistical testing for significant differences in volumes of different brain
regions of tAD subjects at t0 . . . . . . . . . . . . . . . . . . . . . . . . . 166
A.6 Statistical testing for significant differences in volumes of different brain
regions of tAD subjects at 10 years after t0 . . . . . . . . . . . . . . . . . 166
A.7 Statistical testing for significant differences in volumes of different brain
regions between PCA and tAD at -10, 0 and 10 years from t0 . . . . . . . 167

B.1 Comparison of DIVE with two more simplistic models on the ADNI MRI
dataset. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
Chapter 1

Introduction

1.1 Alzheimer’s Disease


Alzheimer’s disease (AD) is a chronic progressive neurodegenerative disorder that ac-
counts for 60% to 70% of all cases of dementia worldwide [7, 8]. In 2010 it was estimated
that up to 35 million people worldwide suffered from AD [8]. It’s symptoms include cog-
nitive dysfunction such as memory loss, language difficulties and psychiatric symptoms
such as depression, hallucinations, delusions and agitation. Diagnosis is usually based on
the person’s medical history, information from relatives and behavioural observations.
In terms of neuroimaging, Magnetic Resonance Imaging (MRI) shows early atrophy
in the medial temporal lobes and fusifom gyrus, which then spreads to the posterior
temporal lobe, parietal lobe, and finally to the frontal lobe [9], with relative sparing
of the sensorimotor cortex, visual cortex and the cerebellum. Imaging with Positron
Emission Tomography (PET) shows reduced metabolism and increased uptake of amyloid
proteins [10]. The underlying disease mechanisms are currently not well understood –
it is currently believed that initial abnormalities in the folding of amyloid-β and/or tau
proteins leads to a cascade of events which results in neurodegeneration and cognitive
decline [11]. These are known as the amyloid and tau hypotheses [11].
There are no treatments that can stop or at least slow down cognitive decline, because
all clinical trials so far have failed to prove any disease modifying effect [11]. One of the
reasons why clinical trials have failed might be due to a lack of understanding of the
underlying mechanisms, which results in wrong drug targets [12]. For example, within
the amyloid and tau hypotheses, it is not precisely understood what is the exact process
underlying the formation of the misfolded amyloid and tau and what might be the cause
of their misfolding [11]. Another reason why clinical trials in AD are believed to have
failed is the late administration of the treatment to patients who were already in the
symptomatic stage [12]. It is currently believed that for clinical trials to be successful in
AD, we need to fully understand the underlying disease mechanisms, in order to identify
the right drug targets, and to administer the treatments early in the pre-symptomatic
stages, and to the right subjects who will otherwise develop dementia in the future [12].

1.2 Posterior Cortical Atrophy


Alzheimer’s disease is a very heterogeneous disease, which has been observed both clin-
ically, with amnestic, visual, executive and aphasic types [13] as well as pathologically,
20 Chapter 1. Introduction

with hippocampal sparing and limbic predominant cases reported in the literature [14].
This heterogeneity can help us understand disease causes and underlying mechanisms,
and identify risk- and protective-factors. For example, it has been observed that differ-
ent speeds of progression can be due to differences in amyloid-β fibrils among subjects
[15]. Another example is that different ages of onset in familial AD are associated with
different underlying mutations in the PSEN1 gene [16].
A notable example of phenotypic heterogeneity in Alzheimer’s disease is given by
Posterior Cortical Atrophy (PCA). PCA, also called Benson’s syndrome [17], is a neu-
rodegenerative disease similar to AD that results in disruptions of the visual and motor
systems. Early symptoms include blurred vision, inability to read, difficulty with depth
perception and problems navigating through space [18, 19], while late-stage symptoms can
include inability to recognise familiar faces and objects as well as visual hallucinations.
Neuroanatomically, PCA is characterised by atrophy in the superior parietal, occipital
and posterior temporal regions [20, 21]. However, due to the rarity of the disease, only a
limited number of small studies have been done in PCA [18].

1.3 Disease Progression Models


For both PCA and typical AD (tAD), in order to understand the underlying disease
mechanisms and to select the right subjects for clinical trials, we need to quantitatively
map their longitudinal evolution. To this end, many biomarkers can be used, which are
based on Magnetic Resonance Imaging (e.g. brain volumes, cortical thickness), Positron
Emission Tomography (e.g. measures of hypometabolism, concentrations of amyloid and
tau proteins), samples from cerebrospinal fluid (CSF) (e.g. concentrations of various
molecular markers) or neuropsychiatric tests. However, no single biomarker is sufficient
for accurate staging and subject prediction, as they are not specific to one disease and
can result in misdiagnosis, can be influenced by variability not related to the disease (e.g.
the cognitive reserve theory [22]), show changes only in limited time windows, and have
inherent noise. Therefore, holistic, quantitative models called disease progression models
are needed, which integrate a variety of biomarker data to estimate the subjects’ disease
stage and future evolution.
A hypothetical model of disease progression has been proposed by [1], describing the
trajectory of key biomarkers along the progression of Alzheimer’s disease. The model sug-
gests that amyloid-beta and tau biomarkers become abnormal long before symptoms ap-
pear, followed by neurodegeneration and cognitive decline. Motivated by this idea, several
data-driven disease progression models have been proposed, that reconstruct biomarker
trajectories and can be used to stage subjects. One such model is the Event-Based Model
[23, 24], which estimates the progression of the disease as a sequence of discrete events,
representing underlying biomarkers switching from a normal to abnormal state. Another
model, the Differential Equation Model (DEM) [25], reconstructs a continuous trajectory
of biomarker measurements from changes in short-term follow-up data, which represent
samples of the slope at different points along the trajectory. Other models such as the
Disease Progression Score (DPS) [26], Self-Modelling Regression [3] or Riemannian mani-
fold techniques [27] have been developed, that build continuous trajectories by ”stitching”
together short-term follow-up data.
While these models have shown great promise at identifying the earliest events in
the Alzheimer’s disease cascade [24, 28], mapping the heterogeneity within Alzheimer’s
1.4. Problem Statement 21

disease [29] and showed increased performance in predictions compared to standard ap-
proaches [30], they have some limitations that need to be addressed. First of all, they
have not been applied to some rare neurodegenerative diseases such as Posterior Cor-
tical Atrophy. Secondly, they are not suitable for modelling the complex dynamics of
biomarkers. This is because they work on extracted features, which generally lack impor-
tant information present in the brain’s morphology; also, they cannot exploit biomarker
relationships shared across related diseases. Third, it is not yet clear how to measure the
performance of such models, and no previous literature study has been done to establish
the comparative performance of such models at different prediction tasks.

1.4 Problem Statement


In the field of Alzheimer’s disease progression, there are several issues that need to be
addressed:

• The longitudinal neuroanatomical progression of Posterior Cortical Atrophy has not


been quantified in a comprehensive study.

• Current disease progression models are not appropriate for modelling the complex
dynamics of biomarker measurements.

• The comparative performance of different models of disease prediction is yet to be


established.

The work I present in this thesis tries to address these three aspects.

1.5 Justification
1.5.1 Longitudinal Modelling of Posterior Cortical Atrophy
The longitudinal neuroanatomical progression of Posterior Cortical Atrophy has not been
quantified in a comprehensive study so far. Several case studies have been published,
which described the brain pathological progression of PCA [31, 32, 33, 34, 35, 36]. The
only longitudinal study of PCA [37] showed widespread gray matter loss in both PCA and
tAD. However, the numbers were small (17 PCA and 16 tAD) and the time interval was
short (1 year). Larger longitudinal studies are therefore required to robustly estimate the
progression of brain pathology in PCA, which is important for understanding underlying
disease mechanisms and for stratification of subjects clinical trials.

1.5.2 Current Disease Progression Models Cannot Model


Complex Dynamics
Current disease progression models are not appropriate for modelling the complex dy-
namics of biomarker measurements. For example, many models such as the event-based
model or the differential equation model cannot be applied to voxelwise biomarker data
such as amyloid load or hypometabolism from PET, or cortical thickness/compression
maps from MRI. While this can be mitigated by averaging these measures over pre-defined
22 Chapter 1. Introduction

regions of interest, it has been shown that patterns of pathology in different types of de-
mentia are dispersed and disconnected, as they follow underlying brain networks [38].
In order to study the link between neuroanatomical pathology and brain networks, we
need to develop spatio-temporal models of disease progression that account for changes
over the brain structure, as well as over the disease timeline. Such spatio-temporal mod-
els can help us understand more complex disease mechanisms and enable more accurate
predictions of disease risk, which can aid stratification in clinical trials.
Another limitation of current disease progression models is that it is challenging to
apply them to study rare types of dementia such as PCA. These models generally require
large multimodal datasets which are often not available for rare dementias. Therefore,
there is a need to develop models that can transfer information from larger multimodal
datasets. In particular for PCA, these transfer-learning approaches can enable us to esti-
mate robust, multimodal biomarker trajectories, and to make more accurate predictions
for each subject.

1.5.3 Comparative Performance of Different Disease


Progression Models
The comparative performance of different models of disease prediction is yet to be es-
tablished. More precisely, there has not been any study comparing the performance of
algorithms and features at longitudinal prediction of subjects at risk of AD. While these
questions are generally answered in the medical image community through grand chal-
lenges, most challenges so far have focused on classification of clinical diagnosis. For
example, the recent CADDementia challenge [39] aimed to predict clinical diagnosis from
MRI scans, while a similar challenge, the ”International challenge for automated predic-
tion of MCI from MRI data” [40], asked participants to predict diagnosis and conversion
status from extracted MRI features. While these challenges are helpful in establishing
which algorithms are best at predicting biomarkers at the current timepoint, they cannot
identify algorithms that are best at predicting the continuous progression of subjects at
risk of AD.

1.6 Thesis Contributions


In this thesis I contributed to the three key aspects mentioned above. My key contribu-
tions for each chapter are described in the following sections.

1.6.1 Longitudinal Neuroanatomical Progression of Posterior


Cortical Atrophy
• I performed the first comprehensive study of longitudinal atrophy progression in
Posterior Cortical Atrophy, and compared it with the atrophy progression in typical
Alzheimer’s disease, using data from the Dementia Research Centre (DRC), UK.
Previous studies were limited to case series, or used small numbers of patients over
short time-frames (1-year interval).

• I estimated the ordering in which brain regions show volume reductions using the
event-based model, and also estimated the rate and extent of volume loss using the
1.6. Thesis Contributions 23

differential equation model. I contrasted these between PCA and tAD, and showed
differences both qualitatively and quantitatively, which were further supported by
statistical tests.
• I showed that three cognitively-defined PCA subgroups show different phenotype-
specific patterns of early atrophy. This was the first study to show quantitative
evidence of heterogeneity within PCA.

1.6.2 DIVE: A Spatiotemporal Progression Model of Brain


Pathology in Neurodegenerative Disorders
• I developed DIVE, a novel spatiotemporal model that estimates fine-grained pat-
terns of pathology at every point on the cortical surface, while also accounting for
subject-specific time-shifts
• I validated DIVE on simulations, in presence of ground truth. More precisely, I
showed that DIVE can accurately estimate the true cluster assignments of each
simulated vertex, biomarker trajectories and subject-specific time-shifts.
• On patient data, I showed that DIVE estimates similar spatial patterns of pathology
in two independent typical AD datasets: the Alzheimer’s Disease Neuroimaging
Initiative (ADNI) and the Dementia Research Centre (DRC), UK.
• On patient data, I showed that DIVE estimates different spatial patterns of pathol-
ogy for distinct diseases (typical Alzheimer’s Disease vs Posterior Cortical Atrophy)
and distinct imaging modalities (MRI vs PET).
• I further validated DIVE on patient data, showing that it is robust under cross-
validation and that the subjects’ latent time-shifts, derived only from imaging data,
are clinically meaningful as they correlate with four different cognitive tests.
• I showed that DIVE has better or similar performance compared to standard ap-
proaches.

1.6.3 Disease Knowledge Transfer across Neurodegenerative


Diseases
• I developed DKT, a novel disease progression model that estimates multimodal
biomarker progressions in rare neurodegenerative diseases even when only limited,
unimodal data is available, by transferring information from larger multimodal
datasets from common neurodegenerative diseases.
• I validated DKT in a simulation in the presence of ground truth, where I showed
that it can accurately estimate biomarker trajectories in one disease, where there is a
complete lack of such data, by exploiting correlations with other known biomarkers.
• I demonstrated DKT on Alzheimer’s variants, where I showed it is able to infer
plausible non-MRI biomarker trajectories in a rare dementia, i.e. Posterior Cortical
Atrophy, by transferring such knowledge from a larger dataset of typical Alzheimer’s
disease.
• I showed that DKT has favourable performance compared to standard models.
24 Chapter 1. Introduction

1.6.4 Novel Extensions to the Event-based Model and


Differential Equation Model
• I made novel extensions to the methodology of two disease progression models, the
event-based model (EBM) and the differential equation model (DEM), which enable
better estimation of their parameters.

• I developed four novel performance metrics that were used to assess the performance
of all the models evaluated.

• I showed that the extended models had better or similar performance compared to
the standard models.

• My results also indicate that the novel performance metrics are more sensitive than
standard approaches based on the prediction accuracy of clinical diagnosis.

1.6.5 TADPOLE Challenge: Prediction of Longitudinal


Evolution in Alzheimer’s Disease
• I helped organise the TADPOLE Challenge, which aims to find algorithms and
features that best predict the evolution of subjects at risk of Alzheimer’s disease.

• I helped build the website and I created the main training dataset.

• I built a leaderboard system that enabled live evaluation of participants’ submissions


based on existing data.

• I promoted the competition at medical imaging conferences, and I organised two


TADPOLE mini-challenges, during the PyConUK 2017 conference and during the
CMIC Medical Imaging Summer School, 2018.

1.7 Thesis Structure


The thesis has the following structure:
• Chapter 2 contains background information on Alzheimer’s disease and Posterior
Cortical Atrophy.

• Chapter 3 contains background information on disease progression models.

• Chapter 4 contains the clinical analysis regarding the progression of Posterior Cor-
tical Atrophy as compared to typical Alzheimer’s disease.

• Chapter 5 presents novel extensions in the event-based model and differential equa-
tion model, which are evaluated against

• Chapter 6 presents the DIVE model formulation and results on four different
datasets, along with model validation.

• Chapter 7 presents the DKT model formulation, along with results on simulated
data and patient data, and model validation. standard implementations based on
performance metrics that I proposed.
1.7. Thesis Structure 25

• Chapter 8 presents the design of the TADPOLE Challenge.

• Chapter 9 presents a summary of the work in this thesis, and proposes directions
for further research.
26 Chapter 1. Introduction
Chapter 2

Background – Alzheimer’s Disease

2.1 Alzheimer’s Disease

Alzheimer’s disease (AD) is a chronic progressive neurodegenerative disease that affects


more than 35 million people worldwide [41], and this number is expected to triple by
2050 (Fig 2.1). Alzheimer’s disease is the most common cause of dementia, accounting
for 60% to 70% of the total cases of dementia [7, 8]. It usually affects people over 65
years of age [7, 8], although early onset forms of the disease also exist. The disease was
first described by German psychiatrist Alois Alzheimer’s in 1906. The worldwide cost of
dementia in 2018 was $818 billion worldwide, which is more than 1% of the aggregate
global gross domestic product (GDP) [42].

2.1.1 Symptoms

Symptoms of AD vary depending on the stage of the disease. Some authors [43] split the
symptoms into several categories: pre-dementia stage, mild dementia, moderate dementia
and severe dementia.

Figure 2.1: Prevalence of dementia around the world, along with forecasts for 2030 and
2050. Source: http://www.worldalzreport2015.org/
28 Chapter 2. Background – Alzheimer’s Disease

2.1.1.1 Pre-dementia Phase

In the pre-dementia stage, the first symptoms are usually attributed to stress and ageing.
Careful neuropsychological investigations may reveal very mild cognitive impairment five
years before the establishment of clinical diagnosis [43]. The performance of complex
tasks might be reduced, and alterations of behaviour including social withdrawal and
depressive dysphoria might also be already present [43].

2.1.1.2 Mild Dementia Stage

In the mild dementia stage, significant impairment of learning and memory are present
[43]. However, short-term and implicit memory are less affected compared to declara-
tive memory. Neuropsychological tests can reveal problems with object naming [44, 45],
semantic difficulties with word generation [44, 45] and inability to draw figures (i.e. con-
structional apraxia) [46]. Non-cognitive disturbances are also present at this stage [47],
where depression has been observed in these mild stages [48].

2.1.1.3 Moderate Dementia Stage

At the moderate dementia stage, the predominant features are severe short-term memory
impairment [49], along with difficulties in logical reasoning, planning, language [50], read-
ing [51] and writing [52]. More complex actions and activities such as using household
appliances, dressing and eating are gradually lost. Vision-related symptoms triggered by
cognitive deficits also develop, such as spatial disorientation, inability to recognise fa-
miliar faces or illusionary misidentification [53]. Around 20% of patients also experience
visual hallucinations, which may be associated with cholinergic deficits [54].
Patients at this stage cannot survive in their community without help from caregivers.
However, hospital or nursing home admission can be delayed if there is a good support
system in place at the patient’s home.

2.1.1.4 Severe Dementia Stage

Specific cognitive dysfunctions cannot be disentangled at this stage, due to widespread


cognitive deficits. Language is reduced to simple phrases. However, emotional signals can
still be received and returned [43]. Patients need support for performing basic functions
such as eating.
The average life expectancy after clinical diagnosis is between three to nine years,
although the speed of progression can vary [41]. Pneumonia, myocardial infarction and
septicaemia are the most frequent causes of death at this stage.

2.1.2 Disease Causes and Mechanisms


The causes for AD are poorly understood and around 70% of them are thought to be
genetic, with many genes involved which include APOE, GSK3β and DYRK1A[55]. Over
the last few decades, several hypotheses have been proposed to explain the mechanisms
of AD: amyloid hypothesis, tau hypothesis, cholinergic hypothesis and neurovascular
hypothesis.
2.1. Alzheimer’s Disease 29

Figure 2.2: Diagram showing the amyloid hypothesis. Amyloid precursor protein is split
by α-secretase resulting in sAPPα, which might have a neuroprotective role. On the
other hand, splitting by β-amyloid cleaving enzyme (BACE) results in amyloid-β, of
which amyloid-β42 is more prone to self-aggregate and lead to pathogenesis. On the left,
many other factors are shown that are believed to influence this pathway and lead to
more pathology. Reproduced with permission from [11].

2.1.2.1 Amyloid Hypothesis


In 1991, Hardy et al. [56] postulated that amyloid-β deposits are a central cause in
the development of AD. The amyloid-β protein, derived from the amyloid precursor
protein (APP), is processed via two distinct pathways: the amyloidogenic pathway which
produces amyloid-β proteins (Fig 2.2) and the non-amyloidogenic pathway which prevents
the formation of amyloid-β and instead produces a secreted form of APP called sAPPα
[11]. The amyloid hypothesis states that dysregulation in APP processing occurs early in
the disease process, causing increased production of the more toxic amyloid-β42 protein,
which aggregates into plaques [11]. The misfolded amyloid-β then causes a chain of events
leading to cognitive impairment, including tau aggregation, phosphorylation, neuronal
damage and brain atrophy. However, the underlying mechanisms through which amyloid-
β induces neurodegeneration are not clear [11].
Different sources of evidence exist to support the amyloid hypothesis. First of all,
mutations in the APP gene cause a rare, early-onset form of familial Alzheimer’s disease
which corresponds clinically and pathologically to AD. This suggests that changes in
APP are an upstream 1 event in the pathological cascade leading to AD [11]. Moreover,
a locus on chromosome 10 which is linked to late onset AD is also associated with increased
amyloid-beta production [57].
Several studies have also established a clear link between amyloid and tau toxicity,
another early event in the AD cascade [58, 59, 60, 61]. Amyloid has been shown to
enhance tau tangle formation in several mice studies [58, 59]. Evidence also exists that
tau pathology is required for amyloid-beta toxicity [60], suggesting that there could be
a feedback loop between amyloid and tau, or that tau pathology is also required for
development of amyloid deficits [61].
There are several aspects of the amyloid hypothesis that indicate it is not complete.
For example, transgenic mouse models carrying the familial AD mutations have showed

1
happening early in the chain of events leading to AD
30 Chapter 2. Background – Alzheimer’s Disease

increases in amyloid toxicity, but no clear evidence of neuronal loss [62, 63] and tau
aggregation as predicted by the amyloid hypothesis [61].

2.1.2.2 Tau Hypothesis


Another key hypothesis about the cause of AD is the tau hypothesis, which suggests that
abnormalities related to the tau proteins initiate the disease cascade [11]. In this case, tau
binding to microtubules is disrupted by phosphorylation, which results in free tau that
aggregates into neurofibrillary tangles. This ends up destroying the cell’s cytoskeleton
which collapses the neuron’s transport system, ultimately resulting in neuronal death.
Support for the tau hypothesis was given by the fact that tau proteins aggregate
and accumulate within neuronal cells and ultimately cause their death. Moreover, the
number of tau tangles has been shown to correlate with cognitive decline [64], especially
in memory-related areas [65, 66]. Furthermore, discoveries of tau aggregation in fronto-
temporal degeneration (FTD) suggest that tau alone can cause degeneration [67].
There is also evidence that the tau hypothesis is incomplete. For example, the fact
that mutations in tau-related genes give rise to tau tangles but no plaques, yet mutations
in the APP gene result in both plaques and tangles suggest that amyloid toxicity might
occur upstream, before tau toxicity [11].

2.1.2.3 Cholinergic Hypothesis


An older hypothesis, on which current AD therapies rely, is the cholinergic hypothe-
sis, which suggests that degeneration of cholinergic neurons and associated disruption
of cholinergic neurotransmission are the main causes of pathology in AD [68]. Support
for the theory came in mid-1970s, where studies provided evidence of deficits in synthe-
sis of neurotransmitter acetylcholine (ACh) and choline acetyltransferase (ChAT) [69].
However, the cholinergic hypothesis lost support due to unsatisfactory results of the
cholinergic drugs [70]. Despite not being disease-modifying, cholinergic drugs have been
shown to provide symptomatic benefits through improved memory and clinical function
[70].

2.1.2.4 Vascular Hypothesis


A vascular hypothesis has also been proposed for AD [71], suggesting that one of the
incipient causes is related to vascular abnormalities, which leads to brain hypoperfusion,
neurodegeneration and cognitive impairment. One study even indicates that this is an
earlier event than amyloid and tau accumulation [28]. Evidence supporting this theory
has been given by the close associations between dementia and stroke [72, 73], cardiac
diseases [74, 75, 76] and atherosclerosis [77].

2.1.2.5 Genetic Causes


Alzheimer’s disease has a strong genetic component, with many genes involved that alter
the risk of developing the disease and the pace of progression. Twin studies show that
disease heritability ranges between 60% to 80% [79, 80]. Currently there are two main
forms of AD: familial AD and sporadic AD. Familial AD is a rare early-onset AD (EOAD)
characterised by autosomal dominant disease transmission, and caused by mutations
in three genes, APP, PSEN1 and PSEN2, which code for amyloid peptide precursor,
2.1. Alzheimer’s Disease 31

Figure 2.3: Diagram showing different genes which increase the risk for AD (y-axis), as
well as their frequency within the population (x-axis). EOAD genes APP, PSEN1 and
PSEN2 (top-left) give a near-certain risk of developing AD, but are found in a very small
minority of the AD population. APOE4 has a moderate risk, while the other genes have a
lower risk, yet are found in a much larger population. Reproduced from [78], CC BY-NC.

presenilin 1 and 2 respectively [81]. Sporadic AD is the most common, late-onset form
of AD (LOAD), characterised by more complex, non-Mendelian transmission.
In familial AD, several genetic risk factors have been identified so far. In 1980s,
the discovery of amyloid-β peptides in AD senile plaques and the identification of these
peptides in the brains of people with Down’s Syndrome, caused by abnormalities in chro-
mosome 21 and where dementia was also observed, led the the hypothesis that mutations
of a gene located on chromosome 21 might cause AD in people without Down’s syndrome
[82]. A few years later, a linkage peak was indeed found on chromosome 21 [83], and the
APP gene was identified [84] and confirmed in EOAD families [85]. However, the amount
of heterogeneity observed in EOAD suggested additional genes were involved, and further
genetic linkage analyses led to the discovery of PSEN1 [86] and PSEN2 genes [87]. As of
March 2014, 40, 197 and 25 mutations were reported in APP, PSEN1 and PSEN2 genes
respectively, all with autosomal dominant transmission with complete penetrance, with
the exception of one mutation in the APP gene [81].
In sporadic, late-onset AD, the genetic landscape is much more complex. The most im-
portant risk factor is given by mutations in genes coding for Alipoprotein E (APOE) [81].
APOE is a protein whose key function is to transport lipids and cholesterols throughout
the body, and has three major isoforms called APOE2, APOE3 and APOE4, correspond-
ing to alleles 2, 3 and 4. Increased risk of AD due to APOE4 has been established in
1993 in three key studies [88, 89, 90]. Until 2005, more than 500 candidate genes other
than APOE have been identified using association studies, with various pathways involved
including tau phosphorylation, vacuolar sorting, glucose and insulin metabolism, nitrous
oxide synthesis, oxidative stress, growth factors, inflammation and lipid-related pathways
[81]. However, after the advent of genome-wide association studies (GWAS) in 2005, the
32 Chapter 2. Background – Alzheimer’s Disease

Figure 2.4: Diagram showing different risk factors for AD related to lifestyle and the
associated level of evidence. Reproduced from [106], CC-BY-NC-ND.

first genes outside the APOE locus were identified in two independent studies [91, 92].
Several genes including CLU [91, 92], CR1 [92], BIN1 [93], PICALM [91], ABCA7 [94]
and CD2AP [95] have been since identified. Moreover, associations were also found with
quantitative endophenotypes, which provide more statistical power than yes/no disease
status, such as early age of onset [96, 97], greater burden of amyloid pathology [98, 99],
abnormal levels of cerebro-spinal fluid (CSF) [100, 101], decrease in total brain volume
[102, 103] and decreased cognitive scores [104, 99, 105].

2.1.3 Other Risk Factors


There are several known risk factors that are associated with AD. The principal risk
factor is age, with incidence rates doubling every 5 years after 65 years of age [41, 107].
Other risk factors include head injuries, depression and hypertension [7]. Lifestyle factors
such as smoking [108] also increase the risk for developing AD. There is also potential
evidence that living in polluted areas increases the risk for AD [109]. Physical exercise
is also associated with lower risk of developing dementia [110]. Other factors influencing
the risk of dementia are shown in Fig. 2.4, and include traumatic brain injury, obesity,
hypertension, diabetes (increases risk), with protective factors including higher levels of
education [106].

2.1.4 Biomarkers
The information in this section has been initially written by me for the TADPOLE
Challenge website2 , with feedback from Esther E. Bron and Daniel C. Alexander. The
material has been subsequently adapted for this thesis.
Over the last decades, various biomarkers have been developed to quantify the severity
of Alzheimer’s disease and track its progression:

• Cognitive tests such as the Mini-Mental State Examination (MMSE) [111] are used
to assess memory and cognitive performance (section 2.1.4.1).

2
https://tadpole.grand-challenge.org/Data/
2.1. Alzheimer’s Disease 33

Figure 2.5: Intercalated pentagons used in the Mini-Mental State Examination (MMSE).
Patients with dementia have difficulty drawing them. Image source: Wikipedia3 CC-SA.

• Magnetic Resonance Imaging (MRI) measures such as cortical volumes, thickness


and atrophy rates detect shrinkage of individual brain areas that is caused by neu-
rodegeneration (section 2.1.4.2).

• Positron Emission Tomography: can be used to measure neuronal metabolism


through Fluorodeoxyglucose (FDG) PET [112], amyloid uptake through the Pitts-
burgh compound B (PiB). [113] and, more recently, tau uptake through AV1451
PET (section 2.1.4.3).

• Cerebro-spinal fluid markers: can be used to measure amyloid plaque deposits [114]
and neurofibrillary tangles through CSF total tau and phosphorylated tau [114]
(section 2.1.4.5).

2.1.4.1 Cognitive Tests


Cognitive tests are neuropsychological tests performed by a clinical expect and can assess
different cognitive domains such as general cognition, memory, language, vision. These
give an overall sense of whether subjects are aware of their symptoms, surrounding en-
vironment and whether they can remember a short list of words, follow instructions and
do simple calculations. For instance, in the Mini-Mental State Examination (MMSE),
patients are asked to draw intercalated pentagons (Fig 2.5). As there are no population
standards, performance in many of these tests is measured relative to a control group,
after adjusting for age, sex and education [111].
Cognitive tests are important in Alzheimer’s disease because they measure cognitive
decline in a direct and quantifiable manner. As a result, they are required for establishing
a clinical diagnosis of probable AD in criteria such as NINCSD-ADRDA [115, 116]. Apart
from establishing the AD diagnosis, these tests are valuable also for establishing patterns
of cognitive impairment, assessing changes over time, comparing drug efficacy and for
establishing correspondences with other imaging, histopathology or molecular biomarkers
[111].
Cognitive tests have several limitations. First of all, they suffer from practice effects,
i.e. patients who undertake the same test several times can learn/remember how to do
it, and thus score higher at a follow-up visit. This limits the usefulness of the test in
assessing dementia. Another limitations is that they have floor or ceiling effects, which
means that many subjects might score the highest/lowest score possible. Finally, they can
also be biased, as each subject is evaluated by a human expert who might be influenced
by prior knowledge of the subject’s cognitive abilities.

3
https://commons.wikimedia.org/wiki/File:InterlockingPentagons.svg
34 Chapter 2. Background – Alzheimer’s Disease

Figure 2.6: Comparison between the MRI brain scan of a healthy subject (left) and sub-
jects with different types of mild cognitive impairment (MCI) (middle-right), showing
different patterns of atrophy for each group. MRI is a widely used technology for mea-
suring the spatial distribution and extent of atrophy and for tracking the progression of
Alzheimer’s disease (AD). Reproduced from [117], CC-BY license.

2.1.4.2 Magnetic Resonance Imaging

Magnetic resonance imaging (MRI) is a technique used to image the anatomy and the
physiological processes of the brain and other body parts. With MRI, brain structures
can be quantified due to different contrast between gray matter (GM), white matter
(WM), cerebrospinal fluid (CSF) and hard tissue such as the skull. The GM is the brain
tissue that consists of the bodies of neurons, while the WM consists of fibres connecting
the neurons. The cerebrospinal fluid is a clear, colourless fluid providing mechanical and
immunological protection to the brain. Within MRI, different types of contrast between
tissues can be obtained through T1, T2, T1-weighted and T2-weighted images.
Brain MRI has been successfully applied to quantify neurodegeneration in Alzheimer’s
disease. Brain atrophy, which is caused by the death of neurons, can be visually assessed
in MRI scans due to shrinkage of the brain (see Fig. 2.6) and can be quantified using
markers of volume, cortical thickness, surface areas, along with changes in these values
between a baseline and a follow-up scan. These quantitative markers can be obtained
with specialised software such as Freesurfer [118].
MRI-derived biomarkers have both advantages and limitations. They are robust and
have less noise compared to cognitive tests, and are non-invasive. Moreover, they are also
a good indicator of progression from MCI to dementia in an individual subject because
they become abnormal slightly earlier than the onset of dementia-specific symptoms [1,
119]. Limitations of these markers are that MRI scans are expensive, require specialised
equipment to be acquired, and can also suffer from motion artefacts.
2.1. Alzheimer’s Disease 35

Figure 2.7: (top) Fluorodeoxyglucose (FDG) PET images for a cognitively normal sub-
ject (left) and a subject with Alzheimer’s disease (right). FDG PET measures cellular
metabolism, which is known to decrease during the development of AD. There is decreased
metabolism in parietal and frontal regions (gray arrows) in the AD subject compared to
the cognitively normal subject. (bottom) Pittsburgh B (PiB) PET image measuring
amyloid uptake in the brain of a healthy control (left) and AD subject (right). There
is widespread amyloid presence in the brain of the AD subject. Image reproduced with
permission from [120].

2.1.4.3 Positron Emission Tomography


Positron Emission Tomography (PET) detects pairs of gamma rays emitted by a ra-
dioactive tracer, which is introduced into the body of a biologically active molecule.
Three-dimensional images of tracer concentration within the body are then constructed
by computer analysis. Before a PET scan, the patient is injected with a contrast agent
(containing the tracer) which spreads throughout the brain and binds to abnormal pro-
teins (amyloid and tau). This enables researchers to track the concentration of these
proteins. PET scans can be of several types, depending on the cellular and molecular
processes that are being measured:

• cell metabolism using Fluorodeoxyglucose (FDG) PET: Neuronal cell metabolism


refers to the the activity going on inside neuronal cells such as the processing of
food and elimination of waste. Metabolic processes use glucose, hence FDG PET
quantifies metabolism by measuring the amount of glucose within each voxel. In
Alzheimer’s disease, neurons that are about to die will show reduced metabolism,
hence FDG PET is also an early indicator of neurodegeneration.

• levels of abnormal proteins such as amyloid-beta through AV45 PET: Amyloid-beta


misfolding (i.e. errors in the construction of its 3D structure) is thought to be one
of the causes of Alzheimer’s disease (see section 2.1.2.1). AV45 PET can be used
to measure the levels of amyloid in the brain, and is hence one of the earliest AD
markers.

• levels of abnormal tau proteins through AV1451 PET: Abnormal phosphorylated


tau (i.e. tau protein and a phosphorus group) that gather together in an insoluble
form eventually cause damage to the neuron’s cytoskeleton, leading to the collapse
36 Chapter 2. Background – Alzheimer’s Disease

Figure 2.8: (Left) Diffusion tensor image of a brain showing white matter fibre con-
nections. The colours represent the direction of the connection (red for left-right, blue
for superior-inferior, and green for anterior-posterior). (Middle) Zoomed image into the
small region of interest (ROI), showing the diffusion tensor ellipses. Each ellipse indicates
the direction where water molecules diffused (i.e. moved). (Right) Diagram showing the
difference between isotropic diffusion (i.e. equal in all directions) versus anisotropic dif-
fusion, along with the diffusivity measures that can be computed. Diagram assembled by
me using images from several sources4 .

of the neuron’s transport system and eventually the neuron’s death (see section
2.1.2.2). AV45 PET can be used to measure the level of misfolded tau proteins and
is also one of the earliest markers in AD.

PET-derived biomarkers are important because they give information about molecular
processes that happen in the brain. These are usually the first to become abnormal in
the cascade of events that lead to Alzheimer’s disease, and are therefore important early
markers of the disease that is about to unfold [1, 119].
PET scans have some limitations that need to be acknowledged. One main limitations
is that the patient is exposed to ionising radiation, which limits the number of scans they
can take in a specific time interval. PET scans also have a much lower spatial resolution
compared to MRI scans. One other caveat with AV1451 PET (tau imaging) is that it is a
very new tracer that is still under research, with some studies indicating evidence of some
off-target binding in some tau conformations found in non-AD tauopathies [121, 122].

2.1.4.4 Diffusion Tensor Imaging


Diffusion tensor imaging (DTI) is an MRI technique that can be used to measure the
degeneration of white matter connections in the brain. This is done by analysing the
diffusion of water molecules along the neuron fibre connections. Molecular diffusion in
tissues is not free, but reflects interactions with many obstacles, such as macromolecules,
fibers, and membranes. When a fiber connection degrades, the diffusion becomes more
isotropic (i.e. equal in every direction), which can be quantified using a measure called
fractional anisotropy (FA). Fig. 2.8 shows a diagram of a DTI image (left) which is

Image sources:
4 http://fmri.uib.no/index.php?option=com_content&view=article&id=68&Itemid=86
https://commons.wikimedia.org/wiki/File:DTI-axial-ellipsoids.jpg
http://www.diffusion-imaging.com/2012/10/voxel-based-versus-track-based.html
2.1. Alzheimer’s Disease 37

Figure 2.9: Diagram showing the cerebro-spinal fluid (CSF) coloured in blue, which is
found in the subarachnoid space around the brain and spinal cord. Source: Wikipedia5 ,
CC license.

made of diffusion tensors estimated at each voxel (middle). Diffusivities parallel and
perpendicular to the fiber direction can then be measured (right).
DTI is important for analysing the progression of Alzheimer’s disease. It has been
shown that AD affects white matter bundles [123]. DTI has also shown great potential
for aiding the diagnosis of dementia [124, 125]. DTI tractography is also important for
building brain structural connectomes which have been shown to be disrupted by different
types of dementias including Alzheimer’s disease [38, 126].
DTI measures have some limitations. As with other MRI modalities, it is susceptible
to motion artefacts and suffers from partial volume effects, i.e. measures at each voxel are
biased due to averaging across many different cells and types of tissue that are contained
in that voxel. Another limitation is that changes in DTI-derived measures such as FA are
not specific, and can be attributed to many changes in the underlying cytoarchitecture,
such as neurite density or dispersion [127].

2.1.4.5 Cerebrospinal Fluid Markers


The cerebrospinal fluid (CSF) is a clear, colourless body fluid found in the brain and
spinal cord. It acts as a cushion or buffer for the brain, providing basic mechanical and
immunological protection to the brain inside the skull. A sample of the CSF can be taken
from patients invasively through lumbar puncture, which involves inserting a needle in
the spinal cord.
Measures of CSF are very important for dementia research. In the CSF, the con-
centration of abnormal proteins such as amyloid-beta and tau is a strong indicator of
AD. Abnormal levels of concentrations in these proteins are some of the earliest signs of
Alzheimer’s disease and can indicate abnormalities many years before symptom onset [1].
The CSF measures have some limitations. One key limitation is that the lumbar
puncture is highly invasive and thus not performed in many studies. The CSF measures
are also not specific to any particular part of the brain.

5
https://en.wikipedia.org/wiki/File:1317_CFS_Circulation.jpg
38 Chapter 2. Background – Alzheimer’s Disease

2.1.5 Diagnosis
A diagnosis of Alzheimer’s disease is usually given based on the person’s medical history,
behaviour and information provided by the relatives. Medical imaging from Magnetic
Resonance Imaging (MRI), Computer Tomography (CT) or Positron Emission Tomog-
raphy (PET) can help exclude other types of brain pathologies or types of dementia.
Memory tests from neuropsychological batteries can help characterise the stage of the
disease [128].
The most commonly used diagnostic criteria are from the National Institute of Neuro-
logical and Communicative Disorders and Stroke (NINCDS) and the Alzheimer’s Disease
and Related Disorders Association (ADRDA) [115, 116]. This criteria, commonly called
NINCSD-ADRDA, require evidence of cognitive impairment through neuropsychological
testing for establishing a clinical diagnosis of probable AD, while histopathologic confir-
mation is required for definite confirmation [115, 116].

2.2 Progression of Alzheimer’s Disease


Several studies have been done so far on Alzheimer’s disease progression [1, 129, 130,
131, 132, 133]. It is currently believed that abnormal changes in amyloid-beta and tau
aggregation happen very early, long before symptoms occur, followed by hypometabolism
and structural atrophy, and then cognitive decline such as memory loss and executive
dysfunction [1, 134]. Neuropathological staging of AD brains showed that the earliest
change in brain structure are in the medial temporal lobe, particularly in the entorhinal
cortex and hippocampus [132]. This has also been confirmed with in-vivo MRI studies,
which showed that even in mild AD the entorhinal area and hippocampus shrink by 20-
25% compared to controls [135, 136, 137, 138]. Results by Schott et al. [133] and Ridha
et al. [129] show that atrophy of the medial temporal lobe precedes the clinical onset of
AD by approximately 3.5 years [133].

2.2.1 Braak Staging


In 1991, Braak and Braak [132] proposed a staging system based on the spatial spread
of amyloid plaques, neurofibrillary tangles (NFT) and neutropil threads (NT). It was the
first attempt to build a staging scale for Alzheimer’s disease from neuropathology, using
a cross-sectional set of brains and without using clinical information.
The amyloid patterns of the Braak staging system proved to be of limited significance
for the differentiation of neuropathological stages [132], but still enabled separation into
three stages. In the initial stage, amyloid deposits are found in the basal portions of the
isocortex, followed by fast spreading in virtually all isocortical association areas by the
middle stage. In the late stage, amyloid plaques are found in all areas of the isocortex,
including sensory and motor fields [132].
The spreading of neurofibrillary tangles allowed separation into six key stages. Stages
I-II were characterised by mild or moderate accumulation in the transentorhinal layer
Pre-α6 [132]. Afterwards, stages III-IV were marked by the spread of NFTs into the
transentorhinal region and proper entorhinal cortex, along with mild involvement of the

6
Pre-α is one of the layers from the principal stratum (Pre) of the entorhinal cortex. It is characterised
by cellular islands of large projection cells, and it’s connections project to the hippocampus.
2.3. Posterior Cortical Atrophy 39

first Ammon’s horn sector [132]. Finally, stages V-VI were marked by the spread of NFTs
and NTs to almost all isocortical association areas. [132]

2.2.2 Neuroimaging
In AD, Magnetic Resonance Imaging (MRI) shows gray matter atrophy throughout the
brain, in particular in the hippocampus and entorhinal cortex [9]. In terms of atrophy
progression, it starts in the medial temporal lobe and fusifom gyrus at least 3 years
before an AD diagnosis, and then spreads to the posterior temporal lobe, parietal lobe,
and finally to the frontal lobe. However, the sensorimotor cortex, visual cortex and the
cerebellum are relatively spared [9].
Imaging with Positron Emission Tomography (PET) shows reduced metabolism (FDG)
and increased uptake of amyloid (e.g. AV45) proteins [10]. In early stages of AD, hy-
pometabolism affects the parietotemporal association areas, the posterior cingulate gyrus
and the precuneus. In later stages, frontal cortices also become affected, while the stria-
tum, thalamus, primary sensorymotor cortices, visual cortices and the cerebellum seem
to be spared [10]. In terms of amyloid deposition through amyloid PET, early deposits
are found in the precuneus, orbitofrontal, inferior temporal and posterior cingulate, later
followed by the entire prefrontal cortex, lateral temporal and parietal lobes [10]. These
patterns have been validated using autopsy studies[139, 140].

2.3 Posterior Cortical Atrophy


Posterior cortical atrophy (PCA) is an early-onset neurodegenerative syndrome that af-
fects the posterior part of the brain, resulting in the disruption of the visual cortex. The
syndrome, also called Benson’s syndrome, was first reported by Benson et al. [17] in 1988
to describe five patients with fairly homogeneous but otherwise unclassified symptoms.

2.3.1 Symptoms
The most common symptoms include general visuospatial and visuoperceptual impair-
ments such as inability to read, blurred vision, light sensitivity, trouble navigating through
space and issues with depth perception [18, 19]. Additional symptoms also include apraxia
(disorder of movement planning), visual agnosia (object recognition deficit) and agraphia
(loss of writing ability) [17, 32]. These symptoms get worse as the disease progresses,
with patients becoming unable to recognise familiar people, objects, difficulty navigating
familiar places and drawing (see Fig. 2.10). Some studies [141, 142, 143] reported visual
hemineglect (difficulty seeing one half of the visual field) to be frequent in PCA patients,
especially if asymmetrical atrophy takes place in the occipital areas.
PCA patients report higher-order visual problems related to object and space per-
ception, compared to more basic visual impairments e.g. in colour and motion, although
impairments in higher order visual functions might be due to lower-level disruption. One
study [144] reported that all PCA subjects showed impairment in at least one low-level
visual process, and that this correlates with higher-order visuospatial and visuoperceptual
functions, but not with non-visual functions of the parietal lobe, including calculations
and spelling.
40 Chapter 2. Background – Alzheimer’s Disease

Figure 2.10: (A) Visual deficits as shown when a 62-year old PCA patient was asked
to copy the intersecting pentagons figure [18]. (B) Structural MRI, FDG PET and PiB
PET scans of the same subject. Structural MRI shows atrophy predominant in the
bilateral parietal, posterior temporal and lateral occipital regions (B, top), FDG PET
shows reduced metabolism in the same regions (B, middle), while PiB-PET shows diffuse
amyloid uptake throughout the entire brain (B, bottom) [18].
2.3. Posterior Cortical Atrophy 41

2.3.2 Causes
The causes of PCA are still unknown, due to the rarity of the disease, gradual onset of
symptoms and no fully accepted diagnostic criteria [19, 18]. The progressive neurode-
generation that characterises PCA is often attributed to Alzheimer’s disease pathology
(i.e. aggregation of amyloid plaques and tau tangles), but alternative causes including
dementia with Lewy bodies, corticobasal degeneration and prion disease have also been
identified [18]. One study reported the PCA syndrome in a 4-sibling family with prion
disease [145], suggesting that prion propagation mechanisms might be involved in PCA.
Genetic factors that underlie PCA are also not well understood [18, 19]. Empirical
findings suggest that there are no significant differences in the number of patients with
a positive family history of PCA and typical AD [18]. Some studies also report no
differences in Alipoprotein E (APOE) genotypes between PCA and typical AD [146, 147,
148, 149], although other studies reported differences in APOE 4 allele status, with
fewer PCA patients being 4-positive [150, 151]. These differences have been attributed
to differences in inclusion criteria of PCA with respect to typical AD [18].

2.3.3 Diagnosis
PCA patients face difficulties in diagnosis due to the young age at onset and the fact that
there are no fully accepted diagnostic criteria. Patients are sometimes misdiagnosed with
depression, anxiety or even malingering in early stages of the disease [18]. They are often
initially referred to opticians and ophthalmologists in the belief that ocular abnormalities
are causing their visual deficits, often leading to unnecessary medical procedures such as
cataract surgery. Neuroimaging modalities such as magnetic resonance imaging (MRI),
positron emission tomography (PET) or single photon emission computed tomography
(SPECT) can aid diagnosis of PCA [152].
There are no widely accepted diagnostic criteria, although two criteria have been pro-
posed so far by Mendez et al. [146] and Tang-Wai et al. [147]. These criteria suggested
presence of visual deficits in absence of other eye diseases, gradual progression, relative
preservation of anterograde memory, absence of stroke or tumour, and other neuropsy-
chological or imaging abnormalities that are related to parietal or occipital functions.
However, these criteria have some limitations. They are yet to be thoroughly val-
idated outside their centres, and need to be linked to underlying pathology, otherwise
inconsistencies between studies and centres will occur. Moreover, the current criteria
provide no guidance to the level of specificity required for a diagnosis of PCA [18]. It has
been suggested that PCA, when caused by underlying AD pathology, lies on a continuum
of phenotypical variation between AD and purely-visual PCA, with no clearly defined
diagnosis boundary [151, 149, 144].

2.3.4 Management
There is no known efficient treatment of PCA that will reverse or stop neurodegeneration
[19]. Patients with PCA are usually treated with the same medication as for AD, namely
cholinesterase inhibitors: tacrine, rivastigmine, galantamine and donepezil [19]. Crutch
et al. [18] suggest that antidepressant drugs might also be appropriate in patients with
low mood, and levodopa or carbidopa could aid individuals with Parkinsonism. However,
there are no studies analysing the efficiency of these drugs in PCA patients [19].
42 Chapter 2. Background – Alzheimer’s Disease

A few non-pharmacological therapies have also been attempted recently in some pa-
tients that included psycho-educative programs [153] or a combination of speech therapy,
occupational therapy and physiotherapy [154].

2.3.5 Neuroimaging
Several MRI studies in PCA have shown damage to posterior brain regions. Studies
by Hof et al. [155] and Tang-Wei et al. [147] show a greater concentration of senile
plaques and neurofibrillary tangles in the occipital and parietal lobes and at the occipito-
temporal junction. Cross-sectional studies using voxel-based morphometry have also
shown significant abnormalities in occipital and parietal lobes, followed by the temporal
lobe [144, 21]. When compared directly to typical AD subjects, PCA have shown greater
atrophy in the right parietal lobe and less in the left temporal and hippocampal regions
[20, 18]. Some DTI studies also seem to suggest white matter damage in posterior regions
[156, 157, 158]. See Fig. 2.10 for MRI scans of a PCA patient, showing the typical
posterior pattern of atrophy.
Non-MRI imaging studies in PCA have also shown similarly posterior abnormality
patterns. Functional imaging studies using single photon emission computer tomography
(SPECT) and FDG PET also show reduced function in occipital and parietal regions
[159, 160, 161, 162]. Amyloid pathology, as measured with PiB-PET, has been found
in occipital and parietal areas, as compared to typical AD subjects [163, 164, 165, 166],
although this finding was not confirmed in two other studies, which found more diffuse
amyloid uptake [148, 167].

2.3.6 Heterogeneity
Some studies [31, 13] have shown that there is considerable heterogeneity within PCA
itself, where three main PCA subgroups have been reported: primary visual (the striate
cortex, caudal), parietal (dorsal) and occipitotemporal (ventral) [31, 13].
Patients with primary visual subtype showed poor vision deficits, with later problems
with memory, attention and presence of visual hallucinations [13, 168]. Imaging showed
reduction in occipital lobe perfusion. In one of the studies, AD diagnosis was confirmed
post-mortem, upon pathological examination [13]. However, evidence for the existence
of this subgroup is very limited, with only two patients identified so far in different case
studies [168, 13], with another study having reported no ”pure” visual deficits within a
cohort of n=21 PCA subjects[144].
Patients with the parietal (dorsal) PCA subtype generally show initial visuospatial
symptoms, agraphia (inability to draw) and dyspraxia, but have preserved visual fields,
basic perceptual abilities, object recognition and reading and show biparietal and occipital
deficits, disrupting the dorsal or ”where” stream [31, 13].
Patients with the occipitotemporal (ventral) PCA subtype generally show symptoms
related to visual distortion, inability to recognise objects, general topography and written
words and show occipitotemporal pathology, disrupting the ventral or ”what” stream
[31, 13].
While all this evidence suggests that there is considerable heterogeneity within the
PCA syndrome, evidence is very limited to a few case studies, with some patients also
having no pathological confirmation of underlying AD pathology. Moreover, some [18,
2.3. Posterior Cortical Atrophy 43

144] have suggested that these subtypes should not be interpreted as distinct groups, but
rather as points on a continuum of phenotypical variation.
44 Chapter 2. Background – Alzheimer’s Disease
Chapter 3

Background – Disease Progression


Models

A disease progression model is a mathematical model that describes the evolution of


biomarkers in a neurodegenerative disease. Such quantitative models promise to en-
able early and precise diagnosis of dementia before symptoms appear, and will enable
stratification of subjects in AD clinical trials. This is important, because it is currently
believed that one of the reason why AD clinical trials failed is because treatments were
not administered early enough, and to the right patients [12]. Moreover, the advent
of large multimodal biomarker datasets containing neuropsychological, imaging, genetic
and molecular data can enable the development of specialised progression models that
accurately predict the evolution of subjects.
Quantitative biomarker signatures estimated through disease progression models have
several other key benefits, which are illustrated in Fig. 3.1. First of all, they enable disease
understanding as well as testing and validation of hypotheses regarding underlying disease
mechanisms. Secondly, they enable staging of patients along the progression axis (x-axis),
along with prognosis estimates, which can be useful in clinical settings. Third, they also
enable differential diagnosis by comparing the fit of the patient’s data to signatures of
different diseases.
In this chapter we review the disease progression models that have been developed
in the literature. We present the hypothetical model by Jack et al. [1] (section 3.1),
followed by early models of progression based on symptomatic groups (section 3.2) and
regression against a clinical marker (section 3.3). We then review data-driven models of
disease progression such as the event-based model 3.5.1, the differential equation model
(section 3.5.2), the disease progression score (section 3.5.3), self-modelling regression (sec-
tion 3.5.4), the manifold-based model (section 3.5.5), the voxelwise mixed effects model
(section 3.6.1) and the network diffusion model (section 3.7.1), as well as discriminative
models used normally in machine learning (section 3.8).
46 Chapter 3. Background – Disease Progression Models

Figure 3.1: Cartoon showing hypothetical biomarker signatures from two diseases, along
with a cross-sectional snapshot of data from a patient (left). For one patient, disease
staging implies finding the optimal time-shift along the horizontal axis that would match
its data. On the negative y-axis, the histogram of possible stages is shown. Differential
diagnosis can performed by evaluating the integral of the distribution of stages on the
negative y-axis, and selecting the disease that has the largest integral. Deriving quan-
titative biomarker signatures using disease progression modelling can help with disease
understanding, staging and differential diagnosis. Image courtesy of Neil Oxtoby and
Daniel Alexander.

3.1 Hypothetical Models


A hypothetical model of disease progression has been proposed by Jack et al. [1, 134],
which describes the trajectory of several key biomarkers during the progression of Alzheimer’s
disease (fig. 3.2). Using aggregated evidence from past literature, the model suggests that
amyloid-β and tau protein biomarkers become abnormal long before the onset of any
dementia symptoms. Afterwards, during the mild cognitive impairment (MCI) phase,
cognitive functions such as memory become abnormal along with brain structure mea-
sured using MRI. These biomarkers continue to be affected in the dementia stage, while
Aβ and tau seem to reach a plateau at this point. This hypothesised model of disease
progression is shaping the current field of AD research [3].
Apart from placing the biomarkers on a single time frame and suggesting the order in
which they become abnormal, the model also made some key observations. First of all,
the sequence of abnormality was not assumed to change then stop, rather some biomark-
ers gradually became abnormal simultaneously, although at different speeds and in an
ordered manner. Secondly, amyloid plaques are necessary but not sufficient to develop
AD pathology [119], with cognitive decline correlating less with amyloid deposition [169]
compared to tau and neurodegenerative markers [170]. Third, the authors suggested
biomarker trajectories follow non-linear curves, hypothesised to be similar in shape to
sigmoids [119, 129, 171]. Fourth, a time-lag exists between evidence of amyloid pathol-
ogy and the appearance of cognitive deficits, probably mediated by brain resilience and
cognitive reserve [119].
The hypothetical model nevertheless has some limitations. First of all, it is a hypothet-
3.2. Models of Progression using Symptomatic Groups 47

Figure 3.2: Dynamic biomarkers of the AD cascade as hypothesised by Jack et al. [1]. Aβ
and tau are thought to become abnormal before the onset of any dementia symptoms,
while brain structure, memory and clinical function are thought to become abnormal
later, during MCI and dementia stages. Reproduced with permission from [1].

ical, theoretical model that is meant to be a guide for future researchers modelling disease
progression in Alzheimer’s disease. Hence, the model is not quantitative and cannot be
used to e.g. stage patients. Another limitation is that the x-axis (disease progression)
and y-axis (biomarker abnormality) are not well-defined. Various implementations that
will be discussed next have made various assumptions about how to define this, such
as computing Z-scores with respect to controls [2], or used percentiles over the observed
biomarker values [3]. In the next sections, we will present the development of quantitative
models that address these limitations.

3.2 Models of Progression using Symptomatic


Groups
Some of the simplest disease progression model are based on symptomatic staging of
patients into a small number of groups, e.g. ”pre-symptomatic”, ”mild”, ”moderate” and
”severe” [23]. They then describe the differences in biomarker measurements among these
groups. Scahill et al. [131] devised such a method that finds changes in brain structure
using voxel-based analysis of serial nonlinear-registered MRI images. Other models based
on symptomatic staging are those of Dickerson et al., 2009 [172] and Thomson et al., 2001
[173] and 2003 [174].
These models have several key limitations. First of all, they rely on clinical assess-
ment, which is usually subjective and biased. Secondly, they offer very limited temporal
resolution and cannot model changes in pre-symptomatic phases of the disease.

3.3 Regression Against One Biomarker


In order to estimate longitudinal biomarker trajectories, some authors have proposed
regressing against a clinical or age-related marker. Sabuncu et al. [175] regressed the
48 Chapter 3. Background – Disease Progression Models

cortical thinning rate against MMSE scores. Jack et al. [176] also used regression against
MMSE to estimate the shape of biomarker trajectories. Doody et al. [177] regressed
biomarkers against time since baseline visit. Driscoll et al. [178] estimated brain volume
trajectories using a mixed effects model against age, using other demographic variables
such as gender and intracranial volume (ICV) as covariates.
These methods have some limitations. Regression methods against clinical markers
are limited by the fact that they cannot estimate biomarker dynamics in pre-clinical
stages. On the other hand, regression against age or time since baseline visit assume that
all subjects have the same age of disease onset or that disease onset is at baseline visit.
Another method for estimating biomarker trajectories, which is popular in familial
AD, performs non-linear regression of mutation carriers’ data against estimated years
from parent’s onset [179, 180]. However, this method can only be applied to dominantly
inherited AD, which represents only a small percentage of the entire AD population.

3.4 Survival Analysis Models


Survival analysis models are a class of models used to predict time until an event, in this
case conversion to mild cognitive impairment (MCI) or AD. One popular type of survival
model in AD is the Cox proportional hazards model, which assumes a multiplicative
increase in the hazard rate with respect to a unit-increase in the covariate. Cox propor-
tional hazards models have been used to estimate the probability of progression to AD in
a variety of studies [24, 181, 182, 183, 184, 185]. A related model is the proportional odds
model, which is more suitable for discrete data, and has also been applied to evaluate
risk of developing in AD [186, 169, 187].
Non-parametric survival models such as the Kaplan-Meyer estimator have also been
used [188, 189, 190, 191]. However, the Kaplan-Meyer method can only use a single,
binary predictor variable as opposed to the Cox regression method.
The main limitation of survival models is that they require accurate and reliable
diagnostic classes, which are not always available and can sometimes be inaccurate due
to human errors.

3.5 Scalar Biomarker Models


Over the last few years, a range of latent-time models of disease progression have also
been proposed, which estimate scalar biomarker trajectories without relying on a-priori
defined cognitive groups, diagnosis or clinical markers. Here, by latent-time models we
mean models that estimate a latent temporal dimension of disease progression in an
unsupervised manner. In this section we will present several such models: the event-
based model (section 3.5.1), the differential equation model (section 3.5.2), the disease
progression score model (section 3.5.3), the self-modelling regression model (section 3.5.4)
and the manifold-based mixed effects model (section 3.5.5). These models use scalar
biomarker values which are assumed to be uncorrelated, as opposed to more complex
spatial data such as brain images or cortical shapes.
The main premise behind there models is that they assume measurements are taken
from subjects who are at various unknown points along the progression of the same
disease. The models attempt to estimate simultaneously differences in the dynamics of
the disease progression while also estimating the time shift and progression speed along
3.5. Scalar Biomarker Models 49

Region 1 Region 2

Patient 1 Patient 2 Patient 3


Region 1 1.1 0.9 0.1
Region 2 0.95 0.0 0.05

Patient 1 Patient 2 Patient 3


Region 1 normal normal abnormal
Region 2 normal abnormal abnormal

Estimated Sequence: Region 2 → Region 1

Figure 3.3: Diagram showing the key concepts behind the event-based model. We assume
a toy dataset (top-left) of two region-of-interest biomarkers from three patients, which
are at different stages along a hypothetical disease progression timeline (bottom-left).
The aim is to estimate which region became abnormal earlier in the disease process. The
event-based model solves this by fitting a mixture model to the data (top-right), where
the two distributions are assumed to represent normal and abnormal biomarker values
respectively. The measurements from each patient are then assessed according to each
distribution (middle-right). Finally, the sequence of abnormality is estimated from these
values, by placing earlier in the sequence the regions/biomarkers for which there are more
abnormal values in the dataset. Diagram made by me.

the disease timeline, also called temporal heterogeneity. Some models go a step further
and also estimate differences that are due to spatial heterogeneity of the subjects, using
random effects estimating deviations from the population trajectory. Such combined
modelling is challenging, as it introduces identifiability issues.

3.5.1 The Event-Based Model


The event-based model was introduced by Fonteijn et al. [23] in 2012 and describes the
disease as a sequence of discrete events. A key diagram describing the EBM is given
in Fig. 3.3. Given a small dataset of biomarker measurements from subjects who are
assumed to lie at unknown shifts along the disease progression timeline (X-axis), the
EBM aims to estimate the order in which brain regions, or more generally any biomarker
measurements, become abnormal as the disease progresses. The disease is modelled as
a sequence of events, where each event represents a change in the patient state, such as
the onset of a new symptom (e.g. ’patient shows a drop in cognitive performance’) or
measurement of tissue pathology (e.g. ’lumbar puncture shows reduced amyloid beta’).
In section 3.5.1.1 we present the theory behind the EBM, in section 3.5.1.2 we present
the methods that are used to estimate the abnormality sequence, in section 3.5.1.3 we
show how to fit the mixture model parameters and in section 3.5.1.4 we show how to
stage the subjects using the EBM.
50 Chapter 3. Background – Disease Progression Models

3.5.1.1 Theory
The event-based model consists of a series of events E1 , E2 , . . . , EN and an ordering
S = [s(1), . . . , s(N )] which is a permutation of the integers 1, . . . , N creating the event
ordering Es(1) , Es(2) , . . . , Es(N ) . The set of events is specified a-priori. Moreover, the
model uses a dataset X which contains a set of Xi measurements for each subject i.
These measurements Xi are defined as Xi = {xi1 , xi2 , . . . , xiN }, where xij represents the
value of biomarker j in subject i and is informative of event Ej sin subject i.
The event-based model makes two key assumptions: first, measurements are mono-
tonic as the disease progresses and secondly, the event ordering is the same across all
patients. The first assumption fits with the hypothetical model presented by Jack et al.
[1] in fig. 3.2. Therefore, a patient for whom event Ej has occurred cannot revert to
a state where event Ej did not occur. This assumption is essential because it ensures
snapshots are informative about the event ordering [23]. The second assumption is nec-
essary to be able to aggregate information about the event ordering from the entire set
of subjects.
The aim of the event-based model is to find the probability density function p(S|X)
of an event ordering given the biomarker data. One starts by fitting a model for the
likelihood function p(xij |Ej ) the likelihood of measuring xij given event Ei occurred. A
similar fit is obtained for p(xij |¬Ej ), the likelihood of measuring xij given event Ej has
not occurred. More information about mixture model fitting can be found in section
3.5.1.3. If a subject i is at stage k in the disease progression, events Es(1) , . . . , Es(k) have
occurred while events Es(k+1) , . . . , Es(N ) have not occurred. We can therefore define the
likelihood of the data from subject i given ordering S as:
k
Y N
 Y 
p(Xi |S, k) = p xi,s(j) |Es(j) p xi,s(j) |¬Es(j) (3.1)
j=1 j=k+1

where measurements xij are assumed to be independent. Since the subject could
potentially be at any stage k in the progression, we integrate over k:
N
X
p(Xi |S) = p(k)p(Xi |S, k) (3.2)
k=0

where p(k) is the prior probability of the subject being at position k in the sequence. A
uniform prior is usually assumed here. Further assuming independence of measurements
across patients we get:
P
Y
p(X|S) = p(Xi |S) (3.3)
i=1

Combining equations 3.1,3.2, 3.3 we get the total likelihood:


P
" N k N
!#
Y X Y  Y 
p(X|S) = p(k) p xi,s(j) |Es(j) p xi,s(j) |¬Es(j) (3.4)
i=1 k=0 i=1 i=k+1

3.5.1.2 Event Sequence Estimation


Applying Bayes’ theorem we can get the posterior on the sequence:
3.5. Scalar Biomarker Models 51

source target source target

S old E1 E2 E3 E4 E5 S old E1 E2 E3 E4 E5

S new E4 E2 E3 E1 E5 S new E2 E3 E4 E1 E5
(a) Fonteijn et al. [23] (b) Young et al. [24]

Figure 3.4: MCMC perturbation rules used by (a) Fonteijn et al. [23] and (b) Young et
al. [24]. Both methods assume randomly selected source and target events. The method
by Fonteijn et al. only swaps the source event (E1) with the target event (E4). On the
other hand, the perturbation used by Young et al. moves a source event after a target
event and slides the other biomarkers accordingly. Diagram made by me.

p(S)p(X|S)
p(S|X) = (3.5)
p(X)
As the marginal distribution p(X) is analytically intractable, one uses a Markov-
chain Monte Carlo (MCMC) algorithm to sample from the posterior distribution p(S|X).
One assumes flat priors on the sequence S as any sequence could be equally likely. In
the MCMC phase, at each iteration the sequence S can be perturbed by swapping two
randomly chosen events. This perturbation rule has been used by Fonteijn et al. [23].
However, another perturbation method used by Young et al. [24] randomly selects a
source and target event and places the source event after the target event, sliding the
other biomarkers accordingly (see Fig. 3.4). The resulting sequence S new is accepted with
probability p = min(1, a) where a = p(X|S new )/p(X|S). Otherwise the old sequence is
stored and the process is repeated. As MCMC depends on accurate initialisation, one also
runs a greedy ascent algorithm in order to find the sequence with the highest likelihood.
The greedy ascent is very similar to the MCMC phase, the only difference being that
a is set to 1 if p(X|S new ) > p(X|S) and to zero otherwise. Depending on the number
of biomarkers, the greedy ascent is run for a few thousand iterations and repeated 10
times, with different random permutations of integers 1, . . . , N as the starting position.
The maximum likelihood sequence obtained from greedy ascent is then used to initialise
MCMC sampling, which usually runs for at least 100,000 iterations, again depending on
problem size.
The resulting MCMC-sampled sequences are usually plotted in a positional variance
matrix M (Fig 3.5), which is a compact way to represent uncertainty in the event ordering.
Each element M (i, j) represents the proportion of times event Es(j) appeared on position
i in the sampled sequences, given some master sequence S. S is usually set to be the
maximum likelihood sequence or the characteristic ordering, which is given by the average
position of the events in the MCMC samples [23].

3.5.1.3 Mixture Models for Data Likelihood


In equation 3.1 we need to model the distributions p (xi,j |Ej ) and p (xi,j |¬Ej ) of abnormal
and normal biomarker values, using the measurements in X. Fonteijn et al. [23] used
a Gaussian distribution for p (xi,j |¬Ej ) and a uniform distribution for p (xi,j |Ej ). The
52 Chapter 3. Background – Disease Progression Models

MCMC samples

1 E2 E1 E4 E3

2 E1 E2 E4 E3
E2

E1

T E2 E1 E3 E4 E4

E3

1 2 3 4
Maximum Likelihood Ordering
Event Position

E2 E1 E4 E3

Figure 3.5: MCMC sampling and positional variance computation. MCMC sampling
finds a series of T samples, which are then used to derive the characteristic ordering,
where events are ordered according to their average position in the MCMC samples.
Entries M (i, j) in the positional variance matrix stores the relative number of times each
event appeared in each position in the sequence. The events in the positional variance
matrix are ordered according to the characteristic ordering. Diagram made by me.

parameters for the Gaussian distribution were set as the mean and standard deviation of
biomarker values corresponding to controls, while the limits of the uniform distribution
were set to be the minimum and maximum observed biomarker values. While this works
in familial AD and Huntington’s disease [23] due to well-defined control populations, this
does not work well in sporadic AD due to the control population being not well-defined
– e.g. some controls can already have abnormal amyloid levels, which could result in the
distribution for normal values encompassing all observed values. Therefore, the approach
of Young et al. [24] for sporadic AD involved optimising the mixture model parameters
based on the subjects’ data in a data-driven manner. In this case, prior constraints
were used on the mixture model parameters, i.e. the mean and standard deviation of
the gaussian distributions, for biomarkers that did not change from healthy to diseased
subjects.

3.5.1.4 Patient Staging and Diagnosis Prediction


After the maximum likelihood sequence has been found using the greedy ascent method
described in section 3.5.1.2, each subject can be assigned a disease stage k as follows:

k
Y N
 Y 
k = arg max p(k)p(Xi |S, k) = p(k) p xi,s(j) |Es(j) p xi,s(j) |¬Es(j) (3.6)
k i=1 i=k+1

As before, the prior p(k) is assumed to be uniform. It should be noted that stages
range from zero to N , the number of events. If a subject is at stage k it means that all
events up to and including k have occurred while the events after k have not occurred.
The event-based model can also be used to classify subjects into controls and AD,
or any other symptomatic subgroups [24]. Given a threshold stage t, one can predict all
subjects having a stage less than or equal to t to be controls and all subjects with stages
3.5. Scalar Biomarker Models 53

greater than t to be patients. The optimal threshold is the one which maximises the
balanced accuracy, defined as follows:
TP + TN
Accuracy = (3.7)
TP + FP + FN + TN
where T P , F P , F N , T N represent the number of true positive, false positive, false
negative and true negative subjects respectively.

3.5.1.5 Discussion
The EBM is a useful tool for modelling the progression of diseases when only limited,
cross-sectional data is available. The model can also be usd to stage subjects, in discreete
units, along the disease progression timeline. The model parameters are estimated using
Markov Chain Monte Carlo sampling, based on optimising a conditional likelihood.

3.5.1.6 Advantages and Limitations


The event-based model by Fonteijn et al. [23] has several advantages. It is a data-driven
progression model which does not use a-priori defined clinical stages, which can often
be unreliable and can limit the temporal resolution of the model. Moreover, it does not
require longitudinal data, which makes it very useful for analysing rare types of dementia
for which comprehensive longitudinal datasets do not exist. The Bayesian framework in
which it is formulated also allows it to estimate uncertainty in the abnormality sequence.
The model can also easily combine data from different modalities.
The current model has several limitations. The trajectory parameters are modelled as
step functions, which is a strong assumption given the continuous nature many biomark-
ers used in AD. Secondly, in the fitting process the conditional probability of the sequence
given a-priori estimated distribution parameters is optimised, instead of the joint distribu-
tion over the sequence and the distribution parameters, which can result in a suboptimal
solution. Third, the model cannot use longitudinal biomarker measurements in order to
enable more precise staging and prognosis estimates. Finally, the model also assumes
that all subjects follow the same progression sequence, which is not the case in heteroge-
neous datasets due to differences in the underlying pathology, genetics and environmental
factors. Identifiability can also be an issue, mostly when, for a certain biomarker, the
distributions of normal and abnormal values overlap – this can result in the biomarker’s
event being placed either towards the beginning or the end of the sequence, even if the
true position of that event is in the middle of the sequence.

3.5.2 Differential Equation Model


The differential equation model (DEM) [192, 193, 175, 25, 30] constructs the biomarker
trajectories from the change in biomarker values between different visits (Fig 3.6). In
many medical settings we only have short-term longitudinal data, hence the biomarker
scores s are observed for each subject over a few visits. By determining how these
scores change (∆s) over time (t) during a specified time interval ∆t, the temporal rate
of progression (∆s/∆t) can be modelled as a function of the mean biomarker value f (s)
[192]:
∆s
≈ f (s) (3.8)
∆t
54 Chapter 3. Background – Disease Progression Models

Forward Model
==========⇒
What we want What we have

∆x δx
lim∆t→
−0 = = f (x)
∆t δt
Solve for x using the
Euler method:

t1 = t0 + δt
x1 = x0 + f (x0 )δt

⇐==========
Inverse Problem
Figure 3.6: Diagram of the Differential Equation Model (DEM). (top-left) Hypothetical
biomarker signature that needs to be reconstructed, along with subject measurements.
(top-middle) To make the model more realistic, each subject is made to follow a slightly
different trajectory due to heterogeneity. (top-right) In practice, we don’t know the
disease stage, so we align the measurements at time since baseline visit. (bottom-right)
The DEM model estimates a rate of change model from the slopes of lines fitted to each
subject’s biomarker data. At least two measurements per subject are required in order
to estimate this slope. (top-middle) The DEM then performs a line integral using the
Euler method to recover the biomarker trajectory (top-right). Diagram made by me.

The model given by f (s) can be parametric (e.g. linear, polynomial) or non-parametric
such as Gaussian Processes (GP). We then perform a line integral along f (s) to recover
s(t). More explicitly, if we take the limit as ∆t →
− 0 from Eq. 3.8, we get that:
∆s δs
lim∆t→
−0 = = f (s) (3.9)
∆t δt
Solving this numerically is done using the Euler method. We set an initial (t0 , s0 ) and
small increment step δt and find the next pair (t1 , s1 ) as follows:

t1 = t0 + δt
s1 = s0 + f (s0 )δt (3.10)
This is repeated until the full curve defined by (t0 , s0 ), (t1 , s1 ), . . . , (tn , sn ) is recon-
structed. Since the DEM model is univariate, the process is repeated independently for
3.5. Scalar Biomarker Models 55

Figure 3.7: Biomarker trajectories estimated by the disease progression model by Jedynak
et al. [2]. Reproduced with permission from [2].

the other biomarkers.

3.5.2.1 Advantages and Limitations


The differential equation model has several advantages. It is a fully data-driven method
that does not require a-priori defined clinical categories. In contrast to the event-based
model, it can estimate non-parametric biomarker trajectories which make minimal as-
sumptions on the shape of the biomarker trajectories. Moreover, the DEM can use any
model to estimate the change in biomarker values. While [30] used Gaussian Processes
to estimate the change in values, others [25] used polynomial functions.
The model has several limitations. First of all, the DEM is univariate, so the biomarker
trajectories are fit independently. This requires alignment on the temporal axis after they
are recovered, and makes it susceptible to noise within that biomarker1 . Secondly, in this
formulation the model does not allow one to directly estimate uncertainty in the trajectory
values along the y-axis. One option for estimating uncertainty is to integrate posterior
samples of the differential model and then align them all on the temporal axis. However,
this does not result in true confidence intervals, since at the anchor point there will be
zero uncertainty.

3.5.3 The Disease Progression Score Model


The disease progression score (DPS) model was proposed by Jedynak et al. [2]. It is
based on three main assumptions:

1
A multivariate model would’ve been able to use information from other biomarkers to help estimate
such a noisy trajectory, hence are more robust in theory.
56 Chapter 3. Background – Disease Progression Models

• Subjects follow a common disease progression but they have a different age at onset
and progression speed.

• Each biomarker trajectory is a monotonic curve that follows a sigmoidal shape

• The speed of progression of each subject is the same across the entire disease time-
course.

Biomarker trajectories estimated by the model for typical AD progression are shown
in Fig. 3.7. The model estimates the optimal shape2 of the biomarker trajectories, while
estimating a disease progression score for each subject, which is the stage along the disease
time course. The disease progression score sij for subject i at visit j is defined as a linear
transformation of age tij :

sij = αi tij + βi (3.11)


where αi and βi represent the speed of progression and time shift (i.e. disease onset) of
subject i.
The DPS model assumes that biomarker measurements are independent and follow a
sigmoidal trajectory f (s) given the disease progression score s. The sigmoidal function
for biomarker k with parameters θk = [ak , bk , ck , dk ] is defined as:
ak
f (s; θk ) = + dk (3.12)
1 + exp(−bk (s − ck ))
where dk is the minimum value, dk + ak is the maximum value, ak bk /4 is the maximum
slope and ck is the inflexion point. Authors choose to model the biomarker trajectory as
parametric sigmoidal curves because they provide a better fit than linear models [175,
194], and can account for floor and ceiling effects. The value yijk of biomarker k from
subject i at visit j is a normally distributed random variable:

p(yijk |αi , βi , θk , σk ) = N (yijk |f (αi tij + βi ; θk ), σk ) (3.13)

We further define I to be the set of all triplets (i, j, k) for which measurements are
available. Assuming independence across all measurements, we get the following model
conditional likelihood:
Y
p(y|α, β, θ, σ) = p(yijk |αi , βi , θk , σk ) (3.14)
(i,j,k)∈I

where y = [yijk ] for (i, j, k) ∈ I. Vectors α = [α1 , . . . , αS ] and β = [β1 , . . . , βS ], where S


is the number of subjects, denote the stacked parameters for the subject shifts. Vectors
θ = [θ1 , . . . , θK ] and σ = [σ1 , . . . , σK ], with K being the number of biomarkers, represent
the stacked parameters for the sigmoidal trajectories and measurement noise specific to
each biomarker.
The parameters of the model are therefore Θ = [α, β, θ, σ] and the log-likelihood
function associated with it is:
X 1
l(α, β, θ, σ) = logσk + 2 (yijk − f (αi tij )) (3.15)
2σk
(i,j,k)∈I

2
within a parametric family, in this case sigmoidal family
3.5. Scalar Biomarker Models 57

3.5.3.1 Model Fitting


Model fitting is done by loopy belief propagation, which alternates between optimising
the sigmoidal parameters σ and the subject specific parameters α, β. Alg. 1 shows the
fitting procedure. In line 1, we initialise αi = 1, βi = 0 for every i. On lines 4 and 5 the
optimal parameters for every biomarker trajectory are optimised. On line 8, the subject
specific shifts and progression speeds are also optimised. On line 14, a transformation is
performed in order to make the model identifiable. For a similar reason, parameters αi
and βi are rescaled for every subject (line 18), so that the disease progression scores of
healthy controls have a mean µN of 0 and a standard deviation σN of 1.

1 Initialise α(0) , β (0) ;


2 for l = 1 to L do
3 for k = 1 to K do
(1) (0) (0)
θk = arg minθk (i,j)∈Ik (yijk − f (αi tij + βi |θk ))2
P
4
(1)2 1
P (0) (0) 2
5 σk = |Ik −2I−4| (i,j)∈Ik (yijk − f (αi tij + βi |θk ))
6 end
7 for i = 1 to I do
(1) (1) P 1 (1) 2
8 αi , βi = arg minαi ,βi (j,k)∈Ii (1) 2 (yijk − f (αi tij + βi |θk ))
σk
9 end
10 α(0) = α(1) , β (0) = β (1)
11 end
12 for k = 1 to K do
13 if bk < 0 then
(1) (1) (1) (1) (1) (1) (1)
14 ak = −ak , bk = −bk , dk = dk + ak
15 end
16 end
17 for i = 1 to I do
18 αi = σαNi , βi = βiσ−µ
N
N

19 end
Algorithm 1: The optimisation procedure for the disease progression score by [2].

3.5.3.2 Advantages and Limitations


The model by Jedynak et al. [2] has several advantages. As opposed to the differential
equation model, the model is multivariate and automatically aligns biomarker trajectories
on the same temporal axis. Furthermore, compared to the event-based model by [23], the
biomarker trajectories are modelled as continuous sigmoidal trajectories instead of step
functions. Moreover, each subject has an associated time shift and progression speed.
Parameter estimation is performed with loopy belief propagation which is very similar to
the Expectation-Maximisation framework [195], but using hard assignments of the latent
variables at each iteration.
The model has several limitations. The main limitation of this model is that the
trajectories are assumed to be sigmoidal, which is not necessarily the case for many
biomarkers such as cognitive tests, which show continuous decline even in late stages of
the disease. Furthermore, each subject is assumed to follow the same progression pattern,
58 Chapter 3. Background – Disease Progression Models

Figure 3.8: Biomarker trajectories estimated using the self-modelling regression approach
by [3]. Reproduced with permission from [3].

which is not true in many heterogeneous datasets such as ADNI. The DPS model can
also suffer identifiability issues when it attempts to stage very early-stage or late-stage
subjects, as in these time-windows the biomarker trajectories are mostly flat. This issue
can generally be addressed by setting priors on the time-shift and progression-speed of
the subjets.

3.5.4 The Self-Modelling Regression Model


Self-modelling regression (SEMOR) is a method that fits several curves under the as-
sumption of a common shape [3]. This approach has been used by Donohue et al. [3] to
estimate non-parametric biomarker trajectories with linear subject-specific effects (Fig
3.8). Compared to the model by Jedynak et al. [2], this model estimates non-parametric
biomarker trajectories as fixed effects and includes subject-specific random effects. No
subject-specific progression speed is modelled in the original formulation [3].
We assume that Yij is the measurement of biomarker j in subject i, gj is a continuously
differentiable monotone function, γi ∼ N (0, σγ2 ) is the time shift for subject i. The model
is defined as follows:
Yij (t) = gj (t + γi ) + α0ij + α1ij t + ij (t) (3.16)
where parameters α0ij , α1ij ∼ N (0, Σj ) model a linear perturbation of the non-parametric
trajectory gj for subject i and biomarker j, t is the time and ij (t) ∼ N (0, σj ) is the
measurement noise.

3.5.4.1 Parameter Fitting


Fitting the model is also done by loopy belief propagation – one iteratively estimates each
set of parameters (gj , γi , α) until convergence of the Residual Sum of Squares (RSS). The
3.5. Scalar Biomarker Models 59

algorithm makes use of the following residuals:


 g  g 
Rij (t) = Yij (t) − α0ij − α1ij t
 E Rij
 α
(t)|gj , t, γi = gj (t + γi )

α (3.17)
Rij (t) = Yij (t) − gj (t + γi ) E Rij (t)|α0ij , α1ij , t = α0ij + α1ij t
 γ  γ
Rij (t) = t − gj−1 (Yij (t)) (t)|γi ≈ gj−1 (gj (t + γi )) − t = γi
 
E Rij

Using the above residuals, the model is fit by initialising γi and iterating the following
steps[3]:

1. Given γi , estimate gj by setting α0ij = α1ij = 0 and iterating the following subrou-
tine:
g
(a) Estimate gj by a monotone curve fit on Rij (t)
α
(b) Estimate α0ij , α1ij using the linear mixed model of Rij (t). Repeat steps a and
b until convergence of each RSSj = it [Yij (t) − gj (t + γi ) − α0ij − α1ij t]2
P

2. Given the estimated gj , set α0ij = α1ij = ij (t) = 0 and estimate each γi with the
average of RijPover all j and t. Steps 1 and 2 are repeated until convergence of the
total RSS = ijt [Yij (t) − gj (t + γi ) − α0ij − α1ij t]2

3.5.4.2 Advantages and limitations


The SEMOR model by [3] has many advantages. It robustly estimates biomarker trajecto-
ries using non-parametric curves, in contrast with previous approaches [2, 23]. Moreover,
it also models unique trajectories for each subject as linear deviations from the average
trajectories using a mixed effects model. Each subject also has it’s own temporal shift
along the disease progression timeline. Parameter estimation is done with loopy belief
propagation, carefully alternating the optimisation of certain groups of parameters in
order to minimise the residual of the cost function.
The model has some limitations. In its basic formulation it does not model different
progression speeds for different individuals. Moreover, estimating all the subject-specific
parameters (α0ij , α1ij , γi ) requires suitable priors and at least two longitudinal measure-
ments per subject, which might not be available in some datasets. The high number
of parameters that need to be estimated, especially the subject-specific parameters, can
result in overfitting of the data, although this is mitigated to some extent by the priors
placed on these parameters. Moreover, it also assumes that biomarker measurements are
independent, which makes it unsuitable for modelling large sets of correlated biomarkers
such as voxelwise measurements. Identifiability can also be an issue with the SEMOR
model when the population trajectory becomes almost linear, such as in the case of
biomarkers that don’t show differences between healthy subejcts and patients.

3.5.5 The Manifold-based Mixed Effects Model


The manifold-based mixed effects model was introduced by Schiratti et al. in 2015 [27].
The model generalises a previous linear mixed effects model by [196] to account for time
shifts and describes it in a Riemannian manifold setting. Each subject i is assumed to
have a trajectory γi which is a deviation from the average population trajectory γ. The
deviation is modelled as a time shift τi and progression speed αi of subject i along the
disease timecourse.
60 Chapter 3. Background – Disease Progression Models

Let us assume we observe p individuals, each having ni observations obtained at times


ti,1 < · · · < ti,ni , each having yi,1 , . . . , yi,ni biomarker measurements. For a geodesic M and
a point p0 ∈ M at time t0 with velocity v0 ∈ Tp0 M , we define γp0 ,t0 ,v0 = Expt0 ,p0 (αi v0 )(.)
as the geodesic which passes through point p0 at time t0 with velocity v0 3 . We also
consider ti,j and yi,j as the age and biomarker measurement for subject i at visit j. This
gives us the following model:

yi,j = Expt0 +τi ,p0 (αi v0 )(ti , j) + i,j (3.18)

where

exp(ηi ), ηi ∼ pi=1 N (0, ση2 )
N
α i = N

τi ∼ pi=1 N (0, στ2 ) (3.19)

i,j ∼ i,j N (0, σ 2 )
 N

The model assumes that each ηi and τi are independent. The parameters of the model
are θ = [p0 , t0 , v0 , ση , στ , σ]. The model above can be re-written as:

yi,j = γp0 ,t0 ,v0 (αi v0 (ti,j − t0 − τi )) + i,j


= γi (ti,j ) + i,j (3.20)

where γi is the subject specific trajectory that is modelled as an affine reparametrisation


of the average trajectory γp0 ,t0 ,v0 . The model described here is univariate. However, an
extension of the model has been published by Schiratti et al. [197] which extends this
framework to a multivariate analysis.
Parameter estimation in the model by Schiratti et al. [27] is performed using maximum
likelihood estimation (MLE) using the Gauss-Hermite quadrature approximation, which
is equivalent to the Laplace approximation of the observed likelihood. Authors used the
Nelder-Mead method for estimating the numerical optimisation.

3.5.5.1 Advantages and Limitations

The model has several strengths and weaknesses. The main strength lies in the flexible
Riemannian manifold framework, that allows one to create different models depending
on how the inner product is defined. Moreover, the model estimates subject specific
trajectories γi , time shifts τi and progression speeds αi . However, one of the limitations
of the model is that it assumes a parametric form of the biomarker trajectories (i.e.
sigmoidal).

3.6 Spatiotemporal Disease Progression Models


Spatiotemporal models of disease progression have been proposed over the last few years.
In the following section, we will present two key spatiotemporal models of progression
that estimate voxelwise patterns of pathology, while also estimating latent subject-specific
time shifts.
3.6. Spatiotemporal Disease Progression Models 61

Figure 3.9: Diagram of the voxelwise disease progression model by Bilgel et al. [4]. The
model places biomarker measurements along a latent ”progression score” axis, and then
models the dynamics of these measurements using linear functions. Reproduced with
permission from [4].

3.6.1 The Voxelwise Disease Progression Model


A voxelwise disease progression model has been introduced in 2016 by Bilgel et al. [4].
This model allows the discovery of patterns of atrophy that are not confined to a given
region of interest (ROI). Since the input data is represented by voxel-wise measurements
such as amyloid burden, a spatial correlation function is used to model correlation between
voxel values. The model is built on the framework of the disease progression score by
Jedynak et al. [2].
A diagram of the model is given in Fig 3.9. The model aligns biomarker measurements
along a latent ”progression score” axis, and then models the dynamics of these measure-
ments using linear functions. Let us assume that tij represents the age for subject i at
visit j. The progression score sij for subject i at visit j is an affine transformation of age
tij :
(
sij = αi tij + βi = qTij ui
(3.21)
ui ∼ N2 (m, V (ν))
where αi , βi are the progression speed and time shift of subject i, qij = [tij , 1]T and
ui = [αi , βi ]T . The prior covariance matrix V is modelled as a 2 × 2 positive definite
matrix that has a Log-Cholesky parametrisation
 given  by ν. More precisely, if we consider
U U
a matrix U such that V = U U T and U = 11 12
, then ν = [logU11 , logU12 , logU22 ]
0 U22
Furthermore, let us denote by yij the K × 1 vector of biomarker measurements for
subject i at visit j. The longitudinal trajectories corresponding to these measurements
are modelled as follows:
(
yij = asij + b + ij
(3.22)
ij ∼ NK (0, R(λ, ρ))
3
In Riemannian geometry, Exp refers to the exponential map, which is a function from the tangent
space Tp M into M itself.
62 Chapter 3. Background – Disease Progression Models

where a = [a1 , . . . , aK ]T , b = [b1 , . . . , bK ]T are the coefficients of the linear model and ij
is the measurement noise that is independent and identically distributed across different
subjects and visits. The matrix R(λ, ρ) is the spatial covariance that is assumed to have
the form R = ΛCΛ, where Λ is a diagonal matrix with diagonal elements λ and C is a
correlation matrix that is parameterised by ρ [4]. This ensures that the matrix R(λ, ρ) is
positive definite. In order to model correlation among voxel measurements, the elements
Ckk0 of matrix C must be a function of the distance d ≡ d(k, k 0 ) between voxels k and k 0 .
Several such options exist:

• Exponential: Ckk0 = exp(−d/ρ)

• Gaussian: Ckk0 = exp(−(d/ρ)2 )


−1
• Exponential: Ckk0 = (1 + (d/ρ)2 )
  3 
3d 1 d
• Spherical: Ckk0 = 1 − 2 ρ + 2 ρ if d < ρ

The model parameters are therefore θ = [m, ν, a, b, λ, ρ]. The model is a mixed effects
model where a, b are the fixed effects and ui are the random effects.

3.6.1.1 Model Fitting


The model is fit using the Expectation-Maximisation (EM), described below. In line with
the standard EM framework [195], the algorithm optimises the expected value of the full
log-likelihood Ep(u|θ0 ) [l(y, u; θ)] given the current estimate of the latent variables u. The
complete data log-likelihood is:
X
l(y, u; θ) = l(yi , ui ; θ) =
i
1X 1X T
− log|2πR| − yij − Zij ui − b
2 i,j 2 i,j
1X 1X
− log|2πV | − (ui − m)T V −1 (ui − m) (3.23)
2 i 2 i

E-step
Let (y, u) be the complete data and θ 0 = [m0 , ν 0 , a0 , b0 , λ0 , ρ0 ] be the parameters es-
timated at the previous EM P iteration. Bilgel et al. [4] show that the E-step integral
0 0
Q(θ, θ ) is proportional to i Φ(ũi ; ûi , Σ0i )l(yi , ũi ; θ)dũi , where Φ is a multivariate nor-
R

mal probability density function with mean:

!−1 !
X X
û0i = Zij0T R0−1 Zij0 + V 0−1 Zij0T R0−1 (yij − b0 ) + V 0−1 m0 (3.24)
j j

P −1
and covariance matrix Σ0i = 0T 0−1 0
j Zij R Zij +V 0−1
. Evaluating the integral gives
the following final form:
3.6. Spatiotemporal Disease Progression Models 63

1X 1X  1X
Q(θ, θ 0 ) = − yij − Zij û0i − b − T r ZijT R−1 Zij Σ0i −

log|R| −
2 ij 2 ij 2 ij
1X 1X 0 T  1X
ûi − m V −1 û0i − m − T r V −1 Σ0i

log|V | − (3.25)
2 i 2 i 2 i

M-step
At the M-step we need to find θ = arg maxθ Q(θ, θ 0 ). The full derivations are given
in [4], yielding the following updates:
P  P  P 
0 0
P
( i νi ) ij yij sij − ij yij ij sij
a= P P  P 2 (3.26)
T 0 02 0
( i νi ) q Σ q
ij ij i ij + s ij − ij ijs
P  P  P  P 
T 0 02 0 0
ij yij ij qij Σi qij + sij − ij yij sij ij sij
b=  P T 0   P 0 2  (3.27)
02
P
( i νi ) q Σ q
ij ij i ij + s ij − s
ij ij

1X 0
m= û (3.28)
n i i

ν = arg max Q(θ, θ 0 ) (3.29)


ν

λ, ρ = arg max Q(θ, θ 0 ) (3.30)


λ,ρ

3.6.1.2 Advantages and Limitations


The model by Bilgel et al. [4] has several advantages. First of all, it is specifically tailored
for dealing with voxelwise measurements such as amyloid load by modelling the spatial
correlations. Secondly, like the disease progression score by Jedynak et al. [2], it estimates
subject specific temporal shifts and progression speeds.
The model has several limitations. First of all, the biomarker trajectories are assumed
to be linear, which is a strong assumption especially for early biomarkers such as amyloid,
which start to plateau when subjects are in the MCI stages. The linearity was required
in order to make the model inference with EM computationally tractable. Moreover, the
biomarker correlation structure based on a spatial distance function does not allow one
to recover fine-grained, disconnected patterns of pathology, as has been found for various
types of dementias due to disruption in underlying brain networks [38]. The model also
doesn’t account for inter-subject differences by estimating deviations from the common
population-wide trajectory.

3.6.2 Cortical Atrophy Progression Model


The cortical atrophy progression model was introduced by Koval et al. [198]. A diagram
of the model is given in Fig. 3.10. The model estimates vertexwise linear trajectories of
cortical thickness over the entire population, accounting for latent subject specific time-
shifts. The equation modelling a biomarker measurement yijk for subject i at visit j for
location k on the brain surface is given as:
64 Chapter 3. Background – Disease Progression Models

Figure 3.10: Diagram of the cortical atrophy progression model by Koval et al. [5]. (top)
The model estimates a unique, linear trajectory for the dynamics of cortical thickness
measurements at each point on the brain cortical surface. (bottom) Subject-specific
trajectories ηi and ηj are modelled by a shift of the population trajectory γ0 through
vectors wi and wj . Reproduced with permission from [5].

yijk = pk + wik + νk αi (tij − τi − t0 ) + ijk (3.31)


where pk and νk are parameters of the linear trajectory over the latent space specific to
location k, wik is a subject and location specific intercept, τi is the time-shift for subject
i, t0 is a time-shift reference and ijk is the Gaussian noise for the yijk measurement.
In order to account for spatial correlation, a set of control nodes V is defined, which
is a subset of all nodes V . Only for these control nodes will parameters pk and νk be
estimated. For the other nodes, the parameters will be an interpolation of the parameters
of the control nodes, weighted by the distance of that node to the control nodes.

3.6.3 Parameter Estimation


Parameter estimation is done using the Monte-Carlos Markov-Chain Stochastic Approx-
imation Expectation Maximisation (MCMC-SAEM) algorithm. This method essentially
approximates the intractable E-step using the MCMC sampler. The optimisation method
is proven to converge if the model belongs to the exponential family [5].

3.6.4 Advantages and Limitations


The model by Koval et al. has several advantages. It can estimate spatiotemporal
patterns of atrophy, and can be extended to other types of voxelwise biomarkers such as
amyloid load or DTI measures. The model also estimates subject-specific latent time-
shifts, accounting for different but unknown ages of disease onset in distinct subjects.
The model also has some limitations that need to be addressed in future work. One
limitation is that the authors need to define a-priori the number of control points, which
can affect the final smoothness level of the estimated patterns of pathology. While the
3.7. Mechanistic Models 65

authors only applied it to a brain surface made of 2,000 nodes, it is unclear whether the
model can scale to higher resolutions.

3.7 Mechanistic Models


3.7.1 The Network Diffusion Model

Figure 3.11: Diagram of the network diffusion model by Raj et al. [6]. The model uses
MRI and DTI data to extract a structural connectome from healthy subjects through
tractography, then computes a connectivity network. Each network is represented as a
graph where nodes represent brain ROIs where there is a certain concentration of toxic
pathogens and edges represent the connectivity strength. Using this matrix, the authors
estimate the eigenvectors of the graph, also called eigenmodes, which are then shown to
correlate with atrophy patterns in normal ageing, AD and bvFTD. More precisely, for
each disease they compute the amount of atrophy within each ROI corresponding to the
graph nodes, and then correlate with the eigenmodes. Reproduced with permission from
[6].

The network diffusion model was introduced by Raj et al. in 2012 [6] and later
extended in 2015 [199]. The model is inspired by evidence that Alzheimer’s disease
pathology spreads along vulnerable pathways in a prion-like manner rather than by spatial
proximity [200, 201, 202]. The model works by simulating the diffusion process of a
pathogenic protein along a structural connectivity graph from healthy controls. Atrophy
and other higher-level pathogenic processes are assumed to be a product of the lower-level
diffusion process. See Fig 3.11 for a diagram of the model.
66 Chapter 3. Background – Disease Progression Models

The diffusion process is modelled on a hypothetical brain network G = {V, E} whose


nodes vi ∈ V represent regions-of-interest and edges (i, j) having weight cij represent
fibre connections between regions i and j. Structures vi are parcellated regions-of-interest
obtained from an atlas, while the connection strength cij is measured using tractography
[203]. If we denote the disease factor in region i as xi , then the flow of the disease from
a region i to another region j in a short time interval δt is β(xi − xj )cij δt, where β is a
diffusivity constant controlling the diffusion speed. As δt → 0, this results in the following
first-order differential equation:
dxj
= βcij (xi − xj ) (3.32)
dt
Now let us denote by x(t) = {x(v, t), v ∈ V } the disease factor at time t in every node of
the network. Eq. 3.32 will then translate into the ”network heat equation” [204]:

dx(t)
= βHx(t) (3.33)
dt
where H is the Laplacian matrix of G defined as:
(P
j 0 6=i cij 0 for i = j
H(i, j) = (3.34)
−cij otherwise
We model the cortical atrophy in region k as the accumulation of the disease process:
Z t
φk (t) = xk (τ )dτ (3.35)
0

Extending this to the whole brain gives:


Z t
Φ(t) = x(τ )dτ (3.36)
0

The solution to the above equation is given by:


x(t) = exp(−βHt)x0 (3.37)
where x0 is the initial disease concentration, where the term exp(−βHt) acts as a
smoothing operator. Performing eigenvalue decomposition on H = U ΛU † , where U =
[u1 , . . . , uN ] is the matrix of eigenmodes, allows to express x(t) as:
N 
X 
x(t) = U exp(−Λβt)U † x0 = exp(−βλi t)u†i x0 ui (3.38)
i=1

The time evolution of atrophy can then be described as:


Z tXN  N

 X 1
Φ(t) = exp(−βλi t)ui x0 ui dt = (1 − exp(−βλi t)) u†i x0 ui (3.39)
0 i=1 i=1
βλi

Raj et al. [6] present evidence to suggest that the eigenmodes ui with the highest cor-
responding eigenvalues λi represent the areas that are normally affected by key neu-
rodegenerative processes or diseases, such as normal ageing, AD and behavioural variant
frontotemporal dementia (bvFTD) respectively. They suggest that these areas are selec-
tively vulnerable to these types of dementia, in line with previous theories in the field
[38, 205, 126].
3.8. Machine Learning Methods 67

3.7.1.1 Advantages and Limitations

The diffusion model by Raj et al. [6] has several advantages. In contrast with the models
presented above, it is able to model the propagation of atrophy along brain connectomes,
which can be used to test prion hypothesis or other related mechanisms. Secondly, this
approach allows one to test for other hypotheses of network-based pathology spread such
as nodal stress, transneuronal spread, trophic failure, and shared vulnerability [126].
The model has several limitations. The model assumes static networks, even though
the network dynamically evolves during the time course of the disease. The model also
assumes a parametric form of the biomarker trajectories, either exponential or sigmoidal.

3.8 Machine Learning Methods


Popular machine learning methods that are normally used for discriminative tasks can
also be extended to model disease progression by estimating continuous variables. One
such method is the Support Vector Machine, which is a non-probabilistic binary linear
classifier that was originally developed by Vladimir N. Vapnik and Alexey Ya. Cher-
vonenkis [206]. They can perform non-linear classification by mapping the input data
into higher-dimensional feature spaces using the kernel trick and finding the hyperplane
that optimally separates the data. They have been successfully used for differential diag-
nosis of different types of dementia using post-mortem confirmed subjects [207]. Other
uses of SVMs on medical images and other high-dimensional data have also been shown
[208, 209, 210, 211].
Other popular classifiers used in the machine learning field are random forests [212,
213]. These work by building a multitude of decision trees on training data, using them
to make predictions on the test data and performing majority voting on the predictions of
individual trees to select the final labels. Random forests have been demonstrated in med-
ical imaging for diagnosis classification [214], image quality transfer [215] or myocardium
segmentation [216].
Yet another machine learning model popular for time-series predictions is the Long
short-term memory (LSTM) network [217]. LSTMs are a type of recurrent artificial neural
network (RNN) that replace the conventional hidden nodes from RNNs with memory
cells, which ensure that the gradient cannot vanish or explode during training. LSTMs
have been applied in medical imaging research for diagnosing between healthy and AD
subjects [218], predicting the onset of diseases [219], predicting diagnoses in electronic
health records (EHR) [220], predicting mortality risk [221] as well as several other clinical
variables [222].

3.8.1 Advantages and Limitations


Discriminative models such as SVMs, random forests or LSTMs have several advantages.
They work for a wide variety of problems, datasets and underlying models and are gen-
erally robust to high-dimensional data. Some of them can also be resistant to overfitting,
especially when used in conjunction with regularisation and data augmentation tech-
niques. A key advantage of LSTMs in particular is that they are suitable for time-series
data, and can be naturally used to predict future biomarker values, although the other
models can also be extended to work with continuous data.
68 Chapter 3. Background – Disease Progression Models

These discriminative models also have several limitations. First of all, they generally
require labelled data, in the form of a-priori defined clinical categories or stages, which
are usually coarse, inaccurate and biased. These limit the temporal resolution of the
model. Moreover, it is also harder to interpret what these models learn from the data,
which limits their use for understanding the disease process. For some models there is
also a lack of mathematical proofs and guarantees regarding their convergence during
training, as well as behaviour while making predictions.

3.9 Summary
In Fig. 3.1 we show a summary of the main features of data-driven disease progres-
sion models, as well as discriminative models. For each model, we show the trajectory
shape, indicate whether models incorporated latent subject-specific time-shifts (in terms
of intercept or intercept + progression speed), subject-specific trajectories in the form of
random effects as well as spatial correlation. For each model, we also indicate the key
limitation.
We can observe several key differences between the models. In terms of time-shifts,
some models such as the DEM or the network diffusion model do not incorporate any
time-shifts, although these could be extended to incorporate such time-shifts. Other
models do not model subject-specific trajectories through random effects. Moreover, only
spatiotemporal or mechanistic models incorporate correlation between different biomarker
measurements.
In conclusion, over the last few years there have been several models of disease pro-
gression that were developed, starting from the early comparisons based on symptomatic
groups and moving on to more data-driven approaches and spatiotemporal models. Fur-
ther work will focus on developing more mechanistic models that enable understanding
of the underlying disease process, and can help guide drug development. One example
of this is the recent work of [223], which models the dynamics of pathogenic proteins
in a neural network and can help understand the effects of such pathogenic proteins in
neurodegeneration. However, validation of such models is required through in vitro and
in vivo studies.
In the following chapters, I will present the application of some of these models to
estimate the progression of Posterior Cortical Atrophy (chapter 4), as well as the devel-
opment of two novel models of disease progression (chapters 6 and 7).
3.9. Summary 69

Model Trajectory Subject Time-shifts Subject- Models Main


shape specific Biomarker Limitation
trajectory Correlation
intercept speed
Event-based step-function X** 7 7 7 discrete time
Model
Differential non- 7* 7* 7 7 univariate
Equation Model parametric
Disease sigmoid X X 7 7 sigmoidal
Progression Score trajectory
assumption
Self-Modelling non- X X X 7 can overfit +
Regression parametric identifiability
Manifold Model linear, sigmoid X X X 7 user-defined
trajectory
assumption
Voxelwise Model linear X X 7 X linear trajectory
assumption
Cortical Atrophy linear X X X X user-defined
Progression trajectory
Model assumption
Network diffusion exponential, 7* 7* 7 X assumes static
Model sigmoidal connectome +
no time-shift

Table 3.1: Comparison of features of various disease progression models. By subject-


specific trajectories, we exclude deviations from the population-wide trajectory only due
to time-shifts. While the models have many limitations, the ones mentioned here are the
main ones according to our own opinion. (*) Models cannot be used for subject staging in
their basic formulation, but extensions to the model can enable them to estimate subject-
specific disease onset and progression speed. Comparison analysis made by me. (**) The
subject-specific time-shift in the EBM is discreete and based on cross-sectional data only.
70 Chapter 3. Background – Disease Progression Models
Chapter 4

Longitudinal Neuroanatomical
Progression of Posterior Cortical
Atrophy

This chapter outlines the clinically applied part of my PhD, which focused on modelling
the progression of Posterior Cortical Atrophy using already developed methods. The con-
tent of this chapter is based on the neuroimaging results from the joint publication below,
where I’ve re-written most of the text for this thesis. I performed all the neuroimaging
work: image pre-processing, statistical analysis with EBM and DEM, and the interpreta-
tion of the results. The data from table 4.1 was gathered by Nicholas Firth. Splitting of
PCA patients into cognitively-defined subgroups was done by Silvia Primativo. Details
in section 4.3.1 regarding patient recruitment, patient numbers, clinical diagnosis and
pathological confirmation along with image acquisition details from section 4.3.2 were
taken from our joint publication.

4.1 Publications
• N. C. Firth*, S. Primativo*, R. V. Marinescu*, T. J. Shakespeare, A. Suarez-
Gonzalez, M. Lehmann, A. Carton, D. Ocal, I. Pavisic, R. W. Paterson, C. F.
Slattery, A. J. M. Foulkes, B. H. Ridha, E. Gil-Nciga, N. P. Oxtoby, A. L. Young, M.
Modat, M. J. Cardoso, S. Ourselin, N. S. Ryan, B. L. Miller, G. D. Rabinovici, E. K.
Warrington, M. N. Rossor, N. C. Fox, J. D. Warren, D. C. Alexander, J. M. Schott,
K. X. X. Yongˆ and S. J. Crutchˆ, Longitudinal neuroanatomical and cognitive
progression of posterior cortical atrophy, Brain, 2019. (*) joint first authors (ˆ)
joint senior authors
In the above manuscript, I preprocessed all the imaging data, performed the mod-
elling and statistical analysis of all the imaging data, and created the figures, tables
and diagrams (including statistical tests in the supplementary). I also drafted the
section of the results which was related to the imaging results. Other authors re-
cruited patients, collected the data, performed the analysis of cognitive tests, and
helped draft the initial version of the manuscript.

• R. V. Marinescu, A. L. Young, Neil P. Oxtoby, N. C. Firth, M. Lorenzi, A. Eshaghi,


V. Wottschel, M. J. Cardoso, M. Modat, K. X. X. Yong, S. Primativo, N. C. Fox,
72 Chapter 4. Longitudinal Neuroanatomical Progression of PCA

M. Lehmann, T. J. Shakespeare, S. J. Crutch, D. C. Alexander, A data-driven


comparison of the progression of brain atrophy in Posterior Cortical Atrophy and
Alzheimer’s disease, AAIC poster, 2016.

• R. V. Marinescu, S. Primativo, A. L. Young, N. P. Oxtoby, N. C. Firth, A. Eshaghi,


S. Garbarino, J. M. Cardoso, K. Yong, N. C. Fox, M. Lehmann, T. J. Shakespeare,
S. J. Crutch, D. C. Alexander, Analysis of the heterogeneity of Posterior Cortical
Atrophy: Data-driven Model Predicts Distinct Atrophy Patterns for three different
Cognitive Subgroups, AAIC poster, 2017

4.2 Introduction
Posterior Cortical Atrophy (PCA), already described in section 2.3, is a progressive neu-
rodegenerative syndrome causing predominantly visuospatial and visuoperceptual impair-
ments [18]. In order to understand complex disease mechanisms underlying PCA, and
design efficient clinical trials for finding treatments of PCA, we need to be able to accu-
rately estimate the temporal progression of atrophy in PCA and contrast it with typical
AD (tAD). Previous neuroimaging studies of PCA have shown more atrophy in the supe-
rior parietal, occipital and posterior temporal regions as compared to typical AD [20, 21].
However, these studies are all cross-sectional and cannot map the continuous longitudi-
nal progression of the disease. One longitudinal study of PCA [37] showed widespread
gray matter loss in both PCA and tAD, but the numbers were small (17 PCA and 16
tAD) and the time interval was short (1 year). Larger longitudinal studies are therefore
required to robustly estimate longitudinal progression patterns of PCA as compared to
tAD. Moreover, a second aspect that needs to be clarified is the heterogeneity within
PCA itself. Some studies have so far reported three dominant subgroups: primary visual
(the striate cortex, caudal), parietal (dorsal) and occipito-temporal (ventral) [31, 13].
However, evidence for the existence of these groups is mainly limited to individual case
reports [31, 13] and no previous study looked at the temporal progression of brain atrophy
in such subgroups.
The aim of this study is to estimate the progression of MRI brain volumes in PCA
as compared to tAD. We used the event-based model (EBM, section 3.5.1) and the
differential equation model (DEM, section 3.5.2) to estimate the progression of brain
volumes in 361 individuals (117 PCA, 106 tAD and 138 controls) from three centres in
the UK, Spain and US. We also use the event-based model to estimate the progression of
atrophy in three cognitively-defined PCA subgroups. Compared to previous studies, our
study is the first comprehensive study of atrophy progression in PCA. We also provide
the first glimpse into the early progression of atrophy within PCA subgroups.

4.3 Methods
4.3.1 Participants
117 patients with PCA were recruited from three specialist centres: 100 from the Demen-
tia Research Centre (DRC) UK, 9 patients from the University Hospital Virgen del Rocio
(HUVR) Memory disorders Unit, Spain and 8 patients from the University of California
San Francisco (UCSF) Memory and Aging Center, USA. All PCA participants met two
4.3. Methods 73

Imaging Neuropsychology
Visits Number Age Visit Interval Number Age Visit Interval

PCA (n=117)
All 89 63.52 ± 6.91 N/A 109 64.49 ± 7.54 N/A
2 46 62.11 ± 6.52 1.03 ± 0.47 70 63.64 ± 7.32 1.18 ± 0.48
3 31 62.75 ± 6.5 0.99 ± 0.47 45 62.73 ± 7.26 1.15 ± 0.45
4 15 61.46 ± 4.44 0.86 ± 0.31 20 63.19 ± 7.00 1.14 ± 0.40
5 9 61.73 ± 4.06 0.81 ± 0.33 7 59.44 ± 4.84 1.06 ± 0.45
6 2 62.35 ± 1.65 0.83 ± 0.24 2 57.22 ± 3.49 1.02 ± 0.35

tAD (n=106)
All 66 66.39 ± 8.58 N/A 58 65.68 ± 7.57 N/A
2 37 66.84 ± 8.83 0.83 ± 1.46 28 64.58 ± 7.08 1.35 ± 0.56
3 21 71.0 ± 6.97 0.53 ± 0.39 5 66.08 ± 2.78 1.26 ± 0.43
4 14 70.89 ± 6.33 0.47 ± 0.33 0 N/A N/A
5 4 72.08 ± 4.81 0.49 ± 0.33 0 N/A N/A
6 1 79.9 ± 0.0 0.58 ± 0.40 0 N/A N/A

Controls (n=138)
All 115 61.87 ± 10.43 N/A 49 63.12 ± 5.90 N/A
2 50 61 ± 12.01 0.79 ± 0.66 18 60.00 ± 5.87 0.91 ± 0.27
3 28 65.75 ± 5.96 0.66 ± 0.52 0 N/A N/A
4 17 66.82 ± 4.88 0.45 ± 0.28 0 N/A N/A
5 8 66.11 ± 4.83 0.44 ± 0.25 0 N/A N/A
6 0 N/A N/A 0 N/A N/A

Table 4.1: Demographic details for participants in the PCA study. Number of participants
(n), mean and standard deviation age of participants at baseline visit and mean and
standard deviation of visit interval is shown per number of visits.

widely-accepted Tang-Wai et al. [147] and Mendez, Ghajarania & Perryman [146] crite-
ria. Participants had no clinical features of other neurodegenerative disorders (e.g. visual
hallucinations, pyramidal signs), hence fulfilling the criteria for PCA-pure [36]. 106 tAD
patients and 138 healthy controls recruited from the DRC UK were also used for this
study. tAD subjects all met criteria for probable AD [224]. Available pathological and
molecular analyses for the patients (45/117 = 38% for PCA, 49/106 = 46% for tAD) all
indicated AD pathology.
Of all the study participants, 270 had undergone at least one T1 MRI scan and 216
at least one cognitive assessment. Available neuroimaging and neuropsychology data,
stratified by the number of visits, are shown in table 4.1. PCA, tAD and healthy controls
were age-matched (65.44 ± 7.51 for PCA, 65.67 ± 7.57 for tAD and 63.13 ± 5.94 for
controls). The gender proportion was as follows: 39% male for PCA, 62% male for tAD
and 50% male for controls. PCA and tAD subjects had a similar level of impairment as
measured by MMSE scores at first assessment: 20.88 ± 5.17 for PCA, 19.58 ± 5.08 for
tAD and 29.02 ± 0.98 for controls.
For analysing the heterogeneity within PCA, we split the dataset into three groups
based on performance on a suite of cognitive tests. For each subject we computed the
74 Chapter 4. Longitudinal Neuroanatomical Progression of PCA

Vision subgroup Space subgroup Object subgroup


(n=30) (n=30) (n=29)
n mean (std) n mean (std) n mean (std)
MMSE 29 22.38 (5.19) 28 19.54 (4.10) 29 21.90 (5.62)
Age 30 64.23 (7.72) 30 63.26 (7.44) 29 65.05 (8.48)
Gender (%male) 30 16% 30 46% 29 41%
PAL 6 15.00 (5.66) 5 5.60 (6.18) 8 10.75 (7.14)
Digitspan forwards 17 6.76 (2.86) 27 6.15 (2.46) 24 7.08 (2.36)
Digitspan backwards 16 3.88 (1.69) 28 2.89 (1.42) 23 3.78 (2.08)
GDA addition 17 1.59 (1.94) 22 0.55 (1.41) 19 1.95 (2.84)
GDA subtraction 17 0.76 (1.16) 22 0.23 (0.67) 19 1.84 (2.87)
Memory
SRMT faces 18 18.78 (4.13) 25 18.88 (4.30) 20 18.10 (3.95)
SRMT words 30 21.70 (2.88) 30 19.43 (3.45) 29 20.38 (2.73)
Visual processing
Shape detection (VOSP) 28 14.43 (3.29) 29 17.17 (2.24) 28 17.18 (2.42)
Shape discrimination 28 12.96 (3.18) 28 15.21 (3.45) 28 16.14 (2.68)
Crowding 22 6.68 (4.23) 28 9.07 (2.34) 28 8.71 (2.23)
Space Perception
Number location 29 2.10 (2.66) 29 2.07 (2.69) 28 4.68 (2.83)
Dot counting (n correct) 29 4.03 (3.27) 29 3.93 (3.15) 29 6.93 (2.64)
A cancellation (time) 28 73.90 (20.24) 29 82.25 (13.95) 28 62.68 (22.65)
A cancellation (n misses) 29 5.76 (5.33) 29 4.00 (4.36) 29 3.62 (4.77)
Object perception
Object decision 30 9.77 (4.51) 30 12.63 (3.95) 29 10.03 (4.13)
Fragmented letters 23 3.57 (4.99) 29 6.34 (5.25) 27 5.19 (5.72)
Unusual views 20 3.05 (3.97) 25 6.16 (5.03) 22 3.36 (3.95)
Usual views 20 9.65 (6.42) 25 15.16 (4.89) 22 12.59 (5.58)
Memory – visual processing
composite score 30 -28.90 (16.82) 30 14.26 (20.33) 29 11.93 (14.79)
discrepancy
Space – object composite
30 -1.62 (26.91) 30 23.06 (16.93) 29 -17.70 (16.30)
score discrepancy

Table 4.2: Baseline population demographics and neuropsychological data for PCA sub-
groups. For every neuropsychological test, we report the number of participants with
available data (n) and the mean and standard deviation of the available measures.
4.3. Methods 75

following composite tests by summing up scores from individual tests:

• Early visual processing: shape detection, shape discrimination and crowding

• Visuoperceptual processing: object decision, fragmented letters, usual and unusual


views.

• Visuospatial processing: number location, dot counting and A cancellation (time –


cut off at 90s)

• Episodic memory: short recognition memory test (sRMT) for words and faces

The score for each of the four categories was computed by standardising each of
the sub-scores on a 0-100 scale, corresponding to the minimum and maximum values
obtained by the participants, and then taking the average within each category. The
subjects were then classified into three groups. The worst 1/3 of subjects (n=30) on the
early visual processing tests as compared to the memory tests (i.e. difference between
early visual and memory tests) were assigned to the vision subgroup. The remaining 2/3
of participants were split into two groups based on the difference between visuoperceptual
and visuospatial tasks: subjects with space < object performance (n=30) were assigned to
the space subgroup while remaining subjects (n=29) were assigned to the object subgroup.
Of all the PCA subjects selected for the subgroup analysis, only 23 (vision), 21 (space)
and 18 (object) had imaging data. Demographics and neuropsychological data of subjects
belonging to the PCA subgroups is shown in table 4.2.

4.3.2 Image Acquisition and Preprocessing


89 PCA, 66 tAD and 115 healthy controls had at least one T1-weighted MRI scan (see
Table4.1). A total of five different scanners were used: two 3T Trio (DRC and UCSF),
1.5T Intera (HUVR) and two 1.5 Signa (DRC). The scans had full brain coverage using
between 124 and 208 coronal or sagittal slices of 1.0 or 1.5mm in thickness.
Estimation of region-of-interest (ROI) volumes was performed using the Geodesic In-
formation Flow (GIF) [225] based on the Neuromorphometrics atlas [226]. The atlas
produced 144 different brain ROIs across the left and right hemispheres. Segmentation
failed for 6 scans belonging to 5 subjects (3 controls, 1 PCA and 1 tAD), which were
subsequently removed. A total of 52 brain ROIs were removed (18 were not part of the
cerebral cortex, 6 had segmentation errors, 28 were grouped into larger ROIs). After
merging the left and right sides of the remaining ROIs, this resulted in a total of 46
ROIs which were further averaged into 8 ROIs corresponding to whole brain, hippocam-
pal, occipital, frontal, entorhinal, parietal and ventricle. GIF-derived ROI volumes were
corrected for total intracranial volume (TIV), age, gender, scanner type and site using a
general linear model and a one-hot encoding scheme.

4.3.3 Statistical Methods


4.3.3.1 The Event-based Model
For finding the order in which brain regions are affected, we ran the standard Event-
Based Model [23] (section 3.5.1) independently on estimated volumes from PCA and tAD,
using the shared control population in both scenarios. For the EBM event distributions,
76 Chapter 4. Longitudinal Neuroanatomical Progression of PCA

(a) (b)

(c) (d)

Figure 4.1: Diagram of the Differential Equation Model. (a) Subject-specific biomarker
rates of change were measured from line of best fit, i.e. line slope. (b) Rate of change
model: the slopes of each fitted line were plotted against the average biomarker value of
each subject (blue crosses). A non-parametric model (Gaussian Process regression, green
line) was then fitted on measurements. (c) Trajectory reconstruction: A line integral
was performed on the rate of change model. (d) Anchoring process: to give an absolute
time reference, the origin t0 was set as the line that best separates controls from patients,
which have been staged along the time axis using their biomarker data. Diagram made
by me.

we chose a Gaussian distribution for likelihood of normal observations, and a uniform


distribution for the likelihood of abnormal observations. We further assumed that the
control population is well-defined, so we fit the Gaussian distribution directly on the
control data. For the uniform distribution, we set the minimum and maximum of the
uniform range to be equal to the smaller and largest observed biomarker values. For
finding the optimal sequence, we used 25 different starting points and performed greedy
ascent for 10,000 biomarker sequences (see section 3.5.1.2). We chose the sequence with
the maximum likelihood as the final sequence.

For estimating uncertainty within the EBM sequence, we used MCMC to take 100,000
samples of the event sequence, starting from the maximum likelihood solution. The
perturbation rule used is described in detail in section 3.5.1.2.
4.3. Methods 77

4.3.3.2 The Differential Equation Model


For estimating the rate and extent of biomarker decline, we applied the Differential
Equation Model [25, 30] (section 3.5.2). The methodology is outlined in Fig. 4.1. The
biomarker measurements for each subject were plotted against time since baseline, and
a line was fit for each subject independently. The slope of these lines was then used
as a measure of the biomarker rate of change (Fig. 4.1a). The slopes of each fitted
line were then plotted against the average biomarker value of each subject (Fig. 4.1b).
A line integral is then performed on the rate of change model (Fig. 4.1c). We repeat
the DEM fitting for 8 ROIs independently: the four main lobes, whole brain, ventricles,
hippocampus and entorhinal cortex. In order to align all images on a common time frame,
we staged the controls and patients based on all their data, and then set an absolute time
reference t0 as the line that best separated (i.e. maximised the balanced classification
accuracy) the controls’ and patients’ stages.1
There were some adaptations that we performed on the DEM model to ensure a good
data fit. In the estimation of the rate of change model (Fig. 4.1b), we did not include the
controls, as there was very little change in their biomarker values. We also normalised
the average biomarker values to z-scores and standardised the rates of change by dividing
them with the average rate of change of all patients. At the line integration step (Fig.
4.1c), the integration limits were defined as the biomarker values where the corresponding
change is zero or the average biomarker value was equal to the minimum or maximum
observed in the dataset.
After the line integration step, we aligned all trajectories on a common time axis
through an anchoring process, where we set the time t0 = 0 to correspond to the value
of that biomarker in patients at baseline, averaged across all patinets. More precisely,
we set fj (t0 ) = avg(Xj ) where fj is the trajectory for biomarker j and Xj are the values
of biomarker j in tAD/PCA patients at baseline visit. After this initial anchoring, we
staged the subjects along their progression and re-set the t0 = 0 to correspond to the
threshold that best separated controls from patients (Fig. 4.1d). After the anchoring
process, we converted all biomarker values to z-scores for comparability (Fig. 4.1d).
To estimate uncertainty in the trajectories, we sampled 20 trajectories from the pos-
terior distribution of the GP, and then anchored them like the mean trajectory. However,
the anchoring would’ve resulted in zero noise interval at the anchor point, so to get real-
istic confidence intervals we added to each trajectory an extra amount of random noise
N (0, σ) on the y-axis, with σ set to the standard deviation of the biomarker measurements
of each subject at baseline visit.

4.3.3.3 Statistical Tests


In order to find out statistically significant differences between the EBM- and DEM-
estimated trajectories, we applied several non-parametric statistical tests.
For EBM results, we tested the effect size of biomarker i becoming abnormal before
another biomarker j both within- and between-group. Within-group differences were
assessed using Wilcoxon signed-rank one-tailed tests for all pairs of biomarkers. Between-
group (PCA vs tAD) and between-subgroup (space vs object, space vs vision and object
vs vision subgroups) differences were assessed using two-tailed Mann-Whitney U tests.

1
The staging of subjects using all their data required an initial trajectory alignment, which we aligned
by initially setting t0 to be the mean biomarker value of patients at baseline.
78 Chapter 4. Longitudinal Neuroanatomical Progression of PCA

We used these non-parametric tests due to non-gaussianity of the data (data is ordinal
representing ranks). The reason for using different tests (Wilcoxon vs Mann-Whitney)
is becuase in one case we compare paired samples (two events within the same sequence
sample), and in the other unpaired (two events in different sequences, e.g. in a randomly
sampled PCA sequence vs a different randomly sampled tAD sequence). We also thinned
the MCMC samples (1/100) due to dependence between consecutive samples.
For DEM results, we tested for differences in estimated biomarker values at different
timepoints (-10, 0 and 10 years from t0 ) both within- and between-groups. For every pair
of ROIs, within-group differences were assessed using two-tailed unpaired t-tests. For
all ROIs and timepoints, between-group (PCA vs tAD) differences were assessed using
similar two-tailed t-tests. For rejecting null hypotheses, we applied Bonferroni-corrected
significance thresholds for all tests performed on EBM and DEM results.

4.4 Results
4.4.1 Progression of PCA and Typical AD
Fig. 4.2 shows the maximum likelihood progression of atrophy estimated by the EBM, for
both PCA and tAD patients. Snapshots of brain atrophy were taken at model stages 4,
8, 16, 24, 32, 40 and 46 (of 46) using the template from Supplementary Fig. A.1. Figure
4.3 shows the maximum likelihood sequence and the variance in the main sequence. PCA
patients show early atrophy in occipital areas, ventricles and the superior parietal region,
while tAD patients show early atrophy in the amygdala, hippocampus and entorhinal
cortex, followed by temporal areas. The ordering is largely preserved under bootstrapping
(Supplementary Fig. A.2), and supported by statistical testing (Supplementary Fig.
A.3). Differences in abnormality sequences between PCA and tAD are also statistically
significant under Bonferroni corrections (Supplementary Figure A.7).
Fig. 4.4 shows the DEM-estimated biomarker trajectories for PCA (left) and tAD
(right). Confidence estimates of the mean trajectory are also given in Fig. 4.5. Amongst
PCA patients, occipital and parietal atrophy was most evident before t0 , and by t0 we
also observe considerable atrophy in the temporal lobe. Between t0 and 10 years after
t0 , we observe a marked increase in the rate of occipital, parietal and temporal atrophy
and ventricular expansion. By contrast, hippocampal, entorhinal and frontal atrophy
never match the extent of tissue loss in posterior and temporal regions. After 10 years
from t0 , atrophy rates in occipital, parietal and temporal lobes seem to slow down, but
limited data in this time window prevents drawing any clear conclusions. Statistical
testing within the PCA cohort also confirms our conclusions – see Supplementary Tables
A.1, A.2 and A.3.
By contrast, before t0 tAD patients showed most extensive tissue loss in the hippocam-
pus, confirmed by significance tests between hippocampal volume and other regions (p <
4e-05, see Supplementary Figs. A.4 and A.5). After t0 , subsequent rates of change are the
highest for temporal atrophy and ventricular expansion. It is of note that within 12 years
from t0 , model estimates of parietal and ventricular abnormality amongst tAD patients
are equivalent to or exceed the relative extent of hippocampal abnormality. Comparing
PCA and tAD trajectories directly (Fig. 4.5), the separation between groups at t0 is
greatest in parietal (PCA > tAD, p < 1e-6) and hippocampal (tAD > PCA, p < 1e-22)
volumes – see Supplementary table A.7 for full statistical testing.
4.4. Results 79

normal abnormal

Posterior Cortical Atrophy


Stage 4 Stage 8 Stage 16 Stage 24 Stage 32 Stage 40 Stage 46

Typical Alzheimer’s Disease


Stage 4 Stage 8 Stage 16 Stage 24 Stage 32 Stage 40 Stage 46

Figure 4.2: Atrophy progression in PCA and tAD patients according to the event-based
model. White regions are within the volume range of healthy controls, while red regions
show abnormally low volumes by the corresponding stage, with shading indicating the
probability of abnormality. By each stage, a number of biomarkers shaded in red became
abnormal. Brain pictures generated using BrainPainter [227]

4.4.2 Progression of PCA Subgroups


Fig. 4.6 shows snapshots of the EBM-estimated atrophy sequence at early stages, in the
three cognitively-defined PCA subgroups. The bottom row in the figure shows the uncer-
tainty in the estimated atrophy sequence. The vision subgroup has initial atrophy in the
inferior occipital lobe, followed by the angular gyrus, middle temporal, precuneus and
superior parietal. The space subgroup shows early atrophy in the superior parietal area
(dorsal pattern), followed by inferior occipital and inferior and middle temporal areas.
Finally, the object subgroup shows initial atrophy in the middle and inferior occipital ar-
eas, with subsequent atrophy in the inferior and middle temporal areas (ventral pattern).
Bonferroni-corrected statistically significant differences in atrophy progression have also
been observed between PCA subgroups – see Supplementary Fig. A.8. Longitudinal
trajectories within PCA subgroups using the DEM could not be estimated due to lack of
sufficient data.
80 Chapter 4. Longitudinal Neuroanatomical Progression of PCA

Posterior Cortical Atrophy

Typical Alzheimer’s Disease

Figure 4.3: Uncertainty in the EBM-estimated atrophy sequences for (top) PCA and
(bottom) tAD from Fig 4.2. The ROIs on the Y-axis are ordered according to the timing
of abnormality, from early abnormalities on the top to late abnormalities on the bottom.
The X-axis shows the position of a biomarker in the abnormality sequence. Each pixel at
position (i, j) shows the probability of biomarker j becoming abnormal at position i, with
darker squares showing higher confidence and whiter squares showing lower confidence.
The biomarker orderings are sampled from the EBM posterior distribution.
4.4. Results 81

Hippocampus Entorhinal
Hippocampus Whole Brain FrontalParietal
Entorhinal Occipital Parietal Whole Brain
Whole Brain Ventricles
Frontal Frontal
Occipital TemporalOccipitalEntorhinal TemporalHippocampus
Ventricles Ventricles Parietal Temporal
2 2 2

Z-score relative to controls


Z-score relative to controls

Z-score relative to controls


0 0 0
2 2 2
4 4 4
PCA
6 PCA 6 Controls 6 Controls
Controls 8 AD
8 8
10
10 15 10 5 0 5 1010 15 20
15 10 5 0 5 10 15 Years
20 since t0 15 10 5 0 5 10 15 20
Years since t0 Years since t0
(a) PCA (b) tAD

Figure 4.4: (a-b) Trajectories of different ROI volumes from the differential equation
model for (a) PCA progression and (b) tAD progression. The x-axis shows the number
of years since t0 , and the y-axis shows the z-score of the ROI volume relative to controls.
The trajectories of the ventricles have been flipped to aid comparison. Overlayed are
histograms of subject stages based on the estimated trajectories.

PCA AD PCA samples AD samples 0 std 1 std

Whole Brain Ventricles Hippocampus Entorhinal


2 2 2 2
0 0 0 0
2 2 2 2
4 4 4 4
6 6 6 6
Z-score of biomarker

8 8 8 8
10 0 10 20 10 0 10 20 10 0 10 20 10 0 10 20
Occipital Temporal Frontal Parietal
2 2 2 2
0 0 0 0
2 2 2 2
4 4 4 4
6 6 6 6
8 8 8 8
10 0 10 20 10 0 10 20 10 0 10 20 10 0 10 20
Years since t0

Figure 4.5: Mean trajectories for ROI volumes for PCA and tAD aligned on the same
temporal scale with samples from the posterior distribution showing the confidence of the
mean trajectory. The axis shows the number of years since t0 , and the y-axis shows the
z-score of the ROI volume relative to controls. The trajectories for the ventricles have
been flipped to aid visual comparison. The 1 std and 0 std horizontal lines represent the
limit of 1 and 0 standard deviations away from the mean values of controls.
82 Chapter 4. Longitudinal Neuroanatomical Progression of PCA

normal abnormal

Vision impairment group


Stage 1 Stage 2 Stage 3 Stage 4 Stage 5 Stage 6 Stage 7

Space perception impairment group


Stage 1 Stage 2 Stage 3 Stage 4 Stage 5 Stage 6 Stage 7

Object perception impairment group


Stage 1 Stage 2 Stage 3 Stage 4 Stage 5 Stage 6 Stage 7

Vision Space Object

Figure 4.6: Early atrophy progression within the three cognitively-defined PCA sub-
groups, as estimated by the EBM. The top figures shows snapshots of the atrophy pat-
terns for the first 7 stages in the EBM, while the last row shows the uncertainty in the
atrophy progression sequence. Brain pictures generated using BrainPainter [227]
4.5. Discussion 83

4.5 Discussion
In this work we performed one of the first longitudinal studies of atrophy progression
in PCA. Results suggest that in PCA occipital and superior parietal areas are the first
to become affected, followed by temporal areas. By 10 years after t0 , there seems to be
widespread atrophy in the occipital, parietal and temporal areas, as well as ventricular
expansion. In contrast, tAD seems to have significant early atrophy in the hippocampus,
with subsequent temporal atrophy and ventricular expansion starting 5 years after t0 .
Regarding PCA heterogeneity, our study also provided the first glimpse into the early
longitudinal patterns of atrophy within three cognitively defined PCA subtypes. We
found early phenotype-specific patterns of atrophy within each cognitively-defined PCA
subgroup. These patterns of pathology overlap with the pathways that are hypothesised
to be affected within each group: striate cortex for the vision subgroup, dorsal pathway
for the space subgroup and ventral pathway for the object subgroup. Nonetheless, among
the subgroups there is considerable variability in these patterns as well as spatial overlap,
which might suggest that these should not necessarily be interpreted as distinct diseases,
but rather that the patients lie on a continuum of phenotypical variation, as suggested
by [144].
Our study has several strengths. First of all, the large number of PCA subjects with
longitudinal neuroimaging and cognitive data allowed us to perform a robust analysis of
PCA atrophy progression. The EBM and DEM methods we used are all data-driven,
don’t require manual biomarker thresholds and don’t rely on diagnostic classes, which
are often noisy and biased. Moreover, the ability of the EBM to work with limited
cross-sectional data allowed us to estimate the progression of PCA subgroups, which are
small and have limited longitudinal data available. An advantage of the DEM method is
its ability to fit continuous, non-parametric biomarker trajectories based on GPs, which
makes it suitable for modelling biomarkers whose trajectories have varying shapes.
Nevertheless, our study has several limitations that need to be addressed. First of
all, since data was acquired over an extended period of time, not all subjects had CSF,
molecular or pathological confirmation for Alzheimer’s disease. This can be a problem,
as previous studies suggested that at least half of patients who receive a diagnosis of
probable AD actually have other non-AD underlying pathologies [228, 229]. Follow-up
studies will need to have a higher proportion of patients with pathological or molecular
confirmation. Moreover, the data was acquired in three different centres using different
scanners and field strengths, although we adjusted for these covariates.
The EBM and DEM models that we employed also have several limitations that
we acknowledge. First of all, both methodologies assume all subjects follow the same
progression sequence. Secondly, the DEM requires longitudinal data, which prevented
us from fitting the DEM to the PCA subgroups, who lacked enough longitudinal data.
Another assumption made by the EBM is that the control population is well-defined, as
we fit the distribution of normal biomarker values directly on the biomarker values of the
control population. The EBM also assumes simplistic, step-wise biomarker trajectories
that switch from a normal to an abnormal value. With respect to the DEM, the approach
requires a reference timepoint, which we took it to be the threshold that best separates
the controls from patients after disease staging.
There are several avenues for future research. Further molecular and pathological
confirmation can be obtained for the remaining patients to ensure they all have a reliable
diagnosis, which will enable an unbiased estimation of the progression sequence. The
84 Chapter 4. Longitudinal Neuroanatomical Progression of PCA

EBM and DEM methodologies can be further extended to allow random effects or to fit
different progression sequences for different sub-populations in a data-driven way, such
as the approach of [230]. Information about the rate and extent of atrophy in the PCA
subgroups can also be computed after enough data has been acquired. A well-defined
control population for the EBM can also be defined by selecting only amyloid-negative
subjects or by other types of stratification. The EBM model can be extended to model
more complex trajectory shapes, while the DEM can be further extended to a multivariate
approach that inherently aligns the biomarker trajectories.
Finally, one of the key directions of future research is to understand the disease mech-
anisms underlying PCA. To this end, several methods can be used to estimate these
mechanisms, such as those based on propagation of pathogenic proteins [6, 223] or the
architecture of brain networks [126]. The influence of genetic factors such as Alipoprotein
E (APOE) status [150, 151] and other factors recently identified [231, 150] from genome-
wide association studies also need to be understood. This research will lead the way
towards drug development in PCA clinical trials and will allow the selection of robust
outcome measures and fine-grained patient stratification in clinical trials in PCA.

4.6 Conclusion
In this work I performed a statistical analysis of the neuroimaging data from PCA and
tAD subjects from the DRC, HUVR and UCSF centres. I pre-processed all the MRI
images and applied the event-based model and the differential equation model on the
PCA and tAD cohorts, as well as on three cognitively-defined PCA subgroups. The
analysis I made gives the first glimpse into the longitudinal progression of atrophy in
PCA subjects, and into the early longitudinal patterns of atrophy in the vision, space
and object subgroups.
In the following chapter, I will present some novel extensions to the EBM and DEM
models that will enable better estimation of the parameters for the EBM and alignment of
the biomarker trajectories for the DEM. These improvements can provide a more accurate
disease signature, and remove the need for ad-hoc methods of estimating parameters.
Chapter 5

Novel Extensions to the Event-based


Model and Differential Equation
Model

5.1 Contributions
In this work I present methodological extensions to the event-based model (EBM) and
differential equation model (DEM) and I evaluate their performance compared to the
standard implementations. In order to assess differences between these methods more
accurately, I also propose novel performance measures based on disease staging consis-
tency and prediction of time elapsed between visits. I formulated and implemented the
novel methodologies, and performed their evaluation. I also pre-processed the DRC MRI
scans. My colleague Alexandra Young pre-processed the ADNI data.

5.2 Introduction
Many data-driven disease progression models (DPMs) that have been presented in chap-
ter 3 make assumptions about the biomarker data and the model parameters, which limit
their usefulness on practical applications. For example, the differential equation model
by [25] is univariate, hence it assumes independence across different biomarkers. In order
to place biomarker trajectories on the same time frame, in the previous chapter we used a
post-hoc anchoring process (see section 4.3.3.2). This anchoring is inaccurate, as it relies
on setting the reference time t0 using biomarker values of a clinical group (i.e. controls
or AD). This anchoring process is challenging because of singularities arising from flat
trajectories1 , and the fact that subjects are at different stages along the disease. Another
limitation of some DPMs is that the fitting algorithm assumes independence between dif-
ferent sets of parameters. While this is done in order to ensure computational tractability,
this yields inaccurate parameter estimates. In particular, the event-based model param-
eter estimation procedure proposed by [23] and [24] assumes that the parameters of the
likelihood models for normal and abnormal values are independent of the abnormality

1
The alignment is performed by setting t0 = 0 so that f (t0 ) = mean(patients). However, if the
trajectory is flat then there are many points t0 that match the mean of patients. Even is the trajectory
is not fully flat, the measurement noise is amplified by the low slope of f .
86 Chapter 5. Novel Extensions to the EBM and DEM

sequence. Some better parameter estimation procedures are therefore needed, which can
ensure a robust data fit.
The evaluation of the performance of disease progression models is another open prob-
lem that has not been addressed so far. While previous studies used accuracy of clinical
status predictions[24], clinical diagnosis is often not reliable without neuropathological
confirmation – one study reported that a clinical diagnosis of probable AD has between
70.9–87.3% accuracy and between 44.3%–70.8% specificity. Therefore, performance met-
rics based on the prediction of clinical diagnosis might not be sufficiently sensitive to
differences in the performance of such algorithms. While [23] computed the number of
subjects with increased staging over time – a performance measure that doesn’t rely on
clinical diagnosis – it does not take model uncertainty of staging into account and it is
specific to discrete models such as the event-based model.
In this chapter we suggest novel extensions in the event-based model and differen-
tial equation model and propose four novel performance measures for evaluating disease
progression models that don’t rely on clinical diagnosis. For the event-based model, we
devise two novel fitting procedures that perform joint optimisation of the parameters of
the normal and abnormal likelihood models, as well as the abnormality sequence. For
the differential equation model, we devise a novel data-driven way to align the biomarker
trajectories to a common axis by estimating trajectory-specific and subject-specific time
shifts. The novel performance measures that we propose exploit uncertainty in the es-
timated stages and are also suitable for evaluating continuous trajectory models. Using
data from the Alzheimer’s Disease Neuroimaging Initiative (ADNI) and the Dementia
Research Centre (DRC), UK, we show that the novel models generally have better or
equal performance compared to standard models. Moreover, we also show that the novel
performance measures that we proposed are more sensitive to changes in models than
standard measures based on the prediction of diagnosis or conversion status.

5.3 Methods
5.3.1 EBM Extensions
In this section we outline two novel methods of parameter fitting for the event-based
model: a blocked MCMC sampling of the distribution parameters and event ordering
(section 5.3.1.1), and an Expectation-Maximisation approach (section 5.3.1.2). Further-
more, we also present a novel methodology performing a data-driven temporal alignment
of the differential equation model trajectories (section 5.3.2).

5.3.1.1 EBM – Joint MCMC Sampling


We present a novel method for fitting the event-based model that jointly optimises the
parameters of the normal and abnormal likelihood models using MCMC sampling. The
full EBM likelihood model, already been described in Eq. 3.4, is:
P
" N k N
!#
Y X Y  Y 
p(X|S) = p(k) p xi,s(j) |Es(j) p xi,s(j) |¬Es(j) (5.1)
i=1 k=0 j=1 j=k+1

where xij represents the value of biomarker j from subject i and is informative of event Ej
in subject i, P is the number of subjects and N is the number of biomarkers. The abnor-
5.3. Methods 87

mality sequence S = [S(1), . . . , S(N )] describes the order in which events E1 , E2 , . . . , EN


become abnormal, and models disease progression.
We now reformulate the EBM likelihood model to explicitly take into account the
parameters of the likelihood models for the normal and abnormal biomarker values. This
will allow joint optimisation of these parameters, along with sequence parameter S. We
assume the likelihood models for normal and abnormal biomarker values are Gaussian
distributions, i.e. p(x|Ej ) ∼ N (µaj , σja ) and p(x|¬Ej ) ∼ N (µnj , σjn ) where µaj and σja model
the distribution of abnormal values for biomarker j (i.e. event Ej occurred), while µnj
and σjn model the distribution of normal values for biomarker j (event Ej did not occur).
 
Thus, the full set of parameters that need to be modelled is θ = [µnj , σjn , µaj , σja ]j=1...N , S .
Therefore, the likelihood in equation 5.1 can be explicitly written as:

p(X|S, [µnj , σjn , µaj , σja ]j=1...N ) =


P
" N k N
!#
Y X Y Y
p xi,S(j) |µaS(j) , σS(j)
a n
p xi,S(j) |µnS(j) , σS(j)
 
p(k) (5.2)
i=1 k=0 j=1 j=k+1

We maximise this likelihood using blocked MCMC sampling, where at each step we
only propose parameters for biomarker j, i.e. [µnj , σjn , µaj , σja ] along with a new sequence
Sjnew where only event Ej changed its position. The distribution parameters for the
other biomarkers and the ordering of the other events i 6= j are kept the same. This
blocked approach can lead to faster convergence because there is strong dependence
between parameters corresponding to the same biomarker and between the position of the
corresponding event in the sequence. The covariance matrix of the proposal distribution
is estimated by taking 100 bootstraps of the dataset and computing the covariance of
[µc , σ c , µp , σ p ], where µc , σ c are the mean and standard deviation of the control group
while µp , σ p are the mean and standard deviation of the patient group.

5.3.1.2 EBM – Expectation Maximisation


The blocked MCMCM approach from the previous section can be challenging to imple-
ment and execute, due to the difficulty of sampling in a high-dimensional space. The
user needs to further tune the covariance matrix in order to get a good acceptance rate,
and ensure enough samples are taken in order to exhaustively explore the space of the
distribution. To mitigate these issues, we further propose a novel parameter estima-
tion procedure for the EBM based on the Expectation Maximisation (EM) framework
[195]. The EM framework is suitable for estimating parameters of models with discrete
latent variables, such as the EBM which has discrete latent variables k representing
the subject-specific stages. The EM framework tries to find the parameters θ∗ that
maximise the expected log-likelihood of the complete data θ∗ = arg maxθ Q(θ|θold ) =
arg maxθ EZ|X,θold [log p(X, Z|θ)]. The key observation to make is that the joint likelihood
over the latent variables Z = [Z1 , ..., ZN ] and X = [X1 , ..., XN ] factorises, giving the
following form:

P X
" zi N
#
X X X
Q(θ|θold ) = p(Zi = zi |Xi , θold ) log p(xij |ES(j) ) + log p(xij |¬ES(j) )
i=1 zi j=1 j=zi +1
(5.3)
88 Chapter 5. Novel Extensions to the EBM and DEM

d
We find the maximum for µnk , the mean of p(x|¬Ek ), by solving dµn
Q(θ|θold ) = 0.
k
This gives the following update equation for µnk :
P
X
µnk = xik win (5.4)
i=1

with weights win defined as:


p(S −1 (k) > Zi |X, θold )
win = PP (5.5)
i=1 p(S −1 (k) > Zi |X, θold )
and p(S −1 (k) > Zi |X, θold ) = K old
P
l=S −1 (k)+1 p(Zi = l|X, θ ). The full derivation is given
in section D.1. Similar update rules are derived in the appendix section D.1 for the other
parameters: σkn , µak , σka .
Optimising the sequence S in the M-step is intractable, so we use MCMC sampling
where at each step of the sampling process we propose a new sequence S new , find the
optimal distribution parameters for each biomarker given S new using the closed-form
EM update rules, and then evaluate the likelihood Q(θ|θold ). We keep performing a
greedy ascent as performed by [23] until convergence. Although this approach might
not guarantee that we truly find the optimal parameters, it still results in an increase
of Q(θ|θold ). This approach, called generalised EM, guarantees that the method will
converge to a local maxima [195]. For parameter initialisation, we use the mean and
standard deviation of the control and patient populations.

5.3.2 DEM – Optimised Trajectory Alignment


We present a novel extension to the DEM (see section 3.5.2) that aims to place the esti-
mated biomarker trajectories on the same temporal axis, in a data-driven way. The main
idea is to use the subjects’ data to find optimal subject-specific and trajectory-specific
time shifts. Since we are mostly interested in estimating population-level trajectories and
to align them on a common time-axis, we do not currently add subject-specific progression
speeds and random-effect deviations from the average trajectories.
Let us denote by X our dataset, where xpb is the measurement of biomarker b in
patient p and fb is the shape of the trajectory for biomarker b. For every biomarker b,
we aim to estimate a temporal shift tb of the trajectory and a measurement noise σb . At
the same time. The log-likelihood for the data Xp from patient p can be expressed as:
B
Y
p(Xp |t1 , . . . , tB , σ1 , . . . , σB , zp ) = N (xpb |fb (zp − tb ), σb ) (5.6)
b=1
where zp is a latent parameter representing the time-shift of subject p. Multiplying
by the prior on zp and summing over all the possible values of zp we get the marginal:
X B
Y
p(Xp |t1 , . . . , tB , σ1 , . . . , σB ) = p(Zp = zp ) N (xpb |fb (zp − tb ), σb ) (5.7)
zp b=1

Assuming the data from each subject is conditionally independent given zp , we get
the full likelihood:
P X
Y B
Y
p(X|t1 , . . . , tB , σ1 , . . . , σB ) = p(Zp = zp ) N (xpb |fb (zp − tb ), σb ) (5.8)
p=1 zp b=1
5.3. Methods 89

This likelihood can be optimised with any method of choice such as MCMC sampling
or gradient methods. We chose to optimise the model using an iterative approach, where
for each biomarker b we optimise it’s trajectory shift tb conditioned on all the other
parameters (Markov blanket), and then estimate it’s measurement noise σb .

5.3.3 Performance Evaluation


We compare the performance of the extended EBM and DEM methods against the stan-
dard implementations, using novel performance metrics that we propose. In section
5.3.3.1, we present metrics which test staging consistency, i.e. that follow-up stages are
greater or equal to baseline stages. We then generalise this concept in section 5.3.3.2 to
test the accuracy of models in predicting the time lapse between two visits of a subject.

5.3.3.1 Staging Consistency


The staging consistency metrics test whether subjects’ stages at follow-up visits are
greater than or equal to the stages at baseline. While such a metric is simple to com-
pute in cases of no uncertainty, we also define a more complex metric that takes staging
uncertainty into account.
Let us consider a set of random variables zti representing the stage of subject i at
timepoint t, where i ∈ [1 . . . N ], t ∈ [1 . . . Ti ], N being the number of subjects and Ti
the number of time points for subject i. For most disease progression models, the EBM
and DEM included, we can find the posterior p(zti |Xi , θ), which we will call the staging
probabilities. Moreover, let Mti = arg maxs p(zti = s) be the maximum likelihood stage
for subject i at time point t. The hard staging consistency Ch counts the proportion of
stages from consecutive visits of every subject where the stage at the later visit must be
greater than the stage at the earlier visit. We define Ch as follows:

Ti
N X
1 X
Ch = PN I[Mti > Mt−1
i
] (5.9)
−N + i=1 Ti i=1 t=2

where the element −N + N


P
i=1 Ti is a normalising constant that represents the number of
pairs of consecutive visits from all subjects and time points in the dataset. The Ch metric
ranges from 0 (no consistent pairs of stages) to 1 (all pairs of stages are consistent).
We further seek to generalise the hard staging consistency, in order to make use of
the full staging probabilities, instead of using only the maximum likelihood stages. We
define the soft staging consistency Csi (t1 , t2 ) for subject i given two time points t1 and t2
as:
X
Csi (t1 , t2 ) = p(zti1 ≤ zti2 ) = p(zti2 = s)p(zti1 ≤ s) (5.10)
s∈S

where S is the set of possible stages in the disease progression model. We then define
the soft staging consistency for the whole population as the mean of subject-specific
consistencies for consecutive timepoints:

Ti
N X
1 X
Cs = PN Csi (t1 , t2 ) (5.11)
−N + i=1 Ti i=1 t=2
90 Chapter 5. Novel Extensions to the EBM and DEM

5.3.3.2 Time-lapse Prediction


Time-lapse prediction is a generalisation of the staging consistency, where the disease
progression model needs to predict how much time passed between two visits of the same
subject, which is the compared against the true time elapsed. This is only possible for
models that estimate continuous latent stage variables, such as the DEM. We define the
hard time-lapse metric Dh as follows:
Ti
N X
1 X i
τ (Mti ) − τ (Mt−1

Dh = PN ) − (ait − ait−1 ) (5.12)
−N + i=1 Ti i=1 t=2

where ait is the age of subject i at timepoint t, Mti is the maximum likelihood stage for
subject i at timepoint t and τ (Mti ) is the estimated time from onset associated with stage
Mti . The equivalent soft time-lapse metric Ds , which uses probabilistic staging variables
zti , is defined as:
Ti
N X
1 X
E[τ (zti ) − τ (zt−1
i

Ds = PN )] − (ait − ait−1 ) (5.13)
−N + i=1 Ti i=1 t=2

5.3.4 Data Preprocessing


5.3.5 The Dementia Research Centre Cohort
The Dementia Research Centre (DRC), UK cohort contains 89 controls, 74 PCA and 67
tAD subjects that have undergone 1.5T/3T T1-weighted MRI scans, with an average of
2-3 visits per subject. The demographics of the DRC dataset is given in table 5.1. More
details on the cohort can be found in the PCA longitudinal study from section 4.3.12 .

Demographics CN PCA AD
Number 89 74 67
Sex M/F 33/56 28/46 35/32
Age (years) 61 ± 11 63 ± 7 66 ± 9
Years from onset - 4.5 ± 2.8 4.8 ± 2.6
Number of visits 2.8 ± 2.5 2.5 ± 1.7 3.0 ± 2.7

Table 5.1: Baseline population demographics for the DRC cohort.

5.3.5.1 Image Processing


The MRI scans were segmented using the Geodesic Information Flows (GIF) algorithm
by [225], which is available as a service at http://cmictig.cs.ucl.ac.uk/niftyweb/.
The atlas that has been used for segmentation is the Neuromorphometrics atlas (provided
by Neuromorphometrics, Inc.), which produced 146 different brain ROIs across the left
and right hemisphere. All brain volumes have been corrected for total intracranial volume
(TIV), age and gender using a general linear model. We summed left and right brain
regions together and further selected a subset of 25 ROIs: (a) whole brain; (b) ventricles;
(c) 2 subcortical regions: hippocampus and amygdala; (d) 5 occipital regions: inferior,
2
This cohort is a subset of the cohort used in the PCA longitudinal study.
5.3. Methods 91

middle and superior occipital and the occipital fusiform and lingual; (e) 5 parietal regions:
superior parietal, angular, precuneus, supramarginal and postcentral; (f) 4 temporal
regions: inferior, middle and superior temporal along with fusiform; (g) 4 frontal regions:
superior, middle and inferior frontal along with precentral; and (h) 3 limbic regions:
entorhinal, parahippocampal and posterior cingulate.

5.3.6 The Alzheimer’s Disease Neuroimaging Initiative Cohort


In this study we also used the ADNI dataset to evaluate our disease progression models.
We used the same biomarker data as previously used by [24], which included all 285
subjects (Controls, MCI and AD) that had a CSF examination at baseline, cognitive
assessment at baseline and MRI scans at baseline and 1 year follow-up. The demographics
of the selected subjects is shown in table 5.2. We also used follow-up imaging, clinical and
CSF data at 12- and 24-months after baseline visit. Clinical diagnoses at baseline, 12-,
24- and 36-months were also used for evaluating performance on diagnosis prediction and
conversion prediction. The CSF total tau and phosphorylated tau were log-transformed
to improve normality.

Demographics CN MCI AD
Number 92 129 64
Sex M/F 48/44 82/47 34/30
Age (years) 75 ± 5 73 ± 7 75 ± 8
Education (years) 15.6 ± 2.9 15.9 ± 3 15 ± 3
APOE +/- 22/70 72/57 45/19

Table 5.2: Baseline population demographics for the ADNI cohort.

5.3.6.1 Image Processing

FreeSurfer Version 4.3 was used to compute regional volumes of the hippocampus, en-
torhinal cortex, middle temporal gyrus, fusiform gyrus, ventricles, whole brain and total
intracranial volume (TIV) at baseline, 12- and 24-month follow-up. All regional volumes
were normalised for each subject by dividing by TIV. Atrophy rates for the whole brain
and hippocampus were estimated using the Boundary Shift Integral (BSI) ([232]) using
the scans at baseline and 12-months follow-up. In particular, volume change for the
whole brain was measured using the KN-BSI method ([233]) and for hippocampus using
the MAPS-HBSI method ([234]).
We used the same biomarker set as the one used by [24], which included 14 biomark-
ers in total: (a) three CSF biomarkers: amyloid-β1−42 , phosphorylated tau and total
tau; (b) 3 cognitive tests: Alzheimer’s Disease Assessment Scale - Cognitive Subscale
(ADAS-Cog), Rey Auditory Verbal Learning Test (RAVLT) and the Mini-Mental State
Examination (MMSE); (c) six regional brain volumes: whole brain, ventricles, hippocam-
pus, entorhinal, middle temporal gyrus and fusiform gyrus; (d) rates of atrophy for two
ROIs: hippocampus and whole brain.
92 Chapter 5. Novel Extensions to the EBM and DEM

5.4 Results
We tested all novel EBM and DEM methods, along with their standard implementations.
We evaluated each model using the staging consistency and time-lapse metrics, using
data from the DRC and ADNI datasets. On the DRC dataset, we also evaluated the
models with respect to diagnosis prediction, while on ADNI we evaluated them based on
prediction of conversion from healthy controls to mild cognitive impairment (MCI) and
from MCI to Alzheimer’s disease.

5.4.1 DRC Results


We ran all the standard and novel methods for the EBM and DEM described above. For
the EBM joint MCMC sampling method, we took 100,000 samples who had an acceptance
rate in the range of 0.29-0.33 (min-max) across all cross-validation folds, which suggested
a good mixing, while the sample autocorrelation was in the range of 0.86-0.93. In order to
get the acceptance rate to this reasonable rate, we estimated the covariance matrix of the
MCMC proposal distribution based on the covariance of maximum likelihood parameter
estimates on bootstrapped subsets of the full dataset. For the EBM-EM method, we
performed 20 iterations as we noticed that the method converged after a maximum of 3-4
iterations.
In table 5.3 we show the staging-based metrics for the PCA cohort. Each entry shows
the mean and standard deviation of the metric calculated over 10 cross-validation folds.
Similar results are shown for the AD cohort in table 5.4. In table 5.5 we show the models’
balanced accuracy in diagnosis prediction on the DRC cohort. For reference, we also show
similar results using a standard Support Vector Machine (SVM) classifier [206]. The SVM
classifier was trained with a linear kernel trained using sequential minimal optimisation
[235] and a box-constraint parameter C = 1. In each entry we show the mean and
standard deviation of the balanced classification accuracy across the 10 cross-validation
folds.

Model Staging Consistency Time-lapse


Hard Soft Hard Soft
Event-based Model
EBM - Standard 0.88 ± 0.12 0.66 ± 0.09 - -
EBM - MCMC 0.96 ± 0.06 0.70 ± 0.06 - -
EBM - EM 0.95 ± 0.10 0.68 ± 0.11 - -

Differential Equation Model


DEM - Standard 0.94 ± 0.06 0.95 ± 0.05 0.54 ± 0.31 0.52 ± 0.29
DEM - Optimised 0.95 ± 0.05 0.95 ± 0.04 0.56 ± 0.28 0.52 ± 0.27

Table 5.3: Model performance according to staging-based metrics on PCA subjects from
the DRC cohort. The mean and standard deviations are calculated for each testing set
in 10-fold cross-validation.
5.4. Results 93

Model Staging Consistency Time-lapse


Hard Soft Hard Soft
Event-based Model
EBM - Standard 0.91 ± 0.16 0.71 ± 0.07 - -
EBM - MCMC 0.96 ± 0.07 0.76 ± 0.10 - -
EBM - EM 0.99 ± 0.01 0.72 ± 0.07 - -

Differential Equation Model


DEM - Standard 0.87 ± 0.10 0.88 ± 0.08 0.72 ± 0.91 0.67 ± 0.92
DEM - Optimised 0.87 ± 0.10 0.88 ± 0.08 0.74 ± 0.92 0.69 ± 0.92

Table 5.4: Model performance according to staging-based metrics on typical AD subjects


from the DRC cohort. The mean and standard deviations are calculated for each testing
set in 10-fold cross-validation.

Model PCA vs AD Controls vs PCA Controls vs AD


Event-based Model
EBM - Standard 0.72 ± 0.13 0.95 ± 0.05 0.90 ± 0.06
EBM - MCMC 0.79 ± 0.09 0.94 ± 0.06 0.90 ± 0.05
EBM - EM 0.80 ± 0.07 0.95 ± 0.05 0.87 ± 0.05

Differential Equation Model


DEM - Standard 0.81 ± 0.07 0.95 ± 0.05 0.90 ± 0.11
DEM - Optimised 0.82 ± 0.09 0.93 ± 0.06 0.88 ± 0.14

Support Vector Machine


SVM 0.79 ± 0.14 0.91 ± 0.06 0.88 ± 0.07

Table 5.5: Model performance at diagnosis prediction on the DRC cohort. Each en-
try shows the mean and standard deviation of the balanced accuracy across the cross-
validation folds.

5.4.2 ADNI Results

In table 5.6 we show the staging-based performance results of the progression models
on the ADNI dataset. As with the DRC results, for each metric we show its mean and
standard deviation over the 10 cross-validation folds. In table 5.7 we also evaluated the
models on how well they predict conversion from MCI to AD at 12-months, 24-months and
36-months from baseline visit. We did not compute results for prediction of conversion
status in controls due to small and very imbalanced datasets (i.e. only ).
94 Chapter 5. Novel Extensions to the EBM and DEM

Model Staging Consistency Time-lapse


Hard Soft Hard Soft
Event-based Model
EBM - Standard 0.83 ± 0.07 0.76 ± 0.05 - -
EBM - MCMC 0.84 ± 0.05 0.76 ± 0.06 - -
EBM - EM 0.84 ± 0.08 0.74 ± 0.06 - -

Differential Equation Model


DEM - Standard 0.87 ± 0.05 0.83 ± 0.08 0.85 ± 0.17 0.85 ± 0.16
DEM - Optimised 0.87 ± 0.05 0.84 ± 0.07 0.86 ± 0.15 0.86 ± 0.16

Table 5.6: Model performance according to staging metrics on ADNI data.

Model Duration between baseline and follow-up


12 months 24 months 36 months
Event-based Model
EBM - Standard 0.69 ± 0.14 0.64 ± 0.11 0.72 ± 0.15
EBM - MCMC 0.66 ± 0.14 0.63 ± 0.10 0.74 ± 0.14
EBM - EM 0.69 ± 0.15 0.63 ± 0.10 0.76 ± 0.15

Differential Equation Model


DEM - Standard 0.73 ± 0.13 0.72 ± 0.14 0.70 ± 0.13
DEM - Optimised 0.64 ± 0.11 0.69 ± 0.12 0.75 ± 0.14

Support Vector Machine


SVM 0.68 ± 0.15 0.70 ± 0.10 0.77 ± 0.08

Table 5.7: Model performance at prediction of conversion from MCI to AD on ADNI


data.

5.5 Discussion
5.5.1 Model Performance on DRC cohort
In the PCA cohort, we notice that the extended EBM methods show better results
compared to the standard EBM method, whereas the extended DEM method has equal
performance compared to the standard method. When comparing EBM vs DEM models,
most EBM models perform as well as the DEM models in terms of hard staging consistency
but relatively worse in soft staging consistency. There is also a drop in EBM staging
consistency when moving from the hard to the soft staging consistency, which can be
explained by the discrete nature of the EBM and by the simplistic biomarker trajectories,
effectively modelled as step-functions, which can result in significant staging uncertainty.
In the AD cohort we again find that the novel EBM methods show improvements
over standard methods, while there is no significant difference between the novel and
standard DEM methods. When comparing EBMs vs DEMs, we notice that the EBM
models actually perform better in terms of hard-, but worse in soft-staging consistency.
5.5. Discussion 95

This could again be due to overly simplistic EBM trajectories that might not offer a good
fit to the data.
In the diagnosis prediction tasks, most disease progression models have similar per-
formance, with only the Standard EBM having a low performance in the PCA vs AD
test. The SVM classifier has slightly worse results compared to the disease progression
models for the Controls vs PCA task, but similar results for the other tasks.

5.5.2 Model Performance on ADNI cohort

In the ADNI cohort, we notice that the extended EBM and DEM methods have similar
performance to the standard methods. There is again a drop in EBM performance on the
soft consistency metric as compared to the hard consistency. The fact that there is no im-
provement in ADNI data between the novel methods and the standard methods suggests
that the standard methods already offered a good fit on this dataset, and further that the
ADNI dataset has different characteristics compared to the DRC dataset. We attribute
this to the fact that the biomarkers present in the ADNI dataset were multimodal and
included both early-stage mollecular markers as well as late-stage cognitive tests, which
enabled even the standard models to robustly estimate the subjects’ disease stages.
The results on conversion prediction in ADNI show that all models have a broadly
similar performance at this task. However, a few clear differences can be noticed in some
models. The model with the best performance at 12-months and 24-months conversion
prediction is the DEM with standard trajectory alignment, while at 36-month conversion
the SVM and the novel EBM and DEM methods perform the best. The fact that different
models have different performance at different durations-of-conversion suggests different
models have better fits on certain time-frames of the disease time course.

5.5.3 Staging-based Metrics

The staging-based performance measures can pick up differences in the performance of


the models in both DRC and ADNI datasets, in particular between different classes
of models such as EBM vs DEM. This is most clear with the soft staging consistency,
which penalises the EBM more than the DEM. This might be the case because the DEM
constructs continuous non-parametric biomarker trajectories which might give a better
fit than the simplistic step-wise trajectories of the EBM.
The staging consistency metrics can also pick up differences across fitting procedures
within the same model, such as in the case of EBMs. On the other hand, there are
generally no statistically significant differences in the DEM between the standard versus
optimised alignment, probably due to the fact that the standard alignment is already
good enough for the two datasets that we tested them on.
There are some differences between the results on the hard- vs soft-staging consistency
metrics. In particular, the soft-staging consistency penalised the EBM models more than
the DEM models, probably due to increased staging uncertainty in the EBM models. In
terms of time-lapse, there were no significant differences between the hard and the soft
versions of this metric.
96 Chapter 5. Novel Extensions to the EBM and DEM

5.5.4 Diagnosis Prediction Metrics


We found that the performance metrics which are based on diagnosis or conversion predic-
tion are less able to discriminate between different types of models or fitting procedures.
Moreover, there was a lot of variability in the values of these performance metrics across
folds and also in the model rankings across experiments, especially in the ADNI cohort,
which made it hard to identify an overall best model. The variability of the diagnosis
prediction metrics can be attributed to inaccuracies and biases in the diagnostic labels,
and to the heterogeneity present in these diseases.

5.6 Summary
In this work we presented several extensions of the EBM and the DEM. We further devised
performance metrics that measure the accuracy of the predicted subject stages and clinical
diagnosis. We evaluated the new methodologies on data from two distinct diseases (PCA
vs tAD), and on two independent datasets (ADNI and DRC). Our results show that
in many situations the novel EBM and DEM fitting methods show improvements with
respect to our performance metrics compared to the standard versions.

5.6.1 Limitations and Future Work


The performance metrics we used for evaluation have certain inherent limitations which
might limit their use for some disease progression models. For example, the staging
consistency metrics are prone to cheating using a specially crafted model that can get
perfect consistency if it simply assigns the same stage to all subjects. However, the time-
difference metric does not suffer from this problem, due to the fact that the model needs
to predict precisely the time that passed between different visits to the clinic.
The staging consistency metrics are based on the idea that the biomarkers evolve
monotonically as the disease progresses. However, this is not the case with some neu-
rological disorders such as multiple sclerosis (MS), where a patient can have an attack
(relapse) followed by a period of steady recovery (remission). For such “non-monotonic“
diseases, other performance measures based on biomarker predictions would be more
suitable, such as the ones used in the TADPOLE Challenge [236].
One limitation of the time difference metrics is that they require the disease pro-
gression model to estimate the time from onset for every stage. This is not normally
modelled in discrete models such as the EBM or other methods based on Markov chains
(e.g. [237]). However, it might be possible to extend these discrete models in order to
estimate time since onset for each of the states.
While we have tested these models only on the DRC and ADNI datasets, their per-
formance might be different on other datasets with different types of neurodegenerative
diseases and biomarker data. Future work should include validation on other datasets,
including well-phenotyped datasets of autosomal-dominant Alzheimer’s disease or genetic
frontotemporal dementia. Moreover, the models should be tested also on other types of
biomarkers, such as non-MRI imaging biomarkers, molecular measurements from cerebro-
spinal fluid or cognitive tests.
Future work can also include performance evaluation of these models on simulated
datasets, in presence of ground truth. This will enable the detection of more subtle
5.7. Conclusion 97

differences in these methodologies, which might not be detectable in patient datasets due
to inherent measurement noise and disease heterogeneity.

5.7 Conclusion
In this chapter I presented methodological extensions in the EBM and DEM, and evalu-
ated their performance based on a set of performance measures, some of which I proposed.
Future work will focus on evaluating other types of disease progression models presented
in chapter 3, or on devising more sensitive performance metrics, for evaluation on both
simulated data as well as patient datasets.
In the next chapter, I will present DIVE: a novel disease progression model that can
estimate fine-grained spatial patterns of brain pathology, and estimate latent subject-
specific time-shifts. Such a model overcomes a some limitations of the EBM and DEM
models, which do not take spatial correlation into account and assume a pre-defined
ROI atlas. DIVE can also help us better understand underlying disease mechanisms by
studying the overlap between spatial patterns of pathology and brain connectomes.
98 Chapter 5. Novel Extensions to the EBM and DEM
Chapter 6

DIVE: A Spatiotemporal
Progression Model of Brain
Pathology in Neurodegenerative
Disorders

In this chapter I present DIVE, a novel spatiotemporal model of disease progression


that I estimates fine-grained spatial patterns of atrophy in the brain. I did the entire
work: model development, mathematical derivations (Supplementary section B), image
pre-processing and analysis and wrote the manuscripts. Arman Eshaghi showed me how
to perform the image processing with Freesurfer. Daniel Alexander and Marco Lorenzi
offered suggestions with modelling and experiment design. All collaborators gave me
feedback on the manuscripts.

6.1 Publications
• R. V. Marinescu, A. Eshaghi, M. Lorenzi, A. L. Young, N. P. Oxtoby, S. Garbarino,
T. J. Shakespeare, S. J. Crutch and D. C. Alexander, A Vertex Clustering Model
for Disease Progression: Application to Cortical Thickness Images, Information
Processing in Medical Imaging, 2017

• R. V. Marinescu, A. Eshaghi, M. Lorenzi, A. L. Young, N. P. Oxtoby, S. Garbarino,


S. J. Crutch, D. C. Alexander, DIVE: A spatiotemporal progression model of brain
pathology in neurodegenerative disorders, NeuroImage, 2019.

6.2 Introduction
Current image-based disease progression models, such as those presented in section 3.5,
estimate the evolution of the disease using a small set of biomarkers corresponding to
pre-defined regions-of-interest (ROI). This ROI parcellation is usually coarse and doesn’t
allow one to find spatially dispersed patterns of atrophy. While spatiotemporal longitudi-
nal models have already been demonstrated [238, 239, 240], these models regress against
pre-defined sets of covariates such as age, time since baseline or clinical markers. This is
problematic because, age-based alignment of subjects assumes all subjects have the same
100 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology

age of disease onset, while for time since baseline, its relationship with disease onset is un-
known. Similarly, clinical markers are noisy, biased, suffer from floor/ceiling and training
effects, are not sensitive in pre-symptomatic phases, and have low test-retest reliability
[241]. Recently, some spatiotemporal models that estimate subject-specific time-shifts
have been developed [4, 5]. However, these models generally cannot recover dispersed
and disconnected pathological patterns, because they assume voxel measurements corre-
late based on spatial distance, either through a distance function or distance from control
points. However, spatially dispersed pathological patterns have been observed in AD and
related dementias and are hypothesised to appear due to the interaction of pathology with
brain networks [38]. Discovering such fine-grained patterns could allow one to understand
underlying mechanisms of pathology propagation along these networks. However, a spa-
tiotemporal disease progression model that allows recovery of dispersed and disconnected
atrophy patterns present in AD, is not currently available.
In this work, we present DIVE: Data-driven Inference of Vertexwise Evolution. DIVE
is a novel disease progression model with single vertex resolution that makes only weak
assumptions on spatial correlation. In contrast to approaches which model temporal
trajectories for a small set of biomarker measures based on a priori defined ROIs, DIVE
models temporal trajectories for each vertex on the cortical surface. DIVE combines
unsupervised learning and disease progression modelling to identify clusters of vertices
on the cortical surface that show a similar trajectory of brain pathology over a particular
patient cohort. This formulation enables us to estimate a fine-grained spatial distribution
of pathology and also provides a novel parcellation of the brain based on temporal change.
We first test DIVE on synthetic data and show that the model can recover known
biomarker trajectories and time-shifts. We then demonstrate the model on both MRI and
PET data from two cohorts: the Alzheimer’s Disease Neuroimaging Initiative (ADNI) and
the Dementia Research Centre (DRC), UK. We use the model to reveal spatiotemporal
patterns of pathology to a much finer resolution than previous models and demonstrate
the ability to assign subjects to stages that predict progression. Finally, we validate
DIVE in terms of how robust are the estimated pathology patterns and how well the
disease progression scores correlate with cognitive tests. Code for DIVE is available
online: https://github.com/mrazvan22/dive.

6.3 Methods
In this section we describe the mathematical formulation of DIVE (section 6.3.1), then
we show how to fit the model using Expectation Maximisation (section 6.3.6) and we
describe further implementation details of the algorithm (section 6.3.7). Afterwards, we
outline the synthetic data-generation process (section 6.3.8) for testing the model in the
presence of ground truth, as well as the pipeline for pre-processing the ADNI and DRC
datasets (section 6.3.9).

6.3.1 DIVE Model


Figure 6.1 illustrates the DIVE aims and implementation. DIVE input measures are
vertexwise or voxelwise biomarker measures in the brain (Fig 6.1A), such as cortical
thickness or amyloid load. A vertex is a location on the cortical surface at which a
biomarker of pathology is quantifiable (e.g. cortical thickness). For each vertex on the
6.3. Methods 101

A Subject 1, visit 2 Subject 2, visit 2 C group vertices into


1 clusters based on
rt ex trajectory dynamics
Ve

extract vertexwise/
B voxelwise measures
Vertex 1 measure

(cortical thickness, iterate until estimate average


PET, DTI) convergence trajectory for each
cluster
E estimate disease D
Vertex measure

Vertex measure
progression scores

Disease
Progression
Score Disease Disease
Progression score Progression score

Figure 6.1: Diagram of the proposed DIVE model. DIVE assumes that biomarkers of
pathology (e.g. cortical thinning) can be measured at many vertices (i.e. locations) on the
cortical surface (A), where each vertex has a distinct trajectory of change during disease
progression (B). In (B), each individual has measurements for vertex 1 at three visits.
DIVE assigns to every cortical vertex one of a small set of temporal trajectories describing
the change in some image-based measurement (e.g. cortical thickness, amyloid PET, DTI
fractional anisotropy measures) from beginning to end of the disease progression. The
estimation process simultaneously estimates the set of clusters, the trajectory defining
each cluster, and the position of each subject along the trajectories, which are defined
on a common timeline. The process iterates assignment of each vertex to clusters (red,
green and blue in this diagram) (C), estimation of the trajectory in each cluster (D) and
estimation of the disease progression score (location along trajectory) for each subject
(E), all within an Expectation-Maximisation framework, until convergence. In particular,
(E) shows how the disease progression score, which is initially set to the individual’s age,
converges to the disease stage of the subject. Diagram made by me.
102 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology

cortical surface (or voxel in the 3D brain volume), we estimate a unique trajectory along
the disease progression timeline (Fig 6.1B), while also estimating subject/visit-specific
disease progression scores (i.e. disease stages). We do that by grouping vertices with
similar biomarker trajectories into clusters (Fig 6.1C), and we estimate a representative
trajectory for every cluster (Fig 6.1D). Each trajectory is a function of subject-/visit-
specific disease progression scores (DPS) (Fig 6.1E). The DPS depends linearly on the
time since baseline visit, but with subject-specific slope and intercept.

6.3.2 Modelling Subject-specific Parameters


The disease progression score sij for subject i at visit j is a latent variable denoting the
current disease stage of the subject at this visit. It defined as a linear transformation of
time since baseline measurement tij (in years):

sij = αi tij + βi (6.1)


where αi and βi represent the speed of progression and time shift (i.e. disease onset) of
subject i respectively.

6.3.3 Modelling Biomarker Trajectory for a Single Vertex


DIVE assumes that the biomarker measure at each vertex on the cortical surface follows
a sigmoidal trajectory f (s; θ) over the disease progression score s and with parameters θ.
We choose a parametric sigmoid function because it is a parsimonious parametric model
that offers better fit compared to linear models, is monotonic, and can account for floor
and ceiling effects [194, 175]. We also assume that vertices are grouped into K clusters
and we model a unique trajectory for each cluster k ∈ 1, ..., K, which will be referred to
as cluster trajectories. The sigmoidal function f (s; θk ) for cluster k is defined as:
ak
f (s; θk ) = + dk (6.2)
1 + exp(−bk (s − ck ))
where s is the disease progression score from Eq. 6.1 and θk = [ak , bk , ck , dk ] are parame-
ters controlling the shape of the trajectory – dk and dk + ak represent the lower and upper
limits of the sigmoidal function, ck represents the inflection point and ak bk /4 represents
the slope at the inflection point.
For a given subject i at visit j, the value Vlij of its biomarker measurement at vertex
l is a random variable that has an associated discrete latent variable Zl ∈ [1, ..., K]
denoting the cluster it was generated from. The value of Vlij given that it was generated
from cluster Zl can be modelled as:

p(Vlij |αi , βi , θZl , σZl , Zl ) = N (Vlij |f (αi tij + βi |θZl ), σZl ) (6.3)
where N (Vlij |f (αi tij +βi |θZl ), σZl ) represents the probability density function (pdf) of the
normal distribution that models the measurement noise along the sigmoidal trajectory
of cluster Zl , having variance σZl . Next, we assume the measurements from different
subjects are independent, while the measurements from the same subject i at different
visits j are linked using the disease progression score from equation 6.1. Moreover, we
also assume a uniform prior on Zl . This gives the following model:
6.3. Methods 103

Y
p(Vl , Zl |α, β, θ, σ) = N (Vlij |f (αi tij + βi |θZl ), σZl ) (6.4)
(i,j)∈I

where I = (i, j) represents the set of all the subjects i and their corresponding visits j.
Furthermore, Vl = [Vlij |(i, j) ∈ I] is the 1D array of all the values for vertex l across every
subject and corresponding visit. Vectors α = [α1 , . . . , αS ] and β = [β1 , . . . , βS ], where
S is the number of subjects, denote the stacked parameters for the subject shifts. If a
subject i has multiple visits, these visits share the same parameters αi and βi . Vectors
θ = [θ1 , . . . , θK ] and σ = [σ1 , . . . , σK ], with K being the number of clusters, represent the
stacked parameters for the sigmoidal trajectories and measurement noise specific to each
cluster.
Due to our main motivation of modelling population trajectories and in order to
ensure robustness and identifiability, we did not add random effects to the trajectories of
specific subjects.

6.3.4 Modelling Biomarker Trajectories for all Vertices


So far we have a model for only one vertex on the brain surface. We therefore extend the
formulation to all the vertices by assuming all these vertex measurements are spatially
independent, giving the complete data likelihood:
L Y
Y
p(V, Z|α, β, θ, σ) = N (Vlij |f (αi tij + βi |θZl ), σZl ) (6.5)
l (i,j)∈I

where V = [V1 , . . . , VL ], Z = [Z1 , . . . , ZL ], L being the total number of vertices on the


cortical surface. The formulation so far assumes spatial independence between measure-
ments in different vertices, but in section 6.3.5 the model is extended to capture spatial
correlations. The full joint distribution is given by:

p(V, Z, α, β, θ, σ) = p(V, Z|α, β, θ, σ)p(α, β, θ, σ) (6.6)


where p(α, β, θ, σ) is an informative prior on the model parameters defined as follows:
Y
p(V, Z, α, β, θ, σ) = p(αi )p(βi )
i
(6.7)
p(αi ) ∼ Γ(αshape , αrate )
p(βi ) ∼ N (βmean , βstd )
where αshape , αrate , βmean , βstd are a-priori defined hyperparameters. The informative
priors on the subject-specific parameters help ensure model identifiability, as the model
otherwise has two extra degrees of freedom. Such informative priors on αi and βi also
help deal with singularities in the objective functions of αi and βi when the biomarker
trajectories are flat.
We get the final model log likelihood for incomplete data by marginalising over the
latent variables Z:
L X
Y K Y
p(V |α, β, θ, σ) = p(Zl = k) N (Vlij |f (αi tij + βi |θk ), σk ) (6.8)
l=1 k=1 (i,j)∈I

Throughout the article, we will use the shorthand zlk = p(Zl = k).
104 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology

6.3.5 Modelling Spatial Correlation


The version of the model presented so far assumes spatial independence between vertex
measurements. However, the regional organisation of the cortex suggests we would expect
spatial correlation1 of the vertex measurements. More precisely, measures of cortical
thickness or other modalities are often similar in neighbouring vertices on the cortical
surface and likely belong to the same cluster. DIVE can be easily extended to include
mild spatial constraints on the correlation of vertex measurements via a Markov Random
Field (MRF), which encourages neighbouring vertices to have the same corresponding
cluster. We hypothesise that incorporating such constraints should reduce the effects of
noise and produce a more stable clustering. However, this does not model correlation
between the actual vertex values, but only between the latent variables Zl , i.e. the
cluster membership of each vertex. The MRF thus has the advantage of not requiring
the use of huge covariance matrices, which are otherwise needed if we want to model
correlation of vertex values directly. Moreover, in contrast to previous methods that use
correlation based on spatial distance [4, 5], we use neighbourhood correlations, which
allow us to estimate fine-grained spatial patterns of pathology. With the MRF, the
full-data likelihood function of the model now becomes:

 
L
Y Y Y
p(V, Z|α, β, θ, σ, λ) =  N (Vlij |f (αi tij + βi |θZl ), σZl ) Ψ(Zl , Zl2 ) (6.9)
l (i,j)∈I l2 ∈Nl

where Ψ(Zl , Zl2 ) is a clique term representing the likelihood of a neighbouring vertex
l2 to have similar label with vertex l. The formula for the clique term is:
(
exp(g(λ)) if k = k2
Ψ(Zl = k, Zl2 = k2 ) = (6.10)
exp(−h(λ)) otherwise
where λ is a parameter controlling how much to penalise neighbouring vertices that
belong to distinct clusters, and g and h are positive, monotonic functions over the λ > 0
range. We choose g(λ) = λ and h(λ) = λ2 , which results in a concave objective function
for λ, ensuring that it can later be optimised (see M-step).
Therefore, the model parameters that need to be estimated are M = [α, β, θ, σ, λ]
where α and β are the subject specific shifting parameters, θ and σ are the cluster specific
trajectory and noise parameters and λ is the clique parameter denoting the penalisation
of spatially non-smooth assignments of latent variables Z.

6.3.6 Fitting the Model using Generalised


Expectation-Maximisation
We choose to fit our model using Expectation-Maximisation (EM), because it offers a
fast convergence given the large number of parameters that need to be estimated and the
huge dimensionality of relevant datasets (e.g. 1973 subjects x 163,842 vertices in ADNI).
In the next two sections we outline the E-step and M-step. While both of these steps
have no closed-form solution, we will solve them using numerical optimisation, which
only results in an increase in the objective function at each iteration. However, the EM
1
By correlation here we mean that these vertex measurements are not statistically independent
6.3. Methods 105

algorithm is still guaranteed to converge, and this approach is called Generalised EM


[195].
Algorithm 6.2 shows the model fitting procedure using the EM algorithm. The pro-
cedure first initialises (line 1) some parameters required to start the EM algorithm: the
subject parameters α and β and the latent parameters zlk which represent the assignment
of vertices to clusters. In the M-step, the method updates the trajectories of each cluster
(lines 4-6), the subjects-specific parameters (line 9) and the clique penalty term λ (line
17). In the E-step, the method computes zlk (line 18) using previously defined functions
that compute zlk given a fixed λ (line 14).

6.3.6.1 E-step
In the Expectation step, at iteration u we seek an estimate of p(Z|V, M (u−1) ), given the
(u−1) (u−1) (u−1) (u−1) (u−1)
current estimates of the parameters M (u−1) = [θk , σk , αi , βi , λi ]. We
perform this using Iterated Conditional Modes [195], which performs coordinate-wise
gradient ascent. This works by conditioning the clique terms Z on the values of Z from
the previous iterations. This approximation gives the following factorisable likelihood:
L
Y h i
(u−1) (u−1)
p(Z|V, M )≈ EZ (u−1) |V ,M p(Zl |Vl , M, ZNl ) (6.11)
Nl l
l

The factorised form allows for tractable computation and memory storage of p(Z).
Let zlk (u) = p(Zl = k|Vl , M (u−1) , Z (u−1) ). After simplifications we reach the following
update rule:

" #
X h i
(u) (u−1)
log zlk ∝ Dlk + log exp(−λ2 ) + zl2 k (exp(λ) − exp(−λ2 )) (6.12)
l2 ∈Nl

where the data-fit term Dlk has the following form:

log (2πσk2 )|I| X 1


Dlk = − − 2
(Vlij − f (αi tij + βi |θk ))2 (6.13)
2 i,j∈I
2σk

The full derivation is given in Supplementary section B.3. In order to enable opti-
misation over λ, a final modification of this step is performed, by considering zlk to be
functions ζlk (λ) over λ. This results in the update equation from Alg. 6.2, line 18 which
is based on pre-defined terms on lines 13-14.

6.3.6.2 M-step
In the Maximisation step we try to estimate the model parameters M = (α, β, θ, σ, λ)
that maximise EZ|V,M (u−1) [log p(V, Z|M )]. We cannot simultaneously optimise all 5 sets
of parameters, so we optimise them independently. In order to get the update rule for the
trajectory parameters θk corresponding to cluster k we need to maximise the expected log
likelihood with respect to θk . The key observation here is that if we assume fixed α, β and
Z, then the trajectory parameters θk for every cluster k are conditionally independent, i.e.
θk ⊥⊥ θm |(Z, α, β, σ) ∀ (k, m), k 6= m. This allows us to maximise every θk independently
using the following equation:
106 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology

(0)
1 Initialise α(0) , β (0) , zlk
2 while θ, σ, α, β or zlk not converged do
; // M-step 1: For each cluster, optimise its trajectory
3 for k = 1 to K do
(u) (u−1) P (u−1) (u−1)
θk = arg minθk Ll=1 zlk ij
|θk ))2 − log p(θk )
P
4 (i,j)∈I (Vl − f (αi tij + βi
(u) (u)
5 θ = make identifiable(θk )
k 2
(u) 1
PL (u−1) P ij (u−1) (u−1) (u) 2
6 σk = |I| l=1 zlk (i,j)∈I (Vl − f (αi tij + βi |θk )) − log p(σk )
7 end
; // M-step 2: For each subject, optimise its time shift αi and progression
speed βi
8 for i = 1 to S do " #
(u−1)
(u) (u) PL PK z ij (u)
βi |θk ))2
P
9 αi , βi = arg minαi ,βi l=1 k=1
lk 2
(u) j∈Ii (Vl − f (αi tij + − log p(αi , βi )
2 σk

10 end
; // E-step 1: Define functions ζlk (λ) computing zlk , the probability of
vertex l being assigned to cluster k, given fixed λ
11 for l = 1 to L do
12 for k = 1 to K do
; // Pre-compute data fit terms Dlk
 2
(u) ij (u) (u) (u) 2
Dlk = − 21 log (2π σk 1
P
13 )|I|−  (u) 2
i,j∈I (Vl −f (αi tij +βi |θk ))
2 σk

14 ζlk (λ)≈ h i
(u−1)
exp(−λ2 ) + zl2 k (exp(λ) − exp(−λ2 ))
P
exp Dlk + l2 ∈Nl log
15 end
16 end
; // M-step 3: optimise clique term λ using above definitions in E-step 1
17 λ(u) =
arg maxλ Ll=1 K
P P  P 2
P 
k=1 ζlk (λ) Dlk + λ l2 ∈Nl ζl2 k (λ) − λ l2 ∈Nl (1 − ζl2 k (λ))
; // E-step 2: Compute next zlk using the best λ
(u)
18 zlk = ζlk (λ(u) )
19 end
(u) (u)
(u) αi (u) βi −µN
20 αi = σN
, βi = σN
; // Re-scale subject shifts

Figure 6.2: The DIVE parameter estimation algorithm. The algorithm, based on
Expectation-Maximisation, iteratively optimises the assignment of vertices to clusters
(E-step) and the parameters for the biomarker trajectories and subject time-shifts (M-
step).
6.3. Methods 107

 
K
X YL Y
θk = arg max p(Z|V, M (u−1) ) log  N (Vlij |f (αi tij + βi |θzl ), σzl ) + log p(θk )
θk z1 ,...,zL l=1 (i,j)∈I
(6.14)
A similar observation of conditional independence can also be observed for the latent
variables Z. This allows us to decompose the joint distribution over Z, and after ex-
panding the noise model we reach the optimisation problem from Alg. 6.2, line 4. See
Supplementary section B.3 for full derivation. This does not have a closed-form solution,
so we use numerical optimisation for finding θk that maximises the equation from Alg.
6.2, line 4.
A similar equation, yet in closed form, is also obtained for σk (Alg. 6.2, line 6). After
estimating θ and σ for every cluster, we use the new values to estimate the subject specific
parameters α and β. For every subject i, we maximise the expected log likelihood with
respect to αi , βi independently, and after simplifications we obtain the update rule from
Alg. 6.2, line 9, which is again solved using numerical optimisation. For the numerical
optimisation of θ we used the Nelder-Mead method for its robustness, while for α and
β we used the second-order Broyden–Fletcher–Goldfarb–Shanno algorithm due to fast
convergence.
The large dimensionality of the dataset (around 163,428 vertices x 400 subjects x 4
timepoints each) makes model fitting extremely difficult from a computational perspec-
tive. Initial optimisation on a smaller subset of around 100 ADNI subjects took around
30h. However, we achieved a significant speed-up in the evaluation of objective functions
by computing a zlk -weighted average of vertex measurements within each cluster (see
Appendix section B.4). This resulted in a final convergence time of around 4-6h depend-
ing on the size of the dataset, using an Intel Xeon E3-1271 @ 3.60GHz CPU. Regarding
memory requirements, loading into memory around 1600 and fitting the model required
around 12GB of RAM. However, we dropped it down by a factor of x4 by using small
16-bit floating representations for the vertexwise biomarkers.
For optimising λ, we again try to optimise in the M-step the expected full data
likelihood under the Z estimates from the previous iteration:

λ(u) = arg max Ep(Z|V,M (u−1) ,λ,Z (u−1) ) [log p(V, Z|M (u−1) )] (6.15)
λ

We simplify the above equation by expanding the likelihood model and approximating
the joint over Z with the product of the marginals zlk over all vertices l. This results in
the update equation from Alg. 6.2 line 17 – see appendix for full derivation. In this final
equation we also replaced zlk with a function ζlk (λ) over λ, which updates zlk based on
the current value of λ being evaluated. This is done to increase convergence, as latent
variables zlk are highly coupled with the value of λ being evaluated.

6.3.7 Implementation Details


6.3.7.1 Parameter Initialisation and Priors
Before starting the fitting process, we need to initialise α, β and the clustering probabil-
ities zlk (Alg. 6.2, line 1). We set αi and βi to be 1 and 0 respectively for each subject,
which sets the initial disease progression score to the time since baseline of the subject
108 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology

at the clinical visit. We initialise zlk using k-means clustering of the vectors Vl . We
also initialise hyperparameters αshape = 16e4, αrate = 16e4, βmean = 0 βstd = 0.1, which
work well in practice as they result in realistic ranges for αi and βi of around [0.3, 3]
and [-15,15] respectively. The reason why we need to give such large numbers of 16e4 is
because there are many vertex measurements (¿100,000) that each drag the subject to
an extremity if most values are above/below the population curve. This can be avoided
in the future by adding subject-specific random effects to the population trajectory.
As already explained in [2], the sigmoid parameters θk are not identifiable because
f (t; ak , bk , ck , dk ) = f (t; −ak , −bk , ck , ak + dk ). We thus need to apply the following trans-
(u) (u) (u) (u) (u) (u) (u) (u)
formation on line 5 of Alg. 6.2: if bk < 0 then ak = −ak ; bk = −bk ; dk = dk −ak .
This ensures model identifiability and is performed at every iteration.

6.3.7.2 Estimating the Optimal Number of Clusters


The EM procedure needs to specify a-priori the number of clusters to fit on the data. We
optimise the number of clusters K using Akaike Information Criterion (AIC), which we
found to better agree with ground truth in simulations than other information criteria
such as the Bayesian Information Criteria (BIC). The number of parameters of the fitted
model is 5K+2S+1, where S is the number of subjects. Note that zlk are not included
as parameters of the model because they are latent variables that are marginalised (see
Eq. 6.8). We repeat the fitting procedure for each K from 2 to 100 clusters and select
the K that minimises the AIC.

6.3.8 Simulation Experiments


6.3.8.1 Motivation
Initial assessment of DIVE performance uses synthetic data, where we know the ground
truth. The aim is to explore how accurately we can recover ground truth parameters as
the problem becomes harder in three different scenarios:

• Scenario 1: as the number of clusters increases, evaluate how well DIVE can esti-
mate the correct number of clusters using AIC and BIC

• Scenario 2: as the trajectories become more similar, test how well we can recover
the assignment of vertices to clusters and the DIVE parameters

• Scenario 3: same as Scenario 2, but for decreasing number of subjects

6.3.8.2 Synthetic Data Generation


We first designed a basic simulation, which the model should be able to fit well since the
trajectories were designed to be well separated and enough subject data was generated
along the disease time course. The data in the basic simulation was generated as follows:

1. Sampled baseline age ai1 and shift parameters αi , βi for 300 subjects with 4 time-
points (each timepoint 1 year apart), with ai1 ∼ U (40, 80), αi ∼ Γ(6.25, 6.25),
βi ∼ N (0, 10). Time since baseline has been obtained for every visit j of subject i
as follows: tij = aij − ai1 .
6.3. Methods 109

2. Generated three sigmoids with different (slope, centre) parameters: [(-0.1, -15), (-
0.1, 2.5), (-0.1, 20)] (Fig. 6.3a, red lines). Upper and lower limits have been set to
1 and 0 respectively.

3. randomly assign every vertex l ∈ {1, . . . , L}, where L = 1000, to a cluster a[l] ∈
{1, 2, 3}

4. Sampled a set of L perturbed trajectories θl from each of the original trajecto-


ries, one for each vertex (Fig. 6.3a, gray lines) using covariance matrix Cθ =
diag([0, 2bk /15, 11.6, 0]).

5. Sampled subject data for every vertex l from its corresponding perturbed trajectory
θl with noise standard deviation σl = 1

From the basic simulation, we generated synthetic data for each of the three scenarios
by varying one parameter at a time and kept the other parameters constant, having the
same values as in the basic simulation. We varied the following parameters:

• Scenario 1: number of clusters - 2, 3, 5, 10, 15, 20, 30 and 40. The cluster centres
were spread evenly across a fixed total DPS range where the data was available.

• Scenario 2: distance between trajectory centres (as proportion of total DPS range
sampled) – 0.33, 0.30, 0.23, 0.17, 0.10, 0.07, 0.03 and 0.02

• Scenario 3: number of subjects - 300, 200, 100, 50, 35, 20, 10 and 5

6.3.8.3 Model Fitting and Evaluation


Since there was no spatial information in the data generation procedure, we used DIVE
without the MRF extension. For Scenario 1, we estimated using AIC and BIC the
optimal number of clusters. For Scenarios 2 and 3, after fitting the parameters of DIVE,
we calculated the agreement between the final clustering probabilities p(Zl ) and the true
clustering assignments a[l]. This
PL agreement, which we will call the clustering agreement,
is defined as ℵ = maxτ (1/L) l=1 p(Zl = τ (a[l])), where τ is any permutation of cluster
labels. We also computed the error in the DPS estimation (sum of squared differences,
SSD) and trajectory estimation (SSD between predicted trajectory and true trajectory
at DPS points of every subject visit).

6.3.9 Data Acquisition and Pre-processing


Data used in this work were obtained from the Alzheimer’s Disease Neuroimaging Initia-
tive (ADNI) database (adni.loni.usc.edu) and from the Dementia Research Centre,
UK. For ADNI, we downloaded all T1 MR images that have undergone gradient warp-
ing, intensity correction, and scaling for gradient drift. We included subjects that had at
least 3 scans, to ensure we get a robust estimate of the subject specific parameters. This
resulted in 138 healthy controls, 235 subjects with mild cognitive impairment (MCI) and
81 subjects with Alzheimer’s disease.
We also downloaded all AV45 PET images from ADNI that were fully pre-processed,
having the following tag: Co-reg, Avg, Std Img and Vox Siz, Uniform Resolution. This
110 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology

Number of Age at baseline


Cohort Diagnosis Number of Subjects
Scans (years)
ADNI Controls 138 4.3 76.3
MRI MCI 235 4.6 74.8
AD 81 3.5 75.8
DRC Controls 31 5.0 66.3
tAD AD 24 5.4 71.2
DRC Controls 31 5.0 66.3
PCA PCA 32 4.1 62.6
Controls 141 2.4 85.5
ADNI SMC 27 2.0 86.1
PET EMCI 149 2.4 85.6
LMCI 104 2.4 86.0
AD 12 2.0 87.3

Table 6.1: Demographics of the four cohorts used in our analysis. ADNI MRI and the
DRC cohorts were used for the cortical thickness analysis, while ADNI PET was used for
the PET AV45 analysis. MCI – mild cognitive impairment, SMC - subjective memory
complaints, EMCI – early MCI, LMCI – late MCI.

meant that the images were co-registered, averaged across the 6 five-minute frames, stan-
dardised with respect to the orientation and voxel size and smoothed to produce a uniform
resolution of 8mm full-width/half-max (FWHM).
The DRC dataset consisted of T1 MRI scans from 31 healthy controls, 32 PCA and
23 typical AD subjects with at least 3 scans each and an average of 5.26 scans per
subject. All PCA patients fulfilled both Tang-Wai [147] and Mendez [146] criteria based
on clinical review. The typical AD patients all met the criteria for probable Alzheimer’s
disease [115, 116].
Given that the ADNI and DRC datasets contained subjects with different modalities
or diseases, we ran DIVE independently on the following four cohorts (see Table 6.1 for
demographics):

1. ADNI MRI: controls, MCI and tAD subjects from ADNI (cortical thickness data)

2. DRC tAD: tAD subjects and controls from the DRC dataset (cortical thickness
data)

3. DRC PCA: PCA subjects and controls from the DRC dataset (cortical thickness
data)

4. ADNI PET: AV45 scans from ADNI containing subjects with following diagnoses:
healthy controls, subjective memory complaints, early MCI, late MCI and Alzheimer’s
disease.

6.3.9.1 MRI Preprocessing


On both datasets, in order to extract reliable cortical thickness measures, we ran the
Freesurfer longitudinal pipeline [118], which first registers the MR scans to an unbiased
within-subject template space using inverse-consistent registration. The longitudinally
6.4. Results 111

registered images were then registered to the average Freesurfer template. No further
smoothing was performed on these images (FWHM level of zero mm). From these
template-registered volumetric images, cortical thickness measurements were computed
at each vertex (i.e. point) on an average 2D cortical surface manifold. For each vertex we
averaged the thickness levels from both hemispheres in order to later ease visualisation
and to obtain a smaller representation of the input data. Each of the final images had a
resolution of 163,842 vertices on the cortical surface.
Finally, we standardised the data by computing Z-scores for each vertex with respect
to the values of that vertex in the control population. This normalisation step ensures that
the model will not be affected by different thicknesses of the cortex at various locations
on the cortical surface. This step is specific for MRI cortical thickness data, and might
not be necessary for other modalities (e.g. PET).

6.3.9.2 PET Preprocessing


We computed amyloid standardised uptake value ratio (SUVR) levels using the PetSurfer
pipeline [242, 243], which is available with Freesurfer version 6. The PetSurfer pipeline
first registers the PET image with the corresponding MRI scan, then applies Partial
Volume Correction, and finally resamples the voxelwise SUVR values onto the cortical
surface. While the final images also had a resolution of 163,842 vertices, the PET data
we obtained from ADNI was inherently more smooth than the MRI cortical thickness
data (8mm FWHM). We did not standardise the SUVR values like we did for cortical
thickness, due to the fact that we did not observe different uptake based on anatomy
within the control population.

6.3.9.3 The MRF Neighbourhood Graph


We estimated the MRF neighbourhood graph based on a Freesurfer triangular mesh
for the fsaverage template. Each vertex was a triangle on the brain surface estimated
with Freesurfer, and we connected the vertices if the corresponding triangles had a shared
edge. For the MRF neighbourhood graph, we used a 3rd degree neighbourhood structure,
meaning that two vertices were considered neighbours if the shortest path between them
was not higher than 3.

6.4 Results
6.4.1 Results on Synthetic Data
In the basic simulation, we obtained a clustering agreement ℵ of 0.97, which suggests
that almost all vertices were assigned to the correct cluster. Fig. 6.3a shows the original
trajectories and the recovered trajectories using our model, plotted against the disease
progression score on the x-axis and the vertex value on the y-axis. In Fig. 6.3b we plotted
the recovered DPS of each subject along with the true DPS. The results for the three
scenarios are shown in Figs. 6.3c-6.3e. In Fig. 6.3c, we show for Scenario 1 the estimated
number of clusters against the true number of clusters using both AIC and BIC criteria.
In Figs. 6.3d-6.3e we show the distributions for ℵ in Scenarios 2 and 3 as the problem
becomes harder in each successive step.
112 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology

(a) (b)

(c) (d) (e)

Figure 6.3: (a-b) Results for the basic simulation, where trajectories are relatively well
separated. (a) Reconstructed temporal trajectories (blue) plotted against the true tra-
jectories (red). The x-axis shows the disease progression score (DPS), while the y-axis
shows the biomarker values of the vertices. (b) Estimated subject-specific DPS scores
compared to the true scores. (C-E) Simulation results for the three scenarios: (c) increas-
ing number of clusters, (d) trajectories becoming similar and (e) decreasing number of
subjects. On the x-axis we show the variable that was changing within the scenario (e.g.
number of clusters), while on the y-axis we show the agreement measure ℵ, representing
the percentage of vertices that were assigned to the correct cluster.

The results show that, in a simple experiment where the trajectories are well sepa-
rated, DIVE can very accurately estimate which clusters generated each vertex. More-
over, the recovered trajectories and DPS scores are close to the true values. The results
of Scenario 1 also suggest that both AIC and BIC are effective at estimating the correct
number of known clusters, with AIC having slightly better performance than BIC for
larger numbers of clusters. On the other hand, the results of the stress test scenarios 2
and 3 show that performance measure ℵ drops when the trajectories become very sim-
ilar with each other or when the number of subjects decreases. This happens because
small differences in trajectories are hard to detect in the presence of measurement noise,
while a small number of subjects doesn’t provide enough data to accurately estimate the
parameters. Similar decreases in performance for scenarios 2 and 3 are observed also for
other measures, such as the error in recovered trajectories or DPS scores (Supplementary
Fig B.1).
6.4. Results 113

6.4.2 Results with ADNI and DRC Datasets


6.4.2.1 Initial Hypotheses
Using ADNI and DRC datasets, we aim to recover the spatial distribution of cortical
atrophy and amyloid pathology, as well as the rate and timing of these pathological
processes. In particular, we hypothesise that these spatial patterns of pathology and
their evolution will be:

• similar on two independent typical AD datasets: ADNI and DRC

• different on distinct diseases: tAD vs PCA

• different in distinct modalities: cortical thickness from MRI vs amyloid load from
AV45 PET.

6.4.2.2 Results
The optimal number of clusters, as estimated with AIC, was three for the ADNI MRI
dataset, three for the DRC tAD dataset, five for the DRC PCA dataset and eighteen for
the ADNI PET dataset. Fig. 6.4a (left) shows the results from the ADNI MRI dataset,
where in the left image we coloured the vertices on the cortical surface according to the
cluster they most likely belong to. We assigned a colour for each cluster (both the brain
figures on the left and the trajectory figures on the right) according to the extent of
pathology of its corresponding trajectory at a DPS score of 1. The cluster colours range
from red (severe pathology) to blue (moderate pathology). In Fig. 6.4a (right), we show
the resulting cluster trajectories with samples from the posterior distribution of each θk .
Similar results are shown for the other three datasets: the DRC tAD dataset (Fig. 6.4b),
DRC PCA dataset (Fig. 6.4c) and the ADNI PET dataset (Fig. 6.4d).
We notice that in tAD subjects using the ADNI datasets (Fig. 6.4a), there is more
severe cortical thinning mainly in the inferior temporal lobe (red cluster), with disperse
atrophy also in parietal and frontal regions (green cluster), with relative sparing of the
inferior frontal and occipital lobes. In tAD subjects from the DRC dataset (Fig. 6.4b),
we see a relatively similar pattern, however with more pronounced atrophy in the supra-
marginal cortex (red cluster) compared to ADNI. This could be due to the younger ages
of controls and tAD subjects in the DRC dataset as compared to ADNI. The spatial
distribution of cortical thinning found with DIVE resembles results from previous longi-
tudinal studies such as [244, 245]. However, in contrast to these approaches, our model
gives insight into the timing and rate of atrophy and is also able to stage subjects across
the disease time course. We also find that the cluster trajectories in the DRC tAD dataset
have similar dynamics to the ADNI MRI dataset, although they show a clearer separation
between each other.
In the PCA subjects (Fig. 6.4c), we find that atrophy is mainly focused on the
posterior part of the brain, with limited spread in the motor cortex, anterior temporal
and frontal areas. This posterior-focused pattern of atrophy is different from the one found
in the tAD datasets, and agrees with previous findings in the literature [18, 20]. However,
as opposed to the results from [20] which showed posterior regions uniformly affected,
we notice that there are two clusters within the posterior region with different pathology
dynamics, with the superior parietal and supramarginal areas affected more that the
remaining posterior regions. This might be attributable to DIVE’s ability to model
114 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology

subjects’ disease onset and progression speed, along with non-linear cortical thinning
dynamics, other differences due to the different subjects analysed, and the merging of left
and right hemispheres could also give such differences.
In ADNI PET (Fig. 6.4d) we see that the regions with the highest amyloid uptake
are more spatially continuous, comprising the precuneus and anterior frontal areas. On
the other hand, the anterior-superior temporal gyrus shows the least uptake of amyloid.
This result closely matches the result by [4], which used a completely different dataset
and modelling technique. These results using AV45 PET are also noticeably different
from results using cortical thickness (e.g. Fig. 6.4a), which have more high-frequency
patterns and only give 3-5 optimal clusters instead of 20. The layers of clusters starting
from the precuneus and frontal lobes, which range from severe to less severe atrophy,
suggest a continuum of variation in vertex trajectories in the case of the PET dataset
(Fig 6.4d, right). These trajectories all start with a low amyloid SUVR, between 0 and
0.25, but in late stages the trajectories for some clusters such as cluster 0 can reach an
SUVR of 1.5. The reason for seeing this continuum might be because the PET images
have a much lower resolution than MR images and were smoothed by ADNI during the
pre-processing steps.

6.4.3 Model Evaluation


6.4.3.1 Motivation
We further tested the robustness and validity of the model as follows:

• Robustness in parameter estimation: test whether similar spatial clustering is esti-


mated for different subsets of the data

• Clinical validity of DPS scores: test whether the subject disease progression scores,
based purely on MRI or PET data, correlate with cognitive tests such as Clinical
Dementia Rating Scale - Sum of Boxes (CDRSOB), Alzheimer’s Disease Assessment
Scale - Cognitive (ADAS-COG), Mini-Mental State Examination (MMSE) and Rey
Auditory and Verbal Learning Test (RAVLT).

• Comparison with other models: to evaluate the benefit of estimating fine-grained


patterns of pathology in DIVE, as well as latent time shifting of subjects, we com-
pared the performance of DIVE with a region-of-interest based method [2] and a
no-staging method that doesn’t estimate subject time shifts. See Supplementary
Section B.2 for precise specifications.

6.4.3.2 Evaluation Procedure


For all scenarios, we ran 10-fold cross-validation (CV) on the ADNI MRI dataset. At
each fold we fit the model using 3 clusters, since this was the optimal number of clusters
found previously on the entire dataset. The trained model was then used to estimate the
DPS of the test subjects.
For the performance comparison of DIVE with other models, we compute two per-
formance metrics: (1) between-subject correlation of the models’ estimated DPS values
with cognitive tests; we estimated a unique DPS for every subject and every visit, which
we then matched with the corresponding cognitive tests at that subject’s visit and (2)
prediction root mean squared error (RMSE) between the predicted vertex-wise values
6.4. Results 115

severe pathology moderate pathology

(a) ADNI MRI

(b) DRC tAD

(c) DRC PCA

(d) ADNI PET

Figure 6.4: (left column) DIVE estimated clusters (left column) and corresponding disease
progression trajectories (right column) on four datasets: (a) ADNI MRI (b) DRC tAD
(c) DRC PCA and (d) ADNI PET. We coloured each cluster according to the extent of
pathology (cortical thickness or amyloid uptake) at DPS=1.
116 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology

f=1 f=2 f=3 f=4 f=5

f=6 f=7 f=8 f=9 f = 10

f=1 f=2 f=3 f=4 f=5

f=6 f=7 f=8 f=9 f=10

Figure 6.5: (top) Clusters estimated from 10-fold cross-validation training sets on the
ADNI MRI dataset. (bottom) Estimated trajectories for each fold.

and actual measurements, averaged over all subjects and all locations on the brain; to
evaluate these predictions, for every subject we use the first n-1 scans for training and
the last scan for testing the prediction.

6.4.3.3 Evaluation Results

Fig. 6.5 shows the brain clusters and corresponding trajectories, estimated for all the
cross-validation folds after fitting the model on the training data. The clusters have been
coloured using a similar colour scheme as in Fig. 6.4. In Fig 6.6 we show scatter plots
of the DPS scores with clinical measures such as CDRSOB, ADAS-COG, MMSE and
RAVLT.
6.4. Results 117

CDRSOB ADAS-COG MMSE RAVLT


(ρ = 0.37, (ρ = 0.37, (ρ = −0.36, (ρ = −0.32,
p < 1e − 65) p < 1e − 64) p < 1e − 63) p < 1e − 49)

Figure 6.6: Scatter plots of the DPS scores estimated from the ADNI MRI dataset,
plotted against four cognitive tests: CDRSOB, ADAS-COG, MMSE and RAVLT. For
each cognitive test we also report the Pearson correlation coefficient and p-value. The
disease progression scores, computed only based on MRI cortical thickness data, correlate
with these cognitive measures, suggesting that the DPS scores are clinically meaningful.

Model CDRSOB (ρ) ADAS13 (ρ) MMSE (ρ) RAVLT (ρ) Prediction (RMSE)
DIVE 0.37 ± 0.09 0.37 ± 0.10 0.36 ± 0.11 0.32 ± 0.12 1.021 ± 0.008
ROI-based model 0.36 ± 0.10 0.35 ± 0.11 0.34 ± 0.13 0.30 ± 0.13 1.019 ± 0.010
No-staging model *0.09 ± 0.06 *0.03 ± 0.09 *0.05 ± 0.06 *0.02 ± 0.06 *1.062 ± 0.024

Table 6.2: Performance evaluation of DIVE and two simplified models on the ADNI MRI
dataset using 10-fold cross-validation. In the middle four columns, we show between-
subject correlations between the DPS scores and several cognitive tests: CDRSOB,
ADAS-Cog13, MMSE and RAVLT. The last column shows the prediction error (RMSE)
of cortical thickness values from follow-up scans. (*) Statistically significant differences
between the model and DIVE, Bonferroni corrected for multiple comparisons.

The results in Fig. 6.5 demonstrate that DIVE is robust in cross-validation, as the
estimated clusters and trajectory parameters are all similar across folds. The average
Dice score overlap across the 10-folds range were 0.77, 0.76 and 0.90 for clusters 0, 1 and
2 respectively. The DIVE-derived DPS scores, which were estimated purely based on
MRI data, are also clinically relevant as they correlate with cognitive tests (Fig. 6.6).

The performance of DIVE in terms of subject staging and biomarker prediction also
compares favourably with simpler no-staging and ROI-based models (Table 6.2). Results
show that DIVE has comparable performance to the ROI-based model, both in terms
of subject staging and cortical thickness prediction. The fact that DIVE has similar
performance to a simpler model which has less parameters is evidence that the estimated
patterns are meaningful. Moreover, DIVE offers qualitative insight into the fine-grained
spatial patterns of pathology and their temporal progression. Furthermore, the No-
staging model performs significantly worse than DIVE, both in terms of subject staging
and for biomarker prediction. This suggests that, when modelling progression of AD, it
is important to account for the fact that patients are at different stages along the disease
time-course.
118 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology

6.5 Discussion
6.5.1 Summary and Key Findings
We presented DIVE, a spatiotemporal model of disease progression that clusters vertex-
or voxel-wise measures of pathology in the brain based on similar temporal dynamics.
The model highlights, for the first time, groups of cortical vertices that exhibit a similar
temporal trajectory over the population. The model also estimates the temporal shift
and progression speed for every subject. We applied the model on cortical thickness
vertex-wise data from three MRI datasets (ADNI, DRC tAD and DRC PCA), as well
as an amyloid PET dataset (ADNI). Our model found qualitatively similar patterns of
cortical thinning in tAD subjects using the two independent datasets (ADNI and DRC).
Moreover, it also found different patterns of pathology dynamics on two distinct diseases
(tAD and PCA) and on different types of data (PET and MRI-derived cortical thickness).
Finally, DIVE also provides a new way to parcellate the brain that is specific to the
temporal trajectory of a particular disease, and enables staging of individuals at risk of
disease, which can potentially help stratification in clinical trials.
The characteristics of the subjects’ data used for training can affect the DIVE out-
put. For instance, in cortical thinning analyses we standardised the data with respect
to controls, which might have already shown cortical thinning due to early pathology.
This can be mitigated through enrichment of the control population to amyloid-negative
individuals. DIVE also relies on subjects spanning the entire disease progression, so in-
clusion of subjects in middle stages is recommended for a robust estimation of trajectories
and spatial patterns. To reliably estimate the subject-specific time shift and progression
speed, multiple follow-up scans are required. We mitigated this by using only subjects
with at least three scans, and further placing informative priors on these parameters.
The DIVE-estimated spatial patterns are patchier in MRI compared to PET scans,
which had lower resolution and were smoothed a-priori. However, we believe MRI images
should not instead smoothed a-priori, as the spatial correlation mechanism within DIVE
enables it to automatically remove high-frequency patterns from MRI that are not mean-
ingful. Moreover, such a-priori smoothing could potentially loose dispersed patterns of
pathology that arise due to underlying disruption of brain networks.

6.5.2 Limitations and future work


DIVE has some limitations that can be addressed. First, we assumed that cluster tra-
jectories follow sigmoidal shapes, which is not the case for many types of biomarkers in
ADNI which do not plateau in later stages. The assumption of sigmoidal trajectories can
be avoided using non-parametric curves such as Gaussian Processes [246], which would be
straightforward to incorporate into the DIVE framework. To get a reliable estimate of the
subject-specific parameters, we only tested DIVE on balanced datasets, where subjects
had at least three scans. However, DIVE can also be applied to less balanced datasets,
by setting stronger priors on these parameters or even fixing the progression speed for
every subject to 1. Another limitation of the model is that it assumes all subjects fol-
low the same disease progression pattern, which might not be the case in heterogeneous
datasets such as ADNI or DRC. This can be a concern, as there might be a pattern of
pathology that occurs in a small set of subjects. However, DIVE can be extended to
account for heterogeneity in the datasets by modelling subject-specific trajectories using
6.6. Conclusion 119

random effects, or different progression dynamics for distinct subgroups, using unsuper-
vised learning methods like the SuStaIn model by [29]. While SuStaIn, just like DIVE,
estimates clusters and trajectories within the dataset, the clusters in SuStaIn are made
of subjects with similar disease progression, while the clusters in DIVE are made of ver-
tices with similar progression. Future work could combine clustering along both subjects
and vertices simultaneously to estimate disease subtypes with distinct spatiotemporal
dynamics at the vertexwise level.
There are several potential future applications of DIVE. One of the advantages of
DIVE is that it can be used to study the link between disconnected patterns of brain
pathology and connectomes extracted from diffusion tractography or functional MRI
(fMRI). Such an analysis would enable further understanding of the exact underlying
mechanisms by which the brain is affected by the disease. Our model, which can es-
timate fine-grained spatial patterns of pathology, is more suitable than standard ROI-
based methods for studying the link between pathology and these structural or functional
connectomes, because white matter or functional connections have a fine-grained and
spatially-varying distribution of endpoints on the cortex.
Apart from studying the link with brain connectomes, there are other potential ap-
plications for DIVE. While we only applied it to vertexwise data, the model can also be
applied to study voxelwise data. Moreover, DIVE can be applied to other modalities or
types of data, including FDG PET, tau PET, DTI or Jacobian compression maps from
MRI. Moreover, the model can also be extended to cluster points on the brain surface
according to a more complex disease signature, that can be made of two or more biomark-
ers. For example, using our cortical thickness and amyloid PET datasets from ADNI, we
could have clustered points on the brain based on both modalities simultaneously. Such
complex disease signatures can offer important insights into the relationships between
different modalities and underlying disease mechanisms.
DIVE is a spatiotemporal model that can be used for accurately predicting and stag-
ing patients across the progression timeline of neurodegenerative diseases. The spatial
patterns of pathology can also be used to test mechanistic hypotheses which consider AD
as a network vulnerability disorder. All these avenues can help towards disease under-
standing, patient prognosis, as well as clinical-trials for assessing efficacy of a putative
treatment for slowing down cognitive decline.

6.6 Conclusion
In this chapter I developed DIVE, a spatiotemporal model of disease progression that
estimates fine-grained spatial patterns of brain pathology, while simultaneously placing
subjects optimally on a disease time axis. I applied it to two typical AD MRI datasets
(ADNI and DRC), one dataset of PCA patients, and one typical AD PET dataset. I also
tested the robustness of the method in simulations, under cross-validation, and I’ve also
compared its performance to simpler feature-based models.
In the next chapter, I will present another model, DKT, that can transfer information
across different types of dementias in order to estimate the progression of rare dementias
from limited, unimodal datasets.
120 Chapter 6. DIVE: A Spatiotemporal Progression Model of Brain Pathology
Chapter 7

Disease Knowledge Transfer across


Neurodegenerative Diseases

7.1 Contributions
In this chapter I present Disease Knowledge Transfer (DKT), a novel method for trans-
ferring biomarker information between related neurodegenerative diseases. I performed
the mathematical modelling, implementation of the DKT method, data pre-processing,
statistical analysis and model validation. The TADPOLE dataset has been assembled by
myself and Neil Oxtoby, with suggestions from the EuroPOND team. The PCA dataset
was acquired by the Dementia Research Centre, UK.
While the original DKT implementation relied on a non-parametric GP disease pro-
gression model by Marco Lorenzi [246] as a building block, for this thesis I chose a
simpler parametric model, due to the complexity of fitting hierarchical, non-parametric,
latent-space models.

7.2 Publications
• R. V. Marinescu, M. Lorenzi, S. B. Blumberg, P. Planell-Morell, A. L. Young, N.
P. Oxtoby, A. Eshaghi, K. X. X. Yong, S. Crutch, D. C. Alexander, arXiv, 2019.

7.3 Introduction
The estimation of accurate biomarker signatures in Alzheimer’s disease (AD) and related
neurodegenerative diseases is crucial for understanding underlying disease mechanisms,
predicting subjects’ progressions, and selecting the right subjects in clinical trials. As a
result, data-driven disease progression models (chapter 3) were proposed that reconstruct
long term biomarker signatures from collections of short term individual measurements.
When applied to large datasets of typical AD, disease progression models have shown im-
portant benefits in understanding the earliest events in the Alzheimer’s disease cascade
[28, 24], the heterogeneity of AD [29], helped discover novel genes involved in AD [247]
and they showed improved predictions over standard approaches [30]. However, by neces-
sity these models require large datasets – in addition they must be both multimodal and
longitudinal. Such data is not available in rare neurodegenerative diseases. In particular,
122 Chapter 7. Disease Knowledge Transfer across Neurodegenerative Diseases

most datasets for rare neurodegenerative diseases come from local clinical centres, are
unimodal (e.g. MRI only) and limited both cross-sectionally and longitudinally – this
makes the application of disease progression models extremely difficult. Moreover, such a
model estimated from common diseases such as typical AD may not generalise to specific
variants. For example, in Posterior Cortical Atrophy – a neurodegenerative syndrome
causing visual disruption – posterior regions such as the occipital lobe and superior pari-
etal regions are affected early, instead of the hippocampus and temporal regions that are
affected early in typical AD.
The problem of limited data in medical imaging has so far been addressed through
transfer learning methods. Such techniques have been successfully used to improve the
accuracy of AD diagnosis [248, 249] or prediction of MCI conversion [250], but have two
key limitations. First, they use deep learning or other machine learning methods, which
are not interpretable and don’t allow us to understand underlying disease mechanisms
that are either specific to rare diseases, or shared across related diseases. Secondly, these
models cannot be used to forecast the future evolution of subjects at risk of dementia,
which is important for selecting the right subjects in clinical trials.
We propose Disease Knowledge Transfer (DKT), a generative joint model that esti-
mates continuous multimodal biomarker progressions for multiple neurodegenerative dis-
eases simultaneously – including rare neurodegenerative diseases – and which inherently
performs transfer learning between the modelled phenotypes. This is achieved by exploit-
ing biomarker relationships that are shared across diseases, whilst accounting for differ-
ences in the spatial distribution of brain pathology. DKT is interpretable, which allows us
to understand underlying disease mechanisms, and can also predict the future evolution
of subjects at risk of diseases. We apply DKT on Alzheimer’s variants and demonstrate
its ability to predict non-MRI trajectories for patients with Posterior Cortical Atrophy,
in lack of such data. This is done by fitting DKT to two datasets simultaneously: (1)
the TADPOLE Challenge [236] dataset containing subjects from the Alzheimer’s Disease
Neuroimaging Initiative (ADNI) with MRI, FDG-PET, DTI, AV45 and AV1451 scans
and (2) MRI scans from patients with Posterior Cortical Atrophy from the Dementia
Research Centre (DRC), UK. We first show that the estimated non-MRI trajectories for
PCA subjects are plausible as they agree with previous literature findings. We finally val-
idate DKT on three datasets: 1) simulated data with known ground truth, 2) TADPOLE
sub-populations with different progressions and 3) 20 DTI scans from controls and PCA
patients from the DRC, showing it yields favourable performance compared to standard
approaches. Code for DKT is available online: https://github.com/mrazvan22/dkt.
7.3. Introduction 123

Disease 1 (e.g. tAD) Disease 2 (e.g. PCA)


Disease Specific

abnormal abnormal
n n
io
ct tio
fu
n
on nc
Dysfunction

Dysfunction
i fu n
D ys n ct y s ct
io
score

fu

score
al lD un
r ys l
ta on ita ys
f l
po lD ra ion
em a i pi cti c ip D po nct
T nt cc un O
c t al m u
Fr
o O ysf on Te ysf
D Fr D

normal Disease 1 normal Disease 2


progression progression
Disease Agnostic

abnormal abnormal
Temporal Unit Occipital Unit
l
ra
po
Biomarker

l
Biomarker

l
te
m ra - ita
value

po d ip
value

d oi al l

Am
yl
oi
u
te
m
po
ral ... l
y it
Am ccip
Ta
u
occ
cc
ip
ita
Ta m o o
te RI
RI M
M
normal Temporal Dysfunction normal Occipital Dysfunction

Figure 7.1: Diagram of the proposed framework for joint modelling of multiple diseases.
We assume that each disease can be modelled as the evolution of abstract dysfunctionality
scores (Y-axis, top row), each one related to different brain regions. Each region-specific
dysfunctionality score then further models (X-axis, bottom row) the progression of sev-
eral modality-specific biomarkers within that same region. For instance, the temporal
dysfunction, modelled as a biomarker in the disease specific model (top row), is the
X-axis in the disease agnostic model (temporal unit, bottom row), which aggregates to-
gether abnormality from amyloid, tau and MR imaging within the temporal lobe. The
biomarker correlations within the bottom units are assumed to be disease agnostic and
shared across all diseases modelled. Disease knowledge transfer can then be achieved via
the disease-agnostic units.
124 Chapter 7. Disease Knowledge Transfer across Neurodegenerative Diseases

7.4 Methods
7.4.1 DKT Framework
Fig. 7.1 shows the overall diagram of our proposed framework for joint modelling of
diseases. We assume that the progression of each disease (X-axis, top row) can be mod-
elled as the evolution of abstract dysfunctionality scores, each one related to different
brain regions (top row). Each dysfunctionality score is then modelled as the progression
of several biomarkers within that same region, but acquired using different noninvasive
imaging modalities (bottom row). Each group of biomarkers in the bottom row will be
called a functional unit, because the correlations between biomarkers are related through
common ”function” in a disease–agnostic way, since they are related to the same under-
lying brain region. Biomarker groupings into functional units are defined a-priori. We
choose to model the correlations within each unit using the disease progression model
(DPM) by Jedynak et al. [2], but any other DPM can also be used. The DPM allows
us to reconstruct unit-specific dysfunction progression manifolds (bottom row, X axis),
which can be used for staging subjects. Finally, we use the same DPM to express the
progression within each disease (Figure 1, top) in terms of the dysfunction scores esti-
mated within each functional unit. More precisely, the X-axis dysfunction scores from
the functional units become Y-axis measurements in the disease specific models.
The model has a concise mathematical formulation. We assume a set of given biomark-
ers measurements Y = [yijk |(i, j, k) ∈ Ω] for subject i at visit j in biomarker k, where Ω is
defined as the set of available biomarker measurements, since subjects can have missing
data at various visits. We assume that each subject i at each visit j has an underlying
disease stage sij = βi + mij , where mij represents the months since baseline visit for
subject i at visit j and βi represents the time shift of subject i. We further denote by θk
the parameters used to represent the trajectory for biomarker k ∈ K within its functional
unit ψ(k), where ψ: {1, ..., K} → Λ is a function that maps each biomarker k to a unique
functional unit l ∈ Λ, where Λ is the set of functional units. Moreover, we denote by
λld the parameters for the trajectory of the dysfunction score corresponding to functional
unit l ∈ Λ in the space of disease d. These definitions allow us to formulate the likelihood
for a single measurement yijk as follows:

ψ(k) ψ(k)
p(yijk |θk , λdi , βi , k ) = N (yijk |g(f (βi + mij ); λdi ; θk ), k ) (7.1)

where g( . ; θk ) represents the trajectory of biomarker k within functional unit ψ(k) and
ψ(k)
f ( . ; λdi ) represents the trajectory of the functional unit ψ(k) within the space of
disease di . To be precise, di ∈ D represents the index of the disease space where subject i
belongs, where D is the set of all diseases modelled. For example, MCI and tAD subjects
from ADNI as well as tAD subjects from the DRC cohort can all be assigned di = 1,
while PCA subjects from the DRC dataset can be assigned di = 2. Healthy controls
can be assigned to either disease space, although a more precise assignment would take
molecular biomarkers into account. Variable k denotes the variance of measurements for
biomarker k.
We extend the above model to multiple subjects, visits and biomarkers to get the full
model likelihood:
Y ψ(k)
p(y|θ, λ, β, ) = p(yijk |θk , λdi , βi ) (7.2)
(i,j,k)∈Ω
7.4. Methods 125

where y = [yijk |∀(i, j, k) ∈ Ω] is the vector of all biomarker measurements, while


θ = [θ1 , ..., θK ] represents the stacked parameters for the trajectories of biomarkers in
functional units, λ = [λld |l ∈ Λ, d ∈ D] are the parameters of the dysfunctionality tra-
jectories within the disease models, β = [β1 , ..., βN ] are the subject-specific time shifts
and  = [k |k ∈ K] estimates biomarker measurement noise. Although we assumed inde-
pendence across different subjects, biomarker measurements and visits are linked using
the latent time-shift βi for each subject. The parameters of the model that need to be
estimated are [θ, λ, β, ]. For model simplicity, we did not account for inter-individual
variability other than that expressed by the time-shift βi , although this could be extended
in future work.

7.4.2 Modelling Biomarker Trajectories


ψ(k)
So far we defined the DKT framework using generic models g( . ; θk ) and f ( . ; λdi )
for the biomarker trajectories within the functional units and the disease models. Now
we choose to implement the f and g models as parametric sigmoidal curves, to enable
a robust optimisation and because these models account for the floor and ceiling effects
normally observed in AD biomarkers [175, 194]. The sigmoidal model for f is defined as:
ak
f (s; θk ) = + dk (7.3)
1 + exp(−bk (s − ck ))
where s is the disease progression score of a subject and θk = [ak , bk , ck , dk ] are pa-
rameters controlling the shape of the trajectory for biomarker k: dk and dk + ak represent
the lower and upper limits of the sigmoidal function, ck represents the inflection point
and ak bk /4 represents the slope at the inflection point. A similar model is used also for
g.

7.4.3 Parameter Estimation


We estimate the model parameters using a two-stage approach. In the first stage, we
perform belief propagation within each functional unit and then within each disease
model. Each functional unit and disease model is assumed to be an independent disease
progression model that we fit by alternatively optimising the fit of biomarker trajectories
and subject-specific time-shifts, using the approach described in [2]. At this stage we
ψ(k) ψ(k)
assume the existence of a latent variable βi = f (βi + mij ; λdi ) representing the
dysfunctionality score of subject i within the functional unit ψ(k), which represents a
time-shift within that functional unit.
In the second stage we jointly optimise across all functional units and disease models
using loopy belief propagation. An overview of the algorithm is given in Figure 7.2. Given
the initial parameters estimated from the first stage (line 1), the algorithm continuously
updates the biomarker trajectories within the functional units (lines 4-5), dysfunctionality
trajectories (line 9) and subject-specific time shifts (line 13) until convergence. The cost
function for all parameters is nearly identical, the main difference being the measurements
(i, j, k) over subjects i, visits j and biomarkers k that are selected for computing the
measurement error. For estimating the trajectory of biomarker k within functional unit
ψ(k), measurements are taken from Ωk representing all measurements of biomarker k from
all subjects and visits. For estimating the dysfunctionality trajectories, Ωd,l represents
the measurement indices from all subjects with disease d (i.e. di = d) and all biomarkers
126 Chapter 7. Disease Knowledge Transfer across Neurodegenerative Diseases

1 Initialise θ (0) , λ(0) , β (0)


2 while θ, λ, β not converged do
; // Estimate biomarker trajectories (disease agnostic)
3 for k = 1 to K do
h  i2
(u) P (u−1) ψ(k),(u−1)
4 θk = arg minθk (i,j)∈Ωk yijk − g f (βi + mij ; λdi ); θk − log p(θk )
h  i 2
(u) (u−1) ψ(k),(u−1) (u)
k = |Ω1k | (i,j)∈Ωk yijk − g f (βi
P
5 + mij ; λdi ); θk
6 end
; // Estimate dysfunctionality trajectories (disease specific)
7 for d = 1 ∈ D do
8 for l = 1 ∈ Λ do
h  i2
l,(u) (u−1) (u)
+ mij ; λld ); θk − log p(λld )
P
9 λd = arg minλld (i,j,k)∈Ωd,l yijk − g f (βi
10 end
11 end
; // Estimate subject-specific time shifts
12 for i = 1 ∈ [1, . . . , S] do
h  i2
(u) P ψ(k),(u) (u)
13 βi = arg minβi (j,k)∈Ωi yijk − g f (βi + mij ; λdi ); θk − log p(βi )
14 end
15 end

Figure 7.2: The algorithm for estimating the DKT parameters. The algorithm succes-
sively updates the biomarker trajectories within the functional units (disease agnostic
models), dysfunctionality trajectories (disease specific) and subject-specific time shifts
until convergence.

k that belong to functional unit l (i.e. ψ(k) = l). Finally, Ωi (line 13) represents all
measurements from subject i, for all biomarkers and visits.
The algorithm we proposed in Figure 7.2 has a complexity of O(I ∗ S), where S is the
number of subjects in the dataset and I is the number of iterations until convergence. In
practice, convergence is achieved after around 10-15 iterations, which takes around 1h on
a Xeon CPU E5-2680 @ 2.5GHz.

7.4.4 Synthetic Experiment


We first test DKT on synthetic data, in order to assess the performance when ground
truth is known. We generate synthetic data from two diseases as follows:

Disease model

• We define two functional units l0 and l1 and 6 biomarkers k0 − k5 , which we allocate


to functional units as follows: l0 : {k0 , k2 , k4 }, l1 : {k1 , k3 , k5 }. Within their units,
we define the trajectory of each biomarker as a sigmoidal curves with the following
θk parameters:

– functional unit l0 : θ0 = (1, 5, 0.2, 0), θ2 = (1, 5, 0.55, 0) and θ4 = (1, 5, 0.9, 0)
– functional unit l1 : θ1 = (1, 10, 0.2, 0), θ3 = (1, 10, 0.55, 0) and θ5 = (1, 10, 0.9, 0)
7.4. Methods 127

• We define two synthetic diseases, ”synthetic AD” (d = 0) and ”synthetic PCA”


(d = 1). For each disease d, each functional unit l has a distinct dysfunctionality
trajectory defined as a sigmoidal curve with parameters λld as follows:

– ”synthetic AD” disease: λ00 = (1, 0.3, −4, 0) and λ10 = (1, 0.2, 6, 0).
– ”synthetic PCA” disease: λ01 = (1, 0.3, 6, 0) and λ11 = (1, 0.2, −4, 0).

Subject model

• We generated time-shifts βi for 100 subjects (disease d0 ) and 50 subjects (disease


d1 ) based on a uniform distribution with ranges (−13, 10) years before/after disease
onset.

• Within each disease, we generated the subjects’ diagnosis (controls/patients) based


on an exponential likelihood model with mean -4.5 (controls)/4.5 (patients) years
before/after disease onset.

• For each subject and each biomarker, we generated data for four consecutive visits,
each visit one year apart, using a noise standard deviation of 0.05.

These trajectory and subject parameters were chosen to mimic the TADPOLE and
DRC cohorts, described below. Before fitting DKT on the synthetic dataset, we discarded
the data from biomarkers k0 , k1 , k4 and k5 for all subjects within the synthetic PCA
cohort, to simulate the lack of multimodal data in these subjects. Remaining biomarkers
k2 and k3 , for which data was still available in the synthetic PCA cohort, are assumed to
be of the same modality (e.g. MRI volume) but to represent measurements from different
brain regions (e.g. temporal and occipital).

7.4.5 Data Acquisition and Preprocessing


We trained DKT on ADNI data from the TADPOLE challenge [236], since it contained
a large number of multimodal biomarkers already pre-processed and aggregated into one
table. From the TADPOLE dataset we selected a subset of 230 subjects which had at
least one FDG PET, AV45, AV1451 or DTI scan. Most subjects also had MRI scans and
cognitive tests. In order to model another disease, we further included 76 PCA subjects
from the DRC in the training set, along with 67 tAD and 87 age-matched controls, all of
which only had MRI scans.
For both datasets, volumetric measures for each subject have been obtained using the
Freesurfer software. For FDG, AV45 and AV1451 PET, we used already extracted SUVR
measures from ADNI. For DTI, we used fractional anisotropy (FA) measures from white-
matter regions adjacent to each lobe. For every lobe, we averaged the biomarker values
for regions of interest within each lobe and regressed out the following covariates: age,
gender, total intracranial volume (TIV) and dataset (ADNI vs DRC dataset). Finally,
we normalised the biomarker values to lie within the [0,1] range.
For validating DKT’s performance at predicting missing biomarkers in PCA, we used
a separate test set of DTI scans from the DRC controls and PCA subjects. As this
validation set was acquired at a centre different from ADNI and on different scanners,
we matched the FA mean and standard deviation of the DRC controls to be equal to the
FA mean and standard deviation of the ADNI controls. No DTI data from PCA subjects
was exposed to the algorithm at training time.
128 Chapter 7. Disease Knowledge Transfer across Neurodegenerative Diseases

7.5 Results
7.5.1 Synthetic Results
Fig. 7.3 shows the true and estimated subject shifts and trajectories for each functional
unit l and biomarker k. In the top-left figures we show scatter plots of the true shifts
(y-axis) against estimated shifts (x-axis), for the ’synthetic AD’ and ’synthetic PCA’
diseases. On the top-right and middle-left figures, we show the trajectories of the func-
tional units within disease d = 0 (synthetic AD) and d = 1 (synthetic PCA). In the
middle-right and bottom-left figures, we show the biomarker trajectories within units l0
and l1 . In Figure 7.4, we show the corresponding trajectories of PCA patients, which as
opposed to Fig. 7.3, are plotted directly against the time-shifts, as it is normally done in
a classical disease progression model. We also show the true trajectories and the data of
the synthetic PCA cohort.
The results in Fig. 7.3 suggest that the DKT-estimated trajectories match closely
(mean absolute error, MAE < 0.058) with the true trajectories, for both the unit-
trajectories within the disease-specific models and the biomarker trajectories within the
disease-agnostic models. Moreover, the subject time-shifts are very close (R2 > 0.98) to
the true time-shifts. When plotted directly against the disease space, the estimated PCA
trajectories also match the true trajectories, even when there is a complete lack of such
data (Fig. 7.4, biomarkers 0,1,4 and 5). There are however small errors in biomarkers
0 and 5 which are due to measurement noise (confirmed by experiments with smaller
noise level – not shown here). The equivalent trajectories estimated for the synthetic AD
cohort also show very good agreement with the true trajectories (Fig. C.1).

7.5.2 Results on TADPOLE and DRC Datasets


Fig. 7.5a shows the estimated biomarker trajectories within the occipital unit plotted
over the dysfunction scores, along with samples from the model posterior and aligned
subject data. The X-axis shows the dysfunctionality scores within the occipital unit,
which represent estimated time-shifts, in months, from an arbitrary reference X=0, while
the Y-axis shows biomarker values normalised to [0,1] range. The model shows an un-
biased data fit (Fig. 7.5a), and we can observe most PCA subjects having abnormal
occipital volumes, thus leading to high occipital dysfunctionality scores, in line with the
current understanding of PCA as affecting posterior regions [18]. We also show the pro-
gression of dysfunctionality scores over the disease stage for typical AD (Fig 7.5b) and
PCA (Fig 7.5c). In typical AD, we see that hippocampal dysfunction becomes abnormal
earliest, while PCA shows early hippocampal dysfunction, which is later exceeded by
the dysfunction in the occipital, temporal and parietal regions, which are known to be
affected in PCA [18, 251].
In Fig. 7.6, we plot the inferred biomarker trajectories for PCA directly across the
disease progression. We do this for five different modalities: MRI Volumes, DTI, FDG,
AV45 and AV1451. The results again recapitulate known patterns in PCA, where poste-
rior regions are predominantly affected in all modalities. However, for MRI volumes and
AV45, we also see early abnormalities, which we attribute to the models underestimating
the biomarker measurement noise.
7.5. Results 129

Subject shifts Subject shifts Dis0 estimated trajectories Dis0 true trajectories
10

dysfunctionality score

dysfunctionality score
R 2 = 0.997 10 R 2 = 0.988 1.0 MAE = 0.057 1.0
5 5
true shifts

true shifts

0 0 0.5 0.5
5 CTL 5 CTL2 Unit0 Unit0
AD PCA 0.0 Unit1 0.0 Unit1
10 10
10 0 10 10 5 0 5 10 0.00 0.25 0.50 0.75 1.00 0.00 0.25 0.50 0.75 1.00
estimated shifts estimated shifts disease progression score disease progression score
Dis1 estimated trajectories Dis1 true trajectories Unit0 estimated trajectories Unit0 true trajectories
dysfunctionality score

dysfunctionality score

1.0 MAE = 0.055 1.0 1.0 MAE = 0.058 1.0


biomarker value

biomarker value
0.5 0.5 0.5 0.5
biomk 0 biomk 0
Unit0 Unit0 biomk 2 biomk 2
0.0 Unit1 0.0 Unit1 0.0 biomk 4 0.0 biomk 4
0.00 0.25 0.50 0.75 1.00 0.00 0.25 0.50 0.75 1.00 0.00 0.25 0.50 0.75 1.00 0.00 0.25 0.50 0.75 1.00
disease progression score disease progression score dysfunctionality score dysfunctionality score
Unit1 estimated trajectories Unit1 true trajectories
1.0 MAE = 0.015 1.0
biomarker value

biomarker value

0.5 0.5
biomk 1 biomk 1
0.0 biomk 3 0.0 biomk 3
biomk 5 biomk 5
0.00 0.25 0.50 0.75 1.00 0.00 0.25 0.50 0.75 1.00
dysfunctionality score dysfunctionality score

Figure 7.3: Comparison between true and DKT-estimated subject time-shifts and
biomarker trajectories. (top-left) Scatter plots of the true shifts (y-axis) against esti-
mated shifts (x-axis), for the ’synthetic AD’ (left) and ’synthetic PCA’ (right) diseases.
We also show the DKT-estimated and true trajectories of the functional units within
the ’synthetic AD’ disease (top-right) and the ’synthetic PCA’ disease (middle-left). For
these figures, the x-axis measures the normalised disease progression score si while the
y-axis measures the dysfunctionality scores f (si ; λld ). Finally, we also show the biomarker
trajectories within unit 0 (middle-right) and unit 1 (bottom), where the x-axis represents
the dysfunctionality scores f (si ; λld ) and the y-axis represents the biomarker value.
130 Chapter 7. Disease Knowledge Transfer across Neurodegenerative Diseases

estimated trajectory true trajectory CTL2 PCA


biomarker 0 biomarker 1 biomarker 2
MAE = 0.048 MAE = 0.021 MAE = 0.035
1.0 1.0 1.0

0.5 0.5 0.5

0.0 0.0 0.0


Biomarker Value

10 0 10 10 0 10 10 0 10
biomarker 3 biomarker 4 biomarker 5
MAE = 0.042 MAE = 0.029 MAE = 0.058
1.0 1.0 1.0

0.5 0.5 0.5

0.0 0.0 0.0

10 0 10 10 0 10 10 0 10
Disease Progression (years)

Figure 7.4: Estimated biomarker trajectories for the ”synthetic PCA” disease, plotted
alongside true trajectories. Estimation of the trajectories in biomarkers 0,1,4 and 5 has
been done without any data from the ”synthetic PCA” disease, only based on the disease-
agnostic correlations with biomarkers 2 and 3.
7.5. Results 131

(a) Occipital unit

(b) tAD (c) PCA

Figure 7.5: (a) DKT-estimated biomarker trajectories in the occipital functional unit.
Subject data from ADNI and our local DRC cohort are also shown. The X-axis, defined
as the occipital dysfunctionality score, represents the time-shifts (in months) of each
subject. (b-c) Progression of DKT-estimated dysfunctionality scores for (b) typical AD
and (c) PCA.
132 Chapter 7. Disease Knowledge Transfer across Neurodegenerative Diseases

Figure 7.6: Estimated multi-modal trajectories for the PCA cohort. The only data that
were available were the MRI volumetric data. The dynamics of the other biomarkers
has been inferred by the model using data from typical AD, and taking into account the
different spatial distribution of pathology in PCA as compared to typical AD.
7.6. Validation on DTI Data in PCA 133

7.6 Validation on DTI Data in PCA


We further validated DKT by predicting unseen DTI data from two patient datasets:

• TADPOLE subjects with a different progression from the training subjects

• A separate test set of 20 DTI scans from controls and PCA patients from our own
cohort.

To split TADPOLE into subgroups with different progression, we used the SuStaIn
model by [29], which resulted into three subgroups: hippocampal, cortical and subcorti-
cal, with prominent early atrophy in the hippocampus, cortical and subcortical regions
respectively. To evaluate prediction accuracy, we computed the rank correlation between
the DKT-predicted biomarker values and the measured values in the test data. We
compute the rank correlation instead of mean squared error as it is not susceptible to
systemic biases of the models when predicting ”unseen data” in a certain disease. We also
compared the performance of DKT at predicting unseen data with four other models:

• Latent stage model : a sigmoidal based disease progression model, as described in


[2]. This model assumes all tAD and PCA subjects follow the same progression.

• Multivariate: A multivariate Gaussian Process model with RBF kernel that predicts
a DTI ROI marker from multiple MRI markers.

• Spline: a univariate cubic spline regression model that predicts the DTI biomarker
based on the corresponding MRI biomarker, independently for each region.

• Linear : Same as above but linear model instead of spline.

Validation results are shown in Table 7.1, for hippocampal to cortical TADPOLE
subgroups, as well as PCA subjects from our DRC cohort. When predicting missing DTI
markers from the TADPOLE cortical subgroup as well as PCA subjects, the DKT corre-
lations are generally high for the cingulate, hippocampus and parietal, and lower for the
frontal lobe. DKT further shows favourable performance compared to the other models,
due to it’s ability to disentangle the progressions of each disease separately. In particular,
it shows the best results for DTI FA prediction in the parietal and temporal lobes on
both datasets and similar performance to the latent-stage model on the PCA dataset for
the cingulate, frontal and hippocampal (differences here are not statistically significant).
Due to the challenging problem of predicting unseen data in these diseases/subtypes,
notice how the models yield bad predictions for the occipital lobe (negative correlations),
most likely due to overfitting.
134 Chapter 7. Disease Knowledge Transfer across Neurodegenerative Diseases

Model Cingulate Frontal Hippocam. Occipital Parietal Temporal


TADPOLE subgroups: Hippocampal subgroup to Cortical subgroup
DKT (ours) 0.56 ± 0.23 0.35 ± 0.17 0.58 ± 0.14 -0.10 ± 0.29 0.71 ± 0.11 0.34 ± 0.26
Latent stage 0.44 ± 0.25 0.34 ± 0.21 0.34 ± 0.24* -0.07 ± 0.22 0.64 ± 0.16 0.08 ± 0.24*
Multivariate 0.60 ± 0.18 0.11 ± 0.22* 0.12 ± 0.29* -0.22 ± 0.22 -0.44 ± 0.14* -0.32 ± 0.29*
Spline -0.24 ± 0.25* -0.06 ± 0.27* 0.58 ± 0.17 -0.16 ± 0.27 0.23 ± 0.25* 0.10 ± 0.25*
Linear -0.24 ± 0.25* 0.20 ± 0.25* 0.58 ± 0.17 -0.16 ± 0.27 0.23 ± 0.25* 0.13 ± 0.23*
typical Alzheimer’s to Posterior Cortical Atrophy
DKT (ours) 0.77 ± 0.11 0.39 ± 0.26 0.75 ± 0.09 0.60 ± 0.14 0.55 ± 0.24 0.35 ± 0.22
Latent stage 0.80 ± 0.09 0.53 ± 0.17 0.80 ± 0.12 0.56 ± 0.18 0.50 ± 0.21 0.32 ± 0.24
Multivariate 0.73 ± 0.09 0.45 ± 0.22 0.71 ± 0.08 -0.28 ± 0.21* 0.53 ± 0.22 0.25 ± 0.23*
Spline 0.52 ± 0.20* -0.03 ± 0.35* 0.66 ± 0.11* 0.09 ± 0.25* 0.53 ± 0.20 0.30 ± 0.21*
Linear 0.52 ± 0.20* 0.34 ± 0.27 0.66 ± 0.11* 0.64 ± 0.17 0.54 ± 0.22 0.30 ± 0.21*

Table 7.1: Performance evaluation of DKT and four other statistical models of decreasing
complexity. We show the rank correlation between predicted biomarkers and measured
biomarkers in (top) TADPOLE subgroups – information transfer from hippocampal sub-
group to cortical subgroup – and (bottom) PCA. (*) Statistically significant difference
in the performance of DKT vs the other models, based on a two-tailed t-test, Bonferroni
corrected.

7.7 Discussion
We presented DKT, a framework that enables, for the first time, joint modelling of
biomarker progressions in multiple neurodegenerative diseases simultaneously. The frame-
work allows the inference of biomarker trajectories in rare diseases, for which there is not
enough data to allow estimation of such trajectories, and accounts for a different spatial
distribution of pathology between distinct types of dementia. This further enables us
to understand the complex mechanisms of rare diseases, as well as mechanisms shared
between different types of related diseases.
We provided an example implementation of DKT using specific models of the biomarker
trajectories, measurement noise and link function (the disease progression score). How-
ever, DKT should be considered as a general framework for joint modelling of biomarker
trajectories within different diseases simultaneously. The actual implementation of DKT
can thus be extended to use non-parametric trajectories, or more complex link functions
that estimate not only subject time-shifts but also progression speed or higher order
terms.
While in this work we have focused on Alzheimer’s variants such as tAD and PCA,
DKT can also be applied to other progressive neurodegenerative diseases of non-Alzheimer’s
type such as tauopathies (e.g. Frontotemporal dementia), synucleinopathies (e.g. Parkin-
son’s disease), other neurodegenerative diseases such as Huntington’s disease or Multiple
Sclerosis, and even the normal ageing process. Cognitive tests can also be included in
the disease-specific sub-models of DKT, or even allocated in the functional units of the
regions that are responsible for those tasks, based on previous voxel-based morphometry
studies. However, some care needs to be exercised when selecting the biomarkers and
grouping them into functional units, as in some diseases the assumption of disease ag-
nostic dynamics might not hold for some groups of molecular biomarkers. For example,
some non-Alzheimer’s tauopathies such as Frontotemporal dementia might show tau ab-
normalities but no corresponding amyloid abnormalities within the same region. In the
case of Frontotemporal dementia, we recommend including higher-level biomarkers such
7.8. Conclusion 135

as glucose metabolism from FDG, white matter degeneration from DTI or volume from
structural MRI, but one should exclude amyloid markers.
Our work has several limitations: 1) DKT assumes all subjects within a disease fol-
low the same trajectory, without considering heterogeneity within the disease population,
2) the allocation of biomarkers into functional units has to be done using a-priori hu-
man knowledge, 3) DKT currently works only on extracted brain features, discarding
important information present in the brain morphometry, 4) for validation, the synthetic
experiment we ran was limited to only one setting of the parameters and 5) the valida-
tion on patient data was also done only on a small set of 20 DTI scans, due to lack of
multimodal data in PCA.
There are several potential avenues for further research: 1) to account for hetero-
geneity, DKT can also be easily extended to include subject-specific effects; 2) improved
schemes for biomarker allocation to functional units can take connectivity into account, or
derive it from the data automatically; 3) to account for brain morphometry and connec-
tivity, DKT can be extended into a fully spatio-temporal model, by estimating continuous
changes in volumetric brain images – in this case, each voxel can have an associated dys-
functionality score that is derived from measurements of various modalities from that
voxel; 4-5) DKT can be further validated on more complex synthetic experiments with
variable parameter settings, and on patient data from ADNI, where the population could
be a-priori split into sub-groups with different progressions. On these subgroups, DKT
can be used to transfer biomarker modalities that have been left out during training.

7.8 Conclusion
In this work I presented DKT, a novel method that can empower studies of rare dementias
with limited biomarker data by leveraging data from larger datasets of related dementias.
When applied to synthetic data with ground truth, I showed that DKT can robustly
recover biomarker trajectories in two distinct diseases and also subject-specific time-shifts.
I also applied DKT to multimodal imaging biomarkers from the TADPOLE Challenge
dataset, where I showed that it can estimate plausible non-MRI biomarker trajectories
for Posterior Cortical Atrophy in lack of such data for this disease. I validated the
performance of DKT on a test set of 20 DTI scans from PCA and controls, and showed
that DKT has similar or better performance compared to simpler models.
In the next chapter, I will present the TADPOLE Challenge, which evaluates the
performance of algorithms and features at predicting the future evolution of subjects at
risk of AD. As opposed to the work performed in this chapter, the TADPOLE challenge
aims to evaluate a much larger set of algorithms and features, comprising regression
techniques, disease progression models, machine learning techniques and even manual
predictions made by clinicians.
136 Chapter 7. Disease Knowledge Transfer across Neurodegenerative Diseases
Chapter 8

TADPOLE Challenge: Prediction of


Longitudinal Evolution in
Alzheimer’s Disease

8.1 Contributions
In this chapter I present the design of The Alzheimer’s Disease Progression Of Longitu-
dinal Evolution (TADPOLE) Challenge, which aims to predict the evolution of subjects
at risk of Alzheimer’s disease. The challenge was organised by the European Progression
of Neurodegenerative (EuroPOND) consortium, in collaboration with the Alzheimer’s
disease Neuroimaging Initiative (ADNI). The key organisers of the challenge were, in
alphabetical order: Daniel Alexander, Frederik Barkhof, Esther Bron, Nick Fox, Stefan
Klein, Razvan Marinescu (myself), Neil Oxtoby and Alexandra Young.
I contributed with suggestions to the challenge design, helped write the website, as-
sembled the TADPOLE D2 longitudinal dataset and the data dictionary, and wrote
benchmark prediction scripts. I also build the leaderboard system which performs live
evaluation of the participants’ submissions. I further helped promote the competition
at several medical imaging conferences, and organised two mini-competitions at the Py-
ConUK conference and at the CMIC summer school, 2018.
Daniel Alexander proposed the main design of the challenge, secured funding, helped
write the website, and wrote simple prediction scripts. Neil Oxtoby contributed to chal-
lenge design, helped me validate the D2 dataset, built the D3 cross-sectional dataset,
helped write the website, organised webinars and promoted the competition. Alexandra
Young contributed to challenge design, helped write the website, performed simulations
to establish which target biomarkers are most suitable and promoted the competition. Es-
ther Bron and Stefan Klein contributed to challenge design and helped write the website.
Nick Fox and Frederik Barkhof provided valuable suggestions on the challenge design.
Arthur Toga and Michael Weiner offered access to the ADNI database.

8.2 Publications
• R. V. Marinescu, N. P. Oxtoby, A. L. Young, E. E. Bron, A. W. Toga, M. W. Weiner,
F. Barkhof, N. C. Fox, S. Klein, D. C. Alexander and the EuroPOND Consortium,
138 Chapter 8. TADPOLE Challenge: Prediction of Evolution in Alzheimer’s Disease

TADPOLE Challenge: Prediction of Longitudinal Evolution in Alzheimer’s Disease,


arXiv, 2018
I wrote this challenge design paper based on text and diagrams from the TADPOLE
website. All collaborators contributed with feedback on the manuscript. The results
of the challenge will be published in a separate paper in 2019, after enough data
has been collected for the final evaluation.

8.3 Introduction
As already mentioned in section 3, early diagnosis of dementia is important in order
to enable the administration of treatments in early disease stages, before the onset of
cognitive decline. While such early and accurate diagnosis of dementia can be challenging,
this can be aided by quantitative biomarker measurements taken from magnetic resonance
imaging (MRI), positron emission tomography (PET), and cerebro-spinal fluid (CSF)
samples extracted from lumbar puncture. It has been hypothesised for AD [1, 119, 252,
253] that all these biomarkers become abnormal at different intervals before symptom
onset, suggesting that together they can be used for accurate prediction of onset and
overall disease progression in individuals. In particular, some of the early biomarkers
become abnormal decades before symptom onset, and can thus facilitate early diagnosis.
Several approaches for predicting AD-related target variables (e.g. clinical diagnosis,
cognitive/imaging biomarkers) have been proposed which leverage multimodal biomarker
data available in AD. Traditional longitudinal approaches based on statistical regression
model the relationship of the target variables with other known variables. Examples in-
clude regression of the target variables against clinical diagnosis [131], cognitive test scores
[193, 175], rate of cognitive decline [177], and retrospectively staging subjects by time
to conversion between diagnoses [254]. Another approach involves supervised machine
learning techniques such as support vector machines, random forests, and artificial neural
networks, which use pattern recognition to learn the relationship between the values of a
set of predictors (biomarkers) and their labels (diagnoses). These approaches have been
used to discriminate AD patients from cognitively normal individuals [207, 255], and for
discriminating at-risk individuals who convert to AD in a certain time frame from those
who do not [256, 257]. The emerging approach of disease progression modelling aims
to reconstruct biomarker trajectories or other disease signatures across the disease pro-
gression timeline, without relying on clinical diagnoses or estimates of time to symptom
onset. Examples include models built on a set of scalar biomarkers to produce discrete
[23, 24] or continuous [2, 3, 25] biomarker trajectories; richer but less comprehensive
models that leverage structure in data such as MR images [258, 259, 4]; and models of
disease mechanisms [38, 126, 6, 28].
These models have shown promise for predicting AD biomarker progression when us-
ing existing test data, but few have been tested on truly unseen future data. Moreover,
different investigators test these models on different datasets (including subsets of a sin-
gle dataset) and use different processing pipelines. Community challenges have proved
effective, in the medical image analysis field and beyond, for providing unbiased com-
parative evaluations of algorithms and tools designed for a particular task. Previous
challenges that focused on prediction of AD progression include the CADDementia chal-
lenge [39], which aimed to predict clinical diagnosis from MRI scans. A similar challenge,
the ”International challenge for automated prediction of MCI from MRI data” [40] asked
8.4. Competition Design 139

participants to predict diagnosis and conversion status from extracted MRI features of
subjects from the ADNI study [260]. Yet another challenge, The Alzheimer’s Disease
Big Data DREAM Challenge [261], asked participants to predict cognitive decline from
genetic and MRI data.
The Alzheimer’s Disease Prediction Of Longitudinal Evolution (TADPOLE) Chal-
lenge aims to identify the data, features and approaches that are the most predictive of
AD progression. In contrast to previous challenges, our motivation is to improve future
clinical trials through identification of patients most likely to benefit from an effective
treatment, i.e., those at early stages of disease who are likely to progress over the short-
to-medium term (1-5 years). Identifying such subjects reliably helps cohort selection by
focusing on groups that highlight positive treatment effects. The challenge thus focuses
on forecasting three key features: clinical status, cognitive decline, and neurodegenera-
tion (brain atrophy), over a five-year timescale. It uses rollover 1 subjects from the ADNI
study for whom a history of measurements is available, and who are expected to continue
in the study, providing future measurements for testing. Since the test data does not exist
at the time of forecast submissions, the challenge provides a completely unbiased basis
for performance comparison. TADPOLE goes beyond previous challenges by drawing
on a vast set of multimodal measurements from ADNI which support prediction of AD
progression.

8.4 Competition Design


The aim of TADPOLE is to predict future outcome measurements of subjects at-risk of
AD, enrolled in the ADNI study. A history of informative measurements from ADNI
(imaging, psychology, demographics, genetics, etc.) from each individual is available to
inform forecasts. TADPOLE participants are required to predict future measurements
from these individuals and submit their predictions before a given submission deadline.
Evaluation of these forecasts occurs post-deadline, after the measurements have been
acquired. A diagram of the TADPOLE flow is shown in Fig 8.1.

Figure 8.1: TADPOLE Challenge design. Participants are required to train a predictive
model on a training dataset (D1 and/or others) and make forecasts for different datasets
(D2, D3) by the submission deadline. Evaluation will be performed on a test dataset
(D4) that is acquired after the submission deadline.

1
i.e. subjects who enrolled in the previous ADNI2 study and who will continue in the third phase.
140 Chapter 8. TADPOLE Challenge: Prediction of Evolution in Alzheimer’s Disease

ADAS ADAS Vent. Vent.


CN MCI AD
RID Month Date ADAS CI CI Vent. CI CI
prob. prob. prob.
lower upper lower upper
A 1 2018-01 0 1 0 30 25 35 0.024 0.021 0.029
B 1 2018-01 3 2 0 25 21 26 0.023 0.021 0.025
C 1 2018-01 0.24 0.38 0.38 40 25 50 0.025 0.023 0.028

Table 8.1: The format of the forecasts for three example subjects. Participants have to
predict, for each subject, the probability of clinical diagnosis (CN/MCI/AD), the ADAS-
Cog13 score and Ventricle volume, as well as the 50% confidence range. RID - Roster ID
is the unique identifier for ADNI subjects, ADAS - ADAS-Cog13, CI - confidence range.
Note that, even if the CN/MCI/AD probabilities don’t sum to one, we will normalise
them anyway.

8.5 Forecasts
Since we do not know the exact time of future data acquisitions for any given individual,
TADPOLE challenge participants are required to make, for every individual, month-by-
month forecasts of three key biomarkers: (1) clinical diagnosis which can be either cog-
nitively normal (CN), mild cognitive impairment (MCI) or probable Alzheimer’s disease
(AD); (2) ADAS-Cog13 (ADAS13) score; and (3) ventricle volume (divided by intra-
cranial volume). Evaluation is performed using forecasts at the months that correspond
to data acquisition. TADPOLE forecasts are required to be probabilistic and some eval-
uation metrics will account for forecast probabilities provided by participants. Methods
or algorithms that do not produce probabilistic estimates can still be used, by setting
binary probabilities (zero or one) and default confidence intervals.
Participants are required to submit forecasts in a standardised format (see Table
8.1). For clinical status, relative likelihoods of each option (CN, MCI, and AD) for
each individual should be provided. These are normalised at evaluation time; negative
likelihoods are set to zero. For ADAS13 and ventricle volume, participants need to
provide a best-guess value as well as a 50% confidence interval for each individual. This
50% confidence interval (as opposed to the more standard 95%) was chosen to provide a
more symmetric and less noisy evaluation of over- and under-estimation of the confidence
interval, because similar sample sizes of data fall inside and outside the interval.

8.6 Data
We provide participants with a standard ADNI-derived dataset (available via the Labora-
tory Of NeuroImaging: LONI) which they can use to train their algorithms, removing the
need to pre-process the ADNI data themselves or merge different spreadsheets. However,
participants are allowed to use a custom training set, by adding any other ADNI data
or data from other studies. The software code used to generate the standard dataset is
openly available in a GitHub repository2 and on the ADNI website, packaged with the
standard dataset in the LONI ADNI database.

2
https://github.com/noxtoby/TADPOLE
8.7. TADPOLE Datasets 141

8.6.1 ADNI Data


Data used in the preparation of this article were obtained from the Alzheimer’s Dis-
ease Neuroimaging Initiative (ADNI) database (adni.loni.usc.edu). The ADNI was
launched in 2003 by the National Institute on Aging (NIA), the National Institute of
Biomedical Imaging and Bioengineering (NIBIB), the Food and Drug Administration
(FDA), private pharmaceutical companies and non-profit organisations, as a $60 million,
5-year public-private partnership. The initial goal of ADNI was to recruit 800 subjects
but ADNI has been followed by ADNI-GO and ADNI-2. To date these three protocols
have recruited over 1500 adults, aged 55 to 90, to participate in the research, consisting
of cognitively normal older individuals, people with early or late MCI, and people with
early AD. The general ADNI inclusion criteria has been described in [262].
The data we used from ADNI consists of: (1) CSF markers of amyloid-beta and tau
deposition; (2) various imaging modalities such as magnetic resonance imaging (MRI),
positron emission tomography (PET) using several tracers: Fluorodeoxyglucose (FDG,
hypometabolism), AV45 (amyloid), AV1451 (tau) as well as diffusion tensor imaging
(DTI); (3) cognitive assessments acquired in the presence of a clinical expert; (4) genetic
information such as Alipoprotein E4 (APOE4) status extracted from DNA samples; and
(5) general demographic information. Extracted features from this data were merged
together into a final spreadsheet and made available on the LONI ADNI website.

8.6.2 Image Preprocessing


The imaging data has been pre-processed with standard ADNI pipelines. For MRI scans,
this included correction for gradient non-linearity, B1 non-uniformity correction and peak
sharpening3 . Meaningful regional features such as volume and cortical thickness were
extracted using the Freesurfer cross-sectional and longitudinal pipelines [118]. Each PET
image (FDG, AV45, AV1451), which consists of a series of dynamic frames, had its
frames co-registered, averaged across the dynamic range, standardised with respect to
the orientation and voxel size, and smoothed to produce a uniform resolution of 8mm
full-width/half-max (FWHM)4 . Standardised uptake value ratio (SUVR) measures for
relevant regions-of-interest were extracted (see [263]) after registering the PET images
to corresponding MR images using the SPM5 software [264]. DTI scans were corrected
for head motion and eddy-current distortion, skull-stripped, EPI-corrected, and finally
aligned to the T1 scans using the pipeline from [265]. Diffusion tensor summary measures
were estimated based on the Eve white-matter atlas by [266].

8.7 TADPOLE Datasets


The TADPOLE Challenge involves three kinds of data sets: (a) a training data set, which
is a collection of measurements with associated outcomes that can be used to fit models
or train algorithms; (b) a prediction data set, which contains only baseline measurements
(possibly longitudinal), without associated outcomes — this is the data that algorithms,
models, or experts use as input to make their forecasts of later patient status and outcome;
3
see MRI analysis on ADNI website: http://adni.loni.usc.edu/methods/mri-analysis/
mri-pre-processing
4
see PET analysis on ADNI website: http://adni.loni.usc.edu/methods/pet-analysis/
pre-processing
142 Chapter 8. TADPOLE Challenge: Prediction of Evolution in Alzheimer’s Disease

Figure 8.2: Venn diagram of the TADPOLE datasets derived from ADNI data, for train-
ing (D1), longitudinal prediction (D2), cross-sectional prediction (D3) and the test set
(D4). D3 is a subset of D2, which in turn is a subset of D1. Other non-ADNI data can
also be used for training.

and (c) a test data set, which contains the patient outcomes against which we will evaluate
forecasts — in TADPOLE, this data did not exist at the time of submitting forecasts.
In order to evaluate the effect of different methodological choices, we prepared three
standard data sets for training and prediction:

• D1: The TADPOLE standard training set draws on longitudinal data from the
entire ADNI history. The data set contains a set of measurements for every in-
dividual that has provided data to ADNI in at least two separate visits (different
dates) across three phases of the study: ADNI1, ADNI GO, and ADNI2.

• D2: The TADPOLE longitudinal prediction set contains as much available data
as we could gather from the ADNI rollover individuals for whom challenge partici-
pants are asked to provide forecasts. D2 includes all available time-points for these
individuals.

• D3: The TADPOLE cross-sectional prediction set contains a single (most recent)
time point and a limited set of variables from each rollover individual in D2. Al-
though we expect worse forecasts from this data set than D2, D3 represents the
information typically available when selecting a cohort for a clinical trial.

The forecasts will be evaluated on future data (D4 – test set) from ADNI3 rollovers, ac-
quired after the challenge submission deadline. In addition to the three standard datasets
(D1, D2 and D3), challenge participants are allowed to use any other data sets that might
serve as useful additional training data.
Fig. 8.2 shows a diagram highlighting the nested structure of datasets D1–D3. Ta-
ble 8.2 shows the proportion of biomarker data available in each dataset. There are a
considerable number of entries with missing data, especially for some biomarkers such as
tau imaging (AV1451). We also estimated the expected number of subjects and avail-
able data for D4, using information from the ADNI3 procedures and using rollovers from
previous ADNI studies (Table 8.2, right-most column) – See E.1 for more information on
D4 estimates. Based on our estimates, we believe the size of D4 (around 330 subjects, 1
visit/subject) should be enough for a reliable evaluation of TADPOLE submissions.
8.8. Submissions 143

Subject statistics D1 D2 D3 D4
Nr. of subjects 1667 896 896 330
Visits per subject 7.6±3.8 8.5±4.2 1.0±0.0 1 .0 ±0 .0
CN 31 38 45 39
Diagnosis* (%) MCI 56 57 39 49
AD 13 5 16 12
Data availability**
Cognitive tests (%) 70 68 84 62
MRI (%) 62 56 75 69
FDG-PET (%) 16 20 0 20
AV45-PET (%) 16 22 0 19
AV1451-PET (%) 0.7 1.1 0 19
DTI (%) 6 8 0 15
CSF (%) 18 19 0 14

Table 8.2: Subject statistics and available data in the TADPOLE datasets D1, D2 and D3.
There is a considerable amount of missing data in some biomarkers such as AV1451. Num-
bers for D4 are estimated based on ADNI3 procedures (see ADNI3 procedures manual)
and rollovers from previous ADNI studies. (*) Diagnosis at baseline visit. (**) Percentage
of all visits (across all subjects) that have measurements for desired biomarker.

8.8 Submissions
There are two kinds of submissions that challenge participants can make. A simple entry
requires a minimal forecast and a description of methods; it makes participants eligible
for the prizes but not co-authorship on the scientific paper documenting the results. A
simple entry can use any training data or prediction sets and forecast at least one of the
target outcome variables (clinical status, ADAS13 score, or ventricle volume). A full entry
entitles participants for consideration as a co-author on the publication documenting the
results. Such a full entry requires a complete forecast for all three outcome variables on
all subjects from the D2 prediction set, along with a description of the methods. Each
individual participant is limited to a maximum of three submissions. This restriction has
been introduced to avoid the risk of participants tuning their method on the test set by
submitting multiple predictions for a range of algorithm settings. Although not required
for a full entry, participants are strongly encouraged to submit predictions also for D3.
Prizes are awarded to the best submissions regardless of the choice of training sets
(D1/custom) and prediction sets (D2/D3). However, the additional submissions support
the key scientific aims of the challenge by allowing us to separate the influence of the
choice of training data, post-processing pipelines, and modelling techniques or prediction
algorithms. The target variables used for evaluation, in particular ventricle volume, will
use the same post-processing pipeline as the standard data sets D1-D3.
Beyond the standard training dataset (D1), participants can include additional fore-
casts from ”custom” (i.e. constructed by the participant) training data or custom post-
processing of the raw data from subjects in the standard training set. The same applies
to the prediction sets D2 and D3, which can be customised by the participants if desired,
e.g. a prediction set with different features from the same individuals as in D2 and D3.
Table 8.3 shows the twelve possible combinations of subject sets, processing and predic-
tion sets, from which a full-entry submission must contain at least one of the first four
144 Chapter 8. TADPOLE Challenge: Prediction of Evolution in Alzheimer’s Disease

(ID 1–4).

ID Training set Prediction set


Subject set Post-processing
1 D1 standard D2
2 D1 custom D2
3 custom standard D2
4 custom custom D2
5 D1 standard D3
6 D1 custom D3
7 custom standard D3
8 custom custom D3
9 D1 standard custom
10 D1 custom custom
11 custom standard custom
12 custom custom custom

Table 8.3: Types of submissions that can be made by participants, for different types of
training sets, prediction sets and post-processing pipelines.

8.9 Forecast Evaluation


8.9.1 Clinical Status Prediction
For evaluation of clinical status predictions, we will use similar metrics to those that
proved effective in the CADDementia challenge [39]: (i) the multiclass area under the
receiver operating curve (mAUC); and (ii) the overall balanced classification accuracy
(BCA). The mAUC is independent of the group sizes and gives an overall measure of
classification ability that accounts for relative likelihoods assigned to each class. The
simpler BCA is also independent of group sizes, but does not exploit the probabilistic
nature of the forecasts.

8.9.1.1 Multiclass Area Under the Receiver Operating Characteristic


(ROC) Curve

The multiclass Area Under the ROC Curve (mAUC) is a simple generalisation of the
area under the ROC curve applicable to problems with more than two classes [267]. The
AUC Â(ci |cj ) for classification of a class ci against another class cj , is:

Si − ni (ni + 1)/2
Â(ci |cj ) = (8.1)
ni nj

where ni and nj are the number of points belonging to classes i and j, respectively; while
Si is the sum of the ranks of the class i test points after ranking all the class i and j
data points in increasing likelihood of belonging to class i. We further define the average
AUC for classes i and j as Â(ci , cj ) = 0.5(Â(ci |cj ) + Â(cj |ci )). The overall mAUC is then
8.9. Forecast Evaluation 145

obtained by averaging Â(ci , cj ) over all pairs of classes:

L i
2 XX
mAU C = Â(ci , cj ) (8.2)
L(L − 1) i=2 j=1

where L is the number of classes. The class probabilities that go into the calculation of
Si in the first equation are pCN , pM CI and pAD , which are derived from the likelihoods of
each ADNI subject being assigned to each diagnostic class, by normalising to have unity
sum.

8.9.1.2 Balanced Classification Accuracy

The Balanced Classification Accuracy (see [268]) is an extension of the classification


accuracy measure that accounts for the imbalance in the numbers of datapoints belonging
to each class. However, the measure is not probabilistic, so in TADPOLE the data points
need to be assigned a hard classification to the class (CN, MCI, or AD) with the highest
likelihood. The balanced accuracy for class i is then:
 
1 TP TN
BCAi = + (8.3)
2 TP + FN TN + FP

where TP, FP, TN, FN represent the number of true positives, false positives, true neg-
atives and false negatives for classification as class i. In this case, true positives are data
points with true label i and correctly classified as such, while the false negatives are the
data points with true label i and incorrectly classified to a different class j 6= i. True
negatives and false positives are defined similarly. The overall BCA is given by the mean
of all the balanced accuracies for every class.

8.9.2 Continuous Feature Predictions


For ADAS13 and ventricle volume, we will use three metrics: mean absolute error (MAE),
weighted error score (WES) and coverage probability accuracy (CPA). The MAE focuses
purely on accuracy of the best-guess prediction ignoring the confidence interval, whereas
the WES incorporates confidence estimates into the error score. The CPA provides an
assessment of the accuracy of the confidence estimates, irrespective of the best-guess
prediction accuracy.

8.9.2.1 Mean Absolute Error

The mean absolute error (MAE) is:

N
1 X
M AE = M̃i − Mi (8.4)

N i=1

where N is the number of data points (forecasts) evaluated, Mi is the actual biomarker
value in individual i in future data, and M̃i is the participant’s best prediction for Mi .
146 Chapter 8. TADPOLE Challenge: Prediction of Evolution in Alzheimer’s Disease

8.9.2.2 Weighted Error Score


The weighted error score is defined as:
PN
C̃ M̃ − M

i=1 i i i
W ES = PN (8.5)
i=1 C̃i

where C̃i is the participant’s relative confidence in their M̃i estimate. We estimate C̃i as
the inverse of the width of the 50% confidence interval of their biomarker estimate:

C̃i = (C+ − C− )−1 (8.6)

where [C−, C+] is the confidence interval provided by the participant.

8.9.2.3 Coverage Probability Accuracy


The coverage probability accuracy is:

CP A = |ACP − N CP | (8.7)

where N CP is the nominal coverage probability, the target for the confidence intervals,
and ACP is the actual coverage probability, defined as the proportion of measurements
that fall within the corresponding confidence interval. In TADPOLE, we set N CP to
be 0.5, which means that ideally only 50% of the measurements would fall inside the
confidence interval. The CPA can take values between 0 and 1, and lower scores are
better.

8.10 Prizes
We are extremely grateful to Alzheimer’s Research UK, The Alzheimer’s Society, and
The Alzheimer’s Association for sponsoring a prize fund of £30,000. At the time of first
submission, we proposed six separate prizes, as outlined in Table 8.4, but reserve the
right to reallocate the prize money depending on the numbers of participants eligible for
each prize. The first four are general categories (open to all challenge participants) and
constitute one prize for the best forecast of each feature as well as one for overall best
performance. The last two prizes are for two different student categories.
8.11. Discussion 147

Prize Performance
Outcome measure Eligibility
amount Metric
£5,000 Clinical status mAUC all
£5,000 ADAS13 MAE all
£5,000 Ventricle volume MAE all
Lowest sum of
£5,000 Overall best all
ranks*
University
£5,000 Clinical status mAUC
teams
High-school
£5,000 Clinical status mAUC
teams

Table 8.4: Prize allocation scheme using funds from Alzheimer’s Research UK, The
Alzheimer’s Society and The Alzheimer’s Association. There are 6 prizes awarded to
different outcome measures, the last two of which are eligible only for university and
high-school teams. (*) The overall best team will be the team that obtains the lowest
sum of ranks in the clinical status, ADAS13 and ventricle volume categories.

8.11 Discussion
We have outlined the design of the TADPOLE Challenge, which aims to identify algo-
rithms and features that can best predict the evolution of Alzheimer’s disease. Challenge
participants use historical data from ADNI in order to predict three key outcomes: clini-
cal diagnosis, ADAS-Cog13 and ventricle volume. Determining which features and algo-
rithms best predict AD evolution can aid refinement of cohorts and endpoint assessment
for clinical trials, and can provide accurate prognostic information in clinical settings.
The TADPOLE Challenge was designed to be transparent and accessible. To this end,
all of our scripts are available in an open repository5 . We also created a public forum6
where we answer participant questions. Finally, in order to enable participants to share
algorithm performance results throughout the competition, we created a leaderboard
system7 that evaluates submissions on an existing test dataset and publishes the results
live on our website.
Going forward, we hope that by November 2018 sufficient data will be available from
ADNI3 rollovers for a first meaningful evaluation of the forecasts. We plan to publish the
results on the website in January 2019, and then submit a publication of the results soon
after. However, we reserve the right to delay evaluation until sufficient data is available.
At that time, we will also evaluate the impact and interest of the first phase of TADPOLE
within the community, to guide decisions on whether to organise further submission and
evaluation phases.
The fact that the D4 test set could have different properties from the training set
is something that can affect the performance of certain algorithms. For example, some
algorithms could perform better on different forecast time windows (short-term vs long-
term) or on subjects with different properties (e.g. those with more follow-up training
data vs those with less data). At the evaluation stage, we thus take into consideration
doing the evaluation on different splits of the test set, in order to understand what kind
5
TADPOLE repository: https://github.com/noxtoby/TADPOLE
6
TADPOLE forum: https://groups.google.com/forum/#!forum/tadpolechallenge
7
Leaderboard: https://tadpole.grand-challenge.org/leaderboard/
148 Chapter 8. TADPOLE Challenge: Prediction of Evolution in Alzheimer’s Disease

of predictions algorithms perform best at.

8.12 Conclusion
In this section I presented the TADPOLE Challenge, which aims to identify algorithms
and features that best predict the evolution of subjects at risk of Alzheimer’s disease.
The outcomes of the challenge will be made available early in 2019, after sufficient data
has been acquired. In the next chapter, I will present future work on the TADPOLE
Challenge, as well as the other chapters of the thesis.
Chapter 9

Conclusions

In this thesis I presented my work on disease progression model applications to typical


Alzheimer’s disease and Posterior Cortical Atrophy, as well as novel methodological de-
velopments. In this chapter I will present a summary of the thesis (section 9.1), along
with future research directions, both for applications to other neurodegenerative diseases
(section 9.2.1) as well as further methodological developments (section 9.2.3).

9.1 Summary
In chapter 2, I first gave an overview of Alzheimer’s disease (section 2.1) by describing its
symptoms, disease causes and mechanisms, various risk factors involved, how it is cur-
rently diagnosed and the key biomarkers available to quantitatively measure Alzheimer’s
disease pathology. Afterwards, in section 2.2 I described the progression of AD biomarkers
and the Braak staging scheme. Finally, in section 2.3 I performed a literature review on
PCA, and described its symptoms, disease causes, diagnosis, management, neuroimaging
and heterogeneity. Throughout the section, I compared and contrasted the differences
between PCA and typical AD.
In chapter 3, I presented the state of the art in disease progression modelling. I
started with the hypothetical model by Jack et al. [1] (section 3.1), then presented early
models of progression based on symptomatic groups (section 2.1.1), then moved to con-
tinuous models which regress against one biomarker (section 3.3) and survival analysis
models that compute time until an event such as clinical conversion occurs (section 3.4).
I then presented state of the art methods that combine multiple biomarker measurements
and generally compute latent time shifts and other hidden variables. I categorised them
into models based on scalar biomarker measurements (section 3.5), spatiotemporal mod-
els (section 3.6) which model changes both in brain structure and over time, as well as
mechanistic models (section 3.7) which can be used to infer underlying disease mecha-
nisms. Finally, I presented a summary of key machine learning methods that have been
frequently used in medical imaging, especially for diagnosis and prognosis (section 3.8).
In chapter 4, I presented a longitudinal comparison of Posterior Cortical Atrophy
with typical Alzheimer’s disease, analysing the progression of atrophy from MRI. I first
presented the demographics (section 4.3.1) of the cohort from the Dementia Research
Centre, UK that I analysed, using data obtained by my collaborators. I then described
the methodology I applied, which involved adaptations of the event-based model and
the differential equation model to this specific dataset (section 4.3.3). I showed that
150 Chapter 9. Conclusions

there were differences in the progression of brain volumes in PCA as opposed to typical
AD, where phenotype-specific areas were affected early in the disease process (section
4.4.1). Moreover, I also showed that there were differences in atrophy progression in
three cognitively-defined PCA subtypes, highlighting the amount of heterogeneity within
PCA (section 4.4.2). Finally, in section 4.5 I discussed the findings of our study, the
strengths and limitations of our methods, and suggested directions for future research.
In chapter 5, I presented methodological advances in two disease progression models,
the event-based model and the differential equation model. In section 5.3.3, I presented
novel performance metrics that I designed, which enable us to compare the performance
of these novel methods against the standard implementations. In section 5.4, I showed
that novel EBM methods perform better than the standard EBM, while the novel DEM
methods performs equally well to the standard method on those datasets. This also
suggested that the novel metrics that we proposed are sensitive to these small changes in
the EBM and DEM methodologies.
In chapter 6 I presented Data-Driven Inference of Vertexwise Evolution (DIVE), a
novel spatiotemporal disease progression model of brain pathology in neurodegenerative
disorders. In section 6.2 I first reviewed existing literature and motivated the need for such
a model, due to the presence of dispersed atrophy patterns in AD caused by disruption
in underlying brain connectomes [38]. I then presented the mathematical formulation of
DIVE in section 6.3. I performed simulations to show that DIVE can reliably estimate
cluster assignments, trajectory parameters and subject time-shifts in the presence of
ground truth (section 6.3.8). Afterwards, I tested DIVE on four different datasets with
distinct diseases (typical AD and PCA) and modalities (MRI and PET), and showed
that it can recover meaningful patterns of pathology, which agree with previous findings
in the literature, but offer us more spatial resolution, along with estimates of biomarker
dynamics and subject-specific time shifts. Finally, in section 6.4.3 I validated DIVE
by showing that the estimated clusters and trajectories are robust under 10-fold cross-
validation, and that it has favourable predictive performance compared to simpler models.
In chapter 7 I presented Disease Knowledge Transfer (DKT), a novel model that
robustly learns patterns of progression from several types of dementia combined. This
allows the inference of biomarker signatures in rare, atypical types of dementia, which
is otherwise difficult due to the lack of multimodal, longitudinal data. In section 7.4, I
presented the DKT framework, which I designed to be flexible, allowing one to plug-in any
disease progression model within each disease-agnostic and disease-specific unit. Using
simulations, I then showed in section 7.5.1 that DKT can accurately estimate biomarker
trajectories in two distinct diseases, and even when there is a lack of data for one of the
diseases, through correlations with other known markers. When applied to patient data
(section 7.5.2), I showed that DKT can estimate plausible biomarker trajectories, and
showed that is has favourable performance compared to standard models. Compared to
previous deep transfer learning approaches, DKT is also interpretable and can predict
the future evolution of subjects at risk of neurodegenerative diseases.
In chapter 8, I presented the design of the TADPOLE Challenge, which aims to
identify algorithms and features that can best predict the progression of subjects at risk
of AD. The challenge was organised jointly by myself and my collaborators, and we
had 33 international teams who made more than 90 submissions. For the challenge, I
helped write the website, assembled the main training dataset, built a live leaderboard
system that allowed instant evaluation of the predictions, and promoted the competition
at various conferences. I also wrote the paper describing the design of the challenge [236].
9.2. Future Research Directions 151

9.2 Future Research Directions


There are several future research directions that can be pursued after this work. In section
9.2.1, I will present further applications of the methods I developed to neurodegenerative
diseases, while in section 9.2.3 I will provide suggestions of improvements to the methods
developed, along with ideas for new methods.

9.2.1 Applications to Neurodegenerative Diseases


The application of the models we developed to different neurodegenerative diseases is
important for several key reasons. First of all, they allow us to understand underlying
mechanisms underpinning phenotypic heterogeneity within PCA and the other diseases,
which can provide more informed drug targets. Secondly, they enable better stratification
and selection of endpoints for clinical trials. Third, they can be used to inform health
policy, by predicting the future evolution of subjects who are at risk of developing such
diseases.

9.2.1.1 Posterior Cortical Atrophy


There are several further questions that need to be answered regarding the progression
of Posterior Cortical Atrophy versus typical Alzheimer’s Disease. To continue the work
presented in chapter 4, one can answer the following questions1 :
• Differences in sub-populations: Are there differences in the estimated biomarker
ordering of abnormality and trajectories for various sub-populations, such as APOE
4 positive vs negative or amyloid positive vs negative?

• Imaging predicting cognition: If we split the PCA population based on the discrep-
ancy between occipital-hippocampal values at baseline, does that predict distinct
patterns of cognitive impairment? One can hypothesise that relatively lower occip-
ital volumes for basic visual-PCA predict early visual deficits, with memory deficits
later on. On the other hand, relatively lower hippocampal volumes would predict
early multi-domain cognitive deficits, with visual deficits later on.

• Relationship between posterior and anterior patterns of atrophy: Does greater in-
ferior posterior atrophy predict greater inferior anterior atrophy, and vice-versa?
Moreover, based on the cognitively-defined subgroups, is atrophy in dorsolateral
prefrontal lobe different in the three cognitive subgroups, in the following manner:
(highest) space > object > vision (lowest)? Similarly, is inferior prefrontal atrophy
different between the three subgroups in the following manner: (highest) object >
space > vision (lowest)?

• Asymmetry analysis: Are the PCA patterns of atrophy asymmetric? Previous


analyses suggested relatively greater atrophy in the right superior parietal lobe,
but this may not be the case for all patients, and cognitive tests suggest at least a
minority have left-predominant atrophy.
The above questions can explored not only using the EBM and DEM, but also using
DIVE and DKT models. Moreover, another research direction would be to apply the
1
The last three questions have been suggested by Sebastian Crutch
152 Chapter 9. Conclusions

models on biomarkers other than MRI brain volumes, such as cortical thickness from
MRI, PET biomarkers (amyloid, tau, FDG), as well as DTI biomarkers (FA, MD, AD).
The multimodal biomarker trajectories estimated in PCA with the EBM, DEM and DIVE
models can also be compared with the ones inferred by DKT.

9.2.1.2 Typical Alzheimer’s disease


Several analyses can also be done to further understand typical Alzheimer’s disease.
For example, using DIVE one could test if the reason why disconnected vertices cluster
together is due to underlying structural or functional connections. Such a hypothesis
could be tested by computing the modularity (or other index of connectivity) of DIVE
clusters on the weighted graph of white-matter connections between different vertices,
where the weights are given by the number of tracts connecting the two vertices. The
modularity coefficient could then be compared against a well-defined null hypothesis.
This would help understand to what extent disruption of underlying connectomes affects
neurodegeneration [38].

9.2.1.3 Familial Alzheimer’s disease


The models and techniques we have developed here can also be applied to familial AD,
using cohorts such as the Dominantly Inherited Alzheimer’s Network (DIAN). So far, the
EBM and DEM models have been applied to study familial AD [30], but more complex
models are yet to be tested.
Some adaptation of our the models should be done when modelling familial AD. As
opposed to sporadic AD, in familial AD we have a relatively reliable estimate of the
subjects’ disease onset based on familial age of onset. Therefore, models such as DIVE
and DKT should be adapted by setting a stronger prior distribution on the subjects’
time-shift, centred on their parental age of onset, and having a standard deviation of
around +/- 5 years, like the approach of [30]. On the other hand, familial AD cohorts
can also be used for model validation, by comparing the subjects’ estimated time-shifts
against the parental age of onset, this time using an uninformative prior on the subjects’
time-shift.

9.2.1.4 Other Alzheimer’s variants


The EBM, DEM, DIVE and DKT methodologies can be further applied to other types
of Alzheimer’s variants, such as the focal temporal lobe dysfunction, pure-amnestic AD
with episodic memory impairment [269] or language variant AD [270, 271, 13].

9.2.1.5 Frontotemporal dementia


Our models can also be applied to study the progression of Fronto-temporal dementia.
Fronto-temporal dementia (FTD) is a clinically and pathologically heterogeneous group
of non-Alzheimer’s dementias that affect frontal and temporal lobes [272]. There are three
main clinical syndromes: behavioural-variant FTD characterised by behavioural changes,
primary progressive aphasia characterised by impaired speech, and semantic dementia,
characterised by impaired semantic memory [272]. FTD also has a strong genetic compo-
nent, due to mutations in the microtubule associated protein tau (MAPT), progranulin
9.2. Future Research Directions 153

(GRN) and C9ORF72 genes. So far, an extension of the event-based model, which es-
timates multiple progression patterns in sub-populations, has been applied to FTD [29].
Applying spatiotemporal models such as DIVE or multi-disease models such as DKT
would help understand the heterogeneity and progression of FTD, find early biomarkers
and allow better stratification in FTD clinical trials. Moreover, the heterogeneity present
in FTD, combined with genetic information, can be used to further validate the DKT
model by checking how robustly it can transfer biomarker trajectories between different
FTD genetic groups.

9.2.1.6 Multiple Sclerosis


Another disease where we can apply our models is Multiple Sclerosis (MS), which is a
chronic autoimmune, inflammatory neurological disease that attacks myelinated axons
and causes neurodegeneration [273]. MS can be of several types: (a) relapsing-remitting
MS, marked by alterating periods of relapses (i.e. exacerbations of symptoms) and remis-
sion, (b) primary-progressive MS characterised by gradual worsening of symptoms, (c)
secondary progressive MS where progressive MS develops in relapsing-remitting patients
and (d) progressive-relapsing MS characterised by progressive disease with intermittent
flare-ups of worsening symptoms [273]. When applying our models to MS data, some care
needs to be taken as all our models assume monotonic biomarker trajectories. If there are
biomarkers that are non-monotonic, the models can be extended to use non-parametric
trajectories that enable non-monotonicity. However, this requires stronger priors on the
subject-specific time-shifts and on the noise variable, otherwise the model will not be
identifiable. In terms of the disease progression models applied on MS, so far the event-
based model has been applied by Eshaghi et al. [274]. However, other data-driven disease
progression models are yet to be tested on MS.

9.2.1.7 Parkinson’s Disease


Our models can also be applied to study the progression of Parkinson’s disease (PD). PD
is a neurodegenerative disease characterised by atrophy in the substantia nigra, dopamine
deficiency and aggregates of α-synuclein [275]. While the disease is diagnosed clinically
based on bradykinesia and other motor features, PD also causes multi-domain cognitive
decline [276, 275]. It would be particularly useful to apply our models to estimate the
progression of PD, help distinguish between PD and other types of degenerative parkin-
sonism, and also to identify early markers in prodromal disease stages, allowing novel
disease-modifying therapies to be started as early as possible [275].

9.2.1.8 Huntington’s Disease


Our models can also be applied to study the progression of Huntington’s disease (HD).
HD is a rare neurodegenerative disease characterised by jerky, involuntary movements, be-
havioural and psychiatric disturbances [277]. The disease is caused by an elongated CAG
repeat (36 repeats or more) on the short arm of chromosome 4p16.3 in the Huntingtin
gene, with longer repeats causing earlier onset of disease [277]. While different types of
MRI [278] and PET [279] imaging modalities have been central in identifying structural
and functional abnormalities, there is still a need to identify quantitative biomarkers for
early disease detection and for mapping its evolution[278]. In terms of data-driven disease
progression models, only the event-based model has been applied to HD [23, 280].
154 Chapter 9. Conclusions

9.2.2 Applications to Clinical Trials


The disease progression models that we have developed can be used to aid clinical trials
in several ways. One area of application is for selecting the right subjects for enrolling
in the clinical trial. For example, based on a few initial measurements such as cognitive
tests and an MRI scan, our models can predict which subjects will develop dementia,
along with the exact type of dementia, within a certain time window. This can help
select a homogenous group of patients for a clinical trial, which are otherwise estimated
to develop the same type of dementia and at the same age/follow-up time.
Another key area of application is evaluation of the effects of putative drugs. The
subject-specific time shifts, estimated based on multimodal imaging measures, could pro-
vide more robust measures of disease stage compared to single imaging or cognitive mark-
ers. Finally, our models could also be applied as a safety endpoint in clinical trials, where
they could detect very early changes that might be due to adverse effects of a drug, before
the appearance of symptoms. Such early detection of adverse effects could be used to
suggest the interruption of the clinical trial [281].

9.2.3 Methodological Developments


Further research can also focus on improvements in the models that I have developed,
along with development of new models. Such methodological improvements are very
important, as they enable understanding more complex disease mechanisms such as as-
sociations with genetics, and will enable more accurate predictions resulting in improved
stratification for clinical trials. In the following sections I will present several key direc-
tions for further work, and will suggest concrete steps towards them.

9.2.3.1 Towards Personalised Predictions


One key direction of these models is to enable them to perform personalised predictions,
which will further enable personalised treatments to be delivered. To enable this in
models such as DIVE or DKT, one can estimate subject-specific trajectories by adding
random effects to the population trajectory. This will enable more accurate predictions
and account for the heterogeneity in the modelled diseases. However, more longitudinal
data is required for such personalised predictions, and model identifiability needs to be
ensured.
Another extension to our models that can aid personalised predictions is to model
distinct progressions for different sub-populations, in a data-driven way as done by the
SuStaIn model [29]. More precisely, one can assume that the population is made of
unknown subgroups with different progressions, and each subject will have an associated
latent variable denoting the subgroup it belongs to. This can still be optimised with the
Expectation-Maximisation framework [195].
While estimating discrete subgroups with different progressions works well for some
diseases such as FTD due to mutations in a few key genes, this might not work that well
for diseases such as PCA or Huntington’s, where it is believed that there is a continuum
of phenotypic variability. In this case, our models should be extended to estimate a
continuous latent dimension of heterogeneity. This can be further extended to more
than one latent dimension, to account for mixed pathologies, where each dimension could
correspond to a different underlying pathology. While some studies have shown that a
large proportion of healthy individuals and patients who receive a clinical diagnosis of AD
9.2. Future Research Directions 155

actually have underlying mixed pathologies [282, 283], this analysis requires both in-vivo
longitudinal data along with post-mortem pathological confirmation. This has recently
become available in ANDI, which now has autopsy data for 56 AD and 52 age-matched
controls [284].

9.2.3.2 Spatio-temporal Modelling


The spatio-temporal DIVE model we proposed can be further extended to cluster points
on the brain based on a multi-modal signature. This could be done by extending the
likelihood model from a sigle univariate Gaussian distribution to a multivariate Gaussian
distribution with a small covariance matrix. Parameter estimation can still be done using
the Expectation-Maximisation framework. Such multimodal clusters could give further
insights into the mechanisms underpinning Alzheimer’s disease and other neurodegener-
ative diseases.
The DKT model that we proposed can also be extended to estimate spatio-temporal
changes in the brain. Such a spatiotemporal DKT model would be able to synthesize,
based on e.g. a structural MRI scan, other types of scans such as PET or DTI, in patients
with rare dementias, where there is a lack of such multimodal data. This could be done
using deep learning methods, where the neural network could have, for each brain region
independently, a shared 3D disease agnostic unit emcompassing multimodal pathology
across all diseases modelled. These shared units will then be used to estimate disease-
specific dynamics by redirecting the training data along different pipelines. The disease
specific parts could be implemented in an unsupervised manner (e.g. with autoencoder)
or in a supervised manner, to predict e.g. cognitive tests.

9.2.3.3 Modelling Disease Mechanisms


One potential direction of research towards understanding underlying disease mechanisms
is to model the dynamics of pathogenic proteins. The work of Raj et al. [6, 199] and
Georgiadis et al. [223] can be used as a starting point. Several concrete steps would be to
extend the network diffusion model [199] to estimate latent subject-specific time-shifts as
in the work of Donohue et al. [3]. The model by [223], while simulating far more complex
dynamics, needs to be validated using in-vitro studies, as well as using amyloid and tau
PET imaging.
One limitation of the diffusion models developed so far [199, 223] is that they assume
a static structural connectome throughout the disease process. To this end, these models
should be extended to account for changes in the connectome structure over the disease
time-course, such as breakdown of key links or nodes, based on different kinds of selective
vulnerabilities, e.g. those suggested by Zhou et al. [126].

9.2.3.4 Incorporating genetic data


Another key direction of further research is to connect our models with genetic data.
In ADNI, genetic data is available to perform genome-wide association studies (GWAS).
In particular, GWAS can help identify novel loci and genes involved in AD by finding
associations with more robust and quantitative endophenotypes derived from imaging
and other types of biomarkers. For example, very recent work by Sclesi et al. [247]
has used the disease progression model by Donohue et al. [3] to estimate a quantitative
multimodal endophenotype from MRI and PET images, which identified a novel locus.
156 Chapter 9. Conclusions

Such associations were not significant for simpler hippocampal volume or cortical amyloid
markers on their own [247]. Extending such work by adding other types of biomarker
data available in ADNI can identify further loci. Moreover, associations can also be found
between genes and various regions in the brain, and even with pathology identified at
voxelwise level from DIVE, using an approach similar to [285].

9.2.3.5 Incorporating data from other datasets


Some of our models can be further extended to incorporate data from other dementia
or normal ageing datasets. This can be done in a manner similar to DKT, but other
transfer-learning approaches can also be used to this end. Some datasets that can be
used include observational studies for sporadic AD (e.g. AIBL [286]) familial AD (e.g.
DIAN [287]), Multiple Sclerosis, Parkinson’s disease (PPMI [288]), Huntington’s disease
(TRACK HD [289]). For normal ageing, datasets such as the Rotterdam study [290] can
also be used.
Our models can also be further extended to use novel biomarker data from wearables
or internet of things (IoT) devices. Data from smart watches or body sensors [291], eye-
tracking devices [292] or speech recorders [293] can be sensitive to dementia, and thus
can be used to identify early signs and track its progression.

9.2.3.6 Estimating better features


An interesting direction of further research is to extract better features for scalar disease
progression models such as EBM, DEM or DKT. This can be done by incorporating
feature learning as part of the model itself, as in the case of DIVE. However, other
methods based on dimensionality reduction using Principal Component Analysis or t-
distributed Stochastic Neighbour Embedding (t-SNE) [294] can also be used to extract
more robust features by projecting the high-dimensional data into a lower-dimensional
space. We hypothesize that models with extracted features could perform better than
vanilla models becuase it is harder for them to overfit the data. Finally, deep learning
approaches can also be used to learn complex non-local features from images, while also
modelling the progression of the disease.

9.2.4 Model Evaluation


Further work can also be done on model evaluation, by extending the TADPOLE Chal-
lenge. While submissions mostly used extracted features from imaging data, more com-
plex non-local features can be extracted and used by spatiotemporal models or deep
learning methods. The TADPOLE Challenge can also be extended to attempt to predict
other types of target variables such as PET or DTI markers. Yet another direction is to
organise a competition similar to TADPOLE for AD related dementias, such as Posterior
Cortical Atrophy, Frontotemporal Dementia, Huntington’s disease, Parkinson’s disease
and Multiple Sclerosis.
Another different research direction in model evaluation is to evaluate models based
on simulated brain images, with more reliable ground truth compared to patient data.
Simulators based on biophysical principles such as [295] can be used to this end, which
generate realistic brain MRI images for a given spatial pattern of atrophy. Models can
be evaluated in several tasks: prediction of future biomarkers or spatial structure at both
population-level and subject-level, differential diagnosis, as well as disease staging.
Appendix A

Longitudinal Neuroanatomical
Progression of Posterior Cortical
Atrophy

Description of the labels shown in figure A.1:


• Frontal Lobe (FL)

– Lateral Surface
∗ Frontal Pole (FRP)
∗ Superior Frontal Gyrus (SFG)
∗ Middle Frontal Gyrus (MFG)
∗ Opercular part of the Inferior Frontal Gyrus (OpIFG)
∗ Orbital part of the Inferior Frontal Gyrus (OrIFG)
∗ Triangular part of the Inferior Frontal Gyrus (TrIFG)
∗ Precentral Gyrus (PrG)
– Medial Surface
∗ Superior Frontal Gyrus, medial segment (MSFG)
∗ Supplementary Motor Cortex (SMC)
∗ Medial Frontal Cortex (MFC)
∗ Gyrus Rectus (GRe)
prcs-med
MP
MP

SMC
rG
splen-ant cs-med

SFG SPL
oG

genu-post

MSFG
MCgG MFG PrG PoG SMG
PCu AnG SOG
PCgG

ant

ACgG FRP
mhos-ant

occ-

OpIFG
mhos-ant

OCP Cun MOG OCP


Calc TrIFG
calc-pos

calc-ant STG
calc-pos

LiG MFC LOr OrIFG


SCA G
GRe
pos-ant

hip-ant

PHG MTG IOG


TMP
TMP

OFuG Ent
FuG
ITG ITG
IOG
STG

Figure A.1: Labels of the different areas analysed in the EBM progression snapshots from
chapter 4.
158 Appendix A. Longitudinal Neuroanatomical Progression of PCA

∗ Subcallosal Area (SCA)


∗ Precentral Gyrus (MPrG)
– Inferior Surface
∗ Anterior Orbital Gyrus (AOrG)
∗ Medial Orbital Gyrus (MOrG)
∗ Lateral Orbital Gyrus (LOrG)
∗ Posterior Orbital Gyrus (POrG)
– Opercular Region
∗ Frontal Operculum (FO)
∗ Central Operculum (CO)
∗ Parietal Operculum (PO)
– Insular Region
∗ Anterior Insula (AIns)
∗ Posterior Insula (PIns)

• Temporal Lobe (TL)

– Lateral Surface
∗ Temporal Pole (TMP)
∗ Superior Temporal Gyrus (STG)
∗ Middle Temporal Gyrus (MTG)
∗ Inferior Temporal Gyrus (ITG)
– Supratemporal Surface
∗ Planum Polare (PP)
∗ Transverse Temporal Gyrus (TTG)
∗ Planum Temporal (PT)
– Inferior Surface
∗ Fusiform Gyrus (FuG)

• Parietal lobe (PL)

– Lateral Surface
∗ Postcentral Gyrus (PoG)
∗ Supramarginal Gyrus (SMG)
∗ Superior Parietal Lobule (SPL)
∗ Angular Gyrus (AnG)
– Medial Surface
∗ Postcentral Gyrus, medial segment (MPoG)
∗ Precuneus (PCu)

• Occipital Lobe (OL)

– Lateral Surface
159

∗ Superior Occipital Gyrus (SOG)


∗ Inferior Occipital Gyrus (IOG)
∗ Middle Occipital Gyrus (MOG)
∗ Occipital Pole (OCP)
– Inferior Surface
∗ Occipital Fusiform Gyrus (OFuG)
– Medial Surface
∗ Cuneus (Cun)
∗ Calcarine Cortex (Calc)
∗ Lingual Gyrus (LiG)

• Limbic Cortex (LC)

– Cingulate Cortex
∗ Anterior cingulate gyrus (ACgG)
∗ Middle cingulate gyrus (MCgG)
∗ Posterior cingulate gyrus (PCgG)
– Medial Temporal Cortex
∗ Parahippocampal Gyrus (PHG)
∗ Entorhinal Area (Ent)
160 Appendix A. Longitudinal Neuroanatomical Progression of PCA

Posterior Cortical Atrophy

Typical Alzheimer’s Disease

Figure A.2: Bootstrap samples of the atrophy sequence as estimated by the event-based
model, for the PCA and typical AD cohorts. The maximum likelihood sequences were
estimated using the EBM from 100 bootstrap datasets, with replacement, stratified by
diagnosis.
161

Posterior Cortical Atrophy

Typical Alzheimer’s Disease

Figure A.3: Hypothesis testing of ordering of events within PCA (top) and typical AD
(bottom). We sampled 10,000 sequences from the EBM posterior using MCMC sampling
and only kept every 1/100 in order to remove correlation between samples. We applied
the non-parametric paired Wilcoxon signed rank test for every pair of biomarkers (x,y).
The null hypothesis is defined as H0: event A (Y-axis) becomes abnormal at the same
time as event B (X-axis), while the alternative hypothesis H1: event A (Y-axis) become
abnormal before event B (X-axis). The black squares show the pair of biomarkers where
the null hypothesis was rejected at alpha=0.05/(N*(N-1)/2), thus surviving Bonferroni
correction.
162 Appendix A. Longitudinal Neuroanatomical Progression of PCA

Basic visual impairment group

Space perception impairment group

Object perception impairment group

Figure A.4: Positional variance diagram estimated by the event-based model, for the
three PCA sugroups: Basic visual impairment group, Space perception impairment and
Object perception impairment.
163

Basic visual impairment group

Space perception impairment group

Object perception impairment group

Figure A.5: Bootstrap samples of the atrophy sequence as estimated by the event-based
model, for the three PCA sugroups: Basic visual impairment group, Space perception
impairment and Object perception impairment.
164 Appendix A. Longitudinal Neuroanatomical Progression of PCA

Basic visual impairment group

Space perception impairment group

Object perception impairment group

Figure A.6: Hypothesis testing of the ordering of events within the three PCA subgroups.
Hypothesis tests were designed as in A.3
165

Region Whole Br. Ventricles Hippo. Entorh. Occipital Temporal Frontal Parietal
Whole Brain - - - - - - - -
Ventricles 1.74e-04* - - - - - - -
Hippo. 1.20e-02 4.95e-02 - - - - - -
Entorhinal 1.61e-12* 1.27e-06* 5.29e-10* - - - - -
Occipital 7.93e-03 4.16e-06* 1.20e-04* 9.44e-12* - - - -
Temporal 2.66e-01 1.17e-02 3.12e-01 5.90e-10* 1.81e-03 - - -
Frontal 9.58e-01 1.57e-04* 1.07e-02 1.52e-12* 8.68e-03 2.49e-01 - -
Parietal 3.45e-04* 1.31e-08* 2.52e-07* 3.17e-15* 8.84e-01 7.91e-05* 4.08e-04* -

Table A.1: Statistical testing for significant differences in volumes of different brain re-
gions of PCA subjects at -10 years before reference t0 . Shown here are p-values from
two-tailed t-tests. (*) Statistically significant differences at significance level = 1.78e-3,
Bonferroni corrected for all 28 comparisons.

Region Whole Br. Ventricles Hippo. Entorh. Occipital Temporal Frontal Parietal
Whole Brain - - - - - - - -
Ventricles 1.52e-16* - - - - - - -
Hippo. 6.03e-13* 8.95e-06* - - - - - -
Entorhinal 4.78e-14* 5.66e-01 9.60e-04* - - - - -
Occipital 1.32e-06* 3.17e-17* 1.45e-14* 5.25e-16* - - - -
Temporal 3.57e-01 1.75e-16* 5.22e-13* 2.90e-14* 1.66e-05* - - -
Frontal 7.31e-12* 1.67e-04* 7.72e-01 4.38e-03 1.62e-14* 3.50e-12* - -
Parietal 1.53e-07* 1.41e-21* 3.30e-19* 2.68e-19* 2.20e-01 8.33e-06* 3.39e-18* -

Table A.2: Statistical testing for significant differences in volumes of different brain re-
gions of PCA subjects at t0 . See Supp. Table A.1 for information on statistical testing.

Region Whole Br. Ventricles Hippo. Entorh. Occipital Temporal Frontal Parietal
Whole Brain - - - - - - - -
Ventricles 5.97e-01 - - - - - - -
Hippo. 7.63e-13* 4.14e-13* - - - - - -
Entorhinal 5.88e-11* 2.34e-11* 1.23e-03* - - - - -
Occipital 4.04e-02 1.44e-01 3.00e-17* 1.06e-15* - - - -
Temporal 2.83e-03 1.22e-02 1.51e-15* 2.66e-14* 1.54e-01 - - -
Frontal 8.90e-15* 5.73e-15* 6.77e-02 7.35e-07* 2.19e-19* 2.99e-17* - -
Parietal 7.38e-02 2.07e-01 1.25e-14* 4.00e-13* 9.44e-01 1.73e-01 1.91e-16* -

Table A.3: Statistical testing for significant differences in volumes of different brain re-
gions of PCA subjects at 10 years after t0 . See Supp. Table A.1 for information on
statistical testing.
166 Appendix A. Longitudinal Neuroanatomical Progression of PCA

Region Whole Br. Ventricles Hippo. Entorh. Occipital Temporal Frontal Parietal
Whole Brain - - - - - - - -
Ventricles 9.26e-03 - - - - - - -
Hippo. 2.04e-10* 2.88e-14* - - - - - -
Entorhinal 2.21e-02 3.40e-01 6.82e-09* - - - - -
Occipital 3.38e-03 2.01e-06* 3.84e-04* 9.98e-04* - - - -
Temporal 3.93e-01 1.04e-01 4.72e-11* 7.64e-02 7.51e-04* - - -
Frontal 8.30e-01 1.04e-02 4.79e-09* 2.41e-02 1.15e-02 3.26e-01 - -
Parietal 4.94e-03 2.13e-06* 3.75e-05* 4.63e-04* 7.35e-01 8.64e-04* 1.57e-02 -

Table A.4: Statistical testing for significant differences in volumes of different brain re-
gions of tAD subjects at -10 years before t0 . See Supp. Table A.1 for information on
statistical testing.

Region Whole Br. Ventricles Hippo. Entorh. Occipital Temporal Frontal Parietal
Whole Brain - - - - - - - -
Ventricles 3.50e-11* - - - - - - -
Hippo. 4.12e-19* 3.61e-25* - - - - - -
Entorhinal 7.83e-02 2.64e-10* 2.93e-13* - - - - -
Occipital 7.65e-02 2.51e-08* 7.95e-10* 7.94e-01 - - - -
Temporal 6.29e-03 4.63e-13* 1.84e-14* 5.12e-01 8.07e-01 - - -
Frontal 2.01e-04* 2.65e-04* 4.16e-20* 2.84e-05* 1.98e-04* 3.31e-07* - -
Parietal 3.56e-03 6.00e-11* 1.94e-10* 2.45e-01 4.77e-01 4.81e-01 2.12e-06* -

Table A.5: Statistical testing for significant differences in volumes of different brain re-
gions of tAD subjects at t0 . See Supp. Table A.1 for information on statistical testing.

Region Whole Br. Ventricles Hippo. Entorh. Occipital Temporal Frontal Parietal
Whole Brain - - - - - - - -
Ventricles 2.92e-02 - - - - - - -
Hippo. 2.83e-01 1.67e-03* - - - - - -
Entorhinal 8.13e-03 6.50e-01 2.63e-04* - - - - -
Occipital 8.40e-02 9.92e-01 1.13e-02 7.28e-01 - - - -
Temporal 2.76e-12* 1.46e-14* 5.41e-11* 6.35e-15* 2.54e-10* - - -
Frontal 1.24e-09* 1.91e-06* 3.87e-11* 4.81e-06* 1.27e-04* 2.43e-19* - -
Parietal 7.92e-01 7.53e-02 1.98e-01 2.51e-02 1.57e-01 5.36e-11* 3.13e-08* -

Table A.6: Statistical testing for significant differences in volumes of different brain re-
gions of tAD subjects at 10 years after t0 . See Supp. Table A.1 for information on
statistical testing.
167

Region T1 = -10 years T2 = 0 years T3 = 10 years


Whole Brain 2.52e-02 4.45e-05* 4.19e-12*
Ventricles 3.74e-05* 3.06e-05* 2.18e-13*
Hippocampus 6.15e-15* 1.13e-23* 5.83e-04*
Entorhinal 5.28e-08* 7.71e-11* 2.68e-03
Occipital 5.72e-01 1.13e-05* 2.44e-12*
Temporal 2.48e-02 1.21e-02 4.02e-11*
Frontal 2.91e-02 6.26e-03 2.12e-01
Parietal 4.22e-01 2.95e-07* 8.33e-12*

Table A.7: Statistical testing for significant differences in volumes of different brain re-
gions between PCA and tAD at -10, 0 and 10 years from t0 . Shown here are p-values from
two-tailed t-tests. (*) Statistically significant differences at significance level = 2.08e-3,
Bonferroni corrected for all 28 comparisons.
168 Appendix A. Longitudinal Neuroanatomical Progression of PCA

Figure A.7: Testing for statistically significant differences in positions of each biomarker
in the EBM abnormality sequences, for both PCA and typical AD. (*) Statistically sig-
nificant differences in position of a biomarker in the EBM sequences for PCA and tAD at
99% confidence, Bonferroni corrected for multiple comparisons (significance level = 5e-5).
A non-parametric Mann-Whitney U test has been applied because of non-gaussianity of
the data, which represents discrete ranks in a sequence. Most biomarkers show significant
differences – it is likely that there are differences in atrophy progression between PCA
and tAD.
169

(a) Vision vs Object subgroups

(b) Object vs Space subgroups

(c) Space vs Vision subgroups

Figure A.8: Testing for statistically significant differences in biomarker positions in the
EBM sequences of PCA subgroups for (a) Object vs Visual (b) Space vs Object and (c)
Visual vs Space subgroups. Only the first 10 biomarkers from the EBM sequence of one
disease (A – visual B – object C – space) are shown. The images also show only the
first 20 positions on the x-axis to aid visualisation. (*) Statistically significant differences
in biomarker positions between pairs of PCA subgroups at 99% confidence, Bonferroni
corrected for multiple comparisons (significance level = 5e-5). A non-parametric Mann-
Whitney U test has again been applied because of data non-gaussianity. All biomarkers
show significant differences – it is likely that there are differences in the progression of
atrophy between PCA subgroups.
170 Appendix A. Longitudinal Neuroanatomical Progression of PCA
Appendix B

DIVE: A Spatiotemporal
Progression Model of Brain
Pathology in Neurodegenerative
Disorders
172 Appendix B. DIVE: A Spatiotemporal Progression Model of Brain Pathology

B.1 Simulations - Error in Estimated Trajectories


and DPS

A B

C D

Figure B.1: Error in DPS scores (A) and trajectory estimation (B) for Scenario 2 in
simulation experiments. (C-D) The same error scores for Scenario 3. We notice that
as the problem becomes more difficult, the errors in the DIVE estimated parameters
increase. Errors were measured as sum of squared differences (SSD) between the true
parameters and estimated parameters. For the trajectories, the SSD was calculated only
based on the sigmoid centres, due to different scaling of the other sigmoidal parameters.

B.2 Comparison Between DIVE and Other Models


B.2.1 Motivation
We were also interested to compare the performance of DIVE with other disease progres-
sion models. In particular, we were interested to test whether:
• Modelling dynamic clusters on the brain surface improves subject staging and
biomarker prediction
• Modelling subject-specific stages with a linear transformation (the αi and βi terms)
improves biomarker prediction
B.3. Derivation of the Generalised EM Algorithm 173

B.2.2 Experiment Design


We compared the performance of our model to two simplified models:

• ROI-based model: groups vertices according to an a-priori defined ROI atlas. This
model is equivalent to the model by Jedynak et al., Neuroimage, 2012 and is a
special case of our model, where the latent variables zlk are fixed instead of being
marginalised as in equation 6.

• No-staging model: This is a model that doesn’t perform any time-shift of patients
along the disease progression timeline. It fixes αi = 1, βi = 0 for every subject,
which means that the disease progression score of every subject is age.

We performed this comparison using 10-fold cross-validation. For each subject in the
test set, we computed their DPS score and correlated all the DPS values with the same
four cognitive tests used previously. We also tested how well the models can predict
the future vertex-wise measurements as follows: for every subject i in the test set, we
used their first two scans to estimate αi = 1, βi = 0 and then used the rest of the
scans to compute the prediction error. For one vertex location on the cortical surface,
the prediction error was computed as the root mean squared error (RMSE) between its
predicted measure and the actual measure. This was then averaged across all subjects
and visits.

B.2.3 Results
Table B.1 shows the results of the model comparison, on ADNI MRI dataset. Each
row represents one model tested, while each column represents a different performance
measure: correlations with four different cognitive tests and accuracy in the prediction of
future vertexwise measurements. In each entry, we give the mean and standard deviation
of the correlation coefficients or RMSE across the 10 cross-validation folds.

Model CDRSOB (ρ) ADAS13 (ρ) MMSE (ρ) RAVLT (ρ) Prediction (RMSE)
DIVE 0.37 +/- 0.09 0.37 +/- 0.10 0.36 +/- 0.11 0.32 +/- 0.12 1.021 +/- 0.008
ROI-based model 0.36 +/- 0.10 0.35 +/- 0.11 0.34 +/- 0.13 0.30 +/- 0.13 1.019 +/- 0.010
No-staging model *0.09 +/- 0.06 *0.03 +/- 0.09 *0.05 +/- 0.06 *0.02 +/- 0.06 *1.062 +/- 0.024

Table B.1: Comparison of DIVE with two more simplistic models on the ADNI MRI
dataset. For each of the three models, we show the correlation of the disease progression
scores (DPS) with respect to several cognitive tests: CDRSOB, ADAS13, MMSE and
RAVLT. The correlation numbers represent the mean correlation across the 10 cross-
validation folds.

B.3 Derivation of the Generalised EM Algorithm


We seek to calculate M (u) = arg maxM EZ|V,M (u−1) [log p(V, Z|M )] + log p(M ) where
M (u) = (α(u) , β (u) , θ(u) , σ (u) , λ(u) ) are the set of model parameters at iteration u of the
EM algorithm. Moreover, p(M (u) ) is a prior on these parameters that is chosen by the
user. Expanding the expected value, we get:
174 Appendix B. DIVE: A Spatiotemporal Progression Model of Brain Pathology

K
X
(u)
M = arg max p(Z = (z1 , ..., zL )|V, M (u−1) ) [log p(V, Z|M )] + log p(M ) (B.1)
M
z1,...,zL

The E-step involves computing p Z = (z1 , ..., zL )|V, M (u−1) , while the M-step com-
prises of solving the above equation.

B.3.1 E-step
In this step we need to estimate p(Z|V, M (u−1) ). For notational simplificy we will drop
the (u − 1) superscript from M

 
L
1 Y Y Y
p(Z|V, M ) = p(V, Z|M ) =  N (Vlij |f (αi tij + βi |θZl ), σZl ) Ψ(Zl , Zl2 )
C l l2 ∈Nl
(i,j)∈I
(B.2)
where Nl is the set of neighbours of vertex l. However, this doesn’t directly factorise
over the vertices l due to the MRF terms Ψ(Zl , Zl2 ). It is however necessary to find
a form that factorises over the vertices, otherwise we won’t be able to represent in
memory the Q joint distribution over all Z variables. If we make the approximation
p(Z|V, M ) ≈ Ll p(Vl |Zl , M ) then we loose out all the MRF terms and the model won’t
account for spatial correlation. We instead do a first-degree approximation by condition-
(u−1)
ing on the values of ZNl , the labels of nearby vertices from the previous iteration. The
approximation is thus:
L
Y h i
(u−1)
p(Z|V, M ) ≈ EZ (u−1) |V ,M p(Zl |Vl , M, ZNl ) (B.3)
Nl l
l

This form allows us to factorise over all the vertices to get p(Zl |Vl , M ):
1 X (u−1) (u−1)
p(Zl |Vl , M ) ≈ p(Vl |Zl , M )p(Zl |ZNl )p(ZNl |Vl , M ) (B.4)
C (u−1)
ZN
l

where C is aQ
normalistion constant that can be dropped. We can now further factorise
(u−1) (u−1)
p(Zl |ZNl ) ≈ m∈{1,...,Nl } p(Zl |M, ZNl (m) = zNl (m) ) and apply a similar factorisation to
(u−1)
the prior p(ZNl |Vl , M ), resulting in:

1 X Y (u−1)
p(Zl |Vl , M ) ≈ p(Vl |Zl , M ) p(Zl |ZNl (m) = zNl (m) )
C z Nl (1) ,..,zNl (|Nl |) m∈{1,...,Nl }
(u−1)
p(ZNl (m) = zNl (m) |Vl , M ) (B.5)

Factorising the summation over zNl ’s we get:

Y X (u−1) (u−1)
p(Zl |Vl , M ) = p(Vl |Zl , M ) p(Zl |Zl2 = zl2 )p(Zl2 = zl2 |Vl , M ) (B.6)
l2 ∈Nl zl2
B.3. Derivation of the Generalised EM Algorithm 175

Replacing zl2 with k2 we get:

Y X (u−1) (u−1)
p(Zl |Vl , M ) = p(Vl |Zl , M ) p(Zl |Zl2 = k2 )p(Zl2 = k2 |Vl , M ) (B.7)
l2 ∈Nl k2

We shall also denote zlk = p(Zl |Vl , M ). Further simplifications result in:

" #" K
#
(u)
Y Y X (u−1)
zlk ∝ N (Vlij |f (αi tij + βi |θk ), σk ) zl2 k2 Ψ(Zl = k, Zl2 = k2 ) (B.8)
i,j∈I l2 ∈Nl k2 =1

" #
(u) log (2πσk2 )
X 1 ij 2
log zlk ∝ − − 2 (Vl − f (αi tij + βi |θk )) +
i,j∈I
2 2σk
" K
#
X X (u−1)
+ log zl2 k2 (δk2 k exp(λ) + (1 − δk2 k ) exp(−λ2 )) (B.9)
l2 ∈Nl k2 =1

We further define the data-fit term Dlk as follows:

log (2πσk2 )|I| X 1


Dlk = − − 2
(Vlij − f (αi tij + βi |θk ))2 (B.10)
2 i,j∈I
2σk

This results in:

 
K
(u) (u−1)
X X
log zlk ∝ Dlk +  log zl2 k2 (δk2 k (exp(λ) − exp(−λ2 )) + exp(−λ2 )) (B.11)
l2 ∈Nl k2 =1

Finally, we simplify the sum over k2 to get the update equation for zlk :

" #
X h i
(u) (u−1)
log zlk ∝ Dlk + log exp(−λ2 ) + zl2 k (exp(λ) − exp(−λ2 )) (B.12)
l2 ∈Nl

In practice, we cannot naively compute the exponential term zlk = exp(log(zlk )) due
to precision loss. However, we go around this by recomputing the exponentiation and
normalisation of zlk simultaneously. Denoting x(k) = log zlk , for k ∈ [1 . . . K], we get:

ex(k) 1
zlk = = (B.13)
ex(1) + ex(2) + · · · + ex(K) ex(1)−x(k) + ex(2)−x(k) + · · · + ex(K)−x(k)

B.3.2 M-step
The M-step itself does not have a closed-form analytical solution. We choose to solve it
by successive refinements of the cluster trajectory parameters and the subject time shifts.
176 Appendix B. DIVE: A Spatiotemporal Progression Model of Brain Pathology

B.3.3 Optimising Trajectory Parameters


Trajectory shape - θ

Taking equation B.1 and fixing the subject time-shifts α, β and measurement noise
σ, we can find its maximum with respect to θ only. More precisely, we want:

K
X
θ = arg max p(Z = (z1 , ..., zL )|V, M (u−1) ) [log p(V, Z|M )] + log p(θ) (B.14)
θ z1,...,zL

We observe that for each cluster the individual θk ’s are conditionally independent,
i.e. θk ⊥⊥ Q
θm |{Z, α, β, σ} ∀k, m. We also assume that the prior factorizes for each θk :
log p(θ) = K k log p(θk ). This allows us to optimise each θk independently:

K
X
θk = arg max p(Z = (z1 , ..., zL )|V, M (u−1) ) [log p(V, Z|M )] + log p(θk ) (B.15)
θk z1,...,zL

Replacing the full data log-likelihood, we get:

K
X
θk = arg max p(Z = (z1 , ..., zL )|V, M (u−1) )
θk z1,...,zL
 
L Y
Y
log  N (Vlij |f (αi tij + βi |θzl ), σzl ) + log p(θk ) (B.16)
l=1 (i,j)∈I

Note that we didn’t include the MRF clique terms, since they are not a function of
θk . We propagate the logarithm inside the products:

K
X L X
X
θk = arg max p(Z = (z1 , ..., zL )|V, M (u−1)
) log N (Vlij |f (αi tij +βi |θzl ), σzl )+
θk z1,...,zL l=1 (i,j)∈I

+ log p(θk ) (B.17)

We next assume that Zl , the hidden cluster assignment for vertex l, is condition-
ally independent of the other vertex assignments Zm , ∀m 6= l (See E-step approxima-
tion from Eq. B.3). This independence assumption induces the following factorization:
p(Z = (z1 , ..., zL )|V, M (u−1) ) = Ll p(Zl = zl |V, M (u−1) ). Propagating this product inside
Q
the sum over the vertices, we get:

L X
X K X
θk = arg max p(Zl = zl |V, M (u−1) ) log N (Vlij |f (αi tij + βi |θzl ), σzl ) + log p(θk )
θk
l=1 zl =1 (i,j)∈I
(B.18)
The terms which don’t contain θk dissapear:
B.3. Derivation of the Generalised EM Algorithm 177

L
X X
θk = arg max p(Zl = k|V, M (u−1) ) log N (Vlij |f (αi tij + βi |θk ), σk ) + log p(θk )
θk
l=1 (i,j)∈I
(B.19)
We further expand the Gaussian noise model:

L
X
θk = arg max p(Zl = k|V, M (u−1) )
θk
l=1
X  1 ij

−1/2 2
log (2πσk ) − 2 (Vl − f (αi tij + βi |θk )) + log p(θk ) (B.20)
2σk
(i,j)∈I

Constants dissapear due to the arg max and we get the final update equation for θk :

L
X X  1 
(u−1) ij 2
θk = arg max p(Zl = k|V, M ) − 2 (Vl − f (αi tij + βi |θk )) + log p(θk )
θk
l=1
2σk
(i,j)∈I
(B.21)
Measurement noise - σ

We first assume a uniform prior on the σ parameters to simplify derivations. Using


a similar approach as with θ, after propagating the product inside the logarithm and
removing the terms which don’t contain σk , we get:

L
X X
σk = arg max p(Zl = k|V, M (u−1) ) log N (Vlij |f (αi tij + βi |θk ), σk ) (B.22)
σk
l=1 (i,j)∈I

Note that, just as for θ above, the MRF clique terms were not included because they
are not a function of σk . Expanding the noise model we get:

L X  
X
(u−1) 2 −1/2 1 ij 2
σk = arg max p(Zl = k|V, M ) log (2πσk ) − 2 (Vl − f (αi tij + βi |θk ))
σk
l=1
2σk
(i,j)∈I
(B.23)
The maximum of a function l(σk ) can be computed by taking the derivative of the
function l and setting it to zero. This is under the assumption that l is differentiable,
which it is but we won’t prove it here. This gives:

L X δ  
δl(σk |.) X (u−1) 2 −1/2 1 ij 2
= p(Zl = k|V, M ) log (2πσk ) − 2 (Vl − f (αi tij + βi |θk ))
δσk l=1
δσ k 2σk
(i,j)∈I
(B.24)
Propagating the differential operator further inside the sums we get:
178 Appendix B. DIVE: A Spatiotemporal Progression Model of Brain Pathology

L X  δ log σ 2 
δl(σk |.) X (u−1) k δ 1 ij 2
= p(Zl = k|V, M ) − − 2
(Vl − f (αi tij + βi |θk ))
δσk l=1
δσ k 2 δσk 2σk
(i,j)∈I
(B.25)
We next perform several small manipulations to reach a more suitable form of the
derivative and then set it to be equal to zero:

L X  1 
δl(σk |.) X (u−1) −2 ij 2
= p(Zl = k|V, M ) − − 3 (Vl − f (αi tij + βi |θk )) (B.26)
δσk l=1
σ k 2σ k
(i,j)∈I

L X  σ2 
δl(σk |.) X (u−1) k 1 ij 2
= p(Zl = k|V, M ) − 3 + 3 (Vl − f (αi tij + βi |θk )) (B.27)
δσk l=1
σk σk
(i,j)∈I

L
δl(σk |.) X X 
p(Zl = k|V, M (u−1) ) −σk2 + (Vlij − f (αi tij + βi |θk ))2 = 0

= (B.28)
δσk l=1 (i,j)∈I

Finally, we solve for σk and get its update equation:


L
1 X X
σk2 = p(Zl = k|V, M (u−1) ) (Vlij − f (αi tij + βi |θk ))2 (B.29)
|I| l=1
(i,j)∈I

B.3.4 Estimating Subject Time Shifts - α, β


For estimating α, β, we adopt a similar strategy as in the case of θ, up to Eq. B.18. This
gives us the following problem:

L X
K
X X 0
αi , βi = arg max p(Zl = k|V, M (u−1) ) log N (Vli j |f (αi0 ti0 j + βi0 |θk ), σk )+
αi ,βi
l=1 k=1 (i0 ,j)∈I

+ log p(αi , βi ) (B.30)


The terms αi0 , βi0 for other subjects i0 6= i dissappear:

L X
X K X
αi , βi = arg max p(Zl = k|V, M (u−1) ) log N (Vlij |f (αi tij +βi |θk ), σk )+log p(αi , βi )
αi ,βi
l=1 k=1 j∈Ii
(B.31)
Expanding the Gaussian noise model we get:

L X
X K
αi , βi = arg max p(Zl = k|V, M (u−1) )
αi ,βi
l=1 k=1
X 1 ij

log (2πσk2 )−1/2 2
− 2 (Vl − f (αi tij + βi |θk )) + log p(αi , βi ) (B.32)
j∈Ii
2σk
B.3. Derivation of the Generalised EM Algorithm 179

After removing constant terms we end up with the final update equation for αi , βi :

" L K #
XX 1 X ij
αi , βi = arg min p(Zl = k|V, M (u−1) ) 2 (Vl − f (αi tij + βi |θk ))2 −
αi ,βi
l=1 k=1
2σ k j∈I i

− log p(αi , βi ) (B.33)

B.3.5 Estimating MRF Clique Term - λ


We optimise λ using the following formula:

λ(u) = arg max Ep(Z|V,M (u−1) ,λ,Z (u−1) ) [log p(V, Z|M (u−1) )] (B.34)
λ

Note that p(Z|V, M (u−1) , λ, Z (u−1) ) is a function of λ, so for each lambda we estimate
zlk through approximate inference. We do this because otherwise the optimisation of λ
will only take into account the clique terms and completely exclude the data terms. We
further simplify the objective function for lambda as follows:

K
X
(u)
λ = arg max p(Z = (z1 , ..., zL )|V, M (u−1) , λ, Z (u−1) )
λ z1,...,zL
 
YL Y L Y
Y
log  N (Vlij |f (αi tij + βi |θzl ), σzl ) Ψ(zl , zl2 ) (B.35)
l=1 (i,j)∈I l=1 l2 ∈Nl

We take the logarithm:

K
X
λ(u) = arg max p(Z = (z1 , ..., zL )|V, M (u−1) , λ, Z (u−1) )
λ z1,...,zL
 
L
X X L
XX
 log N (Vlij |..) + log Ψ(zl , zl2 ) (B.36)
l=1 (i,j)∈I l=1 l2 ∈Nl

Let us denote zlk = p(Zl = k|V, M (u−1) , λ, Z (u−1) ). Assuming independence between
the latent variables Zl we get:

 
L X
X K X
λ = arg max zlk  log N (Vlij |..) +
λ
l=1 k=1 (i,j)∈I

X K
L X K
X X
+ zlk zl2 k log Ψ(Zl = k, Zl2 = k2 ) (B.37)
l=1 k=1 l2 ∈Nl k2 =1

However, we now want to make zlk a function of λ as previously mentioned, so zlk =


ζlk (λ), for some function ζlk . More precisely, using the E-step update from Eq. B.12 we
define for each vertex l and cluster k a function ζlk (λ) as follows:
180 Appendix B. DIVE: A Spatiotemporal Progression Model of Brain Pathology

!
X h i
(u−1)
ζlk (λ) = exp Dlk + log exp(−λ2 ) + zl2 k (exp(λ) − exp(−λ2 )) (B.38)
l2 ∈Nl

where Dlk is as defined in Eq B.10. We then replace zlk with ζlk (λ) and introduce the
chosen MRF clique model to get:

L X
X K XL X
K X X
K
(u)
ζlk (λ)ζl2 k (λ) δkk2 λ + (1 − δkk2 )(−λ2 )
 
λ = arg max ζlk (λ)Dlk +
λ
l=1 k=1 l=1 k l2 ∈Nl k2 =1
(B.39)
We separate the cliques that have matching clusters to the ones that don’t:

L X
K L X X
K
" #
X X X
λ(u) = arg max ζlk (λ)Dlk + ζlk (λ)ζl2 k (λ) λ + ζlk (λ)ζl2 k (λ)(−λ2 )
λ
l=1 k=1 l=1 l2 ∈Nl k k26=k
(B.40)
We also factorise the clique terms:

L X
X K L X X
X K
(u)
λ = arg max ζlk (λ)Dlk + λ ζlk (λ)ζl2 k (λ) +
λ
l=1 k=1 l=1 l2 ∈Nl k
L X X
X K
2
+ (−λ ) ζlk (λ)(1 − ζl2 k (λ)) (B.41)
l=1 l2 ∈Nl k

Finally, we simplify to get the objective function for λ.

L X
K
" #
X X X
λ(u) = arg max ζlk (λ) Dlk + λ ζl2 k (λ) − λ2 (1 − ζl2 k (λ)) (B.42)
λ
l=1 k=1 l2 ∈Nl l2 ∈Nl

For implementation speed-up, data-fit terms Dlk can be pre-computed.

B.4 Fast DIVE Implementation - Proof of


Equivalence
Fitting DIVE can be computationally prohibitive, especially given that the number of
vertices/voxels can be very high, e.g. more than 160,000 in our datasets. We derived a
fast implementtion of DIVE, which is based on the idea that for each subject we compute
a weighted mean of the vertices within a particular cluster, and then compare that mean
with the corresponding trajectory value. This is in contrast with comparing the value at
each vertex with the corresponding trajectory of its cluster. In the next few subsections,
we will present the mathematical formulation of the fast implementation for parameters
[θ, α, β]. Parameter σ already has a closed-form update, while parameter λ has a more
complex update procedure for which this fast implementation doesn’t work. For each
parameter, we will also provide proofs of equivalence.
B.4. Fast DIVE Implementation - Proof of Equivalence 181

B.4.1 Trajectory Parameters - θ


B.4.2 Fast Implementation
The fast implementation for θ implies that, instead of optimising Eq. B.29 we optimise
the following problem:
X
θk = arg min (< V ij >Ẑk −f (αi tij + βi |θk ))2 (B.43)
θk
(i,j)∈I

where < V ij >Ẑk is the mean value of the vertices belonging to cluster k. Math-
ematically, we define Ẑk = [z1k γk , z2k γk , . . . , zLk γk ] where γk = ( Ll=1 zlk )−1 is the
P

normalisation constant. Moreover, we have that < V ij >Ẑk = Ll=1 zlk γk V ij . We take the
P
derivative of the likelihood function lf ast of the fast implementation (Eq. B.43) with
respect to θk and perform several simplifications:
L
!2
δlf ast (θk |.) δ X X ij
= zlk γk V − f (αi tij + βi |θk ) (B.44)
δθk δθk l=1
(i,j)∈I

L
! !
δlf ast (θk |.) X X −δf (.)
= 2 γk zlk V ij − f (αi tij + βi |θk ) (B.45)
δθk l=1
δθk
(i,j)∈I

using the fact that Ll=1 γk zlk = 1 we get:


P

L
!
δlf ast (θk |.) X X −δf (.)
γk zlk V ij − f (αi tij + βi |θk )

= 2 (B.46)
δθk l=1
δθk
(i,j)∈I

L
!
δlf ast (θk |.) X −δf (.) X
ij

= 2γk zlk V − f (αi tij + βi |θk ) (B.47)
δθk δθk l=1
(i,j)∈I

By setting the derivative to zero, the optimal θ is thus a solution of the following
equation:
L
!
X −δf (.) X
zlk V ij − f (αi tij + βi |θk )

=0 (B.48)
δθk l=1
(i,j)∈I

B.4.3 Slow Implementation


We will prove that if theta is a solution of the slow implementation, it is also a solution
of Eq. B.48, which will prove that the fast implementation is equivalent. The slow
implementation is finding θ from the following equation:
L
X X
θk = arg min zlk (Vlij − f (αi tij + βi |θk ))2 (B.49)
θk
l=1 (i,j)∈I

Taking the derivative of the function above (lslow ) with respect to θk we get:
L  
δlslow (θk |.) X X ij δf (.)
= zlk 2(Vl − f (αi tij + βi |θk )) − =0 (B.50)
δθk l=1
δθ k
(i,j)∈I
182 Appendix B. DIVE: A Spatiotemporal Progression Model of Brain Pathology

After swapping terms around and using distributivity we get:

X  δf (.)  XL
− zlk (Vlij − f (αi tij + βi |θk )) = 0 (B.51)
δθk l=1
(i,j)∈I

This is the same optimisation problem as in Eq. B.48, which proves that the two
formulations are equivalent with respect to θ.

B.4.4 Noise Parameter - σ


The noise parameter σ can actually be computed in a closed-form solution for the original
slow model implementation, so there is no benefit in implementing the fast update for
σ. Moreover, the σ in the fast implementation computed the standard deviation in the
mean value of the vertices within a certain cluster, and not the deviation withing the
actual value of the vertices.

B.4.5 Subjects-specific Time Shifts - α, β


B.4.6 Fast Implementation
The equivalent fast formulation for the subject-specific time shifts is similar to the one
for the trajectory parameters. It should be noted however that we need to weight the
sums corresponding to each cluster by γk−1 . This gives the following equation for the fast
formulation:
K
X 1 X
αi , βi = arg min γk−1 2
(< Vlij >Ẑk −f (αi tij + βi |θk ))2 = 0 (B.52)
αi ,βi
k=1
2σk j∈I
i

In order to prove that this is equivalent to the slow version, we need to take the
derivative of the likelihood function (lf ast ) from the above equation with respect to αi ,
βi and set it to zero:

K
δlf ast (αi , βi |.) δ X −1 1 X
= γk 2
(< Vlij >Ẑk −f (αi tij + βi |θk ))2 = 0 (B.53)
δαi , βi δαi , βi k=1 2σk j∈I
i

We expand the average across the vertices and slide the derivative operator inside the
sums:
K L
!
−1 1 −δf (.)
X X X ij
γk 2
2 γk zlk Vl − f (αi tij + βi |θk ) (B.54)
k=1
2σk j∈I l=1
δαi , βi
i

PL
Since l=1 γk zlk = 1 we get:
K L
!
X 1 X −δf (.) X
2 γk−1 2 γk zlk (Vlij − f (αi tij + βi |θk )) (B.55)
k=1
2σk j∈I δαi , βi l=1
i

Removing the factor 2 and sliding γk :


B.4. Fast DIVE Implementation - Proof of Equivalence 183

K L
!
X 1 X −δf (.) X
γk−1 γk 2 zlk (Vlij − f (αi tij + βi |θk )) (B.56)
k=1
2σk j∈I δαi , βi l=1
i
PL
Further sliding l=1 zlk to the left we get the final optimisation problem:
K L
X 1 X X −δf (.) ij
zlk (V − f (αi tij + βi |θk )) (B.57)
k=1
2σk2 l=1 j∈I
δαi , βi l
i

B.4.7 Slow Implementation


In a similar way to the trajectory parameters, we want to prove that solving the problem
from Eq. B.57 (fast implementation) is the same as solving the original slow implemen-
tation problem, which is defined as:
L X
K
X 1 X ij
αi , βi = arg min zlk (V − f (αi tij + βi |θk ))2 (B.58)
αi ,βi
l=1 k=1
2σk2 j∈I l
i

Taking the derivative of the function above with respect to αi , βi , we get:


K L
δlslow (αi , βi |.) X 1 X X −δf (.) ij
= zlk (V − f (αi tij + βi |θk )) (B.59)
δαi , βi k=1
2σk2 l=1 j∈I
δαi , βi l
i

This is the same problem as the fast implementation one from Eq. B.57, thus the fast
model is equivalent to the slow model with respect to α, β.
184 Appendix B. DIVE: A Spatiotemporal Progression Model of Brain Pathology
Appendix C

Disease Knowledge Transfer across


Neurodegenerative Diseases
186 Appendix C. Disease Knowledge Transfer across Neurodegenerative Diseases

estimated trajectory true trajectory CTL AD


biomarker 0 biomarker 1 biomarker 2

1.0 1.0 1.0

0.5 0.5 0.5

0.0 0.0 0.0


Biomarker Value

10 0 10 10 0 10 10 0 10
biomarker 3 biomarker 4 biomarker 5

1.0 1.0 1.0

0.5 0.5 0.5

0.0 0.0 0.0

10 0 10 10 0 10 10 0 10
Disease Progression (years)

Figure C.1: Estimated biomarker trajectories for the ”synthetic AD” disease, plotted
alongside true trajectories. Estimation of the trajectories in biomarkers 0,1,4 and 5 has
been done without any data from the ”synthetic PCA” disease, only based on the disease-
agnostic correlations with biomarkers 2 and 3.
Appendix D

Novel Extensions to the Event-based


Model and Differential Equation
Model

D.1 EBM Fitting using Expectation-Maximisation


Let us assume that p(x|E) and p(x|¬E) follow normal distributions, where p(x|Ek ) ∼
N (µak , σka ) and p(x|¬Ek ) ∼ N (µnk , σkn ). Our vector of parameters is then given by θ =
[[µnk , σkn , µak , σka ]k=1..N , S] where S is the event ordering. Moreover, we further define Z =
[Z1 , Z2 , . . . , ZP ] a vector of (latent) discrete random variables representing the stage of
each subject which can take values 0 . . . N , where N is the number of biomarkers. The
dataset is denoted by X where xij represents the data from subject i for biomarker j
while Xi = [xi1 , . . . , xiN ] is a vector of biomarker data for subject i.

D.1.1 M-step
In the M-step we aim to find the arguments θ∗ that maximise the expected log-likelihood
of the complete data θ∗ = arg maxθ Q(θ|θold ).

Q(θ|θold ) = EZ|X,θold [logp(X, Z|θ)] (D.1)


Assuming a uniform prior on Z, i.e. log p(Z = z) = C and expanding Z we get:
N
X N
X
old
Q(θ|θ ) = C+ ··· p(Z1 = z1 , . . . , ZP = zP |X, θold )log p(X|Z1 = z1 , . . . , ZP = zP , θ)
z1 =0 zP =0
(D.2)
Under the EBM model, the data from each subject is conditionally independent given
the parameters, .i.e Xi ⊥ ⊥ Xj |θ and Xi ⊥⊥ Zj |θ for i 6= j. A similar independence also
holds for the latent variables Zi given the parameters. We therefore get:
P
"P #
X Y Y
Q(θ|θold ) = C + p(Zi = zi |Xi , θold ) log p(Xi |Zi = zi , θ) (D.3)
z1 ,...,zP i=1 i=1

where P is the number of patients. We further factorise the latent variables Z to


obtain:
188 Appendix D. Novel Extensions to the EBM and DEM

P X
" P
#
X Y
Q(θ|θold ) = C + p(Zi = zi |Xi , θold ) log p(Xi |Zi = zi , θ) (D.4)
i=1 zi i=1

After moving the log inside the products and removing the constant C we get:
P X
" z N
#
X X i X
Q(θ|θold ) = p(Zi = zi |Xi , θold ) log p(xij |ES(j) ) + log p(xij |¬ES(j) )
i=1 zi j=1 j=zi +1
(D.5)
Replacing p(x|E) and p(x|¬E) with the pdf of a Gaussian distribution we get:
P X
X P
X
Q(θ|θold ) = p(Zi = zi |Xi , θold )
i=1 zi i=1
" zi N
#
X X
a
log N (xij |µaS(j) , σS(j) )+ n
log N (xij |µnS(j) , σS(j) )
j=1 j=zi +1

(D.6)
The function Q(θ|θold ) is differentiable with respect to all parameters apart from S
(which is discrete). We can thus find θ∗ by solving ∇θ Q(θ|θold ) = 0. We show the
derivation for parameter µnk , which is the solution of dµdn Q(θ|θold ) = 0. Using the result
k
from equation D.6 and moving the derivation operator inside the sums we get:
P
d old
XX
Q(θ|θ ) = p(Zi = zi |Xi , θold )
dµnk i=1 zi
" z N
#
X di X d
n
log N (xij |µaS(j) , σS(j)
a
)+ n
log N (xij |µnS(j) , σS(j)
n
) = 0 (D.7)
j=1
dµk j=z +1
dµk
i

The derivative term cancels all likelihood terms apart from the one where S(j) = k:

P X
" N
#
X X d
p(Zi = zi |Xi , θold ) I[S(j) = k] n log N (xij |µnk , σkn ) = 0 (D.8)
i=1 zi j=z +1
dµk
i

which can be rewritten as:


P X
" N
#
X d X
p(Zi = zi |Xi , θold ) log N (xik |µnk , σkn ) I[j = S −1 (k)] = 0 (D.9)
i=1 zi
dµnk j=z +1 i

P X  
X
old d n n −1
p(Zi = zi |Xi , θ ) log N (xik |µk , σk )I[S (k) > zi ] = 0 (D.10)
i=1 zi
dµnk
Further rearranging the sum terms we get:

P N
X d X
n
log N (x |µ
ik k
n
, σk
n
) I[S −1 (k) > zi ] p(Zi = zi |X, θold ) = 0 (D.11)
i=1
dµ k z =0 i
D.1. EBM Fitting using Expectation-Maximisation 189

P
X d
n
log N (xik |µnk , σkn ) p(S −1 (k) > Zi |X, θold ) = 0 (D.12)
i=1
dµ k

Inserting the formula for the Gaussian pdf we get:


P
X d (xik − µnk )2
n n 2
p(S −1 (k) > Zi |X, θold ) = 0 (D.13)
i=1
dµk 2(σ k )

which results in the update rule for µnk , the mean of p(x|¬Ek )
P
X
µnk = xik win (D.14)
i=1

where
p(S −1 (k) > Zi |X, θold )
win = PP (D.15)
−1 (k) > Z |X, θ old )
i=1 p(S i

and
K
X
−1 old
p(S (k) > Zi |X, θ ) = p(Zi = l|X, θold ) (D.16)
l=S −1 (k)+1

Using a similar approach we get the update rules for σkn , µak , σka :
v
u P
uX
n
σk = t win (xik − µnk )2 (D.17)
i=1

P
X
µak = xik wia (D.18)
i=1

v
u P
uX
a
σk = t wia (xik − µak )2 (D.19)
i=1

where
p(S −1 (k) ≤ Zi |X, θold )
wia = PP (D.20)
−1 (k) ≤ Z |X, θ old )
i=1 p(S i

Solving for S in the M-step is intractable, so we use MCMC sampling where at each
step of the sampling process we propose a new sequence S new , find the optimal distribution
parameters for each biomarker given S new using the EM update rules and then evaluate
the likelihood Q(θ|θold ). The sequence and parameters that maximise the likelihood are
chosen and the EM proceeds to a new iteration. Although this approach might not
guarantee that we truly find the optimal parameters, it still results in an increase of
Q(θ|θold ). This approach, called generalised EM, still guarantees that the method will
still converge to a local maxima [195]. For parameter initialisation, we use the mean and
standard deviation of the control and patient populations.
190 Appendix D. Novel Extensions to the EBM and DEM

D.1.2 E-step
In the E-step, we simply estimate the latent disease stages Zi for every subject i. The
probability p(Zi = l|X, θold ) that subject i is at stage l in the abnormality sequence,
conditioned on the previous parameters θold , has a closed-form solution given by:

Ql a a
QN
old j=1 N (xi,s(j) |µS(j) , σS(j) ) j=l+1 log N (xi,s(j) |µnS(j) , σS(j)
n
)
p(Zi = l|X, θ ) = PK Qm a a
QN
m=0 j=1 N (xi,s(j) |µS(j) , σS(j) ) j=m+1
n
log N (xi,s(j) |µnS(j) , σS(j) )
(D.21)
Appendix E

TADPOLE Challenge: Prediction of


Longitudinal Evolution in
Alzheimer’s Disease

E.1 Expected Number of Subjects and Available


Data for D4
We estimated the number of subjects and available data in D4 (Table 8.2, last column)
using information from the ADNI procedures manual and previous ADNI rollovers. For
estimating the total number of subjects (first row) expected in D4, we computed the
dropout rate (0.36) based on ADNI1 rollovers to ADNI2, then multiplied it by the to-
tal number of subjects in D2 (896). For estimating the proportions of each diagnostic
category (third row), we used the proportion of diagnostic rates in D2 and multiplied
them with conversion rates within 1 year from ADNI1/GO/2 (see website FAQ). For
estimating the average number of visits per subject (mean ± std.) in D4 (second row),
we used the proportions for each diagnostic group and considered one visit per subject
(ADNI procedures). We set the standard deviation to be zero, although in practice this
won’t be the case.
For estimating the available biomarker data (lower half of table), we used a 1-year
time-frame from start of ADNI2 (July 2012 – July 2013) and computed the proportion of
available data in that time frame. For AV1451, we used the same estimate as for AV45,
due to the fact that the scan was introduced later on in ADNI2, and we expect more
subjects to undergo AV1451 scans in ADNI3. A Python script that computes all the data
from Table 8.2 is given in the TADPOLE repository: https://github.com/noxtoby/
TADPOLE/blob/master/statistics/tadpoleStats.py.
192 Appendix E. TADPOLE Challenge: Prediction of Longitudinal Evolution in AD
Appendix F

Bibliography

[1] Clifford R Jack, David S Knopman, William J Jagust, Leslie M Shaw, Paul S Aisen,
Michael W Weiner, Ronald C Petersen, and John Q Trojanowski. Hypothetical
model of dynamic biomarkers of the Alzheimer’s pathological cascade. The Lancet
Neurology, 9(1):119–128, 2010.

[2] Bruno M Jedynak, Andrew Lang, Bo Liu, Elyse Katz, Yanwei Zhang, Bradley T
Wyman, David Raunig, C Pierre Jedynak, Brian Caffo, Jerry L Prince, et al. A com-
putational neurodegenerative disease progression score: method and results with
the Alzheimer’s Disease Neuroimaging Initiative cohort. Neuroimage, 63(3):1478–
1486, 2012.

[3] Michael C Donohue, Hélène Jacqmin-Gadda, Mélanie Le Goff, Ronald G Thomas,


Rema Raman, Anthony C Gamst, Laurel A Beckett, Clifford R Jack, Michael W
Weiner, Jean-François Dartigues, et al. Estimating long-term multivariate progres-
sion from short-term data. Alzheimer’s & Dementia, 10(5):S400–S410, 2014.

[4] Murat Bilgel, Jerry L Prince, Dean F Wong, Susan M Resnick, and Bruno M Jedy-
nak. A multivariate nonlinear mixed effects model for longitudinal image analysis:
Application to amyloid imaging. NeuroImage, 134:658–670, 2016.

[5] Igor Koval, J-B Schiratti, Alexandre Routier, Michael Bacci, Olivier Colliot,
Stéphanie Allassonnière, Stanley Durrleman, and Alzheimers Disease Neuroimag-
ing Initiative. Statistical learning of spatiotemporal patterns from longitudinal
manifold-valued networks. In International Conference on Medical Image Comput-
ing and Computer-Assisted Intervention, pages 451–459. Springer, 2017.

[6] Ashish Raj, Amy Kuceyeski, and Michael Weiner. A network diffusion model of
disease progression in dementia. Neuron, 73(6):1204–1215, 2012.

[7] Alistair Burns and Steve Iliffe. Alzheimer’s disease. BMJ, 338:467–471 ST –
Alzheimer’s disease, 2009.

[8] World Health Organization (WHO) et al. Dementia fact sheet N. 362 2012.

[9] Jennifer L Whitwell. Progression of atrophy in Alzheimers disease and related


disorders. Neurotoxicity research, 18(3-4):339–346, 2010.
194 Appendix F. Bibliography

[10] Charles Marcus, Esther Mena, and Rathan M Subramaniam. Brain PET in the
diagnosis of Alzheimer’s disease. Clinical nuclear medicine, 39(10):e413, 2014.

[11] Amritpal Mudher and Simon Lovestone. Alzheimer’s disease–do tauists and baptists
finally shake hands? Trends in neurosciences, 25(1):22–26, 2002.

[12] Dev Mehta, Robert Jackson, Gaurav Paul, Jiong Shi, and Marwan Sabbagh. Why
do trials for Alzheimer’s disease drugs keep failing? A discontinued drug perspective
for 2010-2015. Expert opinion on investigational drugs, 26(6):735–739, 2017.

[13] Clare J Galton, Karalyn Patterson, John H Xuereb, and John R Hodges. Atypical
and typical presentations of Alzheimer’s disease: a clinical, neuropsychological,
neuroimaging and pathological study of 13 cases. Brain, 123(3):484–498, 2000.

[14] Melissa E Murray, Neill R Graff-Radford, Owen A Ross, Ronald C Petersen, Ranjan
Duara, and Dennis W Dickson. Neuropathologically defined subtypes of Alzheimer’s
disease with distinct clinical characteristics: a retrospective study. The Lancet
Neurology, 10(9):785–796, 2011.

[15] Wei Qiang, Wai-Ming Yau, Jun-Xia Lu, John Collinge, and Robert Tycko. Struc-
tural variation in amyloid-β fibrils from Alzheimer’s disease clinical subtypes. Na-
ture, 541(7636):217, 2017.

[16] AJ Larner and M Doran. Clinical phenotypic heterogeneity of Alzheimer’s dis-


ease associated with mutations of the presenilin–1 gene. Journal of neurology,
253(2):139–158, 2006.

[17] D Frank Benson, R Jeffrey Davis, and Bruce D Snyder. Posterior cortical atrophy.
Archives of neurology, 45(7):789–793, 1988.

[18] Sebastian J Crutch, Manja Lehmann, Jonathan M Schott, Gil D Rabinovici, Mar-
tin N Rossor, and Nick C Fox. Posterior cortical atrophy. The Lancet Neurology,
11(2):170–178, 2012.

[19] François-Xavier Borruat. Posterior cortical atrophy: review of the recent literature.
Current neurology and neuroscience reports, 13(12):1–8, 2013.

[20] Manja Lehmann, Sebastian J Crutch, Gerard R Ridgway, Basil H Ridha, Josephine
Barnes, Elizabeth K Warrington, Martin N Rossor, and Nick C Fox. Cortical
thickness and voxel-based morphometry in posterior cortical atrophy and typical
Alzheimer’s disease. Neurobiology of aging, 32(8):1466–1476, 2011.

[21] Jennifer L Whitwell, Clifford R Jack, Kejal Kantarci, Stephen D Weigand,


Bradley F Boeve, David S Knopman, Daniel A Drubach, David F Tang-Wai,
Ronald C Petersen, and Keith A Josephs. Imaging correlates of posterior corti-
cal atrophy. Neurobiology of aging, 28(7):1051–1061, 2007.

[22] Yaakov Stern. Cognitive reserve in ageing and Alzheimer’s disease. The Lancet
Neurology, 11(11):1006–1012, 2012.
195

[23] Hubert M Fonteijn, Marc Modat, Matthew J Clarkson, Josephine Barnes, Manja
Lehmann, Nicola Z Hobbs, Rachael I Scahill, Sarah J Tabrizi, Sebastien Ourselin,
Nick C Fox, et al. An event-based model for disease progression and its application
in familial Alzheimer’s disease and Huntington’s disease. NeuroImage, 60(3):1880–
1889, 2012.
[24] Alexandra L Young, Neil P Oxtoby, Pankaj Daga, David M Cash, Nick C Fox,
Sebastien Ourselin, Jonathan M Schott, and Daniel C Alexander. A data-driven
model of biomarker changes in sporadic Alzheimer’s disease. Brain, 137(9):2564–
2577, 2014.
[25] Victor L Villemagne, Samantha Burnham, Pierrick Bourgeat, Belinda Brown,
Kathryn A Ellis, Olivier Salvado, Cassandra Szoeke, S Lance Macaulay, Ralph
Martins, Paul Maruff, et al. Amyloid β deposition, neurodegeneration, and cogni-
tive decline in sporadic Alzheimer’s disease: a prospective cohort study. The Lancet
Neurology, 12(4):357–367, 2013.
[26] Bruno M. Jedynak, Andrew Lang, Bo Liu, Elyse Katz, Yanwei Zhang, Bradley T.
Wyman, David Raunig, C. Pierre Jedynak, Brian Caffo, and Jerry L. Prince. A com-
putational neurodegenerative disease progression score: Method and results with
the Alzheimer’s disease neuroimaging initiative cohort. NeuroImage, 63(3):1478–
1486, 2012.
[27] J-B Schiratti, Stéphanie Allassonniere, Alexandre Routier, Olivier Colliot, Stanley
Durrleman, and Alzheimers Disease Neuroimaging Initiative. A mixed-effects model
with time reparametrization for longitudinal univariate manifold-valued data. In
International Conference on Information Processing in Medical Imaging, pages 564–
575. Springer, 2015.
[28] Yasser Iturria-Medina, Roberto C Sotero, Paule J Toussaint, José Marı́a Mateos-
Pérez, Alan C Evans, Michael W Weiner, Paul Aisen, Ronald Petersen, Clifford R
Jack, William Jagust, et al. Early role of vascular dysregulation on late-onset
Alzheimers disease based on multifactorial data-driven analysis. Nature communi-
cations, 7:11934, 2016.
[29] Alexandra L Young et al. Uncovering the heterogeneity and temporal complexity
of neurodegenerative diseases with Subtype and Stage Inference. Nature Commu-
nications, in press, 2018.
[30] Neil P Oxtoby, Alexandra L Young, David M Cash, Tammie LS Benzinger, Anne M
Fagan, John C Morris, Randall J Bateman, Nick C Fox, Jonathan M Schott, and
Daniel C Alexander. Data-driven models of dominantly-inherited Alzheimer’s dis-
ease progression. Brain, 141(5):1529–1544, 2018.
[31] SJ Ross, Naida Graham, Lindsay Stuart-Green, Miriam Prins, John Xuereb, Kar-
alyn Patterson, and John R Hodges. Progressive biparietal atrophy: an atypical
presentation of Alzheimer’s disease. Journal of Neurology, Neurosurgery & Psychi-
atry, 61(4):388–395, 1996.
[32] Maarten Goethals and Patrick Santens. Posterior cortical atrophy. Two case reports
and a review of the literature. Clinical neurology and neurosurgery, 103(2):115–119,
2001.
196 Appendix F. Bibliography

[33] Anna Rita Giovagnoli, Anna Aresi, Fabiola Reati, Alice Riva, Clara Gobbo, and
Alberto Bizzi. The neuropsychological and neuroradiological correlates of slowly
progressive visual agnosia. Neurological sciences, 30(2):123–131, 2009.

[34] Zheng Chang, Paul Lichtenstein, Henrik Larsson, and Seena Fazel. Substance use
disorders, psychiatric disorders, and mortality after release from prison: a nation-
wide longitudinal cohort study. The Lancet Psychiatry, 2(5):422–430, 2015.

[35] Jonathan Kennedy, Manja Lehmann, Magdalena J Sokolska, Hilary Archer, Eliza-
beth K Warrington, Nick C Fox, and Sebastian J Crutch. Visualizing the emergence
of posterior cortical atrophy. Neurocase, 18(3):248–257, 2012.

[36] Sebastian J Crutch, Jonathan M Schott, Gil D Rabinovici, Melissa Murray, Julie S
Snowden, Wiesje M van der Flier, Bradford C Dickerson, Rik Vandenberghe, Sam-
rah Ahmed, Thomas H Bak, et al. Consensus classification of posterior cortical
atrophy. Alzheimer’s & Dementia, 13(8):870–884, 2017.

[37] Manja Lehmann, Josephine Barnes, Gerard R Ridgway, Natalie S Ryan, Eliza-
beth K Warrington, Sebastian J Crutch, and Nick C Fox. Global gray matter
changes in posterior cortical atrophy: a serial imaging study. Alzheimer’s & De-
mentia, 8(6):502–512, 2012.

[38] William W Seeley, Richard K Crawford, Juan Zhou, Bruce L Miller, and Michael D
Greicius. Neurodegenerative diseases target large-scale human brain networks. Neu-
ron, 62(1):42–52, 2009.

[39] Esther E Bron, Marion Smits, Wiesje M Van Der Flier, Hugo Vrenken, Frederik
Barkhof, Philip Scheltens, Janne M Papma, Rebecca ME Steketee, Carolina Méndez
Orellana, Rozanna Meijboom, et al. Standardized evaluation of algorithms for
computer-aided diagnosis of dementia based on structural MRI: the CADDementia
challenge. NeuroImage, 111:562–579, 2015.

[40] Alessia Sarica, Cerasa Antonio, Quattrone Aldo, and Calhoun Vince. A machine
learning neuroimaging challenge for automated diagnosis of Mild Cognitive Impair-
ment. in press, 2018.

[41] Henry W Querfurth and Frank M LaFerla. Mechanisms of disease. New England
Journal of Medicine, 362(4):329–344, 2010.

[42] M Prince, A Wimo, M Guerchet, et al. The global impact of Dementia: an analysis
of prevalence, incidence, cost and trends. 2015. London, UK: Alzheimer’s Disease
International.

[43] Hans Förstl and Alexander Kurz. Clinical features of Alzheimer’s disease. European
archives of psychiatry and clinical neuroscience, 249(6):288–290, 1999.

[44] KL Chobor and JW Brown. Semantic deterioration in Alzheimer’s: the patterns


to expect. Geriatrics (Basel, Switzerland), 45(10):68–70, 1990.

[45] Joseph J Locascio, John H Growdon, and Suzanne Corkin. Cognitive test perfor-
mance in detecting, staging, and tracking Alzheimer’s disease. Archives of neurol-
ogy, 52(11):1087–1099, 1995.
197

[46] Vanessa Moore and Maria A Wyke. Drawing disability in patients with senile
dementia. Psychological Medicine, 14(1):97–105, 1984.

[47] M Haupt, A Kurz, Barbara Romero, et al. Psychopathologische Störungen bei


beginnender Alzheimerscher Krankheit. Fortschritte der Neurologie· Psychiatrie,
60(01):3–7, 1992.

[48] Alastair Burns, Robin Jacoby, and Raymond Levy. Psychiatric phenomena in
Alzheimer’s disease. IV: Disorders of behaviour. The British Journal of Psychi-
atry, 157(1):86–94, 1990.

[49] William W Beatty, David P Salmon, Nelson Butters, William C Heindel, and Eric L
Granholm. Retrograde amnesia in patients with Alzheimer’s disease or Huntington’s
disease. Neurobiology of aging, 9:181–186, 1988.

[50] Barbara Romero, F Pulvermüller, M Haupt, and A Kurz. Pragmatische


Sprachstörungen in frühen Stadien der Alzheimer Krankheit: Analyse der Art und
Ausprägung. Zeitschrift für Neuropsychologie, 6(1):29–42, 1995.

[51] Jeffrey L Cummings, John P Houlihan, and Mary Ann Hill. The pattern of reading
deterioration in dementia of the Alzheimer type: Observations and implications.
Brain and language, 29(2):315–323, 1986.

[52] Jean Neils, Francois Boller, Bernice Gerdeman, and Monroe Cole. Descriptive
writing abilities in Alzheimer’s disease. Journal of Clinical and Experimental neu-
ropsychology, 11(5):692–698, 1989.

[53] Barry Reisberg, Stefanie R Auer, Isabel Monteiro, Istvan Boksay, and Steven G
Sclan. Behavioral disturbances of dementia: an overview of phenomenology and
methodologic concerns. International Psychogeriatrics, 8(S2):169–182, 1996.

[54] EK Perry, J Kerwin, RH Perry, G Blessed, and AF Fairbairn. Visual hallucinations


and the cholinergic system in dementia. Journal of neurology, neurosurgery, and
psychiatry, 53(1):88, 1990.

[55] Clive Ballard, Serge Gauthier, Anne Corbett, Carol Brayne, Dag Aarsland, and
Emma Jones. Alzheimer’s disease. The Lancet, 377(9770):1019–1031, 2011.

[56] John Hardy and David Allsop. Amyloid deposition as the central event in the
aetiology of Alzheimer’s disease. Trends in pharmacological sciences, 12:383–388,
1991.

[57] Nilufer Ertekin-Taner, Neill Graff-Radford, Linda H Younkin, Christopher Eckman,


Matthew Baker, Jennifer Adamson, James Ronald, John Blangero, Michael Hutton,
and Steven G Younkin. Linkage of plasma Aβ42 to a quantitative locus on chromo-
some 10 in late-onset Alzheimer’s disease pedigrees. Science, 290(5500):2303–2304,
2000.

[58] J Götz, F Chen, Jo Van Dorpe, and RM Nitsch. Formation of neurofibrillary tangles
in P301L tau transgenic mice induced by Aβ42 fibrils. Science, 293(5534):1491–
1495, 2001.
198 Appendix F. Bibliography

[59] Jada Lewis, Dennis W Dickson, Wen-Lang Lin, Louise Chisholm, Anthony Corral,
Graham Jones, Shu-Hui Yen, Naruhiko Sahara, Lisa Skipper, Debra Yager, et al.
Enhanced neurofibrillary degeneration in transgenic mice expressing mutant tau
and APP. Science, 293(5534):1487–1491, 2001.

[60] Erik D Roberson, Kimberly Scearce-Levie, Jorge J Palop, Fengrong Yan, Irene H
Cheng, Tiffany Wu, Hilary Gerstein, Gui-Qiu Yu, and Lennart Mucke. Reducing
endogenous tau ameliorates amyloid ß-induced deficits in an Alzheimer’s disease
mouse model. Science, 316(5825):750–754, 2007.

[61] George S Bloom. Amyloid-β and tau: the trigger and bullet in Alzheimer disease
pathogenesis. JAMA neurology, 71(4):505–508, 2014.

[62] Karen K Hsiao, David R Borchelt, Kristine Olson, Rosa Johannsdottir, Cheryl Kitt,
Wael Yunis, Sherry Xu, Chris Eckman, Steven Younkin, Donald Price, et al. Age-
related CNS disorder and early death in transgenic FVB/N mice overexpressing
Alzheimer amyloid precursor proteins. Neuron, 15(5):1203–1218, 1995.

[63] Michael C Irizarry, Megan McNamara, Kerri Fedorchak, Karen Hsiao, and
Bradley T Hyman. APPSw transgenic mice develop age-related Aβ deposits and
neuropil abnormalities, but no neuronal loss in CA1. Journal of Neuropathology &
Experimental Neurology, 56(9):965–973, 1997.

[64] ZS Nagy, MM Esiri, KA Jobst, JH Morris, EM-F King, B McDonald, S Litchfield,


A Smith, L Barnetson, and AD Smith. Relative roles of plaques and tangles in the
dementia of Alzheimer’s disease: correlations using three sets of neuropathological
criteria. Dementia and Geriatric Cognitive Disorders, 6(1):21–31, 1995.

[65] H Braak and E Braak. Evolution of neuronal changes in the course of Alzheimer’s
disease. In Ageing and dementia, pages 127–140. Springer, 1998.

[66] F Braak, Heiko Braak, and E-M Mandelkow. A sequence of cytoskeleton changes
related to the formation of neurofibrillary tangles and neuropil threads. Acta neu-
ropathologica, 87(6):554–567, 1994.

[67] Peter Heutink. Untangling tau-related dementia. Human molecular genetics,


9(6):979–986, 2000.

[68] Paul T Francis, Alan M Palmer, Michael Snape, and Gordon K Wilcock. The
cholinergic hypothesis of Alzheimer’s disease: a review of progress. Journal of
Neurology, Neurosurgery & Psychiatry, 66(2):137–147, 1999.

[69] Peter Davies and AJF Maloney. Selective loss of central cholinergic neurons in
Alzheimer’s disease. The Lancet, 308(8000):1403, 1976.

[70] Alessandro Martorana, Zaira Esposito, and Giacomo Koch. Beyond the cholinergic
hypothesis: do current drugs work in Alzheimer’s disease? CNS neuroscience &
therapeutics, 16(4):235–245, 2010.

[71] Jack C de la Torre. Is Alzheimer’s disease a neurodegenerative or a vascular disor-


der? Data, dogma, and dialectics. The Lancet Neurology, 3(3):184–190, 2004.
199

[72] Raj N Kalaria. Vascular factors in Alzheimer’s disease. International Psychogeri-


atrics, 15(S1):47–52, 2003.

[73] John S Meyer, Gaiane Rauch, Ronald A Rauch, and A Haque. Risk factors for
cerebral hypoperfusion, mild cognitive impairment, and dementia. Neurobiology of
aging, 21(2):161–169, 2000.

[74] Monique Breteler. Vascular involvement in cognitive decline and dementia: epi-
demiologic evidence from the Rotterdam Study and the Rotterdam Scan Study.
Annals of the New York Academy of Sciences, 903(1):457–465, 2000.

[75] MK Aronson, WL Ooi, H Morgenstern, A Hafner, D Masur, H Crystal, WH Frish-


man, D Fisher, and R Katzman. Women, myocardial infarction, and dementia in
the very old. Neurology, 40(7):1102–1102, 1990.

[76] MC Polidori, M Marvardi, A Cherubini, U Senin, and P Mecocci. Heart disease and
vascular risk factors in the cognitively impaired elderly: implications for Alzheimer’s
dementia. Aging Clinical and Experimental Research, 13(3):231–239, 2001.

[77] Albert Hofman, Alewijn Ott, Monique MB Breteler, Michiel L Bots, Arjen JC
Slooter, Frans van Harskamp, Cornelia N van Duijn, Christine Van Broeck-
hoven, and Diederick E Grobbee. Atherosclerosis, apolipoprotein E, and preva-
lence of dementia and Alzheimer’s disease in the Rotterdam Study. The Lancet,
349(9046):151–154, 1997.

[78] Morgan Robinson, Brenda Y Lee, and Francis T Hane. Recent progress in
Alzheimer’s disease research, part 2: genetics and epidemiology. Journal of
Alzheimer’s Disease, 57(2):317–330, 2017.

[79] AL Mina Bergem, Knut Engedal, and Einar Kringlen. The role of heredity in late-
onset Alzheimer disease and vascular dementia: a twin study. Archives of General
Psychiatry, 54(3):264–270, 1997.

[80] Margaret Gatz, Nancy L Pedersen, Stig Berg, Boo Johansson, Kurt Johansson,
James A Mortimer, Samuel F Posner, Matti Viitanen, Bengt Winblad, and Anders
Ahlbom. Heritability for Alzheimer’s disease: the study of dementia in Swedish
twins. The Journals of Gerontology Series A: Biological Sciences and Medical Sci-
ences, 52(2):M117–M125, 1997.

[81] Vincent Chouraki and Sudha Seshadri. Genetics of Alzheimer’s disease. In Advances
in genetics, volume 87, pages 245–294. Elsevier, 2014.

[82] George G Glenner and Caine W Wong. Alzheimer’s disease and Down’s syndrome:
sharing of a unique cerebrovascular amyloid fibril protein. Biochemical and bio-
physical research communications, 122(3):1131–1135, 1984.

[83] Peter H St George-Hyslop, Rudolph E Tanzi, Ronald J Polinsky, Jonathan L Haines,


Linda Nee, Paul C Watkins, Richard H Myers, Robert G Feldman, Daniel Pollen,
David Drachman, et al. The genetic defect causing familial Alzheimer’s disease
maps on chromosome 21. Science, 235(4791):885–890, 1987.
200 Appendix F. Bibliography

[84] Dmitry Goldgaber, Michael I Lerman, O Westley McBride, Umberto Saffiotti, and
D Carleton Gajdusek. Characterization and chromosomal localization of a cDNA
encoding brain amyloid of Alzheimer’s disease. Science, 235(4791):877–880, 1987.

[85] Marie-Christine Chartier-Harlin, Fiona Crawford, Henry Houlden, Andrew Warren,


David Hughes, Liana Fidani, Alison Goate, Martin Rossor, Penelope Roques, John
Hardy, et al. Early-onset Alzheimer’s disease caused by mutations at codon 717 of
the β-amyloid precursor protein gene. Nature, 353(6347):844, 1991.

[86] R Sherrington, EI Rogaev, Y al Liang, EA Rogaeva, G Levesque, M Ikeda, H Chi,


C Lin, G Li, K Holman, et al. Cloning of a gene bearing missense mutations in
early-onset familial Alzheimer’s disease. Nature, 375(6534):754, 1995.

[87] Ephrat Levy-Lahad, Wilma Wasco, Parvoneh Poorkaj, Donna M Romano, Junko
Oshima, Warren H Pettingell, Chang En Yu, Paul D Jondro, Stephen D Schmidt,
Kai Wang, et al. Candidate gene for the chromosome 1 familial Alzheimer’s disease
locus. Science, 269(5226):973–977, 1995.

[88] Warren J Strittmatter, Ann M Saunders, Donald Schmechel, Margaret Pericak-


Vance, Jan Enghild, Guy S Salvesen, and Allen D Roses. Apolipoprotein E: high-
avidity binding to beta-amyloid and increased frequency of type 4 allele in late-
onset familial Alzheimer disease. Proceedings of the National Academy of Sciences,
90(5):1977–1981, 1993.

[89] Ann M Saunders, Warren J Strittmatter, D Schmechel, PH St George-Hyslop,


MA Pericak-Vance, SH Joo, BL Rosi, JF Gusella, DR Crapper-MacLachlan, MJ Al-
berts, et al. Association of apolipoprotein E allele 4 with late-onset familial and
sporadic Alzheimer’s disease. Neurology, 43(8):1467–1467, 1993.

[90] Elizabeth H Corder, Ann M Saunders, Waren J Strittmatter, Donald E Schmechel,


P Craig Gaskell, Gwet Small, Allen D Roses, JL Haines, and Margaret A Pericak-
Vance. Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer’s
disease in late onset families. Science, 261(5123):921–923, 1993.

[91] Denise Harold, Richard Abraham, Paul Hollingworth, Rebecca Sims, Amy Ger-
rish, Marian L Hamshere, Jaspreet Singh Pahwa, Valentina Moskvina, Kimber-
ley Dowzell, Amy Williams, et al. Genome-wide association study identifies vari-
ants at CLU and PICALM associated with Alzheimer’s disease. Nature genetics,
41(10):1088, 2009.

[92] J Lambert, S Heath, G Even, D Campion, K Sleegers, M Hiltunen, O Cambar-


ros, D Zelenika, M Bullido, B Tavernier, et al. Genome-wide association study
indentifies variants at CLU and CR1 associated with Alzheimer’s disease. Nature
Genetics, 41(10):1094–1099, 2009.

[93] Sudha Seshadri, Annette L Fitzpatrick, M Arfan Ikram, Anita L DeStefano, Vil-
mundur Gudnason, Merce Boada, Joshua C Bis, Albert V Smith, Minerva M Car-
rasquillo, Jean Charles Lambert, et al. Genome-wide analysis of genetic loci asso-
ciated with Alzheimer disease. JAMA, 303(18):1832–1840, 2010.
201

[94] Paul Hollingworth, Denise Harold, Rebecca Sims, Amy Gerrish, Jean-Charles
Lambert, Minerva M Carrasquillo, Richard Abraham, Marian L Hamshere,
Jaspreet Singh Pahwa, Valentina Moskvina, et al. Common variants at ABCA7,
MS4A6A/MS4A4E, EPHA1, CD33 and CD2AP are associated with Alzheimer’s
disease. Nature genetics, 43(5):429, 2011.

[95] Adam C Naj, Gyungah Jun, Gary W Beecham, Li-San Wang, Badri Narayan Var-
darajan, Jacqueline Buros, Paul J Gallins, Joseph D Buxbaum, Gail P Jarvik,
Paul K Crane, et al. Common variants at MS4A4/MS4A6E, CD2AP, CD33
and EPHA1 are associated with late-onset Alzheimer’s disease. Nature genetics,
43(5):436, 2011.

[96] Madhav Thambisetty, Yang An, and Toshiko Tanaka. Alzheimer’s disease risk genes
and the age-at-onset phenotype. Neurobiology of aging, 34(11):2696–e1, 2013.

[97] Madhav Thambisetty, Lori L Beason-Held, Yang An, Michael Kraut, Michael Nalls,
Dena G Hernandez, Andrew B Singleton, Alan B Zonderman, Luigi Ferrucci, Simon
Lovestone, et al. Alzheimer risk variant CLU and brain function during aging.
Biological psychiatry, 73(5):399–405, 2013.

[98] A Biffi, JM Shulman, JM Jagiella, L Cortellini, AM Ayres, K Schwab, DL Brown,


SL Silliman, M Selim, BB Worrall, et al. Genetic variation at CR1 increases risk
of cerebral amyloid angiopathy. Neurology, pages WNL–0b013e3182452b40, 2012.

[99] Lori B Chibnik, Joshua M Shulman, Sue E Leurgans, Julie A Schneider, Robert S
Wilson, Dong Tran, Cristin Aubin, Aron S Buchman, Christopher B Heward,
Amanda J Myers, et al. CR1 is associated with amyloid plaque burden and age-
related cognitive decline. Annals of neurology, 69(3):560–569, 2011.

[100] Lyzel S Elias-Sonnenschein, Seppo Helisalmi, Teemu Natunen, Anette Hall, Teemu
Paajanen, Sanna-Kaisa Herukka, Marjo Laitinen, Anne M Remes, Anne M
Koivisto, Kari M Mattila, et al. Genetic loci associated with Alzheimer’s dis-
ease and cerebrospinal fluid biomarkers in a Finnish case-control cohort. PloS one,
8(4):e59676, 2013.

[101] John SK Kauwe, Carlos Cruchaga, Celeste M Karch, Brooke Sadler, Mo Lee, Kevin
Mayo, Wayne Latu, Manti Su’a, Anne M Fagan, David M Holtzman, et al. Fine
mapping of genetic variants in BIN1, CLU, CR1 and PICALM for association with
cerebrospinal fluid biomarkers for Alzheimer’s disease. PloS one, 6(2):e15918, 2011.

[102] Janita Bralten, Barbara Franke, Alejandro Arias-Vásquez, Angelien Heister, Han G
Brunner, Guillén Fernández, and Mark Rijpkema. CR1 genotype is associated
with entorhinal cortex volume in young healthy adults. Neurobiology of aging,
32(11):2106–e7, 2011.

[103] SJ Furney, A Simmons, G Breen, I Pedroso, K Lunnon, P Proitsi, A Hodges,


J Powell, LO Wahlund, I Kloszewska, et al. Genome-wide association with MRI
atrophy measures as a quantitative trait locus for Alzheimer’s disease. Molecular
psychiatry, 16(11):1130, 2011.
202 Appendix F. Bibliography

[104] S Barral, T Bird, A Goate, MR Farlow, R Diaz-Arrastia, DA Bennett, N Graff-


Radford, BF Boeve, RA Sweet, Y Stern, et al. Genotype patterns at PICALM, CR1,
BIN1, CLU, and APOE genes are associated with episodic memory. Neurology,
78(19):1464–1471, 2012.

[105] Jonas Mengel-From, Mikael Thinggaard, Rune Lindahl-Jacobsen, Matt McGue,


Kaare Christensen, and Lene Christiansen. CLU genetic variants and cognitive
decline among elderly and oldest old. PloS one, 8(11):e79105, 2013.

[106] Matthew Baumgart, Heather M Snyder, Maria C Carrillo, Sam Fazio, Hye Kim,
and Harry Johns. Summary of the evidence on modifiable risk factors for cognitive
decline and dementia: a population-based perspective. Alzheimer’s & Dementia,
11(6):718–726, 2015.

[107] Stephen Todd, Stephen Barr, Mark Roberts, and A Peter Passmore. Survival in
dementia and predictors of mortality: a review. International journal of geriatric
psychiatry, 28(11):1109–1124, 2013.

[108] Janine K Cataldo, Judith J Prochaska, and Stanton A Glantz. Cigarette smoking
is a risk factor for Alzheimer’s Disease: an analysis controlling for tobacco industry
affiliation. Journal of Alzheimer’s disease, 19(2):465–480, 2010.

[109] Paula Valencia Moulton and Wei Yang. Air pollution, oxidative stress, and
Alzheimer’s disease. Journal of environmental and public health, 2012, 2012.

[110] J Eric Ahlskog, Yonas E Geda, Neill R Graff-Radford, and Ronald C Petersen.
Physical exercise as a preventive or disease-modifying treatment of dementia and
brain aging. In Mayo Clinic Proceedings, volume 86, pages 876–884. Elsevier, 2011.

[111] Guy McKhann, David Drachman, Marshall Folstein, Robert Katzman, Donald
Price, and Emanuel M Stadlan. Clinical diagnosis of Alzheimer’s disease Report of
the NINCDS-ADRDA Work Group* under the auspices of Department of Health
and Human Services Task Force on Alzheimer’s Disease. Neurology, 34(7):939–939,
1984.

[112] Karl Herholz. Use of FDG PET as an imaging biomarker in clinical trials of
Alzheimer’s disease. Biomarkers in medicine, 6(4):431–439, 2012.

[113] William E Klunk, Henry Engler, Agneta Nordberg, Yanming Wang, Gunnar
Blomqvist, Daniel P Holt, Mats Bergström, Irina Savitcheva, Guo-Feng Huang,
Sergio Estrada, et al. Imaging brain amyloid in Alzheimer’s disease with Pitts-
burgh Compound-B. Annals of neurology, 55(3):306–319, 2004.

[114] Kaj Blennow and Harald Hampel. CSF markers for incipient Alzheimer’s disease.
The Lancet Neurology, 2(10):605–613, 2003.

[115] Bruno Dubois, Howard H Feldman, Claudia Jacova, Steven T DeKosky, Pascale
Barberger-Gateau, Jeffrey Cummings, André Delacourte, Douglas Galasko, Serge
Gauthier, Gregory Jicha, et al. Research criteria for the diagnosis of Alzheimer’s
disease: revising the NINCDS–ADRDA criteria. The Lancet Neurology, 6(8):734–
746, 2007.
203

[116] Bruno Dubois, Howard H Feldman, Claudia Jacova, Jeffrey L Cummings, Steven T
DeKosky, Pascale Barberger-Gateau, André Delacourte, Giovanni Frisoni, Nick C
Fox, Douglas Galasko, et al. Revising the definition of Alzheimer’s disease: a new
lexicon. The Lancet Neurology, 9(11):1118–1127, 2010.

[117] Urban Ekman, Daniel Ferreira, and Eric Westman. The A/T/N biomarker scheme
and patterns of brain atrophy assessed in mild cognitive impairment. Scientific
reports, 8(1):8431, 2018.

[118] Martin Reuter, Nicholas J Schmansky, H Diana Rosas, and Bruce Fischl. Within-
subject template estimation for unbiased longitudinal image analysis. Neuroimage,
61(4):1402–1418, 2012.

[119] Clifford R Jack Jr, David S Knopman, William J Jagust, Ronald C Petersen,
Michael W Weiner, Paul S Aisen, Leslie M Shaw, Prashanthi Vemuri, Heather J
Wiste, Stephen D Weigand, et al. Update on hypothetical model of Alzheimer’s
disease biomarkers. Lancet neurology, 12(2):207, 2013.

[120] Ann D Cohen and William E Klunk. Early detection of Alzheimer’s disease using
PiB and FDG PET. Neurobiology of disease, 72:117–122, 2014.

[121] Val J Lowe, Geoffry Curran, Ping Fang, Amanda M Liesinger, Keith A Josephs,
Joseph E Parisi, Kejal Kantarci, Bradley F Boeve, Mukesh K Pandey, Tyler Bru-
insma, et al. An autoradiographic evaluation of AV-1451 Tau PET in dementia.
Acta neuropathologica communications, 4(1):58, 2016.

[122] Marta Marquié, Marc D Normandin, Charles R Vanderburg, Isabel M Costantino,


Elizabeth A Bien, Lisa G Rycyna, William E Klunk, Chester A Mathis, Milos D
Ikonomovic, Manik L Debnath, et al. Validating novel tau positron emission to-
mography tracer [F-18]-AV-1451 (T807) on postmortem brain tissue. Annals of
neurology, 78(5):787–800, 2015.

[123] Perminder S Sachdev, Lin Zhuang, Nady Braidy, and Wei Wen. Is Alzheimer’s a
disease of the white matter? Current opinion in psychiatry, 26(3):244–251, 2013.

[124] M Bozzali, A Falini, M Franceschi, M Cercignani, M Zuffi, G Scotti, G Comi, and


M Filippi. White matter damage in Alzheimer’s disease assessed in vivo using
diffusion tensor magnetic resonance imaging. Journal of Neurology, Neurosurgery
& Psychiatry, 72(6):742–746, 2002.

[125] Yu Zhang, Norbert Schuff, An-Tao Du, Howard J Rosen, Joel H Kramer,
Maria Luisa Gorno-Tempini, Bruce L Miller, and Michael W Weiner. White matter
damage in frontotemporal dementia and Alzheimer’s disease measured by diffusion
MRI. Brain, 132(9):2579–2592, 2009.

[126] Juan Zhou, Efstathios D Gennatas, Joel H Kramer, Bruce L Miller, and William W
Seeley. Predicting regional neurodegeneration from the healthy brain functional
connectome. Neuron, 73(6):1216–1227, 2012.

[127] Hui Zhang, Torben Schneider, Claudia A Wheeler-Kingshott, and Daniel C Alexan-
der. NODDI: practical in vivo neurite orientation dispersion and density imaging
of the human brain. Neuroimage, 61(4):1000–1016, 2012.
204 Appendix F. Bibliography

[128] G Waldemar, B Dubois, M Emre, J Georges, IG McKeith, M Rossor, P Scheltens,


P Tariska, and B Winblad. Recommendations for the diagnosis and management of
Alzheimer’s disease and other disorders associated with dementia: EFNS guideline.
European Journal of Neurology, 14(1):e1–e26, 2007.

[129] Basil H Ridha, Josephine Barnes, Jonathan W Bartlett, Alison Godbolt, Tracey
Pepple, Martin N Rossor, and Nick C Fox. Tracking atrophy progression in familial
Alzheimer’s disease: a serial MRI study. The Lancet Neurology, 5(10):828–834,
2006.

[130] NC Fox, RI Scahill, WR Crum, and MN Rossor. Correlation between rates of brain
atrophy and cognitive decline in AD. Neurology, 52(8):1687–1687, 1999.

[131] Rachael I Scahill, Jonathan M Schott, John M Stevens, Martin N Rossor, and
Nick C Fox. Mapping the evolution of regional atrophy in Alzheimer’s disease: un-
biased analysis of fluid-registered serial MRI. Proceedings of the National Academy
of Sciences, 99(7):4703–4707, 2002.

[132] H Braak and E Braak. Neuropathological stageing of Alzheimer-related changes.


Acta neuropathologica, 82(4):239–259, 1991.

[133] Jonathan M Schott, Nick C Fox, Chris Frost, Rachael I Scahill, John C Janssen,
Dennis Chan, Rhian Jenkins, and Martin N Rossor. Assessing the onset of structural
change in familial Alzheimer’s disease. Annals of neurology, 53(2):181–188, 2003.

[134] Clifford R Jack, David S Knopman, William J Jagust, Ronald C Petersen,


Michael W Weiner, Paul S Aisen, Leslie M Shaw, Prashanthi Vemuri, Heather J
Wiste, Stephen D Weigand, et al. Tracking pathophysiological processes in
Alzheimer’s disease: an updated hypothetical model of dynamic biomarkers. The
Lancet Neurology, 12(2):207–216, 2013.

[135] Clifford R Jack, Ronald C Petersen, Yue Cheng Xu, Stephen C Waring, Peter C
O’Brien, Eric G Tangalos, Glenn E Smith, Robert J Ivnik, and Emre Kokmen. Me-
dial temporal atrophy on MRI in normal aging and very mild Alzheimer’s disease.
Neurology, 49(3):786–794, 1997.

[136] Stephane Lehericy, Michel Baulac, Jacques Chiras, Laurent Pierot, Nadine Martin,
Bernard Pillon, Bernard Deweer, Bruno Dubois, and Claude Marsault. Amygdalo-
hippocampal MR volume measurements in the early stages of Alzheimer disease.
American Journal of Neuroradiology, 15(5):929–937, 1994.

[137] Kirsi Juottonen, Mikko P Laakso, Kaarina Partanen, and Hilkka Soininen. Com-
parative MR analysis of the entorhinal cortex and hippocampus in diagnosing
Alzheimer disease. American Journal of Neuroradiology, 20(1):139–144, 1999.

[138] M Bobinski, MJ De Leon, J Wegiel, S Desanti, A Convit, LA Saint Louis,


H Rusinek, and HM Wisniewski. The histological validation of post mortem mag-
netic resonance imaging-determined hippocampal volume in Alzheimer’s disease.
Neuroscience, 95(3):721–725, 1999.
205

[139] Christopher M Clark, Julie A Schneider, Barry J Bedell, Thomas G Beach, War-
ren B Bilker, Mark A Mintun, Michael J Pontecorvo, Franz Hefti, Alan P Carpenter,
Matthew L Flitter, et al. Use of florbetapir-PET for imaging β-amyloid pathology.
JAMA, 305(3):275–283, 2011.

[140] Christopher M Clark, Michael J Pontecorvo, Thomas G Beach, Barry J Bedell,


R Edward Coleman, P Murali Doraiswamy, Adam S Fleisher, Eric M Reiman,
Marwan N Sabbagh, Carl H Sadowsky, et al. Cerebral PET with florbetapir com-
pared with neuropathology at autopsy for detection of neuritic amyloid-β plaques:
a prospective cohort study. The Lancet Neurology, 11(8):669–678, 2012.

[141] Katia Andrade, Dalila Samri, Marie Sarazin, Leonardo C de Souza, Laurent Cohen,
Michel T de Schotten, Bruno Dubois, and Paolo Bartolomeo. Visual neglect in
posterior cortical atrophy. BMC neurology, 10(1):68, 2010.

[142] Katia Andrade, Aurélie Kas, Romain Valabrègue, Dalila Samri, Marie Sarazin,
Marie-Odile Habert, Bruno Dubois, and Paolo Bartolomeo. Visuospatial deficits in
posterior cortical atrophy: structural and functional correlates. Journal of Neurol-
ogy, Neurosurgery & Psychiatry, 2012.

[143] Katia Andrade, Aurélie Kas, Dalila Samri, Marie Sarazin, Bruno Dubois, Marie-
Odile Habert, and Paolo Bartolomeo. Visuospatial deficits and hemispheric perfu-
sion asymmetries in posterior cortical atrophy. Cortex, 49(4):940–947, 2013.

[144] Manja Lehmann, Josephine Barnes, Gerard R Ridgway, John Wattam-Bell, Eliza-
beth K Warrington, Nick C Fox, and Sebastian J Crutch. Basic visual function and
cortical thickness patterns in posterior cortical atrophy. Cerebral cortex, 21(9):2122–
2132, 2011.

[145] Raphaël Depaz, Stéphane Haik, Katell Peoch, Danielle Seilhean, David Grabli,
Savine Vicart, Marie Sarazin, Bertrand DeToffol, Catherine Remy, Catherine Fallet-
Bianco, et al. Long-standing prion dementia manifesting as posterior cortical atro-
phy. Alzheimer Disease & Associated Disorders, 26(3):289–292, 2012.

[146] Mario F Mendez, Mehdi Ghajarania, and Kent M Perryman. Posterior cortical
atrophy: clinical characteristics and differences compared to Alzheimer’s disease.
Dementia and geriatric cognitive disorders, 14(1):33–40, 2002.

[147] David F Tang-Wai, NR Graff-Radford, BF Boeve, DW Dickson, JE Parisi, R Crook,


RJ Caselli, DS Knopman, and RC Petersen. Clinical, genetic, and neuropathologic
characteristics of posterior cortical atrophy. Neurology, 63(7):1168–1174, 2004.

[148] MH Rosenbloom, A Alkalay, N Agarwal, SL Baker, JP O’Neil, M Janabi, IV Yen,


M Growdon, J Jang, C Madison, et al. Distinct clinical and metabolic deficits in
PCA and AD are not related to amyloid distribution. Neurology, 76(21):1789–1796,
2011.

[149] Raffaella Migliaccio, Federica Agosta, Katya Rascovsky, Anna Karydas, Stephen
Bonasera, Gil D Rabinovici, BL Miller, and ML Gorno-Tempini. Clinical syndromes
associated with posterior atrophy early age at onset AD spectrum. Neurology,
73(19):1571–1578, 2009.
206 Appendix F. Bibliography

[150] Jonathan M Schott, Basil H Ridha, Sebastian J Crutch, Daniel G Healy, James B
Uphill, Elizabeth K Warrington, Martin N Rossor, and Nick C Fox. Apolipoprotein
e genotype modifies the phenotype of Alzheimer disease. Archives of neurology,
63(1):155–156, 2006.

[151] Julie S Snowden, Cheryl L Stopford, Camille L Julien, Jennifer C Thompson,


Yvonne Davidson, Linda Gibbons, Antonia Pritchard, Corinne L Lendon, Anna M
Richardson, Anoop Varma, et al. Cognitive phenotypes in Alzheimer’s disease and
genetic risk. Cortex, 43(7):835–845, 2007.

[152] Martin A Goldstein, Iliyan Ivanov, and Michael E Silverman. Posterior cortical
atrophy: an exemplar for renovating diagnostic formulation in neuropsychiatry.
Comprehensive psychiatry, 52(3):326–333, 2011.

[153] Hélène Videaud, Frédéric Torny, Leslie Cartz-Piver, N Deschamps-Vergara, and


Philippe Couratier. Impact of drug-free care in posterior cortical atrophy: Prelimi-
nary experience with a psycho-educative program. Revue neurologique, 168(11):861–
867, 2012.

[154] A Weill-Chounlamountry, F Poncet, S Crop, N Hesly, A Mouton, D Samri,


M Sarazin, and P Pradat-Diehl. Physical medicine and rehabilitation multidis-
ciplinary approach in a case of posterior cortical atrophy. Annals of physical and
rehabilitation medicine, 55(6):430–439, 2012.

[155] Patrick R Hof, Brent A Vogt, Constantin Bouras, and John H Morrison. Atypical
form of Alzheimer’s disease with prominent posterior cortical atrophy: a review
of lesion distribution and circuit disconnection in cortical visual pathways. Vision
research, 37(24):3609–3625, 1997.

[156] T Duning, T Warnecke, S Mohammadi, H Lohmann, H Schiffbauer, H Kugel,


S Knecht, EB Ringelstein, and M Deppe. Pattern and progression of white-matter
changes in a case of posterior cortical atrophy using diffusion tensor imaging. Jour-
nal of Neurology, Neurosurgery & Psychiatry, 80(4):432–436, 2009.

[157] Tomokatsu Yoshida, Kensuke Shiga, Kenji Yoshikawa, Kei Yamada, and Masanori
Nakagawa. White matter loss in the splenium of the corpus callosum in a case of
posterior cortical atrophy: a diffusion tensor imaging study. European neurology,
52(2):77–81, 2004.

[158] Raffaella Migliaccio, Federica Agosta, Monica N Toba, Dalila Samri, Fabian Corlier,
Leonardo C De Souza, Marie Chupin, Michael Sharman, Maria L Gorno-Tempini,
Bruno Dubois, et al. Brain networks in posterior cortical atrophy: a single case
tractography study and literature review. Cortex, 48(10):1298–1309, 2012.

[159] Aurelie Kas, Leonardo Cruz De Souza, Dalila Samri, Paolo Bartolomeo, Lucette
Lacomblez, Michel Kalafat, Raffaella Migliaccio, Michel Thiebaut de Schotten, Lau-
rent Cohen, Bruno Dubois, et al. Neural correlates of cognitive impairment in
posterior cortical atrophy. Brain, 134(5):1464–1478, 2011.

[160] Simona Gardini, Letizia Concari, Salvatrice Pagliara, Caterina Ghetti, Annalena
Venneri, and Paolo Caffarra. Visuo-spatial imagery impairment in posterior cortical
207

atrophy: a cognitive and SPECT study. Behavioural neurology, 24(2):123–132,


2011.

[161] Judith Aharon-Peretz, Ora Israel, Dorit Goldsher, and Aharon Peretz. Posterior
cortical atrophy variants of Alzheimer’s disease. Dementia and geriatric cognitive
disorders, 10(6):483–487, 1999.

[162] Pietro Pietrini, Maura L Furey, Neill Graff-Radford, Ulderico Freo, et al. Prefer-
ential metabolic involvement of visual cortical areas in a subtype of Alzheimer’s
disease: clinical implications. The American journal of psychiatry, 153(10):1261,
1996.

[163] Steven Y Ng, Victor L Villemagne, Colin L Masters, and Christopher C Rowe. Eval-
uating Atypical Dementia Syndromes Using Positron Emission Tomography With
Carbon 11–Labeled Pittsburgh Compound B. Archives of neurology, 64(8):1140–
1144, 2007.

[164] Taiki Kambe, Yumiko Motoi, Kenji Ishii, and Nobutaka Hattori. Posterior cortical
atrophy with 11–C Pittsburgh compound B accumulation in the primary visual
cortex. Journal of neurology, 257(3):469–471, 2010.

[165] Olli Tenovuo, Nina Kemppainen, Sargo Aalto, Kjell Någren, and Juha O Rinne.
Posterior cortical atrophy: A rare form of dementia with in vivo evidence of
amyloid-β accumulation. Journal of Alzheimer’s Disease, 15(3):351–355, 2008.

[166] Maı̈té Formaglio, Nicolas Costes, Jérémie Seguin, Yannick Tholance, Didier
Le Bars, Isabelle Roullet-Solignac, Bernadette Mercier, Pierre Krolak-Salmon, and
Alain Vighetto. In vivo demonstration of amyloid burden in posterior cortical atro-
phy: a case series with PET and CSF findings. Journal of neurology, 258(10):1841–
1851, 2011.

[167] Leonardo Cruz De Souza, Fabian Corlier, Marie-Odile Habert, Olga Uspenskaya,
Renaud Maroy, Foudil Lamari, Marie Chupin, Stéphane Lehéricy, Olivier Colliot,
Valérie Hahn-Barma, et al. Similar amyloid-β burden in posterior cortical atrophy
and Alzheimer’s disease. Brain, 134(7):2036–2043, 2011.

[168] David N Levine, John M Lee, and CM Fisher. The visual variant of Alzheimer’s
disease A clinicopathologic case study. Neurology, 43(2):305–305, 1993.

[169] Clifford R Jack Jr, Val J Lowe, Matthew L Senjem, Stephen D Weigand, Bradley J
Kemp, Maria M Shiung, David S Knopman, Bradley F Boeve, William E Klunk,
Chester A Mathis, et al. 11C PiB and structural MRI provide complementary infor-
mation in imaging of Alzheimer’s disease and amnestic mild cognitive impairment.
Brain, 131(3):665–680, 2008.

[170] Bradley T Hyman. Amyloid-dependent and amyloid-independent stages of


Alzheimer disease. Archives of neurology, 68(8):1062–1064, 2011.

[171] CR Jack, Stephen D Weigand, Maria M Shiung, Scott A Przybelski, Peter C


OBrien, Jeffrey L Gunter, David S Knopman, Bradley F Boeve, Glenn E Smith,
and Ronald C Petersen. Atrophy rates accelerate in amnestic mild cognitive im-
pairment. Neurology, 70(19 Part 2):1740–1752, 2008.
208 Appendix F. Bibliography

[172] Bradford C Dickerson, Akram Bakkour, David H Salat, Eric Feczko, Jenni Pacheco,
Douglas N Greve, Fran Grodstein, Christopher I Wright, Deborah Blacker, H Di-
ana Rosas, et al. The cortical signature of Alzheimer’s disease: regionally specific
cortical thinning relates to symptom severity in very mild to mild AD dementia
and is detectable in asymptomatic amyloid-positive individuals. Cerebral cortex,
19(3):497–510, 2009.

[173] Paul M Thompson, Michael S Mega, Roger P Woods, Chris I Zoumalan, Chris J
Lindshield, Rebecca E Blanton, Jacob Moussai, Colin J Holmes, Jeffrey L Cum-
mings, and Arthur W Toga. Cortical change in Alzheimer’s disease detected with
a disease-specific population-based brain atlas. Cerebral Cortex, 11(1):1–16, 2001.

[174] Paul M Thompson, Kiralee M Hayashi, Greig De Zubicaray, Andrew L Janke,


Stephen E Rose, James Semple, David Herman, Michael S Hong, Stephanie S
Dittmer, David M Doddrell, et al. Dynamics of gray matter loss in Alzheimer’s
disease. The Journal of Neuroscience, 23(3):994–1005, 2003.

[175] Mert R Sabuncu, Rahul S Desikan, Jorge Sepulcre, Boon Thye T Yeo, Hesheng
Liu, Nicholas J Schmansky, Martin Reuter, Michael W Weiner, Randy L Buckner,
Reisa A Sperling, et al. The dynamics of cortical and hippocampal atrophy in
Alzheimer disease. Archives of neurology, 68(8):1040–1048, 2011.

[176] Clifford R Jack, Prashanthi Vemuri, Heather J Wiste, Stephen D Weigand, Tim-
othy G Lesnick, Val Lowe, Kejal Kantarci, Matt A Bernstein, Matthew L Sen-
jem, Jeffrey L Gunter, et al. Shapes of the trajectories of 5 major biomarkers of
Alzheimer disease. Archives of neurology, 69(7):856–867, 2012.

[177] Rachelle S Doody, Valory Pavlik, Paul Massman, Susan Rountree, Eveleen Darby,
and Wenyaw Chan. Predicting progression of Alzheimer’s disease. Alzheimer’s
research & therapy, 2(1):2, 2010.

[178] I Driscoll, C Davatzikos, Y An, X Wu, D Shen, M Kraut, and SM Resnick. Lon-
gitudinal pattern of regional brain volume change differentiates normal aging from
MCI. Neurology, 72(22):1906–1913, 2009.

[179] Randall J Bateman, Chengjie Xiong, Tammie LS Benzinger, Anne M Fagan, Alison
Goate, Nick C Fox, Daniel S Marcus, Nigel J Cairns, Xianyun Xie, Tyler M Blazey,
et al. Clinical and biomarker changes in dominantly inherited Alzheimer’s disease.
New England Journal of Medicine, 367(9):795–804, 2012.

[180] Tammie L Benzinger, Tyler Blazey, Clifford R Jack, Robert A Koeppe, Yi Su,
Chengjie Xiong, Marcus E Raichle, Abraham Z Snyder, Beau M Ances, Ran-
dall J Bateman, et al. Regional variability of imaging biomarkers in autosomal
dominant Alzheimer’s disease. Proceedings of the National Academy of Sciences,
110(47):E4502–E4509, 2013.

[181] BC Dickerson, TR Stoub, RC Shah, RA Sperling, RJ Killiany, MS Albert, BT Hy-


man, Deborah Blacker, et al. Alzheimer-signature MRI biomarker predicts AD
dementia in cognitively normal adults. Neurology, 76(16):1395–1402, 2011.
209

[182] FH Bouwman, SNM Schoonenboom, WM van Der Flier, EJ Van Elk, A Kok,
F Barkhof, MA Blankenstein, and Ph Scheltens. CSF biomarkers and medial tem-
poral lobe atrophy predict dementia in mild cognitive impairment. Neurobiology of
aging, 28(7):1070–1074, 2007.

[183] JG Csernansky, L Wang, J Swank, JP Miller, M Gado, D McKeel, MI Miller, and


JC Morris. Preclinical detection of Alzheimer’s disease: hippocampal shape and
volume predict dementia onset in the elderly. Neuroimage, 25(3):783–792, 2005.

[184] Oskar Hansson, Henrik Zetterberg, Peder Buchhave, Elisabet Londos, Kaj Blennow,
and Lennart Minthon. Association between CSF biomarkers and incipient
Alzheimer’s disease in patients with mild cognitive impairment: a follow-up study.
The Lancet Neurology, 5(3):228–234, 2006.

[185] CH Kawas, MM Corrada, R Brookmeyer, A Morrison, SM Resnick, AB Zonderman,


and D Arenberg. Visual memory predicts Alzheimer’s disease more than a decade
before diagnosis. Neurology, 60(7):1089–1093, 2003.

[186] TR Stoub, M Bulgakova, S Leurgans, DA Bennett, D Fleischman, DA Turner,


et al. MRI predictors of risk of incident Alzheimer disease: a longitudinal study.
Neurology, 64(9):1520–1524, 2005.

[187] P Vemuri, HJ Wiste, SD Weigand, LM Shaw, JQ Trojanowski, MW Weiner,


DS Knopman, RC Petersen, CR Jack, et al. MRI and CSF biomarkers in nor-
mal, MCI, and AD subjects: diagnostic discrimination and cognitive correlations.
Neurology, 73(4):287–293, 2009.

[188] John C Morris, Martha Storandt, J Phillip Miller, Daniel W McKeel, Joseph L
Price, Eugene H Rubin, and Leonard Berg. Mild cognitive impairment represents
early-stage Alzheimer disease. Archives of neurology, 58(3):397–405, 2001.

[189] Pedro J Modrego and Jaime Ferrández. Depression in patients with mild cogni-
tive impairment increases the risk of developing dementia of Alzheimer type: a
prospective cohort study. Archives of neurology, 61(8):1290–1293, 2004.

[190] David B Carr, Steven Gray, Jack Baty, and John C Morris. The value of informant
versus individuals complaints of memory impairment in early dementia. Neurology,
55(11):1724–1727, 2000.

[191] Ara S Khachaturian, Christopher D Corcoran, Lawrence S Mayer, Peter P Zandi,


and John CS Breitner. Apolipoprotein E 4 count affects age at onset of Alzheimer
disease, but not lifetime susceptibility: the Cache County Study. Archives of general
psychiatry, 61(5):518–524, 2004.

[192] J Wesson Ashford and Frederick A Schmitt. Modeling the time-course of Alzheimer
dementia. Current psychiatry reports, 3(1):20–28, 2001.

[193] Eric Yang, Michael Farnum, Victor Lobanov, Tim Schultz, Nandini Raghavan, Ma-
hesh N Samtani, Gerald Novak, Vaibhav Narayan, and Allitia DiBernardo. Quanti-
fying the pathophysiological timeline of Alzheimer’s disease. Journal of Alzheimer’s
Disease, 26(4):745–753, 2011.
210 Appendix F. Bibliography

[194] A Caroli, GB Frisoni, and Alzheimer’s Disease Neuroimaging Initiative. The dy-
namics of Alzheimer’s disease biomarkers in the Alzheimer’s Disease Neuroimaging
Initiative cohort. Neurobiology of aging, 31(8):1263–1274, 2010.

[195] C Bishop. Pattern Recognition and Machine Learning (Information Science and
Statistics), 1st edn. 2006. corr. 2nd printing edn, 2007.

[196] Manasi Datar, Prasanna Muralidharan, Abhishek Kumar, Sylvain Gouttard,


Joseph Piven, Guido Gerig, Ross Whitaker, and P Thomas Fletcher. Mixed-effects
shape models for estimating longitudinal changes in anatomy. In International
Workshop on Spatio-temporal Image Analysis for Longitudinal and Time-Series
Image Data, pages 76–87. Springer, 2012.

[197] Jean-Baptiste Schiratti, Stéphanie Allassonniere, Olivier Colliot, and Stanley Dur-
rleman. Learning spatiotemporal trajectories from manifold-valued longitudinal
data. In Advances in Neural Information Processing Systems, pages 2404–2412,
2015.

[198] Igor Koval, Jean-Baptiste Schiratti, Alexandre Routier, Michael Bacci, Olivier Col-
liot, Stephanie Allassonniere, and Stanley Durrleman. Spatiotemporal Propagation
of the Cortical Atrophy: Population and Individual Patterns. Frontiers in Neurol-
ogy, 9, 2018.

[199] Ashish Raj, Eve LoCastro, Amy Kuceyeski, Duygu Tosun, Norman Relkin, Michael
Weiner, and Alzheimer’s Disease Neuroimaging Initiative. Network diffusion
model of progression predicts longitudinal patterns of atrophy and metabolism in
Alzheimer’s disease.

[200] Nicolas Villain, Marine Fouquet, Jean-Claude Baron, Florence Mézenge, Brigitte
Landeau, Vincent de La Sayette, Fausto Viader, Francis Eustache, Béatrice Des-
granges, and Gaël Chételat. Sequential relationships between grey matter and
white matter atrophy and brain metabolic abnormalities in early Alzheimer’s dis-
ease. Brain, 133(11):3301–3314, 2010.

[201] E Englund, A Brun, and C Alling. White matter changes in dementia of Alzheimer’s
type. Brain, 111(6):1425–1439, 1988.

[202] Beth Kuczynski, Elizabeth Targan, Cindee Madison, Michael Weiner, Yu Zhang,
Bruce Reed, Helena C Chui, and William Jagust. White matter integrity and
cortical metabolic associations in aging and dementia. Alzheimer’s & dementia,
6(1):54–62, 2010.

[203] TEJ Behrens, H Johansen Berg, Saad Jbabdi, MFS Rushworth, and MW Woolrich.
Probabilistic diffusion tractography with multiple fibre orientations: What can we
gain? Neuroimage, 34(1):144–155, 2007.

[204] Risi Imre Kondor and John Lafferty. Diffusion kernels on graphs and other discrete
input spaces. In ICML, volume 2, pages 315–322, 2002.

[205] Juan Zhou, Michael D Greicius, Efstathios D Gennatas, Matthew E Growdon,


Jung Y Jang, Gil D Rabinovici, Joel H Kramer, Michael Weiner, Bruce L Miller,
and William W Seeley. Divergent network connectivity changes in behavioural
211

variant frontotemporal dementia and Alzheimer’s disease. Brain, 133(5):1352–1367,


2010.

[206] Vladimir Vapnik. Estimation of dependences based on empirical data. Springer


Science & Business Media, 2006.

[207] Stefan Klöppel, Cynthia M Stonnington, Carlton Chu, Bogdan Draganski,


Rachael I Scahill, Jonathan D Rohrer, Nick C Fox, Clifford R Jack, John Ashburner,
and Richard SJ Frackowiak. Automatic classification of MR scans in Alzheimer’s
disease. Brain, 131(3):681–689, 2008.

[208] Zhiqiang Lao, Dinggang Shen, Zhong Xue, Bilge Karacali, Susan M Resnick, and
Christos Davatzikos. Morphological classification of brains via high-dimensional
shape transformations and machine learning methods. Neuroimage, 21(1):46–57,
2004.

[209] Yong Fan, Dinggang Shen, and Christos Davatzikos. Classification of structural
images via high-dimensional image warping, robust feature extraction, and SVM.
In International Conference on Medical Image Computing and Computer-Assisted
Intervention, pages 1–8. Springer, 2005.

[210] Janaina Mourão-Miranda, Arun LW Bokde, Christine Born, Harald Hampel, and
Martin Stetter. Classifying brain states and determining the discriminating acti-
vation patterns: support vector machine on functional MRI data. NeuroImage,
28(4):980–995, 2005.

[211] Yasuhiro Kawasaki, Michio Suzuki, Ferath Kherif, Tsutomu Takahashi, Shi-
Yu Zhou, Kazue Nakamura, Mie Matsui, Tomiki Sumiyoshi, Hikaru Seto, and
Masayoshi Kurachi. Multivariate voxel-based morphometry successfully differen-
tiates schizophrenia patients from healthy controls. Neuroimage, 34(1):235–242,
2007.

[212] Tin Kam Ho. Random decision forests. In Document Analysis and Recognition,
1995., Proceedings of the Third International Conference on, volume 1, pages 278–
282. IEEE, 1995.

[213] Leo Breiman. Random forests. Machine learning, 45(1):5–32, 2001.

[214] Katherine R Gray, Paul Aljabar, Rolf A Heckemann, Alexander Hammers, Daniel
Rueckert, and Alzheimer’s Disease Neuroimaging Initiative. Random forest-based
similarity measures for multi-modal classification of Alzheimer’s disease. NeuroIm-
age, 65:167–175, 2013.

[215] Daniel C Alexander, Darko Zikic, Jiaying Zhang, Hui Zhang, and Antonio Criminisi.
Image quality transfer via random forest regression: applications in diffusion MRI.
In International Conference on Medical Image Computing and Computer-Assisted
Intervention, pages 225–232. Springer, 2014.

[216] Victor Lempitsky, Michael Verhoek, J Alison Noble, and Andrew Blake. Ran-
dom forest classification for automatic delineation of myocardium in real-time 3D
echocardiography. In International Conference on Functional Imaging and Model-
ing of the Heart, pages 447–456. Springer, 2009.
212 Appendix F. Bibliography

[217] Sepp Hochreiter and Jürgen Schmidhuber. Long short-term memory. Neural com-
putation, 9(8):1735–1780, 1997.

[218] Sweta Karlekar, Tong Niu, and Mohit Bansal. Detecting Linguistic Character-
istics of Alzheimer’s Dementia by Interpreting Neural Models. arXiv preprint
arXiv:1804.06440, 2018.

[219] Narges Razavian, Jake Marcus, and David Sontag. Multi-task prediction of dis-
ease onsets from longitudinal laboratory tests. In Machine Learning for Healthcare
Conference, pages 73–100, 2016.

[220] Zachary C Lipton, David C Kale, Charles Elkan, and Randall Wetzel. Learning to
diagnose with LSTM recurrent neural networks. arXiv preprint arXiv:1511.03677,
2015.

[221] Melissa Aczon, David Ledbetter, L Ho, Alec Gunny, Alysia Flynn, Jon Williams,
and Randall Wetzel. Dynamic mortality risk predictions in pediatric critical care
using recurrent neural networks. arXiv preprint arXiv:1701.06675, 2017.

[222] Hrayr Harutyunyan, Hrant Khachatrian, David C Kale, and Aram Galstyan. Mul-
titask learning and benchmarking with clinical time series data. arXiv preprint
arXiv:1703.07771, 2017.

[223] Konstantinos Georgiadis, Selina Wray, Sébastien Ourselin, Jason D Warren, and
Marc Modat. Computational modelling of pathogenic protein spread in neurode-
generative diseases. PloS one, 13(2):e0192518, 2018.

[224] Guy M McKhann, David S Knopman, Howard Chertkow, Bradley T Hyman, Clif-
ford R Jack Jr, Claudia H Kawas, William E Klunk, Walter J Koroshetz, Jennifer J
Manly, Richard Mayeux, et al. The diagnosis of dementia due to Alzheimer’s dis-
ease: Recommendations from the National Institute on Aging-Alzheimers Associ-
ation workgroups on diagnostic guidelines for Alzheimer’s disease. Alzheimer’s &
dementia, 7(3):263–269, 2011.

[225] M Jorge Cardoso, Marc Modat, Robin Wolz, Andrew Melbourne, David Cash,
Daniel Rueckert, and Sebastien Ourselin. Geodesic Information Flows: Spatially-
Variant Graphs and Their Application to Segmentation and Fusion. 2015.

[226] Neuromorphometrics Brain Atlas. http://www.neuromorphometrics.com/.

[227] Razvan V. Marinescu. BrainPainter - Brain Colouring Software (open-source).


https://github.com/mrazvan22/brain-coloring, 2019.

[228] Julie A Schneider, Zoe Arvanitakis, Woojeong Bang, and David A Bennett. Mixed
brain pathologies account for most dementia cases in community-dwelling older
persons. Neurology, 69(24):2197–2204, 2007.

[229] Julie A Schneider, Zoe Arvanitakis, Sue E Leurgans, and David A Bennett. The
neuropathology of probable Alzheimer disease and mild cognitive impairment. An-
nals of Neurology: Official Journal of the American Neurological Association and
the Child Neurology Society, 66(2):200–208, 2009.
213

[230] Alexandra L Young, Neil P Oxtoby, Jonathan Huang, Razvan V Marinescu, Pankaj
Daga, David M Cash, Nick C Fox, Sebastien Ourselin, Jonathan M Schott, Daniel C
Alexander, et al. Multiple Orderings of Events in Disease Progression. In Informa-
tion Processing in Medical Imaging, pages 711–722. Springer, 2015.

[231] Jonathan M Schott, Sebastian J Crutch, Minerva M Carrasquillo, James Uphill,


Tim J Shakespeare, Natalie S Ryan, Keir X Yong, Manja Lehmann, Nilufer Ertekin-
Taner, Neill R Graff-Radford, et al. Genetic risk factors for the posterior cortical
atrophy variant of Alzheimer’s disease. Alzheimer’s & dementia, 12(8):862–871,
2016.

[232] Peter Freeborough, Nick C Fox, et al. The boundary shift integral: an accurate and
robust measure of cerebral volume changes from registered repeat MRI. Medical
Imaging, IEEE Transactions on, 16(5):623–629, 1997.

[233] Kelvin K Leung, Matthew J Clarkson, Jonathan W Bartlett, Shona Clegg, Clif-
ford R Jack, Michael W Weiner, Nick C Fox, Sébastien Ourselin, and Alzheimer’s
Disease Neuroimaging Initiative. Robust atrophy rate measurement in Alzheimer’s
disease using multi-site serial MRI: tissue-specific intensity normalization and pa-
rameter selection. Neuroimage, 50(2):516–523, 2010.

[234] Kelvin K Leung, Josephine Barnes, Gerard R Ridgway, Jonathan W Bartlett,


Matthew J Clarkson, Kate Macdonald, Norbert Schuff, Nick C Fox, Sebastien
Ourselin, and Alzheimer’s Disease Neuroimaging Initiative. Automated cross-
sectional and longitudinal hippocampal volume measurement in mild cognitive im-
pairment and Alzheimer’s disease. Neuroimage, 51(4):1345–1359, 2010.

[235] John Platt. Sequential minimal optimization: A fast algorithm for training support
vector machines. 1998.

[236] Razvan V Marinescu, Neil P Oxtoby, Alexandra L Young, Esther E Bron, Arthur W
Toga, Michael W Weiner, Frederik Barkhof, Nick C Fox, Stefan Klein, Daniel C
Alexander, et al. TADPOLE Challenge: Prediction of Longitudinal Evolution in
Alzheimer’s Disease. arXiv preprint arXiv:1805.03909, 2018.

[237] Rafid Sukkar, Elyse Katz, Yanwei Zhang, David Raunig, and Bradley T Wyman.
Disease progression modeling using hidden Markov models. In 2012 Annual In-
ternational Conference of the IEEE Engineering in Medicine and Biology Society,
pages 2845–2848. IEEE, 2012.

[238] Gordana Derado, F DuBois Bowman, and Clinton D Kilts. Modeling the spatial
and temporal dependence in fMRI data. Biometrics, 66(3):949–957, 2010.

[239] Jung Won Hyun, Yimei Li, Chao Huang, Martin Styner, Weili Lin, Hongtu Zhu,
and Alzheimer’s Disease Neuroimaging Initiative. STGP: Spatio-temporal Gaussian
process models for longitudinal neuroimaging data. NeuroImage, 134:550–562, 2016.

[240] Marco Lorenzi, Gabriel Ziegler, Daniel C Alexander, and Sebastien Ourselin. Ef-
ficient Gaussian process-based modelling and prediction of image time series. In
International Conference on Information Processing in Medical Imaging, pages 626–
637. Springer, 2015.
214 Appendix F. Bibliography

[241] Keith A Johnson, Nick C Fox, Reisa A Sperling, and William E Klunk. Brain
imaging in Alzheimer disease. Cold Spring Harbor perspectives in medicine, page
a006213, 2012.

[242] Douglas N Greve, Claus Svarer, Patrick M Fisher, Ling Feng, Adam E Hansen,
William Baare, Bruce Rosen, Bruce Fischl, and Gitte M Knudsen. Cortical surface-
based analysis reduces bias and variance in kinetic modeling of brain PET data.
Neuroimage, 92:225–236, 2014.

[243] Douglas N Greve, David H Salat, Spencer L Bowen, David Izquierdo-Garcia,


Aaron P Schultz, Ciprian Catana, J Alex Becker, Claus Svarer, Gitte M Knud-
sen, Reisa A Sperling, et al. Different partial volume correction methods lead to
different conclusions: an 18F-FDG-PET study of aging. Neuroimage, 132:334–343,
2016.

[244] Bradford C Dickerson, Akram Bakkour, David H Salat, Eric Feczko, Jenni Pacheco,
Douglas N Greve, Fran Grodstein, Christopher I Wright, Deborah Blacker, H Di-
ana Rosas, et al. The cortical signature of Alzheimer’s disease: regionally specific
cortical thinning relates to symptom severity in very mild to mild AD dementia
and is detectable in asymptomatic amyloid-positive individuals. Cerebral cortex,
19(3):497–510, 2008.

[245] Vivek Singh, Howard Chertkow, Jason P Lerch, Alan C Evans, Adrienne E Dorr,
and Noor Jehan Kabani. Spatial patterns of cortical thinning in mild cognitive
impairment and Alzheimer’s disease. Brain, 129(11):2885–2893, 2006.

[246] Marco Lorenzi, Maurizio Filippone, Daniel C Alexander, and Sebastien Ourselin.
Disease Progression Modeling and Prediction through Random Effect Gaussian
Processes and Time Transformation. arXiv preprint arXiv:1701.01668, 2017.

[247] Marzia A Scelsi, Raiyan R Khan, Marco Lorenzi, Leigh Christopher, Michael D
Greicius, Jonathan M Schott, Sebastien Ourselin, and Andre Altmann. Genetic
study of multimodal imaging Alzheimer’s disease progression score implicates novel
loci. Brain, 2018.

[248] Marcia Hon and Naimul Khan. Towards Alzheimer’s disease classification through
transfer learning. arXiv preprint arXiv:1711.11117, 2017.

[249] Bo Cheng, Mingxia Liu, Dinggang Shen, Zuoyong Li, Daoqiang Zhang, and
Alzheimers Disease Neuroimaging Initiative. Multi-domain transfer learning for
early diagnosis of alzheimers disease. Neuroinformatics, 15(2):115–132, 2017.

[250] Bo Cheng, Mingxia Liu, Daoqiang Zhang, Brent C Munsell, and Dinggang Shen.
Domain transfer learning for MCI conversion prediction. IEEE Transactions on
Biomedical Engineering, 62(7):1805–1817, 2015.

[251] JC Baron, G Chetelat, B Desgranges, G Perchey, B Landeau, V De La Sayette, and


F Eustache. In vivo mapping of gray matter loss with voxel-based morphometry in
mild Alzheimer’s disease. Neuroimage, 14(2):298–309, 2001.
215

[252] Paul S Aisen, Ronald C Petersen, Michael C Donohue, Anthony Gamst, Rema
Raman, Ronald G Thomas, Sarah Walter, John Q Trojanowski, Leslie M Shaw,
Laurel A Beckett, et al. Clinical Core of the Alzheimer’s Disease Neuroimaging Ini-
tiative: progress and plans. Alzheimer’s & dementia: the journal of the Alzheimer’s
Association, 6(3):239–246, 2010.
[253] Giovanni B Frisoni, Nick C Fox, Clifford R Jack Jr, Philip Scheltens, and Paul M
Thompson. The clinical use of structural MRI in Alzheimer disease. Nature Reviews
Neurology, 6(2):67, 2010.
[254] Ricardo Guerrero, Alexander Schmidt-Richberg, Christian Ledig, Tong Tong,
Robin Wolz, Daniel Rueckert, and ADNI. Instantiated mixed effects modeling
of Alzheimer’s disease markers. NeuroImage, 142:113–125, 2016.
[255] Daoqiang Zhang, Yaping Wang, Luping Zhou, Hong Yuan, Dinggang Shen, and
ADNI. Multimodal classification of Alzheimer’s disease and mild cognitive impair-
ment. Neuroimage, 55(3):856–867, 2011.
[256] Jonathan Young, Marc Modat, Manuel J Cardoso, Alex Mendelson, Dave Cash,
Sebastien Ourselin, and Alzheimer’s Disease Neuroimaging Initiative. Accurate
multimodal probabilistic prediction of conversion to Alzheimer’s disease in patients
with mild cognitive impairment. NeuroImage: Clinical, 2:735–745, 2013.
[257] Jussi Mattila, Juha Koikkalainen, Arho Virkki, Anja Simonsen, Mark van Gils,
Gunhild Waldemar, Hilkka Soininen, and Jyrki Lötjönen. A disease state fingerprint
for evaluation of Alzheimer’s disease. Journal of Alzheimer’s Disease, 27(1):163–
176, 2011.
[258] Stanley Durrleman, Xavier Pennec, Alain Trouvé, José Braga, Guido Gerig, and
Nicholas Ayache. Toward a comprehensive framework for the spatiotemporal statis-
tical analysis of longitudinal shape data. International journal of computer vision,
103(1):22–59, 2013.
[259] Marco Lorenzi, Xavier Pennec, Giovanni B Frisoni, and Nicholas Ayache. Disen-
tangling normal aging from Alzheimer’s disease in structural magnetic resonance
images. Neurobiology of aging, 36:S42–S52, 2015.
[260] Michael W Weiner, Dallas P Veitch, Paul S Aisen, Laurel A Beckett, Nigel J Cairns,
Robert C Green, Danielle Harvey, Clifford R Jack, William Jagust, John C Morris,
et al. Recent publications from the Alzheimer’s Disease Neuroimaging Initiative:
Reviewing progress toward improved AD clinical trials. Alzheimer’s & dementia:
the journal of the Alzheimer’s Association, 13(4):e1–e85, 2017.
[261] Genevera I Allen, Nicola Amoroso, Catalina Anghel, Venkat Balagurusamy,
Christopher J Bare, Derek Beaton, Roberto Bellotti, David A Bennett, Kevin L
Boehme, Paul C Boutros, et al. Crowdsourced estimation of cognitive decline
and resilience in Alzheimer’s disease. Alzheimer’s & dementia: the journal of the
Alzheimer’s Association, 12(6):645–653, 2016.
[262] Ronald Carl Petersen, PS Aisen, LA Beckett, MC Donohue, AC Gamst, DJ Harvey,
CR Jack, WJ Jagust, LM Shaw, AW Toga, et al. Alzheimer’s Disease Neuroimaging
Initiative (ADNI) clinical characterization. Neurology, 74(3):201–209, 2010.
216 Appendix F. Bibliography

[263] William J Jagust, Dan Bandy, Kewei Chen, Norman L Foster, Susan M Landau,
Chester A Mathis, Julie C Price, Eric M Reiman, Daniel Skovronsky, and Robert A
Koeppe. The Alzheimer’s Disease Neuroimaging Initiative positron emission tomog-
raphy core. Alzheimer’s & dementia: the journal of the Alzheimer’s Association,
6(3):221–229, 2010.
[264] John Ashburner. Computational anatomy with the SPM software. Magnetic reso-
nance imaging, 27(8):1163–1174, 2009.
[265] Talia M Nir, Neda Jahanshad, Julio E Villalon-Reina, Arthur W Toga, Clifford R
Jack, Michael W Weiner, Paul M Thompson, and ADNI. Effectiveness of regional
DTI measures in distinguishing Alzheimer’s disease, MCI, and normal aging. Neu-
roImage: clinical, 3:180–195, 2013.
[266] Kenichi Oishi, Andreia Faria, Hangyi Jiang, Xin Li, Kazi Akhter, Jiangyang Zhang,
John T Hsu, Michael I Miller, Peter CM van Zijl, Marilyn Albert, et al. Atlas-
based whole brain white matter analysis using large deformation diffeomorphic
metric mapping: application to normal elderly and Alzheimer’s disease participants.
Neuroimage, 46(2):486–499, 2009.
[267] David J Hand and Robert J Till. A simple generalisation of the area under the ROC
curve for multiple class classification problems. Machine learning, 45(2):171–186,
2001.
[268] Kay Henning Brodersen, Cheng Soon Ong, Klaas Enno Stephan, and Joachim M
Buhmann. The balanced accuracy and its posterior distribution. In International
Conference on Pattern recognition (ICPR), pages 3121–3124. IEEE, 2010.
[269] Meryl A Butters, Oscar L Lopez, and James T Becker. Focal temporal lobe dys-
function in probable Alzheimer’s disease predicts a slow rate of cognitive decline.
Neurology, 46(3):687–692, 1996.
[270] J Green, John C Morris, J Sandson, DW McKeel, and JW Miller. Progressive
aphasia: A precursor of global dementia? Neurology, 40(3 Part 1):423–423, 1990.
[271] John DW Greene, Karalyn Patterson, John Xuereb, and John R Hodges. Alzheimer
disease and nonfluent progressive aphasia. Archives of Neurology, 53(10):1072–1078,
1996.
[272] Jason D Warren, Jonathan D Rohrer, and Martin N Rossor. Frontotemporal de-
mentia. Bmj, 347:f4827, 2013.
[273] Marvin M Goldenberg. Multiple sclerosis review. Pharmacy and Therapeutics,
37(3):175, 2012.
[274] Arman Eshaghi, Razvan V Marinescu, Alexandra L Young, Nicholas C Firth, Fer-
ran Prados, M Jorge Cardoso, Carmen Tur, Floriana De Angelis, Niamh Cawley,
Wallace J Brownlee, et al. Progression of regional grey matter atrophy in multiple
sclerosis. Brain, 141(6):1665–1677, 2018.
[275] Werner Poewe, Klaus Seppi, Caroline M Tanner, Glenda M Halliday, Patrik
Brundin, Jens Volkmann, Anette-Eleonore Schrag, and Anthony E Lang. Parkinson
disease. Nature reviews Disease primers, 3:17013, 2017.
217

[276] Nuria Caballol, Maria J Martı́, and Eduardo Tolosa. Cognitive dysfunction and de-
mentia in Parkinson disease. Movement disorders: official journal of the Movement
Disorder Society, 22(S17):S358–S366, 2007.
[277] Raymund AC Roos. Huntington’s disease: a clinical review. Orphanet journal of
rare diseases, 5(1):40, 2010.
[278] Nellie Georgiou-Karistianis, Anthony J Hannan, and Gary F Egan. Magnetic res-
onance imaging as an approach towards identifying neuropathological biomarkers
for Huntington’s disease. Brain research reviews, 58(1):209–225, 2008.
[279] Andrew Feigin, Klaus L Leenders, James R Moeller, John Missimer, Gabriella
Kuenig, Phoebe Spetsieris, Angelo Antonini, and David Eidelberg. Metabolic Net-
work Abnormalities in Early Huntington’s Disease: An 18F FDG PET Study. Jour-
nal of Nuclear Medicine, 42(11):1591–1595, 2001.
[280] Peter A Wijeratne, Alexandra L Young, Neil P Oxtoby, Razvan V Marinescu,
Nicholas C Firth, Eileanoir B Johnson, Amrita Mohan, Cristina Sampaio, Rachael I
Scahill, Sarah J Tabrizi, et al. An image-based model of brain volume biomarker
changes in Huntington’s disease. Annals of clinical and translational neurology,
5(5):570–582, 2018.
[281] David M Cash, Jonathan D Rohrer, Natalie S Ryan, Sebastien Ourselin, and Nick C
Fox. Imaging endpoints for clinical trials in Alzheimer’s disease. Alzheimer’s re-
search & therapy, 6(9):87, 2014.
[282] Gabor G Kovacs, Ivan Milenkovic, Adelheid Wöhrer, Romana Höftberger, Ellen
Gelpi, Christine Haberler, Selma Hönigschnabl, Angelika Reiner-Concin, Harald
Heinzl, Susanne Jungwirth, et al. Non-Alzheimer neurodegenerative pathologies
and their combinations are more frequent than commonly believed in the elderly
brain: a community-based autopsy series. Acta neuropathologica, 126(3):365–384,
2013.
[283] Bryan D James, Robert S Wilson, Patricia A Boyle, John Q Trojanowski, David A
Bennett, and Julie A Schneider. TDP-43 stage, mixed pathologies, and clinical
Alzheimers-type dementia. Brain, 139(11):2983–2993, 2016.
[284] John Q Trojanowski, Hugo Vandeerstichele, Magdalena Korecka, Christopher M
Clark, Paul S Aisen, Ronald C Petersen, Kaj Blennow, Holly Soares, Adam Simon,
Piotr Lewczuk, et al. Update on the biomarker core of the Alzheimer’s Disease
Neuroimaging Initiative subjects. Alzheimer’s & Dementia, 6(3):230–238, 2010.
[285] Jason L Stein, Xue Hua, Suh Lee, April J Ho, Alex D Leow, Arthur W Toga,
Andrew J Saykin, Li Shen, Tatiana Foroud, Nathan Pankratz, et al. Voxelwise
genome-wide association study (vGWAS). Neuroimage, 53(3):1160–1174, 2010.
[286] Kathryn A Ellis, Ashley I Bush, David Darby, Daniela De Fazio, Jonathan Foster,
Peter Hudson, Nicola T Lautenschlager, Nat Lenzo, Ralph N Martins, Paul Maruff,
et al. The Australian Imaging, Biomarkers and Lifestyle (AIBL) study of aging:
methodology and baseline characteristics of 1112 individuals recruited for a longitu-
dinal study of Alzheimer’s disease. International Psychogeriatrics, 21(4):672–687,
2009.
218 Appendix F. Bibliography

[287] John C Morris, Paul S Aisen, Randall J Bateman, Tammie LS Benzinger, Nigel J
Cairns, Anne M Fagan, Bernardino Ghetti, Alison M Goate, David M Holtzman,
William E Klunk, et al. Developing an international network for Alzheimer research:
the Dominantly Inherited Alzheimer Network. Clinical investigation, 2(10):975,
2012.

[288] Kenneth Marek, Danna Jennings, Shirley Lasch, Andrew Siderowf, Caroline Tan-
ner, Tanya Simuni, Chris Coffey, Karl Kieburtz, Emily Flagg, Sohini Chowdhury,
et al. The parkinson progression marker initiative (PPMI). Progress in neurobiol-
ogy, 95(4):629–635, 2011.

[289] Sarah J Tabrizi, Douglas R Langbehn, Blair R Leavitt, Raymund AC Roos, Alexan-
dra Durr, David Craufurd, Christopher Kennard, Stephen L Hicks, Nick C Fox,
Rachael I Scahill, et al. Biological and clinical manifestations of Huntington’s dis-
ease in the longitudinal TRACK-HD study: cross-sectional analysis of baseline
data. The Lancet Neurology, 8(9):791–801, 2009.

[290] M Arfan Ikram, Guy GO Brusselle, Sarwa Darwish Murad, Cornelia M van Duijn,
Oscar H Franco, André Goedegebure, Caroline CW Klaver, Tamar EC Nijsten,
Robin P Peeters, Bruno H Stricker, et al. The Rotterdam Study: 2018 update on
objectives, design and main results. European Journal of Epidemiology, 32(9):807–
850, 2017.

[291] Yu-Liang Hsu, Pau-Choo Chung, Wei-Hsin Wang, Ming-Chyi Pai, Chun-Yao Wang,
Chien-Wen Lin, Hao-Li Wu, and Jeen-Shing Wang. Gait and balance analysis for
patients with Alzheimer’s disease using an inertial-sensor-based wearable instru-
ment. IEEE journal of biomedical and health informatics, 18(6):1822–1830, 2014.

[292] J Thomas Hutton, JA Nagel, and Ruth B Loewenson. Eye tracking dysfunction in
Alzheimer-type dementia. Neurology, 34(1):99–99, 1984.

[293] Ildikó Hoffmann, Dezso Nemeth, Cristina D Dye, Magdolna Pákáski, Tamás Irinyi,
and János Kálmán. Temporal parameters of spontaneous speech in Alzheimer’s
disease. International journal of speech-language pathology, 12(1):29–34, 2010.

[294] Geoffrey E Hinton and Sam T Roweis. Stochastic neighbor embedding. In Advances
in neural information processing systems, pages 857–864, 2003.

[295] Bishesh Khanal, Marco Lorenzi, Nicholas Ayache, and Xavier Pennec. A biophysical
model of brain deformation to simulate and analyze longitudinal MRIs of patients
with Alzheimer’s disease. NeuroImage, 134:35–52, 2016.

[296] Christopher H Jackson, Linda D Sharples, Simon G Thompson, Stephen W Duffy,


and Elisabeth Couto. Multistate Markov models for disease progression with classi-
fication error. Journal of the Royal Statistical Society: Series D (The Statistician),
52(2):193–209, 2003.

[297] Chantal Guihenneuc-Jouyaux, Sylvia Richardson, and Ira M Longini. Modeling


markers of disease progression by a hidden Markov process: application to charac-
terizing CD4 cell decline. Biometrics, 56(3):733–741, 2000.
219

[298] Clifford R Jack, Heather J Wiste, Timothy G Lesnick, Stephen D Weigand, David S
Knopman, Prashanthi Vemuri, Vernon S Pankratz, Matthew L Senjem, Jeffrey L
Gunter, Michelle M Mielke, et al. Brain β-amyloid load approaches a plateau.
Neurology, 80(10):890–896, 2013.

[299] Neil P Oxtoby, Alexandra L Young, Nick C Fox, Pankaj Daga, David M Cash,
Sebastien Ourselin, Jonathan M Schott, Daniel C Alexander, and Alzheimers Dis-
ease Neuroimaging Initiative. Learning imaging biomarker trajectories from noisy
Alzheimer’s disease data using a Bayesian multilevel model. In Bayesian and grAph-
ical Models for Biomedical Imaging, pages 85–94. Springer, 2014.

[300] Alexandra L Young, Neil P Oxtoby, Jonathan M Schott, and Daniel C Alexander.
Data-driven models of neurodegenerative disease.

[301] Stephen M Stigler. Francis Galton’s account of the invention of correlation. Statis-
tical Science, 4(2):73–79, 1989.

[302] Joseph Lee Rodgers and W Alan Nicewander. Thirteen ways to look at the corre-
lation coefficient. The American Statistician, 42(1):59–66, 1988.

[303] Gabor J Szekely and Maria L Rizzo. Hierarchical clustering via joint between-within
distances: Extending Ward’s minimum variance method. Journal of classification,
22(2):151–183, 2005.

[304] Robert R Sokal. A statistical method for evaluating systematic relationships. Univ
Kans Sci Bull, 38:1409–1438, 1958.

[305] Bernhard E Boser, Isabelle M Guyon, and Vladimir N Vapnik. A training algo-
rithm for optimal margin classifiers. In Proceedings of the fifth annual workshop on
Computational learning theory, pages 144–152. ACM, 1992.

[306] Corinna Cortes and Vladimir Vapnik. Support-vector networks. Machine learning,
20(3):273–297, 1995.

[307] D Luenberger and Y Ye. Linear and nonlinear optimization. Linear and Nonlinear
Optimization, 1984.

[308] A Aizerman, Emmanuel M Braverman, and LI Rozoner. Theoretical foundations


of the potential function method in pattern recognition learning. Automation and
remote control, 25:821–837, 1964.

[309] John Moody and Christian J Darken. Fast learning in networks of locally-tuned
processing units. Neural computation, 1(2):281–294, 1989.

[310] Nicholas Metropolis, Arianna W Rosenbluth, Marshall N Rosenbluth, Augusta H


Teller, and Edward Teller. Equation of state calculations by fast computing ma-
chines. The journal of chemical physics, 21(6):1087–1092, 1953.

[311] W Keith Hastings. Monte Carlo sampling methods using Markov chains and their
applications. Biometrika, 57(1):97–109, 1970.

[312] Vikas Dhikav and Kuljeet Anand. Potential predictors of hippocampal atrophy in
Alzheimer’s disease. Drugs & aging, 28(1):1–11, 2011.
220 Appendix F. Bibliography

[313] Sebastian J Crutch, Jonathan M Schott, Gil D Rabinovici, Bradley F Boeve, Ste-
fano F Cappa, Bradford C Dickerson, Bruno Dubois, Neill R Graff-Radford, Pierre
Krolak-Salmon, Manja Lehmann, et al. Shining a light on posterior cortical atrophy.
Alzheimer’s & Dementia, 9(4):463–465, 2013.

[314] Clifford R Jack, Matt A Bernstein, Nick C Fox, Paul Thompson, Gene Alexander,
Danielle Harvey, Bret Borowski, Paula J Britson, Jennifer L Whitwell, Chadwick
Ward, et al. The Alzheimer’s disease neuroimaging initiative (ADNI): MRI meth-
ods. Journal of Magnetic Resonance Imaging, 27(4):685–691, 2008.

[315] Wilma G Rosen, Richard C Mohs, and Kenneth L Davis. A new rating scale for
Alzheimer’s disease. The American journal of psychiatry, 1984.

[316] Andŕe Rey. L’examen clinique en psychologie. 1958.

[317] Elizabeth K Warrington and Merle James. The visual object and space perception
battery. Thames Valley Test Company Bury St Edmunds, 1991.

[318] Elizabeth K Warrington, Tim Shallice, et al. Category specific semantic impair-
ments. Brain, 107(3):829–853, 1984.

[319] David Wechsler. Wechsler adult intelligence scale–Fourth Edition (WAIS–IV). San
Antonio, TX: NCS Pearson, 2008.

[320] John TE Richardson. Measures of short-term memory: a historical review. Cortex,


43(5):635–650, 2007.

[321] Nick C Fox and Peter A Freeborough. Brain atrophy progression measured from
registered serial MRI: validation and application to Alzheimer’s disease. Journal of
Magnetic Resonance Imaging, 7(6):1069–1075, 1997.

[322] PJ Nestor, D Caine, TD Fryer, J Clarke, and JR Hodges. The topography of


metabolic deficits in posterior cortical atrophy (the visual variant of Alzheimer’s
disease) with FDG-PET. Journal of Neurology, Neurosurgery & Psychiatry,
74(11):1521–1529, 2003.

[323] Klaus Schmidtke, PD Dr Michael Hüll, and Jochen Talazko. Posterior cortical
atrophy: variant of Alzheimer’s disease? Journal of neurology, 252(1):27–35, 2005.

[324] Roger Bullock. Future directions in the treatment of Alzheimer’s disease. Expert
opinion on investigational drugs, 13(4):303–314, 2004.

[325] Hans-Wolfgang Klafki, Matthias Staufenbiel, Johannes Kornhuber, and Jens Wilt-
fang. Therapeutic approaches to Alzheimer’s disease. Brain, 129(11):2840–2855,
2006.

[326] Jeffrey L. Cummings. Alzheimers Disease. New England Journal of Medicine,


351(1):56–67, 2004. PMID: 15229308.

[327] Elizabeth Forsyth and Pamela D Ritzline. An overview of the etiology, diagnosis,
and treatment of Alzheimer disease. Physical therapy, 78(12):1325–1331, 1998.
221

[328] Eric Yang, Michael Farnum, Victor Lobanov, Tim Schultz, R Verbeeck, N Ragha-
van, MN Samtani, G Novak, V Narayan, and A DiBernardo. Quantifying the
pathophysiological timeline of Alzheimer’s disease. Journal of Alzheimer’s disease:
JAD, 26(4):745–753, 2010.

[329] George T Grossberg and Abhilash K Desai. Management of Alzheimer’s disease.


The Journals of Gerontology Series A: Biological Sciences and Medical Sciences,
58(4):M331–M353, 2003.

[330] Eunhee Kim, Yunsoo Lee, Jongkeol Lee, and Seol-Heui Han. A case with
cholinesterase inhibitor responsive asymmetric posterior cortical atrophy. Clini-
cal neurology and neurosurgery, 108(1):97–101, 2005.

[331] M Wakai, H Honda, A Takahashi, T Kato, K Ito, and T Hamanaka. Unusual


findings on PET study of a patient with posterior cortical atrophy. Acta neurologica
scandinavica, 89(6):458–461, 1994.

[332] Alexander Schmidt-Richberg, Ricardo Guerrero, Christian Ledig, Helena Molina-


Abril, Alejandro F Frangi, Daniel Rueckert, and Alzheimers Disease Neuroimaging
Initiative. Multi-stage Biomarker Models for Progression Estimation in Alzheimer’s
Disease. In International Conference on Information Processing in Medical Imaging,
pages 387–398. Springer, 2015.

[333] Alexandra L Young, Neil P Oxtoby, Sebastien Ourselin, Jonathan M Schott,


Daniel C Alexander, and Alzheimer’s Disease Neuroimaging Initiative. A simu-
lation system for biomarker evolution in neurodegenerative disease. Medical image
analysis, 26(1):47–56, 2015.

[334] Antonio Convit, Mony de Leon, Chaim Tarshish, Susan De Santi, Alan Kluger,
Henry Rusinek, and AjaxE George. Hippocampal volume losses in minimally im-
paired elderly. The Lancet, 345(8944):266, 1995.

[335] Y Xu, CR Jack, PC Obrien, E Kokmen, Glenn E Smith, Robert J Ivnik, Bradley F
Boeve, RG Tangalos, and Ronald C Petersen. Usefulness of MRI measures of
entorhinal cortex versus hippocampus in AD. Neurology, 54(9):1760–1767, 2000.

[336] Motohiro Kiyosawa, Thomas M Bosley, John Chawluk, Dara Jamieson, Norman J
Schatz, Peter J Savino, Robert C Sergott, Martin Reivich, and Abass Alavi.
Alzheimer’s disease with prominent visual symptoms: clinical and metabolic eval-
uation. Ophthalmology, 96(7):1077–1086, 1989.

[337] Corina Pennanen, Miia Kivipelto, Susanna Tuomainen, Päivi Hartikainen, Tuomo
Hänninen, Mikko P Laakso, Merja Hallikainen, Matti Vanhanen, Aulikki Nissinen,
Eeva-Liisa Helkala, et al. Hippocampus and entorhinal cortex in mild cognitive
impairment and early AD. Neurobiology of aging, 25(3):303–310, 2004.

[338] Heiko Braak and Eva Braak. Morphological criteria for the recognition of
Alzheimer’s disease and the distribution pattern of cortical changes related to this
disorder. Neurobiology of aging, 15(3):355–356, 1994.
222 Appendix F. Bibliography

[339] Mikko P Laakso, Giovanni B Frisoni, Mervi Könönen, Mia Mikkonen, Alberto Bel-
tramello, Claudia Geroldi, Angelo Bianchetti, Marco Trabucchi, Hilkka Soininen,
and Hannu J Aronen. Hippocampus and entorhinal cortex in frontotemporal de-
mentia and Alzheimer’s disease: a morphometric MRI study. Biological psychiatry,
47(12):1056–1063, 2000.
[340] João Maroco, Dina Silva, Ana Rodrigues, Manuela Guerreiro, Isabel Santana, and
Alexandre de Mendonça. Data mining methods in the prediction of Dementia:
A real-data comparison of the accuracy, sensitivity and specificity of linear dis-
criminant analysis, logistic regression, neural networks, support vector machines,
classification trees and random forests. BMC research notes, 4(1):299, 2011.
[341] Rik Ossenkoppele, Brendan I Cohn-Sheehy, Renaud La Joie, Jacob W Vogel, Chris-
tiane Möller, Manja Lehmann, Bart NM van Berckel, William W Seeley, Yolande A
Pijnenburg, Maria L Gorno-Tempini, et al. Atrophy patterns in early clinical
stages across distinct phenotypes of A lzheimer’s disease. Human brain mapping,
36(11):4421–4437, 2015.
[342] Bengt Winblad, Philippe Amouyel, Sandrine Andrieu, Clive Ballard, Carol Brayne,
Henry Brodaty, Angel Cedazo-Minguez, Bruno Dubois, David Edvardsson, Howard
Feldman, et al. Defeating Alzheimer’s disease and other dementias: a priority for
European science and society. The Lancet Neurology, 15(5):455–532, 2016.
[343] Heiko Braak and Kelly Del Tredici. Potential pathways of abnormal tau and α-
synuclein dissemination in sporadic Alzheimer’s and Parkinson’s diseases. Cold
Spring Harbor perspectives in biology, page a023630, 2016.
[344] Bess Frost and Marc I Diamond. Prion-like mechanisms in neurodegenerative dis-
eases. Nature Reviews Neuroscience, 11(3):155, 2010.
[345] John Hardy and Tamas Revesz. The spread of neurodegenerative disease. New
England Journal of Medicine, 366(22):2126–2128, 2012.
[346] Zeshan Ahmed, Jane Cooper, Tracey K Murray, Katya Garn, Emily McNaughton,
Hannah Clarke, Samira Parhizkar, Mark A Ward, Annalisa Cavallini, Samuel Jack-
son, et al. A novel in vivo model of tau propagation with rapid and progressive
neurofibrillary tangle pathology: the pattern of spread is determined by connectiv-
ity, not proximity. Acta neuropathologica, 127(5):667–683, 2014.
[347] Johannes Brettschneider, Kelly Del Tredici, Virginia M-Y Lee, and John Q Tro-
janowski. Spreading of pathology in neurodegenerative diseases: a focus on human
studies. Nature Reviews Neuroscience, 16(2):109, 2015.
[348] Michel Goedert. Alzheimers and Parkinsons diseases: The prion concept in relation
to assembled Aβ, tau, and α-synuclein. Science, 349(6248):1255555, 2015.
[349] Jeffrey L Cummings. Neurodegenerative Disorders as Proteinopathies: Phenotypic
Relationships. In Genotype–Proteotype–Phenotype Relationships in Neurodegener-
ative Diseases, pages 1–10. Springer, 2005.
[350] Massimo Filippi and Federica Agosta. Structural and functional network connec-
tivity breakdown in Alzheimer’s disease studied with magnetic resonance imaging
techniques. Journal of Alzheimer’s Disease, 24(3):455–474, 2011.
223

[351] Jason D Warren, Jonathan D Rohrer, Jonathan M Schott, Nick C Fox, John Hardy,
and Martin N Rossor. Molecular nexopathies: a new paradigm of neurodegenerative
disease. Trends in neurosciences, 36(10):561–569, 2013.
[352] Patrick R Hof, Constantin Bouras, Jean Constantinidis, and John H Morrison. Se-
lective disconnection of specific visual association pathways in cases of Alzheimer’s
disease presenting with Balint’s syndrome. Journal of neuropathology and experi-
mental neurology, 49(2):168–184, 1990.
[353] DF Tang-Wai, KA Josephs, Bradley F Boeve, DW Dickson, JE Parisi, and RC Pe-
tersen. Pathologically confirmed corticobasal degeneration presenting with visu-
ospatial dysfunction. Neurology, 61(8):1134–1135, 2003.
[354] DF Tang-Wai, KA Josephs, Bradley F Boeve, RC Petersen, JE Parisi, and
DW Dickson. Coexistent Lewy body disease in a case of visual variant of Alzheimer’s
disease. Journal of Neurology, Neurosurgery & Psychiatry, 74(3):389–389, 2003.
[355] JA Renner, JM Burns, CE Hou, DW McKeel, M Storandt, and JC Morris. Progres-
sive posterior cortical dysfunction A clinicopathologic series. Neurology, 63(7):1175–
1180, 2004.
[356] Manja Lehmann, Andrew Melbourne, John C Dickson, Rebekah M Ahmed, Marc
Modat, M Jorge Cardoso, David L Thomas, Enrico De Vita, Sebastian J Crutch,
Jason D Warren, et al. A novel use of arterial spin labelling MRI to demonstrate
focal hypoperfusion in individuals with posterior cortical atrophy: a multimodal
imaging study. J Neurol Neurosurg Psychiatry, pages jnnp–2015, 2016.
[357] Rik Ossenkoppele, Niklas Mattsson, Charlotte E Teunissen, Frederik Barkhof,
Yolande Pijnenburg, Philip Scheltens, Wiesje M van der Flier, and Gil D Rabi-
novici. Cerebrospinal fluid biomarkers and cerebral atrophy in distinct clinical
variants of probable Alzheimer’s disease. Neurobiology of aging, 36(8):2340–2347,
2015.
[358] Rik Ossenkoppele, Daniel R Schonhaut, Michael Schöll, Samuel N Lockhart, Nage-
han Ayakta, Suzanne L Baker, James P ONeil, Mustafa Janabi, Andreas Lazaris,
Averill Cantwell, et al. Tau PET patterns mirror clinical and neuroanatomical
variability in Alzheimer’s disease. Brain, 139(5):1551–1567, 2016.
[359] Guillaume Dorothée, Michel Bottlaender, Edmond Moukari, Leonardo C De Souza,
Renaud Maroy, Fabian Corlier, Olivier Colliot, Marie Chupin, Foudil Lamari,
Stephane Lehéricy, et al. Distinct patterns of antiamyloid-β antibodies in typi-
cal and atypical Alzheimer disease. Archives of neurology, 69(9):1181–1185, 2012.
[360] William C Kreisl, Chul Hyoung Lyoo, Jeih-San Liow, Joseph Snow, Emily Page,
Kimberly J Jenko, Cheryl L Morse, Sami S Zoghbi, Victor W Pike, R Scott Turner,
et al. Distinct patterns of increased translocator protein in posterior cortical atrophy
and amnestic Alzheimer’s disease. Neurobiology of aging, 51:132–140, 2017.
[361] Keir Yong, Kishan Rajdev, Elizabeth Warrington, Jennifer Nicholas, Jason Warren,
and Sebastian Crutch. A longitudinal investigation of the relationship between
crowding and reading: A neurodegenerative approach. Neuropsychologia, 85:127–
136, 2016.
224 Appendix F. Bibliography

[362] Silvia Primativo, Keir XX Yong, Timothy J Shakespeare, and Sebastian J Crutch.
The oral spelling profile of posterior cortical atrophy and the nature of the
graphemic representation. Neuropsychologia, 94:61–74, 2017.

[363] Rik Ossenkoppele, Brendan I Cohn-Sheehy, Renaud La Joie, Jacob W Vogel, Chris-
tiane Möller, Manja Lehmann, Bart NM van Berckel, William W Seeley, Yolande A
Pijnenburg, Maria L Gorno-Tempini, et al. Atrophy patterns in early clinical
stages across distinct phenotypes of A lzheimer’s disease. Human brain mapping,
36(11):4421–4437, 2015.

[364] Flora H Duits, Charlotte E Teunissen, Femke H Bouwman, Pieter-Jelle Visser,


Niklas Mattsson, Henrik Zetterberg, Kaj Blennow, Oskar Hansson, Lennart
Minthon, Niels Andreasen, et al. The cerebrospinal fluid Alzheimer profile: easily
said, but what does it mean? Alzheimer’s & Dementia, 10(6):713–723, 2014.

[365] Leslie M Shaw, Hugo Vanderstichele, Malgorzata Knapik-Czajka, Christopher M


Clark, Paul S Aisen, Ronald C Petersen, Kaj Blennow, Holly Soares, Adam Simon,
Piotr Lewczuk, et al. Cerebrospinal fluid biomarker signature in Alzheimer’s disease
neuroimaging initiative subjects. Annals of neurology, 65(4):403–413, 2009.

[366] Bradford C Dickerson, David A Wolk, and Alzheimer’s Disease Neuroimaging Ini-
tiative. Dysexecutive versus amnesic phenotypes of very mild Alzheimer’s disease
are associated with distinct clinical, genetic and cortical thinning characteristics.
Journal of Neurology, Neurosurgery & Psychiatry, pages jnnp–2009, 2010.

[367] Jennifer L Whitwell, Stephen D Weigand, Bradley F Boeve, Matthew L Senjem, Jef-
frey L Gunter, Mariely DeJesus-Hernandez, Nicola J Rutherford, Matthew Baker,
David S Knopman, Zbigniew K Wszolek, et al. Neuroimaging signatures of fron-
totemporal dementia genetics: C9ORF72, tau, progranulin and sporadics. Brain,
135(3):794–806, 2012.

[368] Helena Chang Chui, Evelyn Lee Teng, Victor W Henderson, and Arthur C Moy.
Clinical subtypes of dementia of the Alzheimer type. Neurology, 35(11):1544–1544,
1985.

[369] Nancy J Fisher, Byron P Rourke, Linas Bieliauskas, Bruno Giordani, Stan-
ley Berent, and Norman L Foster. Neuropsychological subgroups of patients
with Alzheimer’s disease. Journal of clinical and experimental neuropsychology,
18(3):349–370, 1996.

[370] Julene K Johnson, Elizabeth Head, Ronald Kim, Arnold Starr, and Carl W Cot-
man. Clinical and pathological evidence for a frontal variant of Alzheimer disease.
Archives of neurology, 56(10):1233–1239, 1999.

[371] Benjamin Lam, Mario Masellis, Morris Freedman, Donald T Stuss, and Sandra E
Black. Clinical, imaging, and pathological heterogeneity of the Alzheimer’s disease
syndrome. Alzheimer’s research & therapy, 5(1):1, 2013.

[372] Atonella Cappa, ML Calcagni, Giampiero Villa, Alessandro Giordano, Camillo


Marra, Giuseppe De Rossi, M Puopolo, and Guido Gainotti. Brain perfusion abnor-
malities in Alzheimer’s disease: comparison between patients with focal temporal
225

lobe dysfunction and patients with diffuse cognitive impairment. Journal of Neu-
rology, Neurosurgery & Psychiatry, 70(1):22–27, 2001.

[373] Suvarna Alladi, John Xuereb, Thomas Bak, Peter Nestor, Jonathan Knibb, Karalyn
Patterson, and JR Hodges. Focal cortical presentations of Alzheimer’s disease.
Brain, 130(10):2636–2645, 2007.

[374] Răzvan Valentin Marinescu, Arman Eshaghi, Marco Lorenzi, Alexandra L Young,
Neil P Oxtoby, Sara Garbarino, Timothy J Shakespeare, Sebastian J Crutch,
Daniel C Alexander, and Alzheimers Disease Neuroimaging Initiative. A vertex
clustering model for disease progression: application to cortical thickness images.
In International Conference on Information Processing in Medical Imaging, pages
134–145. Springer, 2017.

[375] R Duara, DA Loewenstein, E Potter, J Appel, MT Greig, R Urs, Q Shen, A Raj,


B Small, W Barker, et al. Medial temporal lobe atrophy on MRI scans and the
diagnosis of Alzheimer disease. Neurology, 71(24):1986–1992, 2008.

You might also like