Past, Present and Future Alzheimer and OCT

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 38

Progress in Retinal and Eye Research 83 (2021) 100938

Contents lists available at ScienceDirect

Progress in Retinal and Eye Research


journal homepage: www.elsevier.com/locate/preteyeres

Past, present and future role of retinal imaging in


neurodegenerative disease
Amir H. Kashani a, g, 1, *, Samuel Asanad b, 1, Jane W. Chan c, 1, Maxwell B. Singer a, 1,
Jiong Zhang d, 1, Mona Sharifi d, 1, Maziyar M. Khansari d, 1, Farzan Abdolahi a, 1, Yonggang Shi d, 1,
Alessandro Biffi e, 1, Helena Chui f, 1, John M. Ringman f, 1
a
USC Roski Eye Institute and Department of Ophthalmology, Keck School of Medicine of University of Southern California, Los Angeles, CA, USA
b
Department of Ophthalmology and Visual Sciences, University of Maryland School of Medicine, Baltimore, MD, USA
c
Doheny Eye Institute, University of California, Los Angeles, Los Angeles, CA, USA
d
USC Stevens Neuroimaging and Informatics Institute, Keck School of Medicine of University of Southern California, Los Angeles, CA, USA
e
Department of Neurology, Massachusetts General Hospital, Harvard Medical School, Boston, MA, USA
f
Department of Neurology, Keck School of Medicine of University of Southern California, Los Angeles, CA, USA
g
USC Ginsburg Institute for Biomedical Therapeutics, Keck School of Medicine of University of Southern California, Los Angeles, CA, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Retinal imaging technology is rapidly advancing and can provide ever-increasing amounts of information about the
Retina structure, function and molecular composition of retinal tissue in humans in vivo. Most importantly, this information
Imaging can be obtained rapidly, non-invasively and in many cases using Food and Drug Administration-approved devices
Alzheimer’s disease
that are commercially available. Technologies such as optical coherence tomography have dramatically changed
Parkinson’s disease
our understanding of retinal disease and in many cases have significantly improved their clinical management.
Huntington’s disease
Cerebral small vessel disease Since the retina is an extension of the brain and shares a common embryological origin with the central nervous
system, there has also been intense interest in leveraging the expanding armamentarium of retinal imaging tech­
nology to understand, diagnose and monitor neurological diseases. This is particularly appealing because of the high
spatial resolution, relatively low-cost and wide availability of retinal imaging modalities such as fundus photog­
raphy or OCT compared to brain imaging modalities such as magnetic resonance imaging or positron emission
tomography. The purpose of this article is to review and synthesize current research about retinal imaging in
neurodegenerative disease by providing examples from the literature and elaborating on limitations, challenges and
future directions. We begin by providing a general background of the most relevant retinal imaging modalities to
ensure that the reader has a foundation on which to understand the clinical studies that are subsequently discussed.
We then review the application and results of retinal imaging methodologies to several prevalent neurodegener­
ative diseases where extensive work has been done including sporadic late onset Alzheimer’s Disease, Parkinson’s
Disease and Huntington’s Disease. We also discuss Autosomal Dominant Alzheimer’s Disease and cerebrovascular
small vessel disease, where the application of retinal imaging holds promise but data is currently scarce. Although
cerebrovascular disease is not generally considered a neurodegenerative process, it is both a confounder and
contributor to neurodegenerative disease processes that requires more attention. Finally, we discuss ongoing efforts
to overcome the limitations in the field and unmet clinical and scientific needs.

unique opportunity to study CNS pathology because of the shared


1. Introduction embryological origins, structure and physiology with the retina. Over
the last several decades, many investigators have attempted to leverage
The retina is the only portion of the central nervous system (CNS) the optical accessibility of the retina to better understand, diagnose and
that is optically accessible for high-resolution imaging. This presents a even treat neurodegenerative diseases including sporadic late onset

* Corresponding author. Wilmer Eye Institute, Johns Hopkins School of Medicine, Baltimore, MD, 21287, USA.
E-mail address: akashan1@jhmi.edu (A.H. Kashani).
1
Percentage of work contributed by each author in the production of the manuscript is as follows: Amir H Kashani: 35%, Samuel Asanad: 15%, John Ringman: 5%,
Alessandro Bissi: 5%, Yonggang Shi: 5%, Helena Chui: 5%, Jane Chen: 5%, Jiong Zhang: 5%, Mona Sharifi: 5%, Maziar Khansari: 5%, Maxwell Singer: 5%, Farzan
Abdolahi: 5%.

https://doi.org/10.1016/j.preteyeres.2020.100938
Received 21 August 2020; Received in revised form 11 December 2020; Accepted 17 December 2020
Available online 15 January 2021
1350-9462/© 2021 Elsevier Ltd. All rights reserved.
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

List of abbreviations IPL Inner Plexiform Layer


INL Inner Nuclear Layer
CNS Central Nervous System ONL Outer Nuclear Layer
LOAD Late Onset Alzheimer’s Disease OPL Outer Plexiform Layer
ADAD Autosomal Dominant Alzheimer’s Disease GCIPL Ganglion Cell Layer + Inner Plexiform Layer
HD Huntington’s Disease GCC Retinal Ganglion Cell Complex (mRNFL + GCL + IPL)
PD Parkinson’s Disease PCA Posterior Cortical Atrophy
AD Alzheimer’s Disease NMO Neuromyelitis Optica
FTD Frontotemporal Dementia MOG Myelin Oligodendrocyte Glycoprotein
CSVD Cerebral Small Vessel Disease mRGC Melanopsin containing RGC
MS Multiple Sclerosis CSF Cerebrospinal Fluid
OCT Optical Coherence Tomography VCID Vascular Cognitive Impairment and Dementia
OCTA Optical Coherence Tomography Angiography FAZ Foveal Avascular Zone
PET Positron Emission Tomography ffERG Full Field ERG
MRI Magnetic Resonance Imaging SRL Superficial Retinal Layer
RGC Retinal Ganglion Cells DRL Deep Retinal Layer
LGN Lateral Geniculate Nucleus MoCA Montreal Cognitive Assessment
BRB Blood Retinal Barrier MMSE Mini-Mental Status Exam
BBB Blood Brain Barrier CAG Cytosine-adenine-guanine
RNFL Retinal Nerve Fiber Layer Htt huntington protein
pRNFL Peripapillary RNFL HC Healthy Control
mRNFL Macular RNFL VA Visual Acuity
αsyn Alpha-Synuclein DFE Dilated Fundus Examination
Aβ Beta-Amyloid IOP Intraocular Pressure
ptau Hyperphosphorylated Tau PSEN1 Presenilin 1
CFP Color Fundus Photography PSEN2 Presenilin 2
FAF Fundus Autofluorescence APP Amyloid Precursor Protein
FOV Field-of-View TDP43 Transactive Response DNA Binding Domain Protein 43
RPE Retinal Pigment Epithelium ALS Amyotrophic Lateral Sclerosis
HRI Hyperspectral Retinal Imaging CAA Cerebral Amyloid Angiopathy
FLIO Fluorescence Lifetime Imaging Ophthalmoscopy WMH White Matter Hyperintensity
ERG Electroretinography CMB Cerebral Microbleeds
mfERG Multifocal ERG HTNA Hypertension Associated Angiopathy
PERG Pattern ERG CADASIL Cerebral autosomal dominant arteriopathy with
VEP Visual Evoke Potentials subcortical infarcts and leukoencephalopathy
ON Optic Neuritis

Alzheimer’s Disease (LOAD), Huntington’s Disease (HD), Parkinson’s even primary diagnostic criteria for any neurodegenerative process, but
Disease (PD), Multiple Sclerosis (MS), cerebral small vessel disease there are at least three scenarios in which it might prove useful. First, in
(CSVD) and Frontotemporal Dementia (FTD). In many cases these vivo retinal findings in human subjects with neurodegenerative diseases
studies have demonstrated significant associations of retinal thickness may provide new insight into the underlying disease pathophysiology.
and function with disease severity and have provided tantalizing pos­ Second, in vivo retinal imaging may be a relatively low-cost and widely
sibilities for disease assessment. available screening tool for assessing risk for certain neurodegenerative
The development of imaging techniques such as optical coherence disease if appropriate biomarkers are identified. This would be a valu­
tomography (OCT) and optical coherence tomography angiography able tool in recruiting appropriate populations for clinical trials and for
(OCTA) have reinvigorated the search for retinal changes that are more resource intensive testing such as brain imaging. Lastly, under the
associated with neurodegenerative disease. Although clear clinical ap­ most ideal circumstances, in vivo retinal imaging may be a useful pri­
plications are needed, there are several reasons to continue the search mary or secondary endpoint in clinical trials.
for in vivo retinal changes in neurodegenerative disease. First, diagnosis In this article, we discuss the literature on the most prevalent and
of many neurodegenerative diseases requires significant time, resources well-described neurodegenerative diseases with retinal manifestations
and cost. For example, positron emission tomography (PET) and mag­ and examine how rapidly evolving retinal imaging modalities can be
netic resonance imaging (MRI) are commonly used in the evaluation of leveraged to advance our understanding of these diseases. While there
intracranial pathology but both are time-intensive, not universally are currently few clinical indications for retinal imaging in neurode­
available and come at significant cost. In addition, neither has the spatial generative diseases, the opportunity clearly exists and future advances
resolution to detect sub-millimeter changes in tissue pathology with in our understanding of the diseases and imaging methods may enable
clinically useful reproducibility. In contrast, in vivo OCT imaging of the such indications. By consolidating the knowledge about retinal imaging
retina provides near histologic level resolution of neurosensory tissue findings in this comprehensive review we hope to provide a very useful
and the recent advent of OCTA provides similar information about resource for ophthalmologists, neurologists, researchers and students to
retinal blood vessels, including capillaries. Therefore, there exists the understand the current state-of-the-art and advance the field in new
possibility that in vivo retinal imaging can provide useful biomarkers of directions. It is important to mention that many neurodegenerative
neurodegenerative disease much earlier in the disease process than diseases have extraretinal ophthalmic manifestations (e.g. oculomotor
conventional neuroimaging methods or clinical examination. It is un­ abnormalities) that have been well-described in the neuro-
likely and impractical to think that retinal imaging will be the sole or ophthalmologic literature and which we will not discuss here. We also

2
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

do not attempt to review the vast body of literature from informative neurofibrillary tangles (Leuba and Saini 1995) although the topographic
animal models of neurodegenerative disease so that we can focus on the distribution of these changes does not consistently correspond with the
already vast number of applications of retinal imaging in human patterns of degeneration observed in the retinal nerve fiber layer
neurodegenerative disease in this manuscript. (RNFL). Analogous evidence exists from case reports of occipital lesions
in humans (Meier et al., 2015; Goto et al., 2016). Histopathologic
2. Rationale for retinal imaging in neurodegenerative disease analysis of the visual system in primates with occipital lesions also re­
veals evidence of transsynaptic degeneration of RGCs (Cowey et al.,
2.1. Embryology 2011). Localized defects in the RNFL are also significantly associated
with acute and chronic stroke independent of systemic vascular risk
The retina and optic nerve are derived from the neuroectoderm factors such as hypertension and diabetes mellitus (Wang et al., 2014).
around 23 days of gestation when they invaginate from the dienceph­ RNFL defects are also significantly associated with infarcts and white
alon (Cameron et al., 2017). Therefore the retina shares a common matter lesions on MRI in large population-based studies (Kim et al.,
cellular origin with brain tissue and is considered part of the CNS. 2011; Mauschitz et al., 2018). In subjects with PD, the laterality of
Retinal neurons have many structural and functional similarities to retinal findings seems to correlate with the more affected hemisphere
other neurons in the brain. Retinal ganglion cells (RGC) have axons that suggesting a neuroanatomic link (Pilat et al., 2016).
extend to the lateral geniculate nucleus (LGN) via a nerve fiber tract Lastly, it is likely that a combination of concurrent and sequential
complete with an oligodendrocytic myelin sheath. Retinal neurons are degeneration is present in variable amounts depending on the specific
generally excitable and have synaptic interconnections mediated by pathological process and its duration. We should note that these two
neurotransmitters including acetylcholine, dopamine, glutamate, mechanisms described above are therefore not mutually exclusive.
glycine, and gaba-aminobutyric acid (Gregg et al., 2013). This similarity
led to investigation of the pupillary response to dilute tropicamide, a 3. Imaging methodologies
cholinergic antagonist, as a diagnostic test for LOAD, a disease in which
CNS cholinergic tone is widely depressed (Scinto et al., 1994; Kardon There are numerous methods of imaging the retina ranging from
1998; Iijima et al., 2003). Similar to other axonal pathways, damage to relatively simple color photography to functional imaging of retinal
the optic nerve causes retrograde and anterograde degeneration of the electrical activity. This wealth of imaging technology combined with the
axons, associated cell bodies and target tissues. For example, atrophy of optical accessibility of the retina has been one of the main motivations in
the retina is inversely associated with brain weight in persons with MS exploring the role of retinal pathology in neurodegenerative diseases.
(Green et al., 2010). A broad range of biophysical, pathological, clinical Broadly speaking these imaging modalities can be divided into those
and epidemiologic studies demonstrate that retinal blood vessels share that demonstrate structural features of the retina and those that measure
many similarities with cerebral vessels and undergo similar pathologic some aspect of retinal function. Structural features include size of retinal
changes implicated in cerebral small vessel disease (CSVD) (Zamir vessels, retinal lesions and thickness of retinal layers. Functional fea­
1976a,b; Sherman 1981; Cogan and Kuwabara 1984; Törnquist and Alm tures of the retina that are quantified by imaging methods include
1986; Frank et al., 1990; Ravalico et al., 1996; Trost et al., 2016). Retinal retinal light sensitivity, blood flow and electrical activity. Below we
and cerebral blood vessels share significant similarity in capillary enumerate and describe the key modalities that have been used to study
branching angles and organization (Cogan and Kuwabara 1984), retinal changes in neurodegenerative diseases. In the subsequent sec­
carrier-mediated transport functions of blood-retina barrier (BRB) and tions we discuss the applications of these imaging modalities in disease-
blood brain barrier (BBB) (Törnquist and Alm 1986), structural features specific contexts. It is important to note that most of these modalities are
of capillaries and tight junctions (Trost et al., 2016), and approved for clinical use in ophthalmology practices by the Food and
pericyte-to-endothelial cell ratio (Frank et al., 1990). In addition to these Drug Administration (FDA). Therefore they provide readily available
similarities in health, there are many retinal manifestations of CNS and clinically feasible tools for assessing retinal pathology in clinical
disease as we will discuss below. trials and health care settings.

2.2. Pathophysiological mechanisms of retinal degeneration in


neurodegenerative disease 3.1. Structural imaging

Given the embryological and physiological similarity between the The most widely used and well-accepted modalities in retinal im­
neurosensory retina, retinal vasculature and brain, there are at least two aging provide some measure of retinal structure whether that is the
plausible pathophysiological mechanisms by which neurodegenerative qualitative physical appearance of the retina or a quantitative measure
changes may manifest in the retina. One plausible mechanism is that the such as thickness. The sections below are not meant to be exhaustive
underlying neurodegenerative process concurrently occurs in retinal reviews of the imaging methodologies. Rather, these sections are meant
tissue as well as in the brain. Abnormal intracellular and extracellular to provide an overview of the methods so that the relevance to neuro­
deposition of proteins such as tau, alpha-synuclein (αsyn), and beta- degenerative diseases discussed in subsequent sections can be more
amyloid (Aβ) among others are present in several neurodegenerative easily understood.
disorders and are also found in the aging retina (Löffler et al., 1995;
Leger et al., 2011). There is also evidence that Aβ, tau and αsyn accu­ 3.1.1. Fundus photography
mulate in the retina of subjects with AD (Koronyo-Hamaoui et al., 2011)
and PD (Bodis-Wollner et al., 2014), respectively, and may therefore 3.1.1.1. Color fundus imaging. Color fundus photography (CFP) repre­
mediate the same neurotoxicity in RGCs as in other CNS tissue. In one sents the most common and widely used retinal imaging modality
study, atrophy of specific classes of RGCs was shown to be associated (Yannuzzi et al., 2004; Panwar et al., 2016). CFP was traditionally
with disturbance of the sleep-wake cycle in persons with AD (La Morgia, performed with film but is now available in true color and pseudocolor
Ross-Cisneros et al., 2016). digital format as well as multi-wavelength scanning laser ophthalmos­
A second plausible mechanism is that neurodegenerative changes in copy. These cameras acquire information with relatively low spectral
the CNS cause some form of retrograde degeneration of ganglion cell resolution (red, green and blue color channels) but with high spatial
axons and associated cell bodies located in the retina (Peng et al., 2020). resolution approaching dozens of microns. Most fundus cameras require
There is some evidence that diencephalic and brainstem visual centers pharmacologic pupillary dilation (mydriasis) which can be a barrier to
such as the LGN and superior colliculus manifest senile plaques and use in non-ophthalmic settings. Non-mydriatic cameras are becoming

3
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

more widely available but often have decreased resolution or primarily thought to be a property of the RPE and not the neurosensory
field-of-view (FOV). retina itself, at least in normal subjects. In addition, FAF of the periph­
It is important to note that there can be significant artifactual vari­ eral retina is relatively poorly studied, only recently becoming possible
ations in the spectral patterns (color) of the normal retina using pseu­ with the advent of widefield imaging. Quantitative FAF measurements
docolor digital systems. The color of retinal lesions can be significantly from the retina can vary dramatically, even within the same subject, due
altered by the laser calibration of the camera and media properties of the to relatively subtle changes in the optical media and light sources (Delori
eye. Therefore, depending on the lesion of interest, some cameras may et al., 2011). Therefore, in the absence of carefully calibrated mea­
be more suitable than others. One of the most important variables in CFP surements quantitative FAF measurements with current commercially
is the FOV which can range from 30 degrees of the retinal surface to available devices may not be reliable for longitudinal measurements
more than 100◦ (Panwar et al., 2016). Digital montages of 30–50 degree even within the same subject, at least in the absence of any significant
FOV are also possible to allow for evaluation of peripheral retinal calibration efforts.
findings (Fig. 1). Common uses of CFP include identifying retinal lesions Stimulation of the retina with short wavelength light (in the range of
(e.g. RNFL defects), measuring retinal vascular caliber (e.g. central 400–590 nm) is often referred to as “blue light” FAF and elicits auto­
retinal artery and vein equivalents) and evaluating optic disc appear­ fluorescence between 520 and 800 nm (Fig. 1D). The source of this blue
ance (e.g. pallor) in the posterior pole. Modern devices allow for im­ light FAF originates from the bisretinoid family of compounds often
aging of almost the entire retina and are often referred to as “widefield” referred to as lipofuscin and including A2-
(FOV approximately 90◦ ) or “ultra-widefield” (FOV greater than 90◦ ) glycerophosphoethanolamine, A2E and A2-dihydropyridine-
devices Fig. 2 to contrast with those devices that have more restricted phosphatidylethanolamine (A2-DHP-PE). These compounds represent
fields (ranging between 30 and 50◦ ; Fig. 1). Widefield imaging has the product of irreversible chemical reactions of retinaldehyde and the
particularly useful applications in identifying peripheral retinal pa­ lipid phosphatidylethanolamine that form in the photoreceptors and
thology that is difficult to describe and capture with conventional im­ accumulate in the RPE with age. Normal FAF exhibits a characteristic
aging (Kashani et al., 2014a,b). Therefore, one important parameter hypoautofluorescence in the central macula that results from the
when considering CFP is that the FOV may exclude relevant pathology attenuation of excitation light by macular pigments (xanthophylls).
especially if it is located in the peripheral retina. Outside of the central macula, the pattern of blue-light FAF is homog­
enous except for blood vessels and the disc which lack short-wavelength
3.1.1.2. Fundus autofluorescence imaging. One form of fundus imaging autofluorescence. In addition to these characteristic spatial variations,
takes advantage of the natural autofluorescent characteristics of the short-wavelength autofluorescence shows characteristic age-related
retina and retinal pigment epithelium (RPE). The retina has intrinsic changes in normal subjects (Delori et al., 2001). The relatively
autofluorescence when stimulated by light at several wavelength ranges featureless autofluorescence patterns of normal retina provides a useful
that is commonly referred to as fundus autofluorescence (FAF) (Sparrow background for detecting pathological lesions with potential auto­
et al., 2020). The FAF properties of the retina are studied at “short fluorescent properties such as Aβ (Koronyo et al., 2017).
wavelength” and “long wavelength” ranges (Kellner et al., 2010). FAF is Long-wavelength autofluorescence is commonly generated using
near-infrared light between 700 and 800 nm with excitation in the range

Fig. 1. Examples of retinal imaging modalities


from a 65 year old female illustrate commonly
used methods for evaluation of retinal disease and
retinal changes in neurodegenerative diseases. (A)
Color fundus photograph illustrating the macula,
optic disc and retinal arteries and veins. (B) Digital
collage of color fundus images of the same subject
demonstrating 60–90◦ field of view that includes
the peripheral retina outside the vascular arcades.
(C) Optical coherence tomography angiogram of
the parafoveal area illustrating the capillaries in
the area and the foveal avascular zone. Red and
green pseudocoloring represent the depth of retinal
capillaries in the superficial and deep retinal
layers, respectively. (D) Short wave fundus auto­
fluorescence image of the macula.

4
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Fig. 2. Ultra-Widefield fundus photograph and optical coherence tomography angiogram from a human subject. (A) Optos™ pseudocolor ultra-widefield fundus
photograph illustrates the peripheral retina where traditional color fundus imaging does not typically reach. (B) Widefield optical coherence tomography angiogram
(Zeiss PlexElite™) demonstrates non-invasive imaging of retinal arteries, veins and capillaries beyond the arcades in the same subject.

>800 nm. This form of FAF is primarily from the melanin content of the adenine dinucleotide (NADH), flavin adenine dinucleotide (FAD),
RPE cells and to a lesser degree from the underlying choroidal mela­ retinal, hemoglobin, melanin, collagen, lutein, zeazanthin, and lip­
nocytes but it can also be generated by elevated lipofuscin levels in some ofuscin. Due to the optical properties of the eye and safety consider­
retinal diseases. In contrast to the short-wavelength autofluorescence, ations, not all retinal fluorophores are detectable in vivo. FLIO
long-wavelength autofluorescence has its peak signal intensity in the measurements of normal retina have demonstrated discrete patterns for
central macula. various retinal regions such as the fovea, optic disc, retinal vessels and
non-foveal retina (Dysli et al., 2017a,b). Of particular relevance to the
3.1.1.3. Hyperspectral fundus imaging. Hyperspectral retinal imaging current review is the potential use of FLIO to detect molecules impli­
(HRI) is one of the most recently developed forms of fundus photog­ cated in neurodegenerative disease such as Aβ, tau, and huntingtin. We
raphy. Hyperspectral imaging is based on the principal of spectroscopy will provide a disease-specific discussion of this methodology in the
commonly used for geospatial applications and astronomy. HRI takes relevant sections below.
advantage of the wide-range of spectral features in the eye as well as
advances in spectroscopic imaging over the last several decades (Reshef 3.1.3. Optical coherence tomography
et al., 2020). HRI requires specialized illumination and detection hard­ OCT is a very well-established imaging modality that provides high-
ware that allows the high-resolution collection of a broad range of resolution images of the retinal structure based on interferometric
wavelength information across the visible and near-visible electromag­ methods. OCT was first described in 1991 (Huang et al., 1991) and was
netic spectrum for each pixel in the acquired image. This contrasts with adopted widely in ophthalmology in the subsequent decades because it
more conventional multispectral CFP (see section 5.1.1.1) which ac­ allowed quantitative measurements of retinal thickness, choroidal
quires information with lower spectral resolution. The most common thickness and even sub-layer thickness with micron level precision
application of HRI has been to quantify retinal vascular oxygen content (Asanad et al., 2019a,b) (Fig. 3). These kinds of measurements were
in the form of retinal oximetry (Kashani et al., 2011; Mordant et al., adopted to understand how retinal anatomy can vary in vivo with normal
2011; Jaime et al., 2012). This particular method takes advantage of the aging, gender and ethnicity (Kashani et al., 2010). OCT measurements
well-known spectra of oxy- and deoxyhemoglobin. Several research also provide the diagnostic basis on which anti-vascular endothelial
groups and commercial entities have adopted similar methodology to growth factor agents became widely used in ophthalmology for treat­
examine the spectral features of Aβ in eye tissue (More et al., 2019) as ment of macular edema in a wide range of diseases including age-related
well as the human eye in vivo (Hadoux et al., 2019; Sharafi et al., 2019). macular degeneration and diabetic retinopathy (Drexler and Fujimoto
At least one study has demonstrated a significant correlation between 2008). The high-resolution capability of OCT has progressively
brain PET amyloid burden in humans and retinal spectra associated with improved but even the original time-domain system design was rapidly
Aβ (Hadoux et al., 2019). Further discussion of this imaging modality in adopted by neurologists to study changes in the RNFL in MS (Petzold
disease-specific contexts is provided below and in a recently published et al., 2010) and AD (den Haan et al., 2017). Subsequent studies using
review in the context of AD (Gupta et al., 2020; Santangelo et al., 2020). more advanced spectral domain systems have replicated and expanded
upon the previous studies (den Haan et al., 2017).
3.1.2. Fluorescence lifetime imaging ophthalmoscopy While we will explore the disease specific findings of OCT in subse­
While hyperspectral imaging assesses the breadth of spectral infor­ quent sections of this review, it is important to make a few generalizable
mation available from retinal tissue, fluorescence lifetime imaging points about OCT studies of the retina in neurodegenerative disease.
(FLIO) aims to quantify one specific aspect of the fluorescence of First, most OCT systems image a relatively small portion of the retina
endogenous tissues, namely the duration of fluorescence (Dysli et al., ranging from 3 × 3 to 6 × 6mm regions of the central macula or optic
2017a,b). Of note, FLIO imaging systems are strictly for research and are disc region. Therefore, the majority of retinal tissue is not sampled with
not FDA approved for clinical use. Excitation of endogenous fluo­ standard OCT imaging protocols. More recently, widefield systems can
rophores by monochromatic light will cause excitation of natural fluo­ acquire up to 12 × 12mm FOV (or more) but are costly, not yet
rophores that decay with characteristic lifetimes. The average time commonly available and have lower resolution than smaller scan pat­
between the excitation of the tissue with monochromatic light and the terns. Second, there are numerous manufacturers of OCT systems. It is
return to the ground state is measured using FLIO. The lifetime of flu­ well known that scan patterns, segmentation algorithms and outcome
orophores is both a characteristic of the fluorophore as well as the sur­ measures vary significantly among manufacturers (Garcia-Martin et al.,
rounding molecular environment. Among the most common retinal 2012a,b; Garcia-Martin et al., 2014a,b; Satue et al., 2014; Chan et al.,
fluorophores are phenylalanine, tyrosine, tryptophan, nicotinamide 2019). Therefore, when comparing across studies, it is critical to keep in

5
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Fig. 3. Example of retinal layer and choroidal identification on enhanced depth imaging optical coherence tomography (EDI-OCT) compared to histology. (A) OCT
image depicts retinal layers relative to the corresponding retinal layers on (B) histology in a representative control micrograph stained with hematoxylin and eosin
within the macula. (Reproduced under the Creative Commons Attribution-NonCommercial-NoDerivatives (CC BY-NC-ND) 4.0 International License from Asanad S,
Ross-Cisneros FN, Nassisi M, Barron E, Karanjia R, Sadun AA. The retina in Alzheimer’s disease: histomorphometric analysis of an ophthalmologic biomarker. Invest
Ophthalmol Vis Sci. 2019; 60:1491–1500. https://doi.org/10.1167/iovs.18-25966).

mind the details of the OCT system and scan parameters used. Third, derived from the perfusion state of capillaries. In this sense, OCTA can be
even within the same commercial system, segmentation algorithms may considered an assessment of capillary structure and/or perfusion. This
vary over time and therefore the consistency of measurements made technology has also been extensively reviewed elsewhere (Kashani et al.,
with different generations of the same device may vary in subtle but 2017; Borrelli et al., 2018) but is worth discussing here because it is
significant ways (Chan et al., 2019). Fourth, almost all studies that are uniquely suited to demonstrate microvascular changes in the retina that
published to date use two-dimensional representations of retinal fea­ may correlate with vascular changes in the CNS. OCTA allows
tures such as thickness and vessel density. These two-dimensional rep­ depth-resolved imaging of the retinal vasculature that is far superior to
resentations of three-dimensional biological structures have inherent invasive angiography using fluorescein and approaching histologic
limitations which we discuss at the end of this article with respect to resolution (Matsunaga et al., 2014; Spaide et al., 2015a,b,c). This allows
future directions. separation of the retinal capillaries into at least two (or more) capillary
Lastly, it is important to emphasize that while the resolution of OCT plexi commonly referred to as the superficial and deep retinal layers
devices is on the micron scale, the reproducibility of measurements (Figs. 1C and 2B) (Matsunaga et al., 2014; Campbell et al., 2017). In the
depend on the experience and training of users. In experienced hands, peripapillary region, there is an additional plexus, the peripapillary
the reproducibility of OCT based measurements of retinal thickness can radial capillaries, that feeds the RNFL (Matsunaga et al., 2014). OCTA
be excellent (Intraclass correlation coefficient = 1.0 (Hu et al., 2015) analyses are frequently reported as a density metric such as “vessel
and the coefficient of variation very low (0.5% in (Hu et al., 2015) and density,” “skeleton density” or “perfusion density.” In all cases these
0.4% in (Obis et al., 2020)). To achieve this level of reproducibility and values are reporting the detection of red blood cell flow in patent retinal
precision, significant effort has to be expended for training of users and capillaries. By definition, any value reported using OCTA is a “perfusion
quality control of data. density” metric because capillaries without flow (or with very slow flow)
are not detected. A detailed discussion of OCTA methodology is avail­
able elsewhere (Kashani et al., 2017; Spaide et al., 2018). It is notable
3.2. Functional imaging that most studies performed with OCTA to date only include assessments
of the central macula or parafoveal capillaries (Fig. 1C). However,
There are several imaging modalities that measure functional aspects commercially available wide-field OCTA devices that can image the
of the neurosensory retina and supporting tissues including the RPE and retinal periphery are rapidly becoming available and will likely play an
retinal vasculature. These methods are distinctly different from the important role in the future (Fig. 2). Similar to OCT based measures of
structural imaging modalities discussed above because they measure an retinal thickness the reproducibility of OCTA measures can be excellent
aspect of cellular or tissue function, such as blood flow or electrical with experienced users, appropriate device settings and quality control
activity, rather than provide a static rendering of the retinal structure steps (Chen C et al., 2016, Conti et al., 2018).
such as retinal thickness. Below we will review the most relevant of Capillary level changes detected by OCTA correlate with clinical
these methodologies so that they can be discussed in disease specific disease severity in diabetic retinopathy (Kim et al., 2016a,b; Kashani
contexts in later sections. The sections below are not meant to be et al., 2017), retinal vein occlusion (Kashani et al., 2015; Koulisis et al.,
exhaustive reviews of the methodologies. Rather these sections are 2017) and even inflammatory diseases of the eye (Kim et al., 2016a,b)
meant to provide an overview of the method and relevance to neuro­ among many others (Kashani et al., 2017). Several recent studies have
degenerative disease in particular. demonstrated correlations between capillary features and neurodegen­
erative disease, most notably in AD (O’Bryhim et al., 2018; Querques
3.2.1. Optical coherence tomography angiography et al., 2019). Vascular contributions to cognitive impairment and de­
OCTA is a recent imaging modality based on OCT that was FDA mentia are common and their diagnosis and monitoring represents a
approved in 2015. OCTA provides information based on the movement significant unmet medical need. More than 50% of subjects with LOAD
of red blood cells within retinal capillaries and OCTA output is often pathology have significant comorbid vascular pathology (Schneider
displayed as a static map of retinal capillary density. It is important to et al., 2004; Schneider, Arvanitakis et al. 2007, 2009b; Schneider et al.,
keep in mind that while maps of retinal capillary density are often 2009a). Consensus statements from several leading groups have helped
interpreted as anatomic representation of capillaries, they are actually

6
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

outline the potential relevance of vascular contributions to cognitive eyelids while simultaneous light stimulation is provided to one or both
impairment and dementia (VCID) (Gorelick et al., 2011; Snyder et al., eyes. ERGs can be tuned to record light information preferentially from
2015). These consensus statements called for the development of novel, rods or cones as well as the downstream retinal circuitry that processes
clinically feasible biomarkers of vascular cognitive impairment and photoreceptor information. For example, the a-wave of the standard or
dementia for which OCTA is a promising modality. It is noteworthy that full-field ERG (ffERG) is primarily derived from rods and cones
the same limitations that apply to OCT imaging systems apply in large depending on the dark adaptation status of the retina. The b-wave
part to OCTA imaging (Spaide et al., 2015a,b,c). Further discussion of represents responses primarily from Muller and bipolar cells. The c-
this imaging modality in disease-specific contexts is provided in the wave represents electrical activity of the RPE. Because of the summation
following sections. of electrical activity at the corneal surface, ERGs are not sensitive to
Given the novelty of OCTA it is appropriate to briefly discuss its disease that affects a relatively small retinal area. There are very useful
limitations and the importance of quality control processes in image variations on the ERG methods, such as multifocal ERG (mfERG), which
acquisition and analysis. OCTA images are particularly prone to arti­ measures predominantly cone function in the central 30 degrees of the
facts, even more so than the underlying OCT images, because they are macula and pattern reversal ERG (PERG), which measures predomi­
derived from at least two high-quality and coregistered OCT scans. The nantly ganglion cell activity (Baker et al., 1988; Frishman 2013;
most common OCTA artifacts result from the motion of the eye relative Lachowicz and Lubiński 2018).
to the OCTA device during image acquisition (Spaide et al., 2015a,b,c).
Although efforts have been made to circumvent and correct eye motion 3.2.3. Visual evoked potentials
during image acquisition, this still remains one of the most common Visual evoked potentials (VEP) are not a form of direct retinal im­
artifacts of OCTA images, especially in diseased eyes. Motion correction aging but do reveal useful information about retinal function as well as
technologies attempt to correct these movements but also create unique the afferent arm of the visual pathway. A VEP is a measure of the light-
patterns such as stretch artifact, quilting artifact, or duplication arti­ induced electrical activity recorded from scalp electrodes placed over
facts. Another common type of artifact results from loss of signal the occipital cortex. It is a form of electroencephalography where the
detection by the device. For example, shadow artifacts results from the signal-to-noise ratio can be significantly enhanced by averaging time-
blockage of signal by media opacities like floaters. It has been shown dependent visual stimuli. The most widely used VEP data come from
that media opacities affect OCTA quantitative measures (Spaide et al., measurements of the amplitude and latency of the initial negative peak
2015a,b,c). Blink artifacts results from temporary blockage of signal by (N1), subsequent positive peak (P1 or P100), following negative peak
the eyelid. Advances in eye tracking technology now detect most eye (N2) and ultimate positive peak (P2). Abnormal VEP responses can be
blinks and enable rescanning of the appropriate area. Segmentation due to abnormalities along any portion of the afferent visual pathway
errors are another common type of artifact. Although software errors in from the cornea and lens to the occipital cortex. VEPs are particularly
detection of retinal layer boundaries are seen less frequently in normal useful in persons in whom an assessment of the visual potential is needed
eyes, segmentation errors are common in diseased eyes where layer but the subject is unable to cooperate (e.g. children or developmentally
boundaries are irregular or altogether absent (Spaide et al., 2015a,b,c). delayed adults). Just as with ERG testing, there are useful variations of
Image decentration is another problem that appears especially in sub­ VEP testing, such a pattern-reversal VEP, which can provide more spe­
jects with difficulty with foveal fixation. Although not technically an cific information about particular portions of the visual pathway. VEP
artifact, decentration can significantly impact reproducibility of quan­ abnormalities have been noted in several neurodegenerative diseases
titative measurements. Decentration refers to the misalignment between including MS and PD, suggesting they are not specific for a particular
the foveal avascular zone (FAZ) and the center of the OCTA en face disease process (Regan and Neima 1984). It is reasonable to assume that
image. Decentration severity can be graded based on the distance be­ any process which causes significant retinal dysfunction (e.g. glaucoma)
tween the center of FAZ and center of the en face image. will also cause abnormalities in the downstream tissues impacting VEP.
As mentioned before, some of the artifacts seen in OCTA images are Therefore, VEPs may have some useful role in assessing visual potential
intrinsic to the underlying OCT technology, the most common of them and location of anatomic lesions in neurodegenerative diseases, espe­
being the projection artifact. Projection artifacts occur when the light cially when subjects are unable to cooperate with other forms of testing.
passing through the superficial vessels is altered by reflection or ab­
sorption by the blood cells and surrounding tissue. This light that has 4. Disease specific applications of retinal imaging
passed through the superficial vessels is reflected back by tissues un­
derlying the superficial blood vessels, such as the Retinal Pigment 4.1. Multiple Sclerosis
Epithelium (RPE), and is detected by the device. The reflected light from
the RPE, creates the false impression of moving red blood cells in vessels Multiple Sclerosis (MS) is an autoimmune disease classically char­
within the RPE and, thus, overestimates perfusion density. These pro­ acterized by inflammation and demyelination of the CNS though
jection artifacts are observed in almost all of OCTA images but there are neuronal and axonal degeneration have more recently come to be
commercially available projection removal algorithms that minimize appreciated as critical processes in its pathogenesis (Campbell and
the impact of these artifacts on images (Spaide et al., 2015a,b,c). Mahad 2018). MS impacts about 300,000 people in the United States
and more than 2 million people around the world (Schiess and Calabresi
3.2.2. Electroretinography 2016). The most characteristic lesions of MS are focal demyelinated
Electroretinography (ERG) provides a measure of the electrical ac­ plaques in the CNS that are accompanied by inflammation and gliosis.
tivity in the neurosensory retina and RPE. ERGs provide information Unlike other neurodegenerative disorders, patients with MS often pre­
about neuronal and non-neuronal electrical activity in response to a sents with visual symptoms at the early stages of the disease secondary
light stimulus (Frishman 2013). The electrical activity of the retina, like to retrobulbar optic neuritis. Vision loss is associated with dull retro­
other neuronal tissue, is driven by ionic gradients, primarily sodium and bulbar pain and often aggravated by eye movement. Common visual
potassium, of the cellular components. ERGs provide very useful infor­ symptoms and signs include decreased visual acuity, loss of contrast
mation across a wide-range of retinal diseases and the pathophysiology sensitivity, a worsening red-green color defect, and an enlarging central
of ERG changes, at least for some disease processes, are well understood. scotoma, that usually peak at several days to a week (Group 2008). The
While this review will not attempt a comprehensive review of this topic, Optic Neuritis Treatment Trial evaluated 389 subjects with acute uni­
it is worth noting the key features of ERGs here so that readers can have lateral optic neuritis without a diagnosis of MS (Evangelou et al., 2001)
context when reading disease specific findings in later sections. and followed them for 15 years (Group 2008). This study showed that
ERGs are obtained by surface electrodes placed on the eye and the aggregate cumulative probability of developing MS at 15 years was

7
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

25% for subjects with ocular presentation without MRI lesions and 72% devices have allowed detailed analysis of individual retinal layers with
for patients with one or more intracranial MRI lesions. Therefore, the remarkable insights into the disease. A meta-analysis of 40 studies and
ocular findings in MS are of well-established clinical significance. 5776 eyes, revealed that eyes of subjects with MS but without ON have
Although MS is the more common cause of optic neuritis (ON), ~7 μm thinner pRNFL than controls and those with history of ON have
neuromyelitis optica (NMO) and anti-myelin oligodendrocyte glyco­ ~20 μm thinner pRNFL than controls (Petzold et al., 2017). Further­
protein (MOG) antibody-positivity can also cause ON. It is clinically more, in subjects with clinically isolated syndromes, measurements of
important to differentiate MS from NMO because the prognosis and GCIPL and RNFL can predict visual function and future disease activity
treatment for NMO is different. OCT imaging has played a role in months to years after onset of acute ON (Lambe et al., 2020). Therefore
differentiating these disease entities. Although up to 80% of NMO sub­ OCT can provide detailed quantitative measurements of the RNFL and
jects can be diagnosed by having a positive serum NMO antibody, NMO macular thickness changes associated with clinical features of MS at a
ON typically has more severe peripapillary RNFL (pRNFL) and GCL resolution that is not possible with clinical examination or neuroimaging
thinning than in MS (Filippatou et al., 2020a,b). Furthermore, micro­ methods. Based on these results, it is very likely that OCT could serve in
cystic macular edema is also more common in the inner nuclear layer several roles in the future diagnosis, prognostication and management of
(INL) of NMO subjects than in MS. Recent OCT based studies demon­ MS (Saidha and Naismith 2019).
strate subclinical RNFL and GCL thinning in eyes of subjects with
AQP4-IgG seropositive NMO but without history of ON (Filippatou et al., 4.1.2. Degenerative neurosensory retinal changes in Multiple Sclerosis
2020a,b). Not only do these OCT findings help differentiate NMO from The RNFL consists of the axonal projections of ganglion cells located
MS, they serve as useful outcome measures in clinical trials (Bennett in the macula and throughout the retina. Therefore it is not surprising
et al., 2015). Furthermore, it is clinically important to distinguish MS that histopathological evidence of GCL attenuation is also present in MS.
ON and NMO ON from anti-MOG-positive antibody ON by serum anti­ A meta-analysis of 40 studies and 5776 eyes demonstrated that the
body testing and not by OCT findings. This antibody is present in macular RNFL (mRNFL) was ~2 μm thinner in MS subjects without ON
one-third of all recurrent ON subjects and its presence predicts a better and ~6 μm thinner in those with ON compared to controls (Petzold
visual outcome than those with positive-NMO antibodies (Chen et al., et al., 2017). The GCL and inner plexiform layer (IPL) were ~6 μm
2018). thinner in MS subjects without ON and ~16 μm thinner in those with
ON. Although these differences are small, they are highly significant and
4.1.1. Optic nerve and peripapillary retina in Multiple Sclerosis some were even reproducible in a similar study performed much earlier
Histopathology of the LGN in subjects with MS has demonstrated with Time Domain OCT (Petzold et al., 2010; Petzold et al., 2017).
significant axonal loss (32–45%) in the optic nerve and optic tract as Longitudinal data from several studies demonstrated that a 1 μm loss of
well as parvocellular neuronal loss (Evangelou et al., 2001). In fact, ON pRNFL per year may be detectable in large, longitudinal clinical trials of
is the initial clinical manifestation of MS in ~25% of patients and optic 2–3 years duration aimed at assessing the efficacy of neuroprotective
neuropathy in the form of demyelinating plaques are found in a majority drugs in MS (Petzold et al., 2010; Petzold et al., 2017).
of all postmortem samples, irrespective of any history of ON (Ikuta and
Zimmerman 1976; Toussaint et al., 1983; Kerrison et al., 1994; Group 4.1.3. Inflammatory neurosensory retinal changes in Multiple Sclerosis
2008). Although not consistently performed, T2 weighted MRI images of Retrospective studies demonstrate a correlation between increased
the optic nerve can show hyperintense lesions with gadolinium INL volume and MRI activity in MS (Gelfand et al., 2012; Saidha et al.,
enhancement during acute attacks. In one study, MRI demonstrated 2012). The same meta-analysis of 40 studies and 5776 eyes demon­
hyperintense lesions in 84% of symptomatic and 20% of asymptomatic strated that the INL was about 1 μm thicker in subjects with MS but
MS subjects (Miller et al., 1988). Interestingly, cross-sectional and pro­ without ON than control subjects (Knier et al., 2016). A similar small
spective MRI studies demonstrate that ON can be associated with difference in the combined outer nuclear layer (ONL) and outer plexi­
thickening of the optic nerve in the acute phase followed by atrophy form layer (OPL) was found. Disease modifying therapy has also been
later (Hickman et al., 2004). This supports an initial inflammatory insult correlated with reduction in INL volume especially in cases where dis­
associated with tissue edema followed by atrophy. This pattern is similar ease activity has essentially become undetectable (Knier et al., 2016). It
to the pattern of ophthalmoscopic disc changes that are clinically is notable that an overall decrease in total macular volume is associated
demonstrable if the disease presentation is captured during the acute with increased disease activity. These apparently conflicting findings
phase. In many cases ophthalmoscopic examination of the RNFL can highlight the difficulty of identifying consistent, reproducible and rele­
demonstrate visible attenuation of the RNFL when a history of ON is vant retinal changes for particular aspects of the disease process. The
present (Elbøl and Work 1990). One limitation of clinical examination is opposite changes in RNFL and INL thickness suggest that it is important
that, in general, at least 50% of the RNFL must be lost for clinically to evaluate retinal sublayer thickness in addition to overall retinal
visible findings to be evident (Quigley et al., 1982). Therefore, there is thickness. For example, it has been hypothesized that inflammation of
no doubt that involvement of the optic pathway is of major clinical the INL is associated with increased INL volume while overall loss of
importance in MS but detection of the changes associated with demye­ RNFL explains the decrease in total macular volume. This hypothesis is
linating lesions by clinical examination, MRI and histopathology each supported by histopathology that confirms the presence of inflammatory
have practical limitations. cells in the inner retina of MS patients (Green et al., 2010).
The initial demonstration of peripapillary RNFL (pRNFL) attenuation OCT-demonstrated edema in the INL has also been correlated with dis­
in MS using OCT was of major importance because it provided a much ease severity (Gelfand et al., 2012). These findings are reminiscent of
easier, faster and less expensive method of detecting MS activity (Parisi diffuse tissue infiltration with T-lymphocytes in normal appearing re­
et al., 1999). Numerous studies have replicated and expanded upon the gions of white and gray matter on histopathology (Kutzelnigg et al.,
finding using OCT as well as other non-invasive imaging methods (Steel 2005). Automated segmentation of retinal layers using deep learning
and Waldock 1998; Zaveri et al., 2008). Some degree of pRNFL atten­ methods have demonstrated impressive predictive ability in MS and may
uation occurs in MS subjects regardless of an ON history or changes in become useful in the clinical setting with appropriate human guidance
visual acuity. The attenuation is greater in magnitude for subjects with (He et al., 2019). As we discuss later on (See section 7), novel methods of
ON or measurable visual function deficits. Large meta-analyses of image registration and analysis may help further refine layer specific
Time-Domain OCT (Petzold et al., 2010) and Spectral Domain OCT analyses that may have layer segmentation bias.
(Petzold et al., 2017) have been effectively used to demonstrate that MS
activity and duration correlate with thickness of the pRNFL. More 4.1.4. Retinal vascular changes in Multiple Sclerosis
recently the availability of higher resolution Spectral Domain OCT Retinal vascular changes in the form of periphlebitis (inflammation

8
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

of the veins) are a well-known manifestation of MS (Kerrison et al., threshold level that predicts visual impairment in MS (Costello et al.,
1994). This finding is consistent with subtle perivascular cuffing of 2006). A decrease of one line in low-contrast letter acuity correlates with
mononuclear cells in normal appearing regions of white and gray matter pRNFL thinning of 4 μm. Maximal pRNFL thinning (~10–40 μm)
on histopathology (Kutzelnigg et al., 2005). The relationship between generally occurs after acute ON within 3–6 months and the first
phlebitis and perfusion are unclear. MRI-based methods have been able detectable difference between the fellow eye occurs between 1 and 2
to demonstrate decreased cerebral blood flow within normal appearing months, reflecting axonal degeneration immediately after the primary
areas of white matter and gray matter in MS subjects as compared to demyelinating event. Stabilization of the pRNFL thickness occurs within
controls (Law et al., 2004; Steen, D’haeseleer et al., 2013), but it is 7–12 months from onset of disease (Henderson et al., 2010). Interest­
unclear if the periphlebitis and blood flow abnormalities are related. ingly, an inter-eye difference of 5 μm in pRNFL thickness is predictive of
Attenuated blood flow in MS also has been demonstrated in the retina a previous attack of ON in patients with a history of unilateral ON
independently of obvious periphlebitis. In a study of 16 MS subjects and (Nolan-Kenney et al., 2019). pRNFL measurements also correlate with
17 controls, blood flow velocity and volume in retinal arterioles and self-reported quality of life measures (Garcia-Martin et al., 2013) and
venules was significantly lower in MS subjects compared to controls disease progression (Garcia-Martin et al., 2010; Garcia-Martin et al.,
(Jiang et al., 2016). Interestingly, there was no correlation between 2011; Saidha et al., 2013; Martinez-Lapiscina et al., 2016). These ste­
blood flow and RNFL thickness suggesting that these are independent reotypical findings on OCT have led to the use of OCT based endpoints in
processes in MS. Similar significant reductions in blood flow have been more than two dozen clinical trials in the past few decades, demon­
demonstrated in MS subjects compared to controls in at least one other strating the value of OCT based biomarkers in assessing MS disease ac­
study (Liu et al., 2019a,b). tivity (Lambe et al., 2020).
In two separate cross-sectional studies of relapsing remitting MS
patients compared to healthy controls the density of the retinal capil­ 4.1.6. Retinal electrophysiology in Multiple Sclerosis
laries in the SRL of the macula and in the peripapillary region was Subjects with MS frequently have delayed VEP even without a clear
significantly reduced in those with a history of ON (Murphy et al., 2020; history of ON. This presumably reflects demyelinating lesions in the
Ulusoy et al., 2020). Capillary density in the DRL was also slightly visual pathway but not necessarily in the optic nerve (Halliday, McDo­
decreased in the MS group compared to controls. MS-associated ON was nald et al. 1972, 1973; Matthews et al., 1977). One prospective study of
also associated with atrophy of the inner retinal layers, mainly the RNFL 29 newly diagnosed MS subjects and 32 controls did not show any as­
and ganglion cell-inner plexiform layer (GCIPL). In a study of 68 eyes of sociation of VEP parameters to intraorbital optic nerve enhancement or
45 MS subjects compared to 55 eyes of 32 healthy controls, optic nerve atrophy on MRI (Hickman et al., 2004). However, VEP amplitude and
head blood flow was lowest in MS-associated ON eyes. Although the latency have been shown to correlate with optic nerve volume in MS
RNFL and retinal ganglion cell complex (GCC) thicknesses were reduced patients (Hickman et al., 2002). A significant association exists between
in the eyes with ON, the OCTA measurements did not correlate with the PERG latency (and amplitude) and pRNFL thickness in MS subjects with
structural OCT (Spain et al., 2018). Therefore, it appears that optic nerve a history of ON (Parisi et al., 1999). Despite abnormal VEP amplitude
head blood flow measures an MS effect independent of structural dam­ and latency in subjects with MS with ON compared to controls, the same
age. It is worth noting that OCTA based assessment of retinal capillary study demonstrated no association between VEP latency or amplitude
changes are highly dependent on layer segmentation methods and al­ and pRNFL thickness (Parisi et al., 1999).
gorithms which may vary among devices. As we discuss in section 7,
novel methods of image analysis that are less dependent on layer seg­ 4.2. Late onset Alzheimer’s disease
mentation may help further address the relationship of retinal vascular
changes to neurosensory changes. In addition, more longitudinal studies Alzheimer’s Disease (AD) is a chronic neurodegenerative disorder
will be needed to explain the relationship between perfusion and atro­ and the most common form of dementia affecting over 26 million people
phy over time. In the choriocapillaris layer, higher vessel densities were worldwide (2020). As advancing age is the single strongest risk factor for
associated with disease activity (Feucht et al., 2019), while choroidal AD, the term Late Onset AD (LOAD) is used to distinguish its most
thinning was correlated with disease duration (Esen et al., 2016). These common form from its more rare autosomal dominant genetic form of
findings could represent the vasodilatory effects during active inflam­ younger onset, with a somewhat arbitrary age cut-off of symptom onset
mation and the resultant atrophy, respectively. of 65 years. LOAD is defined by neuropathologic changes including
neuronal loss, gliosis, extracellular accumulation of fibrillar
4.1.5. Clinical correlation of retinal changes in Multiple Sclerosis beta-amyloid (Aβ) and intraneuronal cytoskeletal abnormalities con­
Retinal thickness changes have been correlated with clinical severity sisting in part of hyper-phosphorylated tau (ptau) in the CNS (McKhann
of MS (Gordon-Lipkin et al., 2007; Sepulcre et al., 2007; Toledo et al., et al., 1984; Jack et al., 2011). There are numerous animal models of AD
2008; Jiang et al., 2016). Thinning of the GCIPL seems to be the earliest that mimic various aspects of the disease and of significant interest in
detectable OCT finding in MS and is a more sensitive biomarker for MS understanding the pathophysiology of the disease (Lim et al., 2020; Song
than pRNFL (Garcia-Martin et al., 2012a,b; Walter et al., 2012; et al., 2020). Though deposition of diffuse and fibrillar Aβ in the CNS
González-López et al., 2014). GCIPL thinning occurs within weeks of occurs early during the presymptomatic stage of the disease (as early as
onset of acute ON and can precede the RNFL thinning. GCIPL thinning 20 years prior to the development of overt dementia), ptau accumula­
within the first month of onset of ON is predictive of visual impairment tion in the form of neurofibrillary tangles is most strongly correlated
by 6 months (Gabilondo et al., 2015). An inter-eye difference of 4 μm in with neuronal death and clinical symptoms (Nelson et al., 2012). The
GCL thickness is predictive of a previous attack of ON in patients with a diagnosis of AD can be made based on clinical parameters alone
history of unilateral ON (Nolan-Kenney et al., 2019). The largest and (“clinically probable AD”) for which specificity and sensitivity can vary
most robust differences between the eyes of MS and control eyes were between 80 and 90%, or based on post-mortem diagnosis or on
found in the pRNFL and GCIPL. While these changes are all relatively AD-specific CSF and PET biomarkers for which the accuracy of detecting
small, it is very likely that in the correct clinical setting OCT can provide AD pathology during life is generally higher (Jack et al., 2011). For in
confirmatory or additional evidence of underlying disease activity. vivo human studies, papers prior to 2000 tended to use clinical diagnoses
Physiological pRNFL thinning due to age occurs at 0.017% per year while AD-specific PET imaging and cerebrospinal fluid biomarkers
starting at 18 years of age, with a 10–20 μm loss over 60 years (Kana­ became available in the early 2000’s (Klunk et al., 2004; Jack et al.,
mori et al., 2003). Above and beyond this, pRNFL thinning correlates 2018). It is important therefore to note what diagnostic criteria are used
with impaired visual function and pRNFL thickness may predict visual when interpreting ophthalmologic studies in AD. A detailed review of
recovery after ON. pRNFL thinning at 75–80 μm has been found to be the animal models as well as molecular and cellular findings associated with

9
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

retinal changes in AD has been published recently (Gupta et al., 2020; macular volume (− 0.23 μm, p = 0.0003), total macular thickness (range
Qin et al., 2020). -9μm to − 14μm; p < 0.0001 for all retinal quadrants) and pRNFL
It is worth noting that ophthalmologic impairments in human (− 5.99 μm, p < 0.0001) in clinically-defined LOAD subjects compared to
contrast sensitivity, color discrimination, motion perception, reading controls (Chan et al., 2019). Yet another meta-analysis of 887 LOAD
speed and visual fields have been reported in LOAD (Sadun et al., 1987; subjects and 864 controls showed standardized mean reductions in
Katz and Rimmer 1989; Gilmore et al., 2006; Boucart et al., 2015). Some pRNFL of LOAD subjects (0.98 μm, p < 0.0001) compared to controls.
visual dysfunction in LOAD has been associated with degeneration of This study also shows a standardized mean reduction in total macular
anterior visual pathways, namely the optic nerve and retina (Hinton thickness of 0.88 μm (p = 0.0001) (den Haan et al., 2017).
et al., 1986). However, it is unclear how early and specifically these The macula contains more than 50% of the RGCs in the entire retina
ophthalmologic manifestations can be reliably demonstrated in subjects and RGC bodies are 10–20 times larger than the diameter of their axons.
with LOAD (Uhlmann et al., 1991). It would not be surprising if there Therefore, macular measurements of GCIPL may be more sensitive than
was objective retinal pathology underlying these changes that was pRNFL measures (Cheung et al., 2015). Specifically, attenuation of the
detectable by retinal imaging methods. Below we summarize and syn­ mRNFL, GCL, and IPL collectively known as the retinal ganglion cell
thesize a large body of literature describing retinal imaging findings in complex (GCC) has been reported by several research groups (Kesler
LOAD and implications of these findings in future studies. The numerous et al., 2011; Marziani et al., 2013; Thomson et al., 2015; Cunha et al.,
references in this section are also summarized in Table 1. 2016; Garcia-Martin et al., 2016; Shao et al., 2018). A large metanalysis
comparing OCT measurements of macular thickness in 467
4.2.1. Optic nerve changes in AD clinically-defined LOAD subjects and 518 controls has further confirmed
In 1986, Hinton et al. provided the first histopathological evidence of significant reductions in the GCIPL, GCC, macular full-thickness, and
optic neuropathy in LOAD after observing diffuse RGC loss and axonal macular volume (Chan et al., 2019).
atrophy in postmortem optic nerve tissues derived from severe LOAD The findings above, while very encouraging must be considered in
patients (Hinton et al., 1986). These findings were independently the context of at least a few well-conducted studies that have found no
corroborated on CFP (increased cup-to-disc ratios) and using red free significant thinning among subjects that meet clear and objective
optic disc photographs that demonstrated RNFL defects in LOAD sub­ criteria for probable LOAD (Haan et al., 2019a,b; Sánchez et al., 2020).
jects (Tsai et al., 1991). Subsequent histopathology studies further For example, in subjects with posterior cortical atrophy (PCA), a variant
demonstrated degeneration of the RNFL and the RGC, most severe su­ of LOAD disproportionately affecting parietal and occipital cortex, there
periorly and inferiorly with respect to the optic nerve (Blanks et al., was no reported association between any measure of retinal thickness
1989; Bassi and Sadun 1990; Blanks, Torigoe et al. 1991, 1996a,b; and LOAD status (Haan et al., 2019a,b). On the other hand, these in­
Curcio and Drucker 1993; Blanks et al., 1996a,b; La Morgia, vestigators found a correlation between atrophy in parietal cortex and
Ross-Cisneros et al., 2016). RNFL thickness, regardless of the presence of fibrillar Aβ on PET scan
OCT has been widely used to quantify retinal and retinal sublayer (tau-PET was not performed), supporting a transsynaptic mechanism for
thinning in living human subjects with LOAD. In a recent meta-analysis RNFL thinning independent of amyloid pathology (den Haan et al.,
comparing 1061 clinically-defined LOAD and 1130 control subjects, 2018a,b). These findings most directly support the idea that trans­
there was significant pRNFL thinning in LOAD for all retinal quadrants, synaptic retrograde degeneration of neurons is the cause of retinal
most pronounced superiorly and inferiorly (Chan et al., 2019). These thinning although further research is needed. They also emphasize the
OCT findings of superior and inferior retinal atrophy are consistent with need for pathology or biomarker-based diagnoses to minimize con­
retinal histopathological findings in postmortem LOAD eyes described founding variables such as underlying vascular disease. As we discuss in
above. The pattern of RNFL loss in LOAD seems to be specific in com­ sections 6.6 and 7, assessment of cerebral vascular pathology in the
parison to some other neurodegenerative diseases (Sadun et al., 1987; La context of neurodegenerative disease is a major area of unmet medical
Morgia, Di Vito et al., 2017). Specifically, the superior and inferior RNFL and research need. Advances in both understanding of the pathophysi­
thinning in LOAD is consistent with loss of M-type RGCs that comprise ology of cerebrovascular disease as well as qualitative and quantitative
the magnocellular pathway (M-cells), are mainly located in the extra­ assessment of capillary level changes may add significant insight into
macular retina, and are more involved in low-resolution peripheral this confounding variable.
vision (Sadun and Bassi 1990). This is in contrast to the P-type RGCs that
populate the papillomacular bundle in the temporal region of the optic 4.2.3. Degenerative neurosensory retinal changes in MCI and preclinical
disc and mediate central vision. This difference may explain the pre­ AD
served visual acuity that is observed in subjects with advanced LOAD One particularly enticing application of retinal imaging in LOAD is to
(Sadun et al., 1987) and may be associated with differences in vulner­ detect incremental progression of disease or subclinical disease changes
ability to LOAD pathology among RGC subtypes (La Morgia, Di Vito that are otherwise undetectable with conventional neuroimaging. For
et al., 2017). example, OCT measured changes in retinal structure and function in
subjects with cognitive impairment but without definitive dementia
4.2.2. Degenerative neurosensory retinal thickness changes in AD could serve as a useful biomarker in disease prognosis and preventative
Several lines of evidence demonstrate changes throughout the clinical trials. It is thought that the neurodegenerative process associ­
neurosensory retina in subjects with LOAD. Postmortem histology from ated with LOAD precedes the onset of symptoms by 20 years or more
subjects with neuropathologically confirmed severe LOAD (Braak Stages (Villemagne et al., 2013). In this context, the use of retinal imaging
V-VI) demonstrate a gradient of retinal thickness reduction whereby markers may define early changes in neurodegenerative entities
thinning was greatest for the RNFL and RGC followed by the INL and including LOAD when interventions may be effective in preserving
ONL in a superotemporal and superonasal pattern with respect to the neuronal tissue. In this section we will review the relevant literature that
optic nerve (Asanad et al., 2019a,b) (Fig. 4). These changes have been suggests retinal based biomarkers may be useful in identifying early
corroborated by in vivo OCT studies in LOAD. A meta-analysis of 380 changes associated with mild cognitive impairment (MCI) and/or pre­
clinically-defined LOAD patients and 293 controls from 11 clinical LOAD. Biomarker-based identification of AD pathology is
cross-sectional studies showed a weighted mean difference of − 15.95 particularly important in this group in whom the diagnosis is not yet
μm (p < 0.0001) in RNFL of LOAD subjects compared to controls clear.
(Coppola et al., 2015). Another meta-analysis of 1257 LOAD subjects A consistent and significant decrease in pRNFL and macular thick­
and 1460 controls from 30 cross-sectional studies showing weighted ness parameters has been demonstrated in subjects with MCI by several
reductions in the thickness of macular GCIPL (− 3.66 μm, p = 0.01), cross-sectional studies and reviewed in at least two meta-analyses

10
Table 1

A.H. Kashani et al.


Summary of retinal findings in human subjects with late onset Alzheimer’s disease.
Finding Preclinical No Difference MCI/Early AD vs HC No Difference Moderate-severe AD vs HC No Difference Biomarker confirmed (Aβ1; Complete eye exam* excluding
AD vs HC Tau2) confounding disease

OCT/Histology
pRNFL Asanad van de Kreeke Coppola et al., den Haan Tsai et al., (1991); Coppola et al., Haan et al., 2019a,b; van de Kreeke et al. (2019)1; Asanad et al., 2020; van de Kreeke
thinning et al., 2020 et al. (2019) (2015); Chan et al., et al., 2018a,b; (2015); La Morgia et al., (2016); Sadun Sánchez et al., Asanad et al., 20201,2; La et al., (2019); den Haan et al. (2018a,
(2019); den Haan Lad et al., et al., (1987), Chan et al., 2019; (2020); Lad et al., Morgia et al. (2016)1; b); Kesler et al., (2011); Ascaso et al.,
et al. (2018a,b); (2018) Asanad et al., 2019a,b; Cipollini et al., (2018); Ho et al., Koronyo et al. (2017)1; den (2014); den Haan et al., 2018a,b; Lad
Kesler et al., (2011); 2020; Koronyo et al., (2017); Asanad (2014), Blanks et al., Haan et al. (2018a,b)1,2; et al., (2018); Salobrar-García et al.,
Paquet et al., (2007); et al., 2019b (1989); Williams Asanad et al., 2019b1; Haan 2019; Sánchez et al., (2020)
Ascaso et al., (2014) et al., (2017) et al., 2019a,b1,2; Alves et al.,
20191,2
mRNFL Santos et al. van de Kreeke Coppola et al., Sánchez et al., Marziani et al., (2013); Cunha, lopes. Haan et al., 2019a,b; Santos et al. (2018)1; Sadda van de Kreeke et al., (2019); Sánchez
thinning/ (2018) et al. (2019) (2015); Chan et al., (2020); Alves 2016; Garcia-Martin et al., (2016); Sánchez et al., (2020) et al. (2019)1,2; Haan et al., et al., (2020); Marziani et al., (2013);
thickening (2019); den Haan et al., 2019 Kesler et al., (2011); Shao et al., 2019a,b1;van de Kreeke et al. Cunha, lopes. 2016; Haan et al., 2019a,
et al. (2018a,b); (2018); Thomson et al., (2015); (2019)1; den Haan et al. b; Sadda et al., (2019); den Haan et al.
Kesler et al., (2011); Salobrar-García et al., (2019) (2018a,b)1,2 2018a,b; Kesler et al., (2011); Ascaso
Paquet et al., (2007); et al., (2014); Garcia-Martin et al.,
Ascaso et al., (2014) (2016); Salobrar-García et al., (2019)
RGC-IPL Sadda et al. van de Kreeke Lad et al., Hinton et al., (1986); La Morgia et al., Haan et al., 2019a,b; van de Kreeke et al. (2019)1; Sadda et al., (2019); van de Kreeke
thinning (2019) et al., (2019); (2018); (2016); Blanks et al., (1989); Blanks Sánchez et al., Asanad et al., 20201,2; La et al., (2019); Asanad et al., 2020; Lad
Asanad et al., Sánchez et al., et al., 1996a,b; Bassi and Sadun (2020); Lad et al., Morgia et al. (2016)1; Haan, et al., (2018); Sánchez et al., (2020);
2020 (2020); Alves (1990); Blanks et al., (1991); Curcio (2018) et al. 2019a,b1; den Haan Marziani et al., (2013); Cunha, lopes
et al., 2019 and Drucker (1993); den Haan et al., et al. 2018a,b1,2; Asanad et al. et al., 2016; Garcia-Martin et al.,
(2018a,b); den Haan et al. 2018a,b; 20191; Curcio and Drucker (2016); Kesler et al., (2011);
Asanad et al., 2019a,b; Chan et al., (1993)1,2 Salobrar-García et al., (2019)
2019; Marziani et al., (2013); Cunha,
lopes et al., 2016; Garcia-Martin et al.,
(2016); Kesler et al., (2011); Shao
11

et al., (2018); Thomson et al., (2015);


Salobrar-García et al., (2019)
Total macular van de Kreeke Chan et al., (2019); den Haan et al. den Haan et al. 2018a,b/ Sánchez et al. (2020) den Haan et al. 2018a,b1,2; van de Kreeke et al., (2019); Asanad
thinning/ et al., (2019); Shao et al., (2018) 2018a,b; Salobrar-García et al., (2019) den Haan et al., (2018a,b)1,2; et al., 2020; den Haan et al., 2018a,b;
thickening Asanad et al., Sánchez et al., van de Kreeke et al. (2019)1; Sánchez et al., (2020); den Haan et al.,
2020 (2020) Asanad et al., 20201,2 (2018a,b); Salobrar-García et al.,
(2019)
Choroidal Bulut et al., (2018); Chan et al., 2019/ Haan et al., (2019a,b) Asanad et al. 20191; Haan Bulut et al., (2018); Asanad et al.,
thinning/ Asanad et al., 2019 et al., 2019a,b1,2 2019a,b; Haan et al., 2019a,b
thickening
Vascular
Capillary Sadda et al., Yoon et al., (2019); Bulut et al., (2018); Jiang et al., 2018a, Querques et al. Sadda et al. (2019)1,2; Haan Sadda et al., (2019); Haan et al., 2019a,
density (2019); Haan Jiang et al., 2018a,b b (2019) et al., 2019a,b1,2; Querques b; Querques et al., (2019); Jiang et al.,

Progress in Retinal and Eye Research 83 (2021) 100938


decrease et al., 2019a,b et al. (2019)1,2 2018a,b; Yoon et al., (2019)
Enlarged FAZ O’Bryhim Bulut et al. (2018) O’Bryhim et al. (2018)1,2 O’Bryhim et al., (2018); Bulut et al.,
et al. (2018) (2018)
Decreased Feke et al., (2015); Berisha et al., Feke et al., (2015); Berisha et al.,
blood flow (2007); Szegedi et al., (2020); Jiang (2007); Szegedi et al., (2020); Jiang
et al., 2018a,b et al., 2018a,b
Decreased Querques et al. (2019) Szegedi et al. (2020) Querques et al. (2019)1,2 Querques et al., (2019); Szegedi et al.,
vessel (2020)
reactivity
Decreased Berisha et al., (2007); Cheung et al., Berisha et al., (2007); Cheung et al.,
vessel (2014); Szegedi et al., (2020) (2014); Szegedi et al., (2020)
caliber
ERG
(continued on next page)
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

(Paquet et al., 2007; Kesler et al., 2011; Ascaso et al., 2014; Coppola

Abbreviations: OCT = Optical Coherence Tomography, ERG = Electroretinography, AD = Alzheimer’s Disease, HC = Healthy Controls, pRNFL = peripapillary retinal nerve fiber layer, mRNFL = macular retinal nerve fiber
layer, MCI = mild cognitive impairment, RGC = Retinal Ganglion Cells, IPL = Inner Plexiform Layer, FAZ = Foveal Avascular Zone, Blank cells indicate no data or study was available for that parameters. (*) Complete eye
et al., 2015; den Haan et al., 2017; Chan et al., 2019). A meta-analysis
Krasodomska et al., (2010); Sartucci
of 68 MCI patients and 293 controls from 11 cross-sectional studies
showed a weighted mean difference of − 13.39 μm (p = 0.013) in RNFL
Complete eye exam* excluding

Katz B, S Rimmer et al., 1989


of clinically-defined MCI subjects compared to controls(Coppola et al.,
2015). Another meta-analysis of 305 clinically-defined MCI subjects
confounding disease

and 1460 controls from 30 cross-sectional studies showed weighted


reductions in the thickness of macular GCIPL (− 10.19 μm, p = 0.05) in
et al., (2010)

MCI subjects compared to controls (Chan et al., 2019). Notably there


was no difference in macular volume or pRNFL except for a significant
weighted increase in nasal peripapillary thickness. A third
meta-analysis of 216 clinically-defined MCI subjects and 864 controls
showed standardized mean reductions in pRNFL of 0.71 μm in MCI
subjects (p = 0.008) compared to controls. This study also showed a
Biomarker confirmed (Aβ1;

standardized mean reduction in macular thickness of 0.88 μm (p =


0.0001) (den Haan et al., 2017).
These studies support the notion that subclinical retinal changes
accompany MCI and may be useful in detecting disease progression in
longitudinal studies. For example, a prospective study was reported in
27 cognitively healthy participants with pathologic cerebrospinal fluid
Tau2)

Aβ42/tau ratios consistent with the presence of AD pathology, and 16


cognitively healthy controls with normal Aβ42/tau ratios. Mean RNFL
was not only significantly thinner in asymptomatic LOAD participants
Katz B, S Rimmer

relative to controls, but also demonstrated high sensitivity (87%) but


No Difference

modest specificity (56%) in classifying cognitively healthy individuals


et al., 1989

with elevated CSF Aβ42/Tau ratios (Asanad et al., 2020) (Fig. 5). These
studies have set the stage for larger, multicenter, prospective studies
that could formalize these associations.

4.2.4. Inflammatory neurosensory retinal changes in MCI and preclinical


et al., (2010); Sartucci et al., (2010)
Parisi et al., (2001); Krasodomska

AD
In contrast to the above findings, a significant number of well-
Moderate-severe AD vs HC

performed studies report essentially no change or even an increase in


retinal thickness in association with early stages of dementia or MCI
(Lad et al., 2018; Shao et al., 2018; Alves et al., 2019; Hadoux et al.,
Parisi et al. (2001)

2019; Salobrar-García et al., 2019; van de Kreeke et al., 2019; Marquié


et al., 2020). For example, one study evaluated retinal thickness in
exam includes visual acuity, intraocular pressure measurement, and dilated fundus examination.

relation to amyloid accumulation in 165 cognitively healthy mono­


zygotic twins among which 18 were Aβ (+). OCT demonstrated no
thinning of the pRNFL, mRNFL, GCL, IPL, or total macula. In addition,
no significant associations were found between retinal thickness and
No Difference

PET Aβ levels following correction for multiple testing (van de Kreeke


et al., 2019). The reason for these negative findings is not clear but may
be related to the use of Aβ, which may represent earlier stages of
preclinical disease, rather than a neurodegenerative marker.
Shao et al. evaluated macular thickness in 25 clinically-defined
MCI/Early AD vs HC

LOAD, 24 MCI, and 21 control subjects. Relative to controls, the MCI


and LOAD cohort demonstrated macular thinning but also significant
ONL and photoreceptor thickening (Shao et al., 2018). Lad et al.
analyzed 15 clinically-defined mild-moderate LOAD, 15 MCI, and 18
control subjects. Average pRNFL and GCIPL thicknesses did not
significantly differ between groups but areas of thickening were found
adjacent to areas of thinning in the pRNFL and GCIPL (Lad et al., 2018).
No Difference

More recently, Marquie et al. conducted a 2-year longitudinal study to


investigate the relationships between retinal thickness and PET Aβ
levels. OCT showed nasal macular thickening which positively corre­
lated with amyloid uptake (Marquié et al., 2020). Alves et al. evaluated
morphological changes in retina and in brain white matter integrity
et al. (2012)
Preclinical
AD vs HC

using OCT and diffusion tensor imaging, respectively, among 17 early


Moschos

amyloid-PET defined LOAD and 23 control subjects and noted that INL
Table 1 (continued )

thickness was positively associated with fractional anisotropy in early


LOAD (Alves et al., 2019). Salobrar-Garcia et al. analyzed 39 mild and
Dysfunction
dysfunction
Outer retinal

21 moderate clinically-defined LOAD patients, and 40 control subjects.


Inner retinal

Relative to controls, mild LOAD patients showed thinning in the central


Finding

macula, whereas moderate LOAD patients showed thickening in this


region. Both mild and moderate LOAD cohorts showed pronounced

12
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Fig. 4. Retinal histopathology of AD and control subjects. Light microscopy depicts (A) supero-temporal RNFL (black arrows) in control and (B) AD postmortem
tissue. Qualitative assessment of the supero-temporal RGCL, INL, and ONL (marked by red boxes) in (C) representative control and (D) subject with AD depicts
supero-temporal RGCL, INL, and ONL thinning most pronounced in the macular region of the subject with AD. All stains are hematoxylin and eosin. (Reproduced
under the Creative Commons Attribution-NonCommercial-NoDerivatives (CC BY-NC-ND) 4.0 International License from Asanad S, Ross-Cisneros FN, Nassisi M,
Barron E, Karanjia R, Sadun AA. The retina in Alzheimer’s disease: histomorphometric analysis of an ophthalmologic biomarker. Invest Ophthalmol Vis Sci. 2019;
60:1491–1500. https://doi.org/10.1167/iovs.18-25966).

thinning of the mRNFL, GCL, and OPL. However, the ONL was signifi­ overall retinal thickness until the degenerative process overtakes the
cantly thicker in mild and moderate LOAD relative to controls (Salo­ inflammatory stage of the disease. This possibility would explain the
brar-García et al., 2019). A common interpretation of these results is that absence of any significant change in retinal thickness in a number of
there is inflammation early in the neurodegenerative process that could studies discussed above. Careful retinal sublayer analysis in larger,
manifest as thickening of specific retinal sublayers. A similar observa­ prospective studies should help answer this question. In addition, more
tion has been proposed and demonstrated in MS (Green et al., 2010; sophisticated methods of OCT analysis will enable further exploration of
Gelfand et al., 2012). layer specific changes and volume changes independently of layer seg­
One particularly confounding possibility is that degenerative mentation (see section 7).
changes in one retinal layer are offset by thickening from inflammatory
changes in other layers. This would result in no detectable change in

Fig. 5. Peripapillary retinal nerve fiber


layer thickness is reduced in subjects with
preclinical AD. (A) The thicknesses of the
retinal nerve fiber layer (temporal, superior,
nasal, inferior) was measured using OCT in
the regions outlined in black. (B) Depicts
least-squares mean (95% CI) total retinal
nerve fiber layer thickness adjusted for side
and region between cognitively healthy
controls (blue) and cognitively healthy par­
ticipants with pathologic CSF Aβ42/Tau
levels (red). (Adapted from Asanad S, Fan­
tini M, Sultan W, Nassisi M, Felix CM, Wu J
et al. (2020) Retinal nerve fiber layer
thickness predicts CSF amyloid/tau before
cognitive decline. PLoS ONE 15(5):
e0232785. https://doi.org/10.1371/journal.
pone.0232785 under the terms of the
Creative Commons Attribution License).

13
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

4.2.5. Neurosensory retinal amyloid beta deposition in LOAD Aβ accumulation, one plausible reason is the relatively limited
A pathologic hallmark of LOAD is deposition of Aβ and ptau in the cross-sectional sampling of the retina in comparison to the retinal whole
CNS (McKhann et al., 1984; Montine et al., 2016). Therefore, identifi­ mounts used in studies with positive Aβ findings in the retina (Alber
cation of these proteins in the retina has been an area of intense interest. et al., 2020).
Retinal Aβ protein deposition in postmortem LOAD retinal tissue and One potentially exciting avenue for detection of retinal Aβ is the use
transgenic mice (APPSWE/PS1E9) has been demonstrated following cur­ of autofluorescence, hyperspectral imaging and fluorescence lifetime
cumin staining (Koronyo-Hamaoui et al., 2011; Chibhabha et al., 2020; imaging ophthalmoscopy (FLIO). Pilot studies demonstrate significant
Sidiqi et al., 2020). These findings have been corroborated by recent differences in the fluorescence lifetimes of retinal fluorophores between
human studies, which have similarly demonstrated Aβ plaque accumu­ 7 CSF-defined preclinical LOAD participants and 8 controls using FLIO.
lation highest in the mid-to far-peripheral regions of superior and infe­ Fluorescence lifetimes were not only significantly prolonged in pre­
rior hemiretinas (Alexandrov et al., 2011; La Morgia, Ross-Cisneros clinical LOAD retinas, but also correlated with CSF Aβ and tau levels and
et al., 2016; Koronyo et al., 2017). Immunohistochemistry studies GCIPL thickness measured by OCT (Sadda et al., 2019). Hadoux and
illustrated Aβ distribution localized to the RGC in postmortem LOAD colleagues demonstrated remarkable differences in retinal reflectance
eyes with selective loss of intrinsically photosensitive subtypes of RGCs patterns using hyperspectral imaging. In addition, retinal hyperspectral
known as melanopsin-containing RGCs (mRGCs) (La Morgia, scores were notably associated with brain Aβ burden as measured by
Ross-Cisneros et al., 2016). Since mRGCs drive circadian photoentrain­ PET (Hadoux et al., 2019). The pathogenesis of these spectral differences
ment, the authors hypothesized that mRGC loss may also contribute to are not clear and much further investigation is needed. However, the
circadian dysfunction characteristically seen in LOAD (La Morgia, general concept of leveraging the spectral properties of Aβ for detection
Ross-Cisneros et al., 2016). Koronyo et al. analyzed Aβ plaque deposition in the retina is promising. A clever variation on this theme is the use of
in whole mount retinas from 8 LOAD and 7 control postmortem eyes. exogenous agents to label amyloid in vivo. For example, Koronyo et al.
LOAD retinas showed a 4.7-fold increase in plaque aggregation which demonstrate putative amyloid containing lesions in the retina of 10
correlated with Aβ burden in both the primary visual cortex and ento­ human subjects with AD compared to 6 healthy controls using a cur­
rhinal cortex (Koronyo et al., 2017). cumin labelling scheme and fundus autofluorescence imaging (Koronyo
Despite these findings, there is some controversy regarding the et al., 2017). It is not yet clear how quantitative this method can be but
detection of these proteinopathies in human retina in LOAD (Blanks prospective clinical studies are in progress.
et al., 1989; Ho et al., 2014; Williams et al., 2017; den Haan et al.,
2018a,b). Ho et al. and Williams et al. were unable to identify Aβ and 4.2.6. Subretinal amyloid beta deposition in age-related macular
ptau deposits in retinal cross-sections of autopsy-confirmed LOAD cases degeneration and LOAD
relative to age-matched controls (Ho et al., 2014; Williams et al., 2017). Aβ is also known to be present in pathologic and non-pathologic age-
Haan et al. reported ptau accumulation in the IPL and OPL in 6 post­ related changes called drusen (Anderson et al., 2004; Isas et al., 2010).
mortem LOAD eyes as well as diffuse Aβ deposition throughout the inner Drusen typically occur in the subretinal space with the greatest preva­
retina in both LOAD cases and controls (den Haan et al., 2018a,b). While lence in the posterior pole and show some physiologic age-related
it is not exactly clear why these studies did not detect pathologic retinal accumulation (Hoh Kam et al. 2010). The most common pathologic

Fig. 6. Color fundus photographs and positron


emission computed tomograms (PET) demon­
strating clinical correlation of retinal and brain
imaging findings in (A,B) 74 year old subject
without dementia and (C,D) 69 year old subject
without dementia. (A) Color fundus photograph of
the macula demonstrates numerous drusen (area
7.35 mm2) and (B) F-AV45 (florbetapir) PET yields
positive standard uptake values (SUVR) of 1.15
(values greater than or equal to 1.10 are consid­
ered positive). (C,D) Similar images for a 69 year
old female subject without dementia. (C) This
subject had minimal drusen on fundus images and
(D) similarly negative SUVR ratio of 1.04. SUVR
values indicate the ratio of cortical to cerebellar A-
beta accumulation. (Adapted with permission from
Shoda C et al., Journal of Alzheimer’s Disease 62
(2018) 239–245 (doi 10.3233/JAD-170956).

14
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

association of drusen is age-related macular degeneration but numerous contributions to cognitive impairment and dementia is significant, many
studies have examined the potential role of Aβ accumulation in drusen of these studies cannot rule-out the confounding role of cardiovascular
and LOAD (Ohno-Matsui 2011). Unlike intraretinal Aβ deposits, drusen risk factors versus amyloid angiopathy in the correlation with OCTA
are easily detectable on clinical examination and with a variety of based metrics. Notably in a study where MRI and CSF amyloid levels are
ancillary diagnostic imaging methods. Interestingly, macular drusen used to characterize LOAD patients, the correlation between retinal
area on CFP independently correlates with cerebral Aβ accumulation as capillary density metrics and LOAD status is less clear (Querques et al.,
measured by florbetapir PET in elderly people without dementia (Fig. 6) 2019). In this study, a significant impairment of arteriolar reactivity, but
(Shoda et al., 2018). Histopathology of specimens from subjects with not capillary density, was noted in response to flicker stimulation of the
neuropathologically diagnosed LOAD demonstrates increased numbers retina in subjects with MCI and LOAD compared to normal. This study
of drusen compared with age-matched controls (Ukalovic et al., 2018). also accounted for vascular comorbidity. In addition, vascular reactivity
These findings are supported by epidemiological findings correlating impairment was inversely correlated with CSF Aβ levels and directly
decreased MMSE and late stages of age-related macular degeneration in correlated with GCL thickness (Querques et al., 2019). A reduction in
the Blue Mountain Eye Study (Pham et al., 2006). However, histopath­ RNFL thickness, retinal arteriolar diameter and blood flow was also
ological assessment of 157 autopsy eyes from subjects greater than 75 recently reported in subjects with clinically-defined LOAD or MCI using
years of age did not support an increase in prevalence of age-related a different imaging methodology but flicker induced reactivity was not
macular degeneration among subjects with LOAD (Schwaber et al., observed (Doppler OCT and Retinal Vessel Analyzer) (Szegedi et al.,
2020). This suggests that the pathophysiology of Aβ accumulation in 2020).
drusen is likely independent of the severity of age-related macular There is controversy regarding retinal capillary changes in preclini­
degeneration. cal LOAD. One study of 32 subjects with biomarker positive LOAD
(positive CSF Aβ42 or Florbetapir or Pittsburgh compound B PET li­
4.2.7. Retinal vascular changes in MCI and AD gands) but CDR score of 0 showed increasing foveal avascular zone and
Several lines of evidence in human LOAD have also provided evi­ decreased retinal thickness in biomarker positive subjects compared to
dence of retinal vascular pathology. The retina is supplied by two biomarker negative controls (O’Bryhim et al., 2018). Notably this study
separate vascular structures. The central retinal artery branches from did not report a change in capillary density. In contrast, Haan et al.
the ophthalmic artery and feeds the retina through additional branches showed no significant decrease in capillary density or increase in foveal
from the optic nerve head directly into the neurosensory retina. This avascular zone in LOAD participants with Aβ PET and CSF positivity,
“retinal” blood supply is further subdivided into capillary layers or plexi after adjusting for age and sex. There were no associations between any
depending on the depth of the capillary bed within the neurosensory retinal vascular parameters and CSF biomarkers or MMSE scores (Haan
retina and the topographic location within the retina. The retinal cir­ et al., 2019a,b). However, there was a positive and significant associa­
culation is responsible for supplying the inner two-thirds of the neuro­ tion between retinal vessel density in the macula and Fazekas score
sensory retina. In contrast, the posterior ciliary arteries also branch from suggesting some form of underlying vascular association (Haan et al.,
the ophthalmic artery but supply the choroidal circulation which is 2019a,b). One limitation of these studies is that assessment of retinal
anatomically distinct from the retina and immediately posterior to it. capillary density is limited by layer segmentation methods that vary
The choroidal circulation provides the nutrient supply to the outer one- across commercial devices and analysis methods. As we discuss in Sec­
third of the retina. Here we will review relevant findings in both vascular tion 7, volume based image analysis methods may help overcome this
systems in subjects with AD. difficulty.
Numerous studies have demonstrated changes in LOAD in the large
caliber retinal vessels (100–200 μm) emanating from the optic disc 4.2.8. Choroidal vascular changes in AD
including: narrowed vessel caliber (Berisha et al., 2007; Frost et al., In contrast to the retinal circulation, there are fewer studies exam­
2013; Cheung et al., 2014; Szegedi et al., 2020), increased tortuosity ining the choroidal circulation. This may be partly because the choroidal
(Frost et al., 2013; Cheung et al., 2014; Williams et al., 2015), reduced circulation is anatomically separate from the retina, does not share the
blood flow (Berisha et al., 2007; Feke et al., 2015; Szegedi et al., 2020), common embryological origins of the CNS vasculature, and does not
and altered blood oxygen saturation (Einarsdottir et al., 2016; Stefáns­ have a BRB. Therefore, there is less reason to suspect a direct correlation
son et al., 2017; Szegedi et al., 2020). Notably increase in arteriolar between changes in the choroidal circulation and CNS pathology.
tortuosity and decrease in caliber have been demonstrated in the cortical Nevertheless, the choroidal blood supply is an extremely important
microvasculature of transgenic mouse model of AD (Dorr et al., 2012). source of nutrition for the outer retina and may, at least, reflect changes
In addition to the large caliber vessel changes there are a number of in the metabolic activity of the retina. In the population based Beijing
studies demonstrating capillary level changes in the retina of LOAD Eye Study on 3009 subjects, there was a significant positive association
subjects. Decreased blood flow rates in the precapillary arterioles and between OCT measured choroidal thickness and MMSE score even after
post-capillary venules in the macula have been demonstrated (Jiang adjusting for age, axial length, gender, anterior chamber depth, lens
et al., 2018a,b) corresponding to the decreased flow rates observed in thickness, visual acuity and depression score (Jonas et al., 2016).
the larger retinal vessels at the disc (Berisha et al., 2007; Feke et al., In a meta-analysis of 203 clinically-defined AD subjects and 307
2015). Aβ accumulation and subsequent plaque formation in vessel controls from 5 cross-sectional studies using OCT, subfoveal choroidal
walls may occlude and thus impair blood flow (Berisha et al., 2007; Dorr thickness was significantly thinner in LOAD subjects than controls
et al., 2012). Schultz et al. reported significant reduction in capillary (standard mean reduction − 1.03 μm, p < 0.001) (Chan et al., 2019). It
pericyte number between retinal sections derived from LOAD and should be noted that some studies have reported conflicting results for
non-demented patient cases (Schultz et al., 2018) reminiscent of the loss choroidal thickness measurements as well (Gharbiya et al., 2014,
of brain pericytes reported in the AD cortex and hippocampus (Sengillo Bayhanet al., 2015; Bulut et al., 2016; Trebbastoni et al., 2017; Bulut
et al., 2013). et al., 2018; Haan et al., 2019a,b; López-de-Eguileta et al., 2020).
Several studies have demonstrated significant attenuation of capil­ A recent histopathological study of choroidal thickness and vascu­
lary density or perfusion in the retina of subjects with dementia (Bulut larity in 8 severe LOAD (Braak V-VI) and 11 control postmortem eyes
et al., 2018; Jiang et al., 2018a,b) or MCI (Jiang et al., 2018a,b; Yoon demonstrated global choroidal thinning in the superior hemi-retinas.
et al., 2019). Due to the high cost and limited availability of PET and CSF Intriguingly, regional analysis of the central macular choroid demon­
studies, most of these correlations are often made with only MMSE as strated significant thickening in CSF-biomarker defined LOAD, which
measure of disease status and little information is available about co­ strongly correlated with stromal vessel number (Asanad et al., 2019a,b)
morbid vascular disease. Since the prevalence of comorbid vascular (Fig. 7). Quantification of choroidal vasculature using OCT and OCTA is

15
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Fig. 7. Qualitative assessment of the choroid in controls and subjects with Alzheimer’s Disease. Representative light microscopy revealed qualitative (A) superonasal
choroidal thinning in patients with AD relative to (B) controls. Representative light microscopy revealed qualitative (C) superotemporal macular choroidal thickening
with increased vascularity in patients with AD relative to (D) controls. All stains are hematoxylin and eosin. Abbreviation: AD, Alzheimer’s disease. (Reproduced
under the Creative Commons Attribution-NonCommercial-NoDerivatives (CC BY-NC-ND) 4.0 International License from Asanad et al., Alzheimer’s and Dementia:
Diagnosis, Assessment and Disease Monitoring 11 (2019a) 775–783; doi.org/10.1016/j.dadm.2019.09.005).

much less developed and more subjective than quantification of retinal pathology is of interest. Though impairment in episodic memory is the
circulation (Ferrara et al., 2016) and there is a significant amount of most prominent symptom in early AD, executive function is also
work to be done in assessing choroidal changes in LOAD. Specifically, affected. Far less commonly, language and visuospatial deficits may be
analysis of choroidal changes on OCT is confounded by the even more the presenting symptom. Considering this clinical heterogeneity, the
obscure borders and anatomy of the choroid and choriocapillaris (in large variety of assessment instruments used and the absence of
comparison to the less obscure retinal boundaries). As we discuss in appropriate norms across diverse populations, the use of cognitive
Section 7 below, volume based assessments of the choroid and chorio­ impairment and decline as outcome measures in retinal studies poses
capillaris may help address this technical challenge. significant challenges.
One notable trend in the literature assessing retinal findings in sub­
4.2.9. Visual pathway electrophysiology of Alzheimer’s disease jects with dementia is the widespread use of the MMSE as an outcome
Several studies have provided evidence of retinal electrophysiologic measure. The MMSE is a screening instrument. It is easy to administer
dysfunction in LOAD that is complementary to the histopathologic and is reported in the majority of studies as an assessment of the severity
findings of GCL, RNFL and optic nerve atrophy described above. For of cognitive impairment in subjects with probable or possible LOAD. In
example, in the retina, multifocal ERG amplitudes were significantly many cases the complete characterization of the AD status of subjects is
reduced and implicit times were delayed in subjects with AD (Sen et al., not presented. The MMSE is a relatively easy screening test (low ceiling)
2020a,b). In addition, decreased N95 amplitude and increased implicit which does not assess all cognitive domains (e.g. does not assess exec­
time by PERG examination significantly correlated with pRNFL thick­ utive function). It exhibits poor sensitivity in relatively healthy partic­
ness reduction in clinically-defined LOAD (Parisi et al., 2001; Kraso­ ipants (Spencer et al., 2013) and in MCI (Arevalo-Rodriguez et al., 2015;
domska et al., 2010; Moschos et al., 2012). PERG measured ganglion cell Tsoi et al., 2015).
activity demonstrates significant amplitude reduction while PVEP It is therefore not surprising that there are mixed reports on the as­
assessment of retinocortical conduction time and ffERG are normal in sociation of MMSE and retinal thickness parameters. For example,
LOAD subjects compared to controls. This suggests that signal conduc­ Cipollini et al. illustrated significant thinning in global pRNFL, GCC and
tion velocity in the primary visual pathways may be spared from LOAD total macular volume in 25 clinically-defined LOAD patients in com­
pathology at least in some stages of the disease (Rimmer et al., 1989). parison with 17 controls. However, OCT-derived measurements did not
Variations on PERG and VEP stimulation paradigms to further isolate the correlate with MMSE scores but did correlate with measures of
magnocellular visual pathways in LOAD subjects seem to suggest that constructional praxis and processing speed (Cipollini et al., 2020).
they are more affected (Sartucci et al., 2010) in support of histopatho­ Similarly, in a study of 57 CSF amyloid positive and PET amyloid pos­
logic findings discussed above demonstrating preferential atrophy of itive LOAD subjects, MMSE did not correlate with any retinal thickness
larger ganglion cell axonal profiles also indicative of magnocellular parameter (Haan et al., 2019a,b). In subjects with PCA, a variant of AD
neurons (Sadun and Bassi 1990). largely affecting the visual cortex, there is no reported association be­
tween any measure of retinal thickness and LOAD status including
4.2.10. Cognitive impairment in AD and retinal thickness MMSE score (Haan et al., 2019a,b). In contrast, other studies show some
As cognitive decline is the most relevant and salient concomitant of association between retinal thickness parameters and MMSE in subjects
the severity of brain pathology in AD, its correlation with retinal with LOAD. For example, Cunha et al. show a correlation of 0.33 and

16
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

0.43 for average macular thickness and foveal thickness, respectively, and CSF tau in pre-symptomatic mutation carriers (Ringman et al.,
with MMSE in a univariate analysis of 24 clinically-defined LOAD pa­ 2008). Clinical, biochemical, and radiologic changes in ADAD have
tients (mean MMSE = 17) and 24 controls (Cunha et al., 2016). It is well-established temporal associations demonstrating that the disease
likely that more sensitive and comprehensive measures of cognition, process is active decades before clinical symptoms present (Bateman
perhaps composite measures emphasizing episodic memory and execu­ et al., 2012) and have been shown to parallel those seen in LOAD
tive function, might be better suited for correlation with retinal changes (Dubois et al., 2016). For example, while clinical disease activity in
in AD (Guzmán-Vélez et al., 2018). ADAD is detectable using the Clinical Dementia Rating sum of boxes
testing within 1 year of overt dementia, the earliest change seen is a
4.2.11. Longitudinal studies of retinal imaging in AD and dementia decrease in Aβ in the CSF ~25 years before presentation. The precuneus
Longitudinal data seem to provide complementary evidence to sup­ shows elevated levels of amyloid deposition on PET using Pittsburgh
port a potential causal association between neurodegenerative changes compound B approximately 25 years before disease onset as well a
and retinal atrophy. One longitudinal study of 3289 Dutch adults reports decline in glucose metabolism as measured by [F18] fluorodeoxyglucose
the association of retinal thickness with both incident and prevalent (FDG) PET 17 years before disease onset. Performance on a composite of
dementia, including LOAD specifically (Mutlu et al., 2018). As expected neuropsychological assessments, including episodic memory, complex
this study reports that thinner RNFL is associated with an increased risk attention, processing speed, and a general cognitive screen, are
of dementia overall and more specifically with incident dementia abnormal several years before clinically manifest disease. Hippocampal
(Hazard Ratio per standard deviation increase in RNFL 1.44 CI volume attenuation is detectable a few years before disease onset. This
1.19–1.75 for dementia and 1.43 CI 1.15–1.78 for LOAD) but not ordering of biomarkers and clinical tests strengthens the hypothesis that
prevalent dementia. Thinner GCIPL was associated with prevalent de­ clinical diagnosis of AD is made late in the biological cascade and tar­
mentia (odds ratio per standard deviation decrease in GCIPL 1.37 CI geting both diagnosis and pharmacologic manipulation of Aβ earlier in
0.99–1.9) (Mutlu et al., 2018). A second study of 32,038 subjects in the the course of the disease may lead to better clinical outcomes. In this
United Kingdom without clinically diagnosed neurodegenerative dis­ context, retinal studies in asymptomatic carriers of ADAD mutations are
ease demonstrated that thinner RNFL was associated with worse underway and are of significant potential interest. A recent study
cognitive performance at baseline and progressively worse cognitive showed decreased retinal thickness in 10 presymptomatic carriers of a
testing 3 years later (Odds Ratio 1.92 CI 1.29–2.85, p < 0.001) (Ko et al., common PSEN1 mutation relative to 10 matched non-carriers. Differ­
2018). A 27-month longitudinal study conducted by Santos et al. re­ ences were most evident in the outer nuclear layer. Interestingly, there
ported significant mRNFL thinning in preclinical LOAD. Although IPL was no effect of age on retinal thickness nor were there any differences
and ONL thinning were also observed, these thickness changes were in retinal vascular parameters assessed on fundus photography (number
reportedly attributed to age-related decline, as expected over the time­ of branch points, tortuosity, or fractal dimension) (Armstrong et al.,
span of the study. Notably, only the mRNFL significantly correlated with 2020). In another recent study, early stage (asymptomatic) carriers of
PET Aβ levels (Santos et al., 2018). One challenge with longitudinal ADAD causing mutations demonstrated increased capillary blood flow
studies is coregistration of images across subjects and across time so that and blood flow heterogeneity compared to controls or late stage
subtle subclinical changes can be detected. As we discuss later on (See (symptomatic) carriers of the same mutations (Singer M, Ringman J
section 7.2), novel methods of image registration and analysis may help et al. in press).
address this challenge.
4.4. Parkinson’s disease
4.3. Autosomal Dominant Alzheimer’s disease
Parkinson’s Disease (PD) affects more than 10 million people
As discussed in the previous section, the vast majority of AD cases worldwide and is the second most common neurodegenerative disease
present late in life with some heritable component (estimated at in the developed world (Archibald et al., 2009; Veys et al., 2019) and
58–79%) due to common genetic variants with small effects, and are numerous animal models have informed out understanding of the dis­
termed “sporadic” late onset AD (LOAD) (Ringman et al., 2014). How­ ease pathophysiology (Pingale and Gupta 2020). The pathologic hall­
ever, around 1% of cases are inherited as Autosomal Dominant AD or mark of PD is the abnormal deposition of cytoplasmic inclusions
ADAD (Moulder et al., 2013). Mutations in three genes cause ADAD: comprising the α-synuclein (αsyn) protein leading to the progressive loss
presenilin 1 (PSEN1), presenilin 2 (PSEN2), and amyloid precursor of dopaminergic neurons. In particular, dopamine depletion following
protein (APP). All pathogenic mutations increase the relative production neuronal loss in the nigrostriatal pathway leads to characteristic motor
of longer-length Aβ species,a finding not consistently observed in LOAD symptoms in PD including bradykinesia, resting tremor, rigidity, and
(Szaruga et al., 2015). Mutations in PSEN1 are the most common, ac­ postural instability. However, there are numerous non-motor manifes­
counting for about two-thirds of ADAD cases, APP the second most tation of PD including alterations in mood and sleep, autonomic nervous
common, and PSEN2 the least common. For persons carrying pathogenic system disturbances, visual hallucinations, and dementia
mutations, the future development of AD can be reliably predicted. (Ortuño-Lizarán et al., 2018). Consistent with this clinical spectrum of
Furthermore, the age of disease onset can be estimated as this age tends extra-motor symptoms in PD, studies have demonstrated αsyn neuro­
to be consistent among persons with the same mutation (Ryman et al., pathology in the CNS, peripheral nervous system and various end-organs
2014). PSEN1 has the youngest age of onset (as early as the 20’s but (Beach et al., 2010). Dopamine and dopaminergic receptors have been
more commonly in the 40’s or 50’s). Overall, the average age of onset is demonstrated in the retina via radioligand binding analyses (Borbe
45.7 ± 6.8 years, so there is a lower incidence of the comorbidities of et al., 1982) and histopathology (Archibald et al., 2009). Evidence of
aging such as cardiovascular disease and diabetes that frequently dopaminergic loss in retina of subjects with PD confirms the presence of
confound the diagnosis of LOAD and its underlying pathophysiology. neurodegenerative changes in the retina (Archibald et al., 2009). Clin­
Clinical and pathological phenotypes of ADAD share numerous simi­ ical impairments reported in subjects with PD include reading difficulty,
larities with LOAD (namely progressive dementia associated with impaired visual acuity, color discrimination, motion perception, and
accumulation of β-amyloid plaques and ptau neurofibrillary tangles), contrast (Archibald et al., 2009; Weil et al., 2016). Notably, the
but may differ in the mechanisms leading to pathologic accumulation of impairment in at least some of these clinical measures, such as contrast
Aβ (Ringman et al., 2014). Nevertheless, ADAD can serve as a model to sensitivity, is reversible with administration of levodopa (Bulens et al.,
study AD pathophysiology, biomarkers, and potential disease modifying 1987; Hutton et al., 1993; Giaschi et al., 1997). Though symptoms and
therapies in the absence of confounding age-related disease. signs of extrapyramidal dysfunction and responsiveness to L-dopa make
Early studies in subjects with ADAD identified elevated plasma Aβ the clinical diagnosis of PD more reliable and not as

17
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Table 2
Summary of retinal imaging studies characterizing human subjects with Parkinson’s disease.
Finding PD vs Controls No Difference Biomarker Complete eye exam* excluding confounding
confirmed (αsyn) disease

OCT/Histology
pRNFL Inzelberg et al., 2004; Altintaş et al., 2008; Aaker et al., 2010; Archibald et al., 2011; Bodis-Wollner et al., Inzelberg et al., 2004; Altintaş et al., 2008;
thinning Moschos et al., 2011; Garcia-Martin et al., Tsironi et al., 2012; Albrecht et al., 2012; (2014); Beach et al., Moschos et al., 2011; Garcia-Martin et al.,
2012a,b; La Morgia et al., 2013; Kirbas Roth et al., 2014; Mailankody et al., 2015; (2014) 2012a,b; La Morgia et al., 2013; Kirbas
et al., 2013; Garcia-Martin et al. (2012a,b); Bittersohl et al., 2015; Nowacka et al., et al., 2013; Garcia-Martin et al. (2012a,b);
Rohani et al., 2013; Jiménez et al., 2014; 2015; Stemplewitz et al., 2015; Pillai Rohani et al., 2013; Jiménez et al., 2014;
Satue et al., 2014; Garcia-Martin et al. et al., 2016; Polo et al., 2016; Gulmez Satue et al., 2014; Garcia-Martin et al.
(2014a,b); Bodis-Wollner et al., (2014); Sevim et al., 2019 (2014a,b); Kaur et al., 2015; Sari et al.,
Beach et al., (2014); Garcia-Martin et al. 2015; Pilat et al., 2016; Ucak et al., 2016;
(2014a,b); Bayhan et al., 2014; Yu et al., Satue et al., 2017; Moschos and Chatziralli,
2014; Kaur et al., 2015; Sari et al., 2015; 2018; Visser et al., 2018; Sung et al., 2019
Pilat et al., 2016; Ucak et al., 2016; Satue Aaker et al., 2010; Archibald et al., 2011;
et al., 2017; Moschos and Chatziralli, 2018; Albrecht et al., 2012; Mailankody et al.,
Visser et al., 2018; Chrysou et al., 2019; 2015; Bittersohl et al., 2015; Nowacka
Sung et al., 2019 et al., 2015; Stemplewitz et al., 2015; Polo
et al., 2016; Gulmez Sevim et al., 2019
mRNFL Pilat et al., 2016; Garcia-Martin et al. Lee et al. 2014; Schneider et al., 2014; Garcia-Martin et al. (2014a,b); Pilat et al.,
thinning (2014a,b) Ahn et al., 2018 Müller et al., 2014 2016; Ahn et al., 2018; Lee et al. 2014
RGC-IPL Hajee et al., 2009; Shrier et al., 2012 Adam Albrecht et al., 2012; Lee et al. 2014; Roth Bodis-Wollner et al., Hajee et al., 2009; Shrier et al., 2012;
thinning et al., 2013; Garcia-Martin et al. (2014a,b) et al., 2014; Schneider et al., 2014; Müller (2014); Beach et al., Garcia-Martin et al. (2014a,b); Moschos
Bayhan et al., 2014; Bodis-Wollner et al., et al., 2014; Mailankody et al., 2015; Pilat (2014) and Chatziralli, 2018; Ahn et al., 2018;
(2014) Kaur et al., 2015; Ucak et al., 2016; et al., 2016; Polo et al., 2016; Gulmez Adam et al., 2013; Gulmez Sevim et al.,
Moschos and Chatziralli, 2018; Ahn et al., Sevim et al., 2019 2019; Mailankody et al., 2015; Kaur et al.,
2018; Chrysou et al., 2019; Sung et al., 2015; Pilat et al., 2016; Ucak et al., 2016;
2019 Sung et al., 2019
Albrecht et al., 2012; Lee et al. 2014;
Mailankody et al., 2015; Pilat et al., 2016;
Polo et al., 2016; Gulmez Sevim et al., 2019
Total macular Altintaş et al., 2008; Aaker et al., 2010; Archibald et al., 2011; Shrier et al., 2012 Altintaş et al., 2008; Aaker et al., 2010;
thinning Cubo et al., 2010; Satue et al., 2014; Albrecht et al., 2012; Lee et al. 2014; Roth Shrier et al., 2012; Satue et al., 2014;
Garcia-Martin et al. (2014a,b); et al., 2014; Schneider et al., 2014; Kaur Garcia-Martin et al. (2014a,b);
Garcia-Martin et al. (2014a,b); Mailankody et al., 2015; Bittersohl et al., 2015; Garcia-Martin et al. (2014a,b); Mailankody
et al., 2015; Stemplewitz et al., 2015; Pilat Nowacka et al., 2015; Ucak et al., 2016; et al., 2015; Stemplewitz et al., 2015; Pilat
et al., 2016; Satue et al., 2017; Ahn et al., Pillai et al., 2016; Polo et al., 2016; et al., 2016; Satue et al., 2017; Ahn et al.,
2018; Huang et al., 2018; Chrysou et al., Gulmez Sevim et al., 2019; Uchida et al., 2018; Huang et al., 2018
2019 2018 Archibald et al., 2011; Shrier et al., 2012;
Albrecht et al., 2012; Lee et al. 2014; Kaur
et al., 2015; Bittersohl et al., 2015;
Nowacka et al., 2015; Ucak et al., 2016;
Pillai et al., 2016; Polo et al., 2016; Gulmez
Sevim et al., 2019
Choroidal Miri et al., 2015; Moschos and Chatziralli, Miri et al., 2015; Moschos and Chatziralli,
thinning 2018 2018
Vascular
Capillary Kwapong et al., 2018 Kwapong et al., 2018
density
decrease
FAZ decrease Miri et al., 2015 Miri et al., 2015
ERG
Inner retinal Ikeda et al., 1994; Peppe et al., 1995; Ikeda et al., 1994; Peppe et al., 1995;
dysfunction Tagliati et al., 1996; Garcia-Martin et al., Garcia-Martin et al., 2014a,b; Huang et al.,
2014a,b; Huang et al., 2018 2018
Outer retinal
Dysfunction

Abbreviations: = Optical Coherence Tomography, ERG = Electroretinography, PD = Parkinson’s Disease, pRNFL = peripapillary retinal nerve fiber layer, mRNFL =
macular retinal nerve fiber layer, RGC = Retinal Ganglion Cells, IPL = Inner Plexiform Layer, FAZ = Foveal Avascular Zone, Blank cells indicate no data or study was
available for that parameters. (*) Complete eye exam includes visual acuity, intraocular pressure measurement, and dilated fundus examination.

biomarker-dependent as that of AD, it is more difficult to differentiate immunoreactivity in the retina of 5 subjects with PD (Nguyen-Legros
dementia with Lewy bodies from AD. In elderly persons with PD or 1988; Witkovsky 2004). This finding has been corroborated and
dementia with Lewy bodies, concomitant diseases of aging should be expanded on by several subsequent studies. Histopathology of 4 PD eyes
considered as confounders when assessing retinal changes. Below we and 12 control eyes demonstrated αsyn immunoreactivity localized
summarize and synthesize a large body of literature describing retinal within the GCL, IPL and INL corresponding with the anatomic distri­
imaging findings in PD and implications of these findings in future bution of dopaminergic amacrine cells (Beach et al., 2014; Bod­
studies. The numerous references in this section are also summarized in is-Wollner et al., 2014). Prominent accumulation of phosphorylated
Table 2. αsyn deposition has been demonstrated in the RGC of subjects with PD
compared to controls and significantly correlates with cortical p-synu­
4.4.1. Parkinson’s disease retinal histopathology cleinopathy, disease severity, and functional motor scores
The initial evidence for retinal involvement in PD came from histo­ (Ortuño-Lizarán et al., 2018). It is notable that at least one study has
pathologic evidence demonstrating reduced tyrosine hydroxylase demonstrated more diffuse αsyn deposition throughout the retina of PD

18
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

(Ho et al., 2014). et al., 2016). As we discuss later on (See section 7.2), novel methods of
These findings are complemented by high performance liquid chro­ image registration and analysis may help circumvent these limitations.
matography demonstrating modulation of retinal dopamine levels in
subjects receiving recent L-DOPA therapy compared to those who did 4.4.4. Retinal vascular changes in Parkinson’s disease
not receive similar therapy (Harnois and Di Paolo 1990). Similar data Few studies have evaluated the retinal microvasculature in PD. Miri
from rodent and primate models of PD using a neurotoxin for dopami­ et al. assessed the perifoveal capillary network in 23 PD patients and 13
nergic cells, 1-methyl-4-pheny-1,2,3,6-tetrahydropyridine (MPTP), healthy controls using invasive injection of contrast dye (fluorescein
demonstrates a dose-dependent reduction in tyrosine hydroxylase ac­ angiography). PD patients exhibited significant reduction in FAZ area
tivity in retinal amacrine cells (Ghilardi et al., 1988a,b; Tatton et al., and foveal thickness. In addition, foveal thinning was highly associated
1990). with disease duration and motor impairment in PD patients (Miri et al.,
2015). Choroidal thickness reduction in the subfoveal area as well as in
4.4.2. Peripapillary RNFL changes in Parkinson’s disease the inner and outer quadrants in PD retina has also been demonstrated
In a meta-analysis of 1916 subjects with PD and 2006 controls there (Moschos and Chatziralli 2018). More recently, Kwapong et al. analyzed
was a significant association between diagnosis of PD and thinning of the SRL and DRL capillary densities in 38 early PD and 28 control sub­
the pRNFL (Cohen’s d = − 0.42, CI -0.54 to − 0.29) and the GCIPL (d = jects using OCTA. SRL density was markedly reduced relative to healthy
− 0.40, CI -0.72 to − 0.07) (Chrysou et al., 2019). There was no detect­ controls and correlated with GCIPL thickness, suggesting a vascular
able association with sectoral thinning, duration of disease or severity of contribution or consequence of retinal degeneration in PD (Kwapong
symptoms in this study. A number of studies have revealed preferential et al., 2018).
thinning of the inferior and temporal pRNFL quadrants (Inzelberg et al.,
2004; Moschos et al., 2011; Garcia-Martin et al., 2012a,b; Kirbas et al., 4.4.5. Clinical correlation of retinal changes in PD
2013; La Morgia, Barboni et al., 2013; Sari et al., 2015; Satue et al., Researchers have hypothesized that retinal thinning in PD reflects a
2017; Visser et al., 2018; Sung et al., 2019). However, results have been primary loss of dopaminergic amacrine cells and a decrease in dopamine
mixed among other studies (Rohani et al., 2013, Bayhanet al., 2014; neurotransmitter levels in the retina that is presumably reflective of
Garcia-Martin et al., 2014a,b; Jiménez et al., 2014; Yu et al., 2014; Kaur similar changes in the CNS. There are functional implications of this
et al., 2015; Pilat et al., 2016; Pillai et al., 2016; Ucak et al., 2016; hypothesis including correlation of retinal changes with abnormalities
Hasanov et al., 2019). Despite the differences in sectoral location, it is in visual function, motor function and cognitive function. A large meta-
clear that retinal thinning, predominantly in the inner retina, is present analysis of 36 studies (n = 1916 subjects with PD and 2006 controls)
in subjects with PD. Careful evaluation of device parameters, segmen­ found significant thinning of the inner retinal layers in PD but only a few
tation algorithms and reliability as well as disease stage may explain significant correlations were found between retinal thickness and clin­
why some studies do not report significant retinal thickness changes in ical severity, perhaps due to the heterogeneity in the underlying data
the pRNFL (Archibald et al., 2009; Albrecht et al., 2012; Tsironi et al., and methodology among the studies (Chrysou et al., 2019). Nevertheless
2012; Roth et al., 2014; Mailankody et al., 2015). For example, it is there are some interesting studies that are worth mentioning. For
possible that changes in macular thickness precede detectable RNFL example, in one study of 14 PD subjects and 14 controls there was a
changes of the ganglion cell body, dendritic field and synaptic connec­ significant association between impaired contrast sensitivity and inner
tions are the primary site of cell damage. Interestingly, some studies retinal thickness among normal subjects that was absent in subjects with
report changes in functional parameters in the absence of structural PD (Adam et al., 2013). Direct evidence for a common pathophysio­
thinning suggesting that functional changes may precede detectable logical mechanism in the retina and CNS of PD subjects also exists.
attenuation in pRNFL thickness (Tsironi et al., 2012). Namely, a significant association has been shown between the thickness
of the inner retina and dopamine aminotransferase activity in the sub­
4.4.3. Macular OCT changes in Parkinson’s disease stantia nigra via PET imaging (Ahn et al., 2018).
In a meta-analysis of 1916 subjects with PD and 2006 controls there Although in most studies the heterogeneity of PD duration, severity
was a significant association between diagnosis of PD and thinning of and clinical stage precludes rigorous analysis of clinical PD staging and
most sectors of the macula (Cohen’s d range − 0.37 to − 0.57) (Chrysou retinal thickness measures (Chrysou et al., 2019) cross-sectional asso­
et al., 2019). The majority of studies investigating the macula in PD have ciations between the severity of PD and retinal thinning in the macula
reported significant reduction in one or more layers, especially the (Bayhanet al., 2014; Sari et al., 2015; Pilat et al., 2016; Ahn et al., 2018;
mRNFL, GCL, and IPL (Hajee et al., 2009, Cubo et al., 2010; Kaur et al., Sung et al., 2019) and pRNFL (Jiménez et al., 2014; Mailankody et al.,
2015; Sari et al., 2015; Ucak et al., 2016; Satue et al., 2017; Ahn et al., 2015; Pilat et al., 2016) have been reported. In a few studies, the severity
2018; Moschos and Chatziralli 2018; Sung et al., 2019). These OCT and duration of PD has been shown to correlate with the GCIPL layer
findings are complementary to the histopathologic findings described in thickness (Sari et al., 2015). There are very few longitudinal studies to
the inner retinal layers (Bodis-Wollner et al., 2014). It has also been support these cross-sectional associations. At least in one study that
proposed that the slope or shape of the foveal pit may have some utility evaluated the same 30 PD subjects and 30 controls over a 5 year period,
as a marker of PD-specific changes (Shrier et al., 2012; Pilat et al., 2016) the attenuation in the temporal and supratemporal pRNFL was signifi­
although there is significant normal variation in foveal morphology that cantly greater in subjects with PD than controls (Satue et al., 2017). In
needs be taken into account when considering such an association. this same study, PD progression using the Hoehn and Yahr scale also
There is significant disagreement in the literature regarding retinal correlated significantly with thinning in the supratemporal pRNFL
thickness changes in the outer retina in subjects with PD. Significant (Satue et al., 2017). Nevertheless, in some well-performed studies there
thickness changes of the outer retina including the OPL, photoreceptor were no reliable correlations between retinal thickness measures of any
layer, and RPE layer have been reported in some studies (Müller et al., kind and severity or duration of motor symptoms using consensus PD
2014; Roth et al., 2014; Pilat et al., 2016) and refuted in others (Uchida assessment tools such as the Unified Parkinson’s Disease Rating Scale
et al., 2018; Chrysou et al., 2019). Very similar to the layer specific (Albrecht et al., 2012).
changes reported in subjects with AD, there appear to be layer specific Cognitive impairment is one of several clinical features of PD that
changes in PD which may offset each other (Garcia-Martin et al., 2014a, significantly impacts the quality of life of PD subjects. Most PD patients
b). This requires further investigation and may explain the findings from develop dementia and as many as 26% of non-demented subjects have
a large number of studies that fail to detect changes in overall retinal MCI (Litvan et al., 2011). There are generally very few studies that show
thickness (Aaker et al., 2010; Archibald et al., 2011; Tsironi et al., 2012; associations between cognitive impairment in PD and retinal thickness
Schneider et al., 2014; Nowacka et al., 2015; Pillai et al., 2016; Polo parameters. However, cognitive assessment in PD is subject to the same

19
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

challenges as those in AD described above. In many of these studies the 0.38 in 100,000, respectively (Pringsheim et al., 2012). Numerous ani­
mean MMSE is above 27 suggesting that a representative spectrum of mal models have informed our understanding of this disease in humans
subjects with cognitive impairment was not available for analysis (Howland et al., 2020). Though somewhat variable, the age of symptom
(Mailankody et al., 2015; Visser et al., 2018). Even in one study of 61 PD onset among carriers of pathogenic CAG repeat expansions is earlier
subjects and 30 controls with mean MMSE of 24, there was no correla­ with longer expansions (Andrew et al., 1993). The histopathological
tion between cognitive impairment and retinal thickness parameters hallmarks of HD include abnormal aggregations of intranuclear and
(Lee et al., 2014). Notably, a significant correlation between MoCA score cytoplasmic inclusions of the huntingtin (htt) protein, which lead to
(which includes measures of executive function and is more sensitive for progressive cortical and striatal atrophy. Patients classically present
detecting MCI) and thickness of the macular GCIPL has been demon­ with a clinical triad of motor, cognitive, and behavioral abnormalities
strated (Sung et al., 2019). As mentioned earlier, since the GCIPL is the (Ross et al., 2014). Motor symptoms characteristically include invol­
location of the dopaminergic amacrine cells in the retina as well as untary, jerky, or slow writhing movements (chorea), most obviously
ganglion cell bodies, this may suggest that detection of early changes in involving the upper extremities. Ocular motility deficits such as saccade
retinal structure require targeted sublayer analyses with careful atten­ apraxia, saccade hypometria, and smooth pursuit impairment have also
tion to segmentation methods. As we discuss later on (See section 7.2), been described in various stages of HD and are reviewed extensively
novel methods of image registration and analysis may help circumvent elsewhere (Anderson and MacAskill 2013). Below we summarize and
these limitations. synthesize the available body of literature describing retinal imaging
findings in HD and implications of these findings in future studies. The
4.4.6. Visual pathway electrophysiology in Parkinson’s disease numerous references in this section are also summarized in Table 3.
Some of the earliest work implicating electrophysiological abnor­
malities in the visual pathway in PD comes from studies using VEP in PD 4.5.1. Huntington’s Disease retinal histopathology
subjects. Specifically, increased VEP latency in response to mid-spatial Autopsy-derived brain and retinal tissue samples from 1 HD and 2
frequency stimuli have been demonstrated in PD subjects by several control subjects showed the expected findings of significant cortical
studies (Bodis-Wollner and Yahr 1978; Onofrj et al., 1986; Ikeda et al., atrophy associated with htt and ubiquitin deposits in the caudate nu­
1994). These abnormalities are at least partially reversible during cleus and putamen but the retina was devoid of protein aggregations or
L-dopa therapy (Bodis-Wollner and Yahr 1978; Peppe et al., 1995) and degenerative changes both macroscopically and histologically (Pet­
are inducible with dopamine antagonists (Onofrj et al., 1986). In at least rasch-Parwez et al., 2005). This is in contrast to reports in animal
one study, progression of ERG and PERG changes has been correlated models, which have demonstrated progressive retinal degeneration in
with clinical disease progression (Ikeda et al., 1994). Studies comparing R6/2 HD mice (Helmlinger et al., 2002). Due to the paucity of human
VEP latency and measures of retinal function such as PERG (Nightingale histopathology in HD and the discrepancy between human and animal
et al., 1986) suggest that at least part of the VEP abnormality is sec­ studies, it is possible that under-sampling in human tissue resulted in
ondary to retinal dysfunction. PERG abnormalities in the retina of sub­ false negative finding. It is also possible that retinal manifestations of HD
jects with PD are also responsive to administration of L-dopa (Peppe pathology are different than other CNS manifestations.
et al., 1995; Tagliati et al., 1996) and can be induced in non-PD subjects
and monkey models with selective D2-receptor antagonists I-sulpiride 4.5.2. Huntington’s Disease peripapillary RNFL
(Tagliati et al., 1994). Primates treated with the MPTP demonstrate Several researchers have shown clinical evidence of pRNFL thinning
decreased retinal dopamine levels, abnormal PVEP latencies and PERG in HD preferentially of the temporal quadrant (Kersten et al., 2015;
amplitudes as well as systemic signs of PD (Ghilardi et al., 1988a,b). Gatto et al., 2018, Gulmez Sevim et al. 2019). Temporal thinning in the
Notably the changes in PVEP and PERG were transiently reversible with pRNFL corresponds with the small axonal fibers comprising the papil­
administration of levodopa-carbidopa in the primate model (Ghilardi lomacular bundle which is a pattern of retinal atrophy classically seen in
et al., 1988a,b). mitochondrial disorders (Carelli et al., 2009). Therefore, despite the
Structure-function correlation of the retina support the notion that absence of histopathologic evidence of retinal involvement there is
the electrophysiologic changes described above are part of a common significant in vivo evidence suggesting pRNFL thinning in HD.
degenerative process of PD. For example, in a group of 46 PD subjects
and 33 controls, foveal thickness measurements and PERG N95 ampli­ 4.5.3. Huntington’s Disease macular RNFL
tude each were remarkably predictive of disease severity and quality of Although several studies failed to demonstrate any changes or cor­
life scores in PD patients (Garcia-Martin et al., 2014a,b). Similarly a relations between total macular thickness and HD status (Kersten et al.,
combined mfERG and OCT study comprising 53 PD and 41 control 2015; Gatto et al., 2018) retinal sublayer analysis has revealed signifi­
subjects found diminished amplitude and prolonged implicit time of the cant thinning for all of the inner retinal layers including the mRNFL,
P1 mfERG wave component in addition to macular thinning in PD pa­ GCL, IPL, and INL. These atrophic changes in the inner retina are
tients relative to controls. The diagnostic efficacy of ERG and structural coincident with thickening in the outer retinal layers that may explain
parameters when combined greatly outweighed the performances of the absence of significant changes in total retinal thickness measures
these biomarkers individually (Huang et al., 2018). However, the as­ (Andrade et al., 2016, Gulmez Sevim et al. 2019). Among retinal layers,
sociation between macular volume, foveal thickness, RNFL thickness however, thinning of the GCL correlated the strongest with HD and
and VEP latency appears poor (Altintaş et al., 2008). As we discuss measures of disease progression which suggest earlier involvement of
previously, this may be because measurement of specific retinal layers is the ganglion cell bodies in HD as compared to the axonal fibers (Satue
necessary to detect the relevant pathological process and segmentation et al., 2016, Gulmez Sevim et al. 2019).
errors or variances can easily obscure such subtle changes. As we discuss
later on (See section 7.2), novel methods of image registration and 4.5.4. Functional impairment and Huntington’s Disease retinal findings
analysis may help circumvent these limitations. Several studies have reported significant correlations between
retinal thickness measures and clinical assessments of HD severity. For
4.5. Huntington’s Disease example total mean RNFL thickness significantly correlated with total
functional capacity scores among 14 genetically confirmed HD subjects
Huntington’s disease (HD) is an autosomal dominant neurodegen­ and 13 matched controls (Gatto et al., 2018). Both duration of disease
erative disorder caused by expansion of the cytosine-adenine-guanine (R2 = − 0.57, p = 0.006) and Unified Huntington’s Disease Rating Scale
(CAG) polyglutamine triplet repeat in the HTT gene on chromosome (R2 = − 0.56, p = 0.01) were significantly and inversely correlated with
4p16.3.(1993) HD has a prevalence and annual incidence of 2.71 and macular volume in 26 HD patients and 29 controls (Kersten et al., 2015).

20
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Table 3
Summary of retinal imaging studies characterizing human subjects with Huntington’s disease.
Finding HD vs Controls No Difference Genetically confirmed (CAG) Complete eye exam* excluding
confounding disease

OCT
pRNFL thinning Kersten et al., (2015), Gatto et al., Kersten et al., (2015), Gatto et al., Kersten et al., (2015), Gatto et al.,
(2018), Gulmez Sevim et al., 2019 (2018), Gulmez Sevim et al., 2019 (2018), Gulmez Sevim et al., 2019
mRNFL thinning Gulmez Sevim et al., 2019, Andrade Gulmez Sevim et al., 2019 Gulmez Sevim et al. 2019, Andrade et al.,
et al., (2016) (2016)
RGC-IPL thinning Gulmez Sevim et al., 2019, Andrade
et al., (2016)
Total macular Kersten et al., (2015), Kersten et al., (2015), Gatto et al., Kersten et al., (2015), Gatto et al., (2018)
thinning Gatto et al., (2018) (2018)
Choroidal Andrade et al. (2016) Andrade et al. (2016)
thinning
ERG
Inner retinal dysfunction
Outer retinal Pearl et al., (2017); Knapp et al., Pearl et al., (2017); Knapp et al., (2018) Pearl et al., (2017); Knapp et al., (2018)
Dysfunction (2018)

Abbreviations: = Optical Coherence Tomography, ERG = Electroretinography, HD = Huntington’s Disease, pRNFL = peripapillary retinal nerve fiber layer, mRNFL =
macular retinal nerve fiber layer, RGC = Retinal Ganglion Cells, IPL = Inner Plexiform Layer, Blank cells indicate no data or study was available for that parameters. (*)
Complete eye exam includes visual acuity, intraocular pressure measurement, and dilated fundus examination.

Notably, these correlations increased in strength following exclusion of study (Josiassen et al., 1984). Nevertheless, VEP amplitudes were still
presymptomatic patients from the analysis. Other sublayers including significantly lower for both medicated and unmedicated HD patients
the GCL, IPL and mRNFL thicknesses significantly correlate with HD relative to controls. Importantly, similar VEP irregularities are also seen
duration, CAG repeat number, and Unified Huntington’s Disease Rating in psychiatric diseases and are not specific for HD (Josiassen et al.,
Scale scores (Gulmez Sevim et al. 2019). Similar associations have been 1984). Notably, at least one study has not been able to demonstrate
reported between clinical measures of HD severity and one or more similar findings using PVEP over a two year period (Ehle et al., 1984).
macular thickness measures in other studies (Andrade et al., 2016).

4.5.5. Visual pathway electrophysiology in Huntington’s Disease 4.6. Cerebral small vessel disease
In a recent study comprising 18 HD and 10 control subjects, signif­
icantly increased b-wave amplitudes in HD at light- and dark-adapted Cerebral small vessel disease (CSVD) refers to a group of human
red flash intensities were reported (Pearl et al., 2017). In addition, cerebrovascular diseases that affect small intraparenchymal arterioles in
greater numbers of CAG triplet repeats correlated with higher b-wave the CNS (Pantoni 2010; Iadecola 2013; Cannistraro et al., 2019) and
amplitudes. These results expand upon the initial foveal blue light many aspects of these diseases have been replicated in animal models
studies by Paulus et al. demonstrating increased photoreceptor thresh­ (Tuo et al., 2020). Common causes of human CSVD include hypertensive
olds in HD (Paulus et al., 1993), and suggest electrophysiologic arteriolosclerosis, cerebral amyloidosis, lupus vasculitis, and cerebral
dysfunction in retinal cone photoreceptor pathways in HD. More autosomal dominant arteriopathy with subcortical infarcts and leu­
recently, Knapp et al. evaluated retinal structure and function in a koencephalopathy (CADASIL). CSVD is responsible for approximately
25-year-old man with presymptomatic HD (Knapp et al., 2018). 25% of strokes worldwide (both ischemic and hemorrhagic), and is
Although retinal structure appeared normal on OCT, ERG amplitudes associated with increased risk of recurrent stroke (Rensma et al., 2018).
were decreased in dark- and light-adapted a- and b-waves as well as in It is also a primary contributor to cognitive decline, with up to 45% of
light-adapted 30 Hz flicker testing (Knapp et al., 2018). In addition, dementia cases in the general population being associated with CSVD
mfERG showed a decrease in response amplitude suggesting predomi­ (Gorelick et al., 2011; Gorelick and Pantoni 2013; Sachdev et al., 2014).
nant involvement of cone photoreceptors and downstream pathways. Current clinical and research practices for identification of CSVD rely on
These findings in presymptomatic disease may also suggest early neuroimaging (usually MRI) evidence of parenchymal brain injury
involvement of the photoreceptor pathway in HD that precedes retinal caused by CSVD (Wardlaw et al., 2013; Cannistraro et al., 2019; Das
morphologic changes measured by OCT. This is also consistent with the et al., 2019). These biomarker findings include, white matter hyper­
retinotopic pattern of degeneration involving the parvocellular neurons intensity (WMH), lacunar infarcts, cerebral microbleeds (CMB), con­
of the papillomacular bundle described above (La Morgia, Di Vito et al., vexal subarachnoid hemorrhage (cSAH), cortical superficial siderosis
2017). This is at least consistent with the R6/2 mouse model of HD that (cSS), large hemorrhagic stroke lesions, and enlarged perivascular space
develops cone-rod degeneration measurable by ERG (Ragauskas et al., (PVS) (Fig. 8) (Wardlaw et al., 2013).
2014). Nevertheless, these results should be interpreted with caution Population-based series including imaging and autopsy data identi­
given the small numbers of subjects involved. fied two primary CVSD subtypes that account for the vast majority of
cases (Cannistraro et al., 2019). Hypertension-related Angiopathy
4.5.6. Huntington’s Disease visual evoked potentials (HTNA) is a form of acquired CVSD and represents a small vessel dis­
The majority of reports have illustrated decreased amplitude with order caused by prolonged uncontrolled hypertension. Rather than a
normal peak latency in the VEP responses of HD patients (Ellenberger single pathological entity, HTNA represents a group of hypertensive
et al., 1978; Rizzo et al., 1980; Oepen et al., 1981; Josiassen et al., 1984). cerebral angiopathies that frequently overlap and include hyaline arte­
The magnitude of the amplitude decrease seems to correlate with the riolosclerosis, hyperplastic arteriolosclerosis, segmental arterial disor­
severity or duration of the disease (Ellenberger et al., 1978). One ganization, and microaneurysms (Pantoni 2010). The changes in lumen
potentially confounding variable in these assessments is the effect of morphology and biological properties associated with these conditions
medication on observed waveform abnormalities. In particular, VEP are thought to lead to a cascade of secondary injury processes including
amplitudes were lower in HD patients receiving antipsychotic treatment decreased cerebral blood flow (up to and including vascular occlusion
as compared to patients who were not receiving treatment in at least one and acute ischemic infarction), leakage of blood products from micro­
aneurysm (results in hemorrhagic lesions ranging in severity from CMBs

21
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

to large hypertensive hemorrhages), and chronic ischemia of oligoden­ retinal changes associated with hypertension have been consistently
drocytes dysfunction leading to loss of myelin integrity (manifesting as associated with risk for cerebrovascular disease (Henderson et al.,
white matter disease) (Wardlaw et al., 2013; Promjunyakul et al., 2018). 2011). Additional studies of retinal histopathology and CSVD are
Cerebral amyloid angiopathy (CAA) is another form of CVSD that is required to identify tissue changes specific to CSVD, and correlate them
acquired and caused by deposition of Aβ in the walls of small-to-medium with characteristics neuroimaging findings and disease progression.
cortical and leptomeningeal arterioles and arteries (Biffi and Greenberg
2011). In turn, the development of amyloid-related angiopathy results in 4.6.2. Neurosensory and peripapillary retinal changes in cerebral small
loss of vessel compliance and decreased cerebral vascular reactivity. vessel disease
Ultimately these changes in vessel physiology lead to accumulation of Retinal structural changes have so far received limited attention in
ischemic (WMH) and hemorrhagic (cSAH, CSS, large cortical hemor­ CSVD investigative efforts. RNFL defects on fundus photography were
rhages) lesions, leading to acute stroke and progressive diffuse cortical shown to be associated with WMH in a large study including Korean
atrophy (Pantoni 2010). Because of shared links to Aβ deposition CAA is community-dwelling individuals who participated in health checkups
usually associated with AD, with the two conditions representing spe­ (Kim et al., 2011). RNFL and GCL were shown to be associated with
cific manifestations (vascular vs. parenchymal) of an overarching dis­ CSVD, primarily in terms of WMH prevalence and severity (Qu et al.,
order of amyloid cerebral production, processing and clearance 2020). Additional studies, ideally including longitudinal evaluation of
(Greenberg et al., 2020). both retinal neurosensory changes in relation to the progression of
capillary and arteriolar pathology, are warranted at this time. As we
4.6.1. Retinal histopathology in cerebral small vessel disease discuss in sections 5 and 7, high-resolution capillary imaging of the
There have been to date no dedicated studies of retinal histopa­ retina and novel image analysis and registration methods will provide
thology in CVSD, either in the CAA or HTNA form. This in part relates to excellent tools for assessing longitudinal and subclinical changes.
the ease of acquiring detailed information on vascular structures of in­
terest in vivo by a variety of techniques (fundus photography and OCTA 4.6.3. Retinal vascular changes in cerebral small vessel disease
being the most commonly implemented, see below). It is worth For the purposes of our discussion, the retinal vasculature can be
mentioning that retinal vascular changes identified in amnestic MCI and divided into macrovascular components that are generally greater than
AD patients may partially reflect separate contributions from CSVD 100 μm in diameter (retinal arteries and venules) and microvascular
(either CAA or HTNA), which is comorbid in a large proportion of cases components (arterioles, venules and capillaries) that are less than 100
(Matej et al., 2019). Histopathological studies of AD retinas showed μm in diameter. Many studies classify retinal vascular changes as
pathogenetic Aβ deposits accumulating within and along retinal vascu­ “retinopathy” which is a generic term that applies to almost any vascular
lature in a manner highly reminiscent of CAA (La Morgia, Ross-Cisneros changes in the retina. In the context of this manuscript, we will use
et al., 2016; Koronyo et al., 2017). As for HTNA, histopathological “retinopathy” to refer specifically to end-stage capillary changes

Fig. 8. Representative MRI findings in cerebral small vessel disease.

22
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

(microaneurysms, cotton wool spots, retinal hemorrhages, hard exu­ determined via Doppler) was notable for lower end-diastolic and mean
dates) that are secondary changes resulting from obliteration or signif­ velocities, and higher pulsatility and resistive indexes among CSVD
icant structural abnormality of the retinal capillaries. These distinctions patients compared to controls. Doppler flow velocities were also asso­
are important because color fundus photography is not able to resolve ciated with WMH severity on MRI (Hiroki et al., 2003). While limited,
end-arteriolar and capillary changes in the retina but can clearly show evidence from these studies suggest that a multimodal approach to
the secondary results of chronic capillary pathology such as micro­ retinal imaging may increase the yield of current CSVD research ap­
aneurysms, cotton wool spots, hemorrhages and hard exudates. In proaches, focused primarily on retinal imaging.
contrast, studies using OCTA are specifically designed to look directly at
capillary changes and not secondary effects of capillary damage. 4.6.4. Visual pathway electrophysiology in cerebral small vessel disease
Multiple studies have investigated the association between mea­ There have been to date no studies leveraging visual pathway elec­
surements of neuronal and retinal macrovascular integrity on fundus trophysiology to study CSVD. However, multiple studies investigated
photography and CSVD (Wu et al., 2017). In addition, retinopathy on response in cerebral blood flow to visual pathway stimulation, via either
fundus imaging is associated with prevalence and progression of cere­ Transcranial Doppler ultrasound or functional MRI. Patients with CAA
bral infarcts and WMH in population-based studies (Cooper et al., 2006; demonstrate reduced cerebral vascular reactivity to visual pathway
Longstreth et al., 2007; Cheung et al., 2010; Hanff et al., 2014). These stimulation, consistent with imaging and histopathology studies indi­
retinal photography findings were also associated with CSVD neuro­ cating greater disease severity in the occipital lobe (Smith et al., 2008;
imaging markers among patients presenting with acute stroke, and Peca et al., 2013).
preferentially identified patients presenting with CSVD-related stroke
subtypes (primary intracerebral hemorrhage and lacunar infarcts) 4.6.5. Autosomal dominant cerebral small vessel disease
(Lindley et al., 2009; Ong et al., 2013; Liew et al., 2014; Wei et al., Thus far there are few studies of retinal changes in genetically-
2016). Among quantitative retinal vasculature metrics arteriolar mean determined CSVD but this is a rapidly growing area of investigation
diameters and fractal dimensions (as derived from high-resolution with exciting initial findings. Cerebral autosomal dominant arteriopathy
fundus images) were repeatedly found to associate with CSVD MRI with subcortical infarcts and leukoencephalopathy (CADASIL) is the
markers, especially WMH and lacunar infarcts (Kwa, van der Sande most common genetic form of CSVD (Federico et al., 2012). CADASIL is
et al., 2002; Doubal et al., 2010; McGrory et al., 2019). In the Rotterdam an autosomal dominant genetic disorder caused by mutations in the
Study retinal venular diameter did not associate with baseline CSVD NOTCH3 gene, encoding a transmembrane receptor expressed almost
severity, but did correlate with progression on follow-up MRI scans exclusively in vascular smooth cells and pericytes (Monet-Leprêtre et al.,
(progression of WMH and new lacunar infarcts) (Ikram et al., 2006). A 2009). Accumulation of the mutated protein product leads to formation
recent study using OCTA has further demonstrated associations between of toxic extracellular deposits, leading to alterations in small vessel
markers of retinal capillary health and function with intracranial physiology similar to those observed in CAA. Among
vascular health and function (Ashimatey, D’Orazio et al., 2020). In this genetically-determined forms of CSVD, CADASIL has been the topic of
study, lower retinal capillary perfusion density was significantly asso­ most studies looking into potential application of retinal imaging tech­
ciated with worse MRI measures of cerebrovascular reactivity, lower niques. In an early study of 10 consecutive subjects with skin biopsy or
perfusion in the middle cerebral artery perfusion territory as well as genetically confirmed CADASIL, color fundus photographs and
impaired cognition (Ashimatey, D’Orazio et al., 2020) (Fig. 9). Collec­ dye-based retinal angiography demonstrated bilateral arteriolar
tively these data strongly suggest that retinal capillary changes mirror sheathing in 30% of subjects, arteriolar narrowing in 80% of subjects
vascular pathology in the brain at all levels of the vascular tree. and arteriovenous nicking in 90% of subjects (Haritoglou et al., 2004).
One of the advantages of MRI for CSVD is the ability to distinguish Notably, the prevalence of focal arteriolar narrowing and arteriovenous
between subtypes, which in turn has direct implications for associated nicking in population-based studies is much lower that reported in this
clinical manifestations and prognosis (Cannistraro et al., 2019). To date, small cohort (~14%) (Wong et al., 2007). Subsequent studies focused
few studies distinguished between different forms of CSVD when primarily on OCT and OCTA. Genetically confirmed CADASIL cases
exploring associations with retinal imaging. Retinopathy lesions on were shown to have significantly increased OCT-based mean arterial and
fundus imaging were found to be more common among survivors of venous diameters compared to controls (Alten et al., 2014). OCT mea­
CSVD-related stroke compared to other stroke subtypes (Baker et al., surements of arterial and venous mean diameters were also found to
2010a,b; Gobron et al., 2014). Among subtypes of CSVD-related stroke, correlate with lacunar infarcts and cerebral microbleeds on MRI (Fang
cases of amyloid-related lobar intracerebral hemorrhage demonstrated et al., 2017). In one study of OCTA in CADASIL, vessel density of the DRL
the highest prevalence of microvascular wall signs and retinopathy was found to be significantly decreased in affected patients compared to
findings on fundus imaging (Baker et al., 2010a,b). A study in rural healthy controls (Nelis et al., 2018). Larger and longitudinal studies will
Ecuador found associations between hypertensive retinopathy and be ultimately needed to validated and expand upon these findings.
CSVD findings on MRI, but did not distinguish between hypertensive Additional evidence on the use of retinal imaging has also recently
and amyloid subtypes (Del Brutto, Mera et al., 2016). Another study emerged for disorders with mixed small and large vessel CNS manifes­
found reduction in arteriolar fractal dimensions to be more common in tations. In Fabry disease, retinal and conjunctival vessel tortuosity rep­
CSVD-related vascular impairment compared to AD, but did not further resents an important diagnostic finding. Analysis of retinal and
distinguish between different small vessel disease etiologies (Jung et al., conjunctival vasculature demonstrated increased vessels tortuosity in
2019). While more evidence is needed to definitively assess whether Fabry disease patients (Sodi et al., 2019a,b). Evidence of abnormal RNFL
retinal imaging modalities can accurately subtype CSVD, this potential and vessel appearance on OCT have also been reported in smaller studies
application would directly impact research and clinical care practices. of Mitochondrial Encephalopathy, Lactic acidosis, and Stroke-like epi­
In addition to OCT/OCTA, other retinal vascular imaging modalities sodes (MELAS) syndrome (Cho and Yu 2015; Sultan et al., 2017) Taken
were investigated for association with CSVD. In one study, the diagnosis together, these findings suggest that retinal imaging of neuronal and
of Nonarteritic Anterior Ischemic Optic Neuropathy (NAION) was vascular structures may aid in early diagnosis and longitudinal moni­
associated with an almost five-fold increase in the prevalence of CSVD toring of rarer genetic forms of CSVD.
(as defined by MRI evaluation) compared to NAION-free controls (Kim
et al., 2019a,b). Decreased retinal vasoreactivity determined via Dy­ 4.7. Frontotemporal Dementia
namic Vessel Analyzer was associated with impaired cerebral vaso­
reactivity (assessed via transcranial Doppler), a key marker of CSVD Frontotemporal Dementia (FTD) is a heterogeneous group of
(Bettermann et al., 2017). Central retinal artery blood flow (as neurodegenerative diseases characterized by neuronal and glial

23
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Fig. 9. Optical coherence tomography angiography (OCTA) images and quantification of retinal capillary changes in a subject with cognitive impairment and
control. The figure shows 3 × 3mm2 parafoveal OCTA images of two subjects with CDR-SOB scores = 0 and CDR-SOB = 1, and ages 76 and 73 years, respectively.
Neither subject has diabetes but both have hypertension. Images on the second and third columns show the corresponding skeletonized images and pseudo-colored
maps of the capillary density. The subject with CDR-SOB = 0 has more localized areas of higher capillary density than the age and medical condition similar subject
with CDR-SOB >0. (Reproduced under the Creative Commons Attribution-NonCommercial-NoDerivatives (CC BY-NC-ND) 4.0 International License from Ashimatey
et al., Alzheimer’s and Dementia: Diagnosis, Assessment and Disease Monitoring in press 2020).

proteinaceous inclusions preferentially involving the prefrontal and implications of these findings in future studies. The numerous references
anterior temporal lobes of the brain. The age of FTD onset is typically in this section are also summarized in Table 4.
younger than in LOAD (peaking in the 60’s) with an incidence ranging
from 5 to 22/100,000 person-years with a rough estimate of 4.7.1. Neurosensory retinal thickness in Frontotemporal Dementia
20,000–30,000 prevalent cases in the United States (Neary et al., 2005; Studies of the retina in FTD are very few. Ferrari et al. compared
Knopman and Roberts 2011). FTD is associated with two principal retinal thickness between 17 FTD patients and 49 healthy controls using
proteinopathies including microtubule-associated protein tau and OCT (Ferrari et al., 2017). This cross-sectional study showed significant
transactive response DNA-binding protein 43 (TDP-43) and this het­ pRNFL and GCIPL thinning in FTD relative to controls (Ferrari et al.,
erogeneity is captured with the neuropathological term “frontotemporal 2017). However, retinal thickness did not correlate with MMSE scores
degenerations” (Irwin et al., 2015). FTD and ALS occur on a spectrum in (Ferrari et al., 2017). Kim et al. measured macular thickness in 38 FTD
some families, most commonly in association with C9orf72 expansions. patients with probable tau proteinopathy and 44 healthy controls using
FTD is predominantly a behavioral disorder featuring significant OCT (Kim et al., 2017). After adjusting for age, race and sex, significant
changes in personality and social conduct. Notably, however, patients thinning of the outer retinal photoreceptor complex, most pronounced
may also present with language impairment, cognitive deficits in exec­ in the ONL and in the IS/OS junction (Ellipsoid Zone), was observed in
utive function, and motor symptoms (Neary et al., 2005). Therefore, subjects with probable tauopathy but not in those with other subtypes of
FTD is more commonly referred to as a group of 3 clinical syndromes: disease. Specifically, the ONL was about 10% thinner in FTD associated
behavioral variant, progressive non-fluent aphasia, and semantic de­ with tauopathy than in controls. In contrast, the inner retinal layers
mentia(Neary et al., 2005; Ferrari et al., 2017). These syndromes relate comprising the GCC were relatively unaffected. Outer retinal thickness
more to the areas of the brain involved rather than the subtype of protein also correlated significantly with MMSE scores in the whole cohort of
accumulation. The clinical presentation correlates poorly with the un­ FTD subjects (Kim et al., 2017).
derlying proteinopathy and clinical diagnosis remains especially chal­ A 16-month longitudinal study to further investigate disease pro­
lenging given the phenotypic variability of this disease (Neary et al., gression in the retina showed persistent outer retinal thinning in FTD
2005). Interestingly, a significant fraction of subjects with presumed without significant involvement of the inner retinal layers (Kim et al.,
FTD are later diagnosed with AD by neuropathology (Kertesz et al., 2019a,b). Outer retinal thickness similarly correlated with MMSE scores
2005; Knopman et al., 2005; Forman et al., 2006). While the specific prospectively, suggesting a possible relationship between the retina and
pathology of the frontotemporal degenerations has not been histopath­ the severity of disease. The investigators hypothesize that since
ologically demonstrated in the human retina, abnormal accumulation of microtubule-associated tau is found in the retina as a photoreceptor
tau has been demonstrated in the aging retina (Leger et al., 2011). ciliary protein, a tau-based mechanism for outer retinal atrophy may be
Therefore, non-invasive retinal imaging may provide a potentially useful responsible for the observations (Kim et al., 2019a,b). Importantly, Kim
biomarker for this disease though studies of retinal changes in FTD must et al. also performed objective CSF testing to exclude participants with
take into account the considerable genetic and pathological heteroge­ confounding AD (Kim et al., 2017). Additional morphologic analyses of
neity in this group of disorders. Below we summarize and synthesize a postmortem eyes correlated with retinal changes in well-characterized,
large body of literature describing retinal imaging findings in FTD and homogeneous FTD cohorts are necessary to improve our understanding

24
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

of retinal markers for this spectrum of disease. resolution often varies inversely with field-of-view and therefore subtle
findings (e.g. capillary level findings) may be difficulty to resolve in
5. Future directions and conclusions larger images. While the wide-field capabilities of these devices are just
beginning to be explored it is very likely that technological development
We have mentioned at several points throughout this article that and clinical demand will push the development of wide-field OCT and
there is a significant need for longitudinal studies to address the cau­ OCTA in the next several years. Implementation of these technologies in
sality and temporal relationship of retinal imaging findings in the dis­ studies of retinal changes and neurodegenerative disease will provide
eases we have reviewed. Beyond this there are a number of limitations data on aspects of retinal anatomy that were not previously available. By
that are associated with retinal imaging methodologies and studies that imaging more of the peripheral retina, future studies may better explain
are worth further addressing because they significantly impact the the differences between histopathologic studies and in vivo imaging
conclusions that may be derived from the retina and retinal imaging studies.
about neurological processes. In addition, these limitations provide
significant opportunity for methodological development in future
studies that we will discuss in the next section. For the purposes of this 5.2. Image registration
article and discussion we categorize these limitations and future di­
rections in the following groups: field-of-view size, image registration, 2- The detection of subtle, subclinical changes in retinal thickness,
dimensional versus 3-dimensional analyses, imaging artifacts, and study capillary density and especially sublayer thickness requires precise
design. We believe that one or more of these categories underly the often measurements that are reliable between and among subjects. For
contradictory or inconsistent findings in the literature. example, at least one cross-sectional study that did not find significant
differences in average RNFL or GCIPL of subjects with LOAD (or MCI)
compared to controls, did report focal regions of RNFL thickening
5.1. Wide-field retinal imaging immediately adjacent to areas of thinning (Lad et al., 2018). This sug­
gests that quantifying retinal thickness or sublayer thickness across
In general, most studies of retinal findings in neurodegenerative relatively large regions of retina may obscure subtle changes that cancel
disease are limited by the field-of-view of the imaging modality whether each other out. This problem is also encountered in the analysis of
that is fundus photography or OCT based imaging. In the vast majority of retinal images from subjects with retinal disease (e.g. macular edema or
cases the field-of-view is limited to the macula or optic disc. Specifically, retinal atrophy). Therefore, one important consideration in the analysis
for OCT imaging the central 6 × 6mm or even 3 × 3mm parafoveal of retinal imaging data is image registration. Image registration is crit­
region of the macula is the most commonly imaged region of the retina ical to ensure comparison of analogous retinal regions of the same
and for fundus photographs the 30–50◦ (approximately 10–15 mm; subject from different time points in longitudinal studies or analogous
Fig. 1A) immediately adjacent to and temporal to the optic nerve head is retinal regions from different subjects in cross-sectional studies. This is
the most commonly imaged. It is important to note that by virtue of the particularly relevant in OCT studies of retinal thickness but also appli­
study design and imaging device, large areas of the peripheral retina are cable to other imaging modalities. Because accurate image registration
excluded from analysis in these studies (compare Figs. 1A and 2A). is difficult, almost all the studies we describe in this review report
While it is not clear exactly how important the peripheral retina is average retinal thickness values across large regions (often pre­
compared to the central retina, it is worth keeping in mind the sampling determined by commercially available software) of the OCT scan. Image
bias introduced by this limitation. For example, some histopathological registration between or among subjects is generally not considered at
studies suggest that retinal amyloid deposition occurs primarily in the all. This is a significant area of improvement which is particularly
peripheral (extramacular) retina (Koronyo-Hamaoui et al., 2011; Kor­ applicable to longitudinal studies of the retina in subjects with neuro­
onyo et al., 2017) and would not be detected by studies that are only degenerative disease especially since effect sizes are usually small and
assessing central macular images. may be masked by slight changes in image sampling area.
The adoption of wide-field imaging methods that we discuss in sec­ Image registration is a computational technique for finding one-to-
tion 5.1.1 will change this bias in the future and is particularly relevant one correspondence between pairs of images and useful for image
to studies of neurodegenerative disease. Wide-field color fundus imag­ analysis techniques including quantitative measures of retinal changes
ing is rapidly becoming the standard-of-care throughout ophthalmology (thickness, capillary density), image fusion or stitching. Multiple image
practices and will grow in popularity in research settings as well (Pan­ registration techniques have been described in the field of medical
war et al., 2016). Wide-field OCT and OCTA are also a subject of intense image processing (Toga and Thompson 2001; Murphy et al., 2008; Teng
investigation and technological development (Eastline et al., 2019) et al., 2010; Gavaghan et al., 2011; Oliveira and Tavares 2014). These
(Fig. 2). One important point to keep in mind in evaluating data from techniques have been used for registration of OCT and OCTA images to
different field-of-view OCT and OCTA images is that the image evaluate treatment efficiency (Lee et al., 2015) understand retinal

Table 4
Summary of retinal imaging studies characterizing human subjects with frontotemporal dementia.
Finding FTD vs Controls No Difference Biomarker confirmed (tauopathy, TAR DNA–binding Complete eye exam* excluding
protein 43) confounding disease

OCT
pRNFL thinning Ferrari et al.
(2017)
mRNFL thinning Kim et al., (2017); Kim et al., Kim et al., (2017); Kim et al., 2019a,b
2019a,b
RGC-IPL thinning Kim et al., (2017); Kim et al., Kim et al., (2017); Kim et al., 2019a,b
2019a,b
Total macular Kim et al., (2017); Kim et al., Kim et al., (2017); Kim et al., 2019a,b
thinning 2019a,b

Abbreviations: = Optical Coherence Tomography, FTD = Frontotemporal Dementia, pRNFL = peripapillary retinal nerve fiber layer, mRNFL = macular retinal nerve
fiber layer, RGC = Retinal Ganglion Cells, IPL = Inner Plexiform Layer, Blank cells indicate no data or study was available for that parameters. (*) Complete eye exam
includes visual acuity, intraocular pressure measurement, and dilated fundus examination.

25
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

diseases (Chen et al., 2014; Lee et al., 2015; Antony et al., 2016), correct 5.3. 2D vs 3D OCTA analysis
motion artifacts (Kraus et al., 2012) and assist with layer segmentation
(Duan et al., 2018). The majority of these methods have been performed Another major limitation of retinal imaging that is particularly
on 2D images because of the technical challenges and computation cost relevant in studying neurodegenerative disease is that most data (fundus
of 3D image registration. However, the latter is more likely to provide photographs) are available and analyzed only in 2-dimensional (2D)
anatomically accurate information for both qualitative and quantitative format. However, OCT and OCTA data that are available in 3D volumes
analyses. More importantly, it is possible and likely that 3D based are still also largely analyzed in 2D formats. This is partly because of the
measurements would capture more subtle retinal changes that are so ease with which 2D analyses can be performed. However, 2D repre­
prevalently (but inconsistently) reported in the neurodegeneration sentations (e.g. en face images) of 3D retinal structures are inherently
literature. Furthermore, many of the layer specific findings (or lack limited because they confound vascular and tissue relationships that are
thereof) among current studies would benefit from a method that likely to be important in detecting pathological changes, especially
allowed qualitative and quantitative evaluation of retinal sections subtle ones that are characteristic of neurodegenerative diseases. Un­
independently of layer segmentation. fortunately, there are significant technical challenges associated with
Techniques for within and between subject image registration have rendering and analyzing 3D data. For example, OCTA data are often
been proposed previously. Within subject registration of OCTA image compromised by high noise levels, vessel discontinuity and low vessel
volumes has been reported to align vascular pattern of repeated scans visibility that mainly affects the appearance of small capillaries. Addi­
(Zhang et al., 2019). However, between-subject OCTA registration is not tionally, OCTA imaging artifacts such as motion and projection of large
feasible due to lack of features that can be reliably used for registration. vessels and challenges in image registration negatively impact the true
In fact, vascular patterns in OCTA are the dominant features and are not visualization and geometry of the retinal vessel network (Spaide et al.,
expected to match across subjects. However, since OCTA is generated 2015a,b,c; Holmen et al., 2020). To enable true 3D OCTA analyses,
from more than one OCT image, the same registration for OCT volumes multi-scale and multi-orientation curvilinear enhancement techniques
is applicable to OCTA images. In OCT, retinal layer boundaries and need to be developed to separate blood flow from static tissues and
foveal pit are considered as distinctive features across subjects (Khansari obtain a high-quality vessel and tissue maps. Recent studies have
et al., 2020). These features can be used for reliable registration of be­ demonstrated the potential advantages of 3D OCTA analysis in retinal
tween and within subjects OCT image volume. Several methods have images. For example, several studies have (Chen et al., 2016; Spaide
been developed for cross-subject OCT registration in 2D (Gibson et al., 2016; Spaide et al., 2017) shown that the correlation of intraretinal fluid
2010; Chen et al., 2014; Lee et al., 2015). and retinal capillaries can be represented effectively by volume
Khansari et al. developed an automated 3D registration technique for rendering analysis of 3D-OCTA data. OCTA 3D vessel density can better
cross-subject registration of OCT image volumes in subjects with retinal quantify foveal ischemia in diabetic retinopathy compared to the 2D
vascular disease (Khansari et al., 2020). This technique consisted of an vessel density (Wang et al., 2019).
initial restricted affine transformation to define anatomically consistent To reduce speckle noise in 2D-OCT images, numerous algorithmic
volumes of interest. Affine transformation is similar to rigid trans­ approaches (Adler et al., 2004; Jian et al., 2009; Wong et al., 2010; Fang
formation which also corrects shear and scale between pairs of images. et al., 2012; Mayer et al., 2012; Xu et al., 2012; Boroomand et al., 2013;
Afterwards, an efficient B-spline transformation using stochastic Cameron et al., 2013; Luan and Wu 2013) have been developed. Since
gradient descent is performed to align layers boundaries and foveal pit OCTA is constructed by subtracting several OCT images, some of the
within the VOI. B-spline finds local deformation between pair of images proposed techniques for OCT speckle noise reduction that can incorpo­
and is particularly powerful for 3D registration. The authors showed rate the retinal microvascular structures of various size and orientation
high accuracy of the registration in terms of aligning eight different into the denoising framework are also applicable to OCTA. Among the
retinal layer boundaries in healthy and subjects with diabetic retinop­ exiting techniques, the 3D curvelet transform has been applied suc­
athy complications. This method is particularly unique because the cessfully to both structural OCT (Jian et al., 2010) and OCTA (Shi et al.,
determinant of Jacobian matrix which is partial derivative of the 2017), but it may be more appropriate for OCTA vessel enhancement.
vector-field at each voxel for the non-linear deformation was used to Curvelet is a model-based filtering approach with high directional
visualize tissue expansion and contraction independently of layer seg­ sensitivity and anisotropy characteristics and it can provide an efficient
mentation (Fig. 10). Additionally, tensor-based morphometry was per­ representation of edges and other singularities along curves. Compared
formed for detection of group-wised localized structural changes in to conventional filtering methods, curvelet can avoid excessive denois­
different stages of DR. Tensor based morphometry measures significance ing and preserve capillary details in OCTA (Fig. 11). A novel 3D shape
of local structural differences between groups of images based on voxel modeling framework was recently developed based on curvelet
level statistics. In Khansari et al., 2020, for each voxel, t-test was per­ modeling to produce a novel and high-quality 3D OCTA microvascular
formed on Jacobian values between healthy and diabetic subjects to visualization which had not been previously demonstrated and might
generate a p-value map. Voxels in the p-value map with values less than provide a unique tool for clinical and scientific analysis (Zhang et al.,
a threshold (i.e. p < 0.05) were considered to be significantly deformed. 2019). This framework includes a surface reconstruction process that is
Fig. 10 shows example non-linear registration of OCT of subjects with applied to transform 3D discrete OCTA volume representations to 3D
diabetic retinopathy to a normative image which is generated by continuous triangular surface representations (Fig. 12). The algorithmic
registering and averaging multiple healthy OCTs. Color-coded Jacobian advances would make 3D tissue analyses much more feasible and
maps show the magnitude of local retinal expansion and constriction potentially enable more accurate OCT and OCTA based metrics.
independently of retinal layer segmentation. Once OCT of different Hessian-based methods (Frangi et al., 1998; Sato et al., 1998) have
subjects are registered, the transformation can be applied to the corre­ been also widely used to identify and enhance tubular structures in both
sponding OCTA image volume of each subjects. This brings OCTA im­ 2D and 3D medical images. The Frangi filter (Frangi et al., 1998) is one
ages into a common space allowing meaningful layer-based analysis of of the most common multiscale hessian-based methods that has been
vascular morphology between and among subjects (Sarabi et al., 2019). applied to OCTA enface images (Yousefi et al., 2015). To construct the
Future applications of this method can be performed in a layer seg­ vesselness measure, Frangi uses the eigenvalues ratio of the Hessian
mentation independent manner and will likely be more sensitive for matrix (second gradient). The tubular vessel shape is usually repre­
detecting the subclinical retinal changes described in various neurode­ sented by the two large eigen values and one small eigen value. How­
generative diseases. ever, due to the OCTA projection artifact, the 3D deformed geometry of
the large vessels won’t be properly detected by the default Frangi ves­
selness measure. Recently, researchers proposed a projection resolved

26
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Fig. 10. Representative 3D registration and


Jacobian maps of OCT image volumes in two
subjects with diabetic retinopathy using
non-linear registration. Images are a cut
through the volume. (A) OCT volume of the
one subject with visible diabetic macular
edema and alterations in the shape of the
foveal pit. (B) Cut through the normative
atlas image demonstrating retinal structure
of a healthy subject. (C) Color-coded Jaco­
bian map demonstrating magnitude of
localized contraction and expansion in
diseased subject versus healthy subject in­
dependent of layer segmentation. (D) OCT
volume of another subject with visible tissue
loss. (E) Cut through the atlas image which
represents the retinal structure of a healthy
subject. (F) Color-coded Jacobian map
demonstrating magnitude of localized
contraction and expansion in diseased sub­
ject versus healthy subject independent of
layer segmentation. Red indicates areas of
expansion (edema). Blue indicates areas of
contraction (tissue loss).

technique based on the 3D Frangi vesselness measure for suppressing the the OCTA binary vessel segmentation that is obtained by automatic
tail artifact of large vessels in OCTA (Liu et al., 2019a,b). As part of their thresholding on OOF vesselness map using Otsu’s global thresholding
method, the authors suggested the plateness measure by modifying the (Otsu 1979). The above 3D-OCTA preprocessing framework has been
3D Frangi vesselness to measure the probability that a structure belongs employed to develop 3D-OCTA microvascular shape analysis (Zhang
to elongated vessels. In addition to the previously mentioned tech­ et al., 2019) and study the morphological and topological vessel changes
niques, other filter-based and hybrid methods have been also developed in subjects with diabetic retinopathy (Fig. 12). Future applications of
for enhancement of the OCTA vessels. Several groups presented an these methods to other data sets from subjects with neurodegeneration
automatic algorithm by combining a joint Markov-Gibbs model, a Naïve may yield novel insights into neurosensory and retinal vascular changes
Bayes (NB) classifier, and a 2D connectivity filter for segmentation of 2D in those diseases without the confounding effects of layer segmentation
superficial and deep retinal maps of normal and diabetic eyes (Chu et al., and 2D analyses (Fig. 11).
2016; Eladawi et al., 2017; Chlebiej et al., 2019). Others have developed
an OCT amplitude-decorrelation algorithm to enhance the SNR of flow
5.4. Study design
detection and microvascular network connectivity in cross-sectional
images (Jia et al., 2012).
One of the most significant limitations of studies that attempt to
Recently, an effective OCTA 3D vessel enhancement module has also
assess any retinal feature in neurodegenerative disease is the study
been demonstrated by combining curvelet denoising (Shi et al., 2017)
design. It is critical to state clearly from the outset of any study what the
and optimally orientated flux (OOF) (Law and Chung 2008; Sarabi et al.,
purpose of the proposed retinal findings are intended to be as this will
2019). OOF is a localized multi-scale curvilinear structure detector that
guide the essential details of the study design. For example, retinal
computes the vesselness measure based on the projected image gradient
findings may serve in early detection of subclinical disease, as bio­
at the boundary of a spherical region centered at every image voxel. The
markers of disease progression/regression, to characterize in vivo disease
main advantage of OOF is providing robust vessel detection response in
pathophysiology, or may be only associated clinical findings. Some
presence of closely located structures which makes it an ideal tool for
studies may aim to impact clinical care or enable drug development
vessel enhancement in OCTA dense microvascular networks. Fig. 11
while other studies may aim to make contributions to our understanding
demonstrates pre-processing results of a normal subject using the com­
of disease pathophysiology. In any case, given the small effect sizes that
bined curvelet and OOF approach. The results show that the combina­
are anticipated in most retinal imaging studies it is critical to power
tion of curvelet and OOF techniques resolved both the vessel
studies appropriately from the outset.
discontinuity and noise adjacent to tiny vessels while preserving the
Because of the optical accessibility of the retinal tissue and the high
microvasculature geometry in 3D-OCTA. In addition, Fig. 11 illustrates
resolution imaging methods that are available, it would seem logical to

Fig. 11. New Fig 12 Demonstration of the 3D OCTA surface reconstruction and the subsequent shape and Reeb analysis that may be used for future quantitative
analyses of retinal vessels. (A) The original 3D OCTA volume is processed to reconstruct (B) the 3D vessels surface representation. Using (B), we can perform intrinsic
shape analysis for (C) large and small vessels classification, and (D) Reeb graph analysis to quantify valuable 3D vessel geometry and topology.

27
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Fig. 12. New Fig 11 Example of OCTA images from a normal subject before and after preprocessing algorithm for image registration and optimization for further
analysis. (A) Original 3D-OCTA image volume. (B) Enhanced OCTA volume obtained by applying 3D curvelet denoising on previous panel. (C) OCTA vesselness map
generated by computing the optimally oriented flux response of previous panel. (D) OCTA binary vessel mask obtained by applying Otsu’s global thresholding on
previous panel. (E–H) Selected enface views of images in above panels.

target the use of retinal findings as risk factors for future development of axial length of the eye that occurs in myopia induces magnification error
disease. In this context, relatively low cost and rapid retinal screening in many types of retinal images that require careful quantification and
tests can be used to guide more resource intensive and likely specific correction (Bennett et al., 1994; Lee et al., 2018; Llanas et al., 2019). In
screening efforts (such as PET or CSF studies). The relatively high spatial addition, myopia is associated with pathologic thinning of the choroid
resolution of retinal imaging modalities may also be useful to assess for which may impact studies assessing choroidal thickness changes in
subclinical disease onset and subclinical disease progression. In this neurodegenerative disease (Ang et al., 2019).
context, a retinal biomarker of early disease or risk must be validated Glaucoma is another ocular disease characterized by subtle, but
using objective measures of the underlying neurodegenerative disease significant, thinning of the RNFL (Weinreb et al., 2014) and attenuation
status (e.g. PET, CSF and blood) because clinical findings can be sub­ of the retinal peripapillary capillaries (Zabel et al., 2019). Glaucoma is
jective, subtle or altogether absent. In many cases, these objective highly prevalent especially in older demographics and among Asian and
measures are not used because of ethical, financial or logistical reasons. African American races (Tham et al., 2014). Glaucoma is also highly
For example the availability of PET scanners is relatively limited and underdiagnosed (Quigley 2011; Tham et al., 2014) because it is largely
lumbar puncture may seem like a relatively risky proposition for an asymptomatic disease until the disease is very advanced. By some
otherwise asymptomatic subjects. Nevertheless, studies should strive to estimates, more than 110 million people worldwide will have some form
use objective diagnostic criteria from blood, brain imaging, CSF, or ge­ of glaucoma by 2040. Most importantly, many of these people may not
netic testing to ensure the most accurate associations with retinal find­ be diagnosed with or know they have the disease for many years.
ings. In addition, there are few or no studies that assess retinal findings Therefore, some form of glaucoma screening is essential in studies un­
across a sufficiently large sample and sufficiently broad spectrum of dertaking any evaluation of retinal thickness in subjects with possible or
disease severity to rigorously validate the association of the retinal actual neurodegenerative disease.
findings with the underlying neurodegenerative disease. In 2014 an estimated 422 million people worldwide had diabetes
Another major methodological limitation of many studies is the mellitus (WHO Fact Sheet; https://www.who.int/news-room/fact-shee
challenge of performing comprehensive ophthalmologic evaluations to ts/detail/diabetes). Diabetic retinopathy is estimated to effect 29% of
establish clear inclusion/exclusion criteria. There are confounding and adults in the United States with diabetes (Zhang et al., 2010). Diabetic
highly prevalent ophthalmic disorders that can significantly impact retinopathy can cause many retinal manifestations including retinal
retinal measures of thickness, perfusion and neurosensory function. thickening, retinal thinning and attenuation of retinal capillary density.
Appropriately anticipating, quantifying and controlling for these con­ As with glaucoma, diabetic retinopathy can be essentially asymptomatic
founding variables is possible and requires close collaboration between for many years or even decades. Therefore, it is critical to assess for the
ophthalmologists and neurologists. While an extensive discussion of presence of diabetes mellitus in study subjects as well as some assess­
each of these diseases is out of the scope of this review article we review ment of diabetic retinopathy severity. It is equally as important to assess
the pertinent findings and rationale for carefully considering some of the the duration and control of diabetes mellitus.
most prevalent comorbid ocular disease (diabetic retinopathy, glaucoma
and myopia) in the exclusion and inclusion criteria of future studies
below. There is also the possibility that these disease processes are not 5.5. Retinal vascular reactivity
confounding. It has been suggested that there are common underlying
pathophysiological mechanisms between what are considered primarily Many of the studies of retinal vascular changes in neurodegenerative
retinal diseases and neurodegenerative diseases (Lynch and Abràmoff disease focus on the static aspects of the retinal vasculature (e.g. capil­
2017; Sen et al., 2020a,b). lary density, vessel tortuosity, retinal hemorrhages etc). While these are
Myopia was estimated to impact 23% (1.4 billion people) of the interesting metrics of vascular structure, they are only indirect measures
world population in 2000 and is expected to increase in prevalence to of vascular function, if any measure of function at all. Improvements and
50% (4.7 billion people) by 2050 (Holden et al., 2016). Myopia can be advances in retinal imaging now allow measurement of capillary level
progressive in nature with age or relatively stationary. The increased changes associated with physiologic stimuli For example, OCTA based
measurements of retinal capillary blood flow and/or caliber can be

28
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

reliably made during physiologic fluctuations in inhaled oxygen and disease (i.e. CSF and PET amyloid etc).
carbon dioxide (Ashimatey et al., 2019; Kushner-Lenhoff et al., 2020). Retinal imaging presents a particularly unique opportunity in at least
This method has been employed to demonstrate preclinical changes in two largely unexplored fronts, Autosomal Dominant Alzheimer’s Dis­
retinal vascular reactivity in subjects with diabetes even before the onset ease and Cerebral Small Vessel Disease. The confounding role of
of clinically evidence retinopathy (Ashimatey et al., 2019; Singer et al., Vascular Cognitive Impairment and Dementia in assessing neurode­
2020). Similar physiologic changes have been measured using non-FDA generative changes in the aged population is particularly challenging to
approved methods such as AO-SLO (Duan, Bedggood et al. 2016, 2017). address. Ongoing prospective studies (e.g. MarkVCID) using optical
These methods will allow quantification of retinal vascular function coherence tomography angiography have particular promise in detect­
which has not been possible in the past. This may be advantageous for ing and quantifying retinal vascular changes at the capillary level that
several reasons including having larger and reversible effect sizes that may be indicative of similar changes in the brain vasculature. Overall,
can be more reliably measured. In addition, changes in retinal vascular the potential of retinal imaging in understanding of neurodegenerative
function likely precede changes in retinal vascular structure. Future disease and in its clinical management is very promising but remains to
application of these methods in cross-sectional and longitudinal studies be fully realized. Novel imaging methods (wide-field imaging, optical
will provide novel and potentially interesting insights into the patho­ coherence tomography angiography and hyperspectral imaging) and
physiology of neurodegenerative diseases, especially cerebrovascular analysis algorithms (3D versus 2D) will provide more opportunities to
disease and Alzheimer’s Disease. When coupled with measures of retinal apply retinal findings to neurodegenerative disease with very exciting
function, these studies also have the potential to provide new insight implications scientifically and clinically.
into neurovascular coupling.
Author contributions
6. Conclusions
Amir H Kashani 35% (conceptualization, methodology, data cura­
The notion that neurodegenerative changes in the brain are present tion, project administration, resources, supervision, original draft
in the retina is supported by extensive evidence from in vivo retinal writing, review and editing).
imaging studies as well as evidence from histopathology specimens. This Samuel Asanad 15% (data curation, original draft writing, review
evidence is particularly strong and well-established from histopatho­ and editing).
logic and cross-sectional studies of subjects with Multiple Sclerosis and John Ringman 5% (original draft writing, review and editing).
Alzheimer’s Disease. These studies show significant changes in inner Alessandro Biffi, 5% (original draft writing, review and editing).
retinal layers and most profoundly in the retinal nerve fiber layer. Yonggang Shi, 5% (review, supervision and editing).
Interestingly, layer specific findings involving the ganglion cell layer, Helena Chui, 5% (review, supervision and editing).
inner nuclear layer and photoreceptor layers have been demonstrated in Jane W. Chan 5% (original draft writing, review and editing).
different neurodegenerative diseases. This suggests that clinical stages of Jiong Zhang 5% (original draft writing, review and editing).
these chronic diseases may manifest in the retina in differing, and often Mona Sharifi 5% (original draft writing, review and editing).
very subtle ways. For example, the findings that a subtle increase in Maziar Khansari 5% (original draft writing, review and editing).
inner nuclear layer thickness correlates with MRI activity in Multiple Maxwell Singer 5% (original draft writing, review and editing).
Sclerosis suggests that inflammation attributable to the underlying dis­ Farzan Abolahi 5% (original draft writing, review and editing).
ease process is present in the retina. This hypothesis seems to be sup­
ported by histopathologic evidence of inflammatory cells in the inner Funding sources
retina of subjects with Multiple Sclerosis. Advances in OCT analysis al­
gorithms using 3D metrics may provide additional insight into these This work was supported by the National Eye Institute (USA,
subtle layer specific changes. More recent findings of amyloid deposition K08EY027006 to AHK) and National Institute for Neurological Diseases
in the neurosensory retina and retinal vasculature suggest that the am­ (USA, UH3NS100614 to AHK), an unrestricted grant from Research to
yloid pathology that correlates with Alzheimer’s Disease status is also Prevent Blindness (USA) to the USC Department of Ophthalmology,
present in the retina. Use of wide-field and hyperspectral imaging Brightfocus Foundation (USA, CA2020004 to AHK)), National Institutes
methods presents a novel and exciting opportunity in detecting retinal of Health (USA, R01AG062007 and P30AG066530 to JMR).
amyloid lesions in vivo. Advances in OCT metrics and hyperspectral
imaging are also potentially applicable to Parkinson’s Disease and
Declaration of competing interest
Huntington’s Disease where cross-sectional and histopathological evi­
dence for retinal involvement also exists.
The authors have no competing interests to declare. AHK received
Unfortunately, the clinical utility of retinal findings in neurodegen­
research funding and honoraria from Carl Zeiss Meditec (Dublin, CA,
erative disease has been limited because, in most cases, the cross-
USA). Funding sources did not have any role in the conception, writing
sectional data and associations are from subjects with relatively
or editing of this manuscript.
advanced disease. Multiple Sclerosis is one exception where optic nerve
findings and OCT retinal thickness measures have demonstrated clinical
References
utility because the initial manifestations of the disease can be primarily
ocular. With the growing evidence that subtle retinal findings are Aaker, G.D., Myung, J.S., Ehrlich, J.R., Mohammed, M., Henchcliffe, C., Kiss, S., 2010.
detectable during the presymptomatic stages of Alzheimer’s Disease (as Detection of retinal changes in Parkinson’s disease with spectral-domain optical
well as other neurodegenerative diseases), reliable and more sensitive coherence tomography. Clin. Ophthalmol. 4, 1427–1432.
Adam, C.R., Shrier, E., Ding, Y., Glazman, S., Bodis-Wollner, I., 2013. Correlation of
retinal imaging methods would present a significant clinical benefit in inner retinal thickness evaluated by spectral-domain optical coherence tomography
detection, monitoring and possibly even therapeutic development. At and contrast sensitivity in Parkinson disease. J. Neuro Ophthalmol. 33 (2), 137–142.
the least, there should be awareness among ophthalmologists, neurol­ Adler, D.C., Ko, T.H., Fujimoto, J.G., 2004. Speckle reduction in optical coherence
tomography images by use of a spatially adaptive wavelet filter. Opt. Lett. 29 (24),
ogists and clinician-scientists about the growing number of clinical trials
2878–2880.
that are recruiting subjects at risk for neurodegenerative disease as well Ahn, J., Lee, J.Y., Kim, T.W., Yoon, E.J., Oh, S., Kim, Y.K., Kim, J.M., Woo, S.J., Kim, K.
as the potential for ocular manifestations of these diseases. W., Jeon, B., 2018. Retinal thinning associates with nigral dopaminergic loss in de
In addition, large scale and prospective studies are also needed to novo Parkinson disease. Neurology 91 (11), e1003–e1012.
Alber, J., Goldfarb, D., Thompson, L.I., Arthur, E., Hernandez, K., Cheng, D., DeBuc, D.C.,
assess the temporal relationship between retinal imaging findings and Cordeiro, F., Provetti-Cunha, L., den Haan, J., Van Stavern, G.P., Salloway, S.P.,
known biomarkers associated with preclinical neurodegenerative Sinoff, S., Snyder, P.J., 2020. Developing retinal biomarkers for the earliest stages of

29
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Alzheimer’s disease: What we know, what we don’t, and how to move forward. Group, 2010b. Differential associations of cortical and subcortical cerebral atrophy
Alzheimers Dement. 16 (1), 229–243. with retinal vascular signs in patients with acute stroke. Stroke 41 (10), 2143–2150.
Albrecht, P., Müller, A.K., Südmeyer, M., Ferrea, S., Ringelstein, M., Cohn, E., Aktas, O., Bassi, C.J., Sadun, A.A., 1990. Alzheimer’s disease and vision: correlation versus
Dietlein, T., Lappas, A., Foerster, A., Hartung, H.P., Schnitzler, A., Methner, A., 2012. causation, revisited. Ophthalmology 97 (4), 395–397.
Optical coherence tomography in parkinsonian syndromes. PloS One 7 (4), e34891. Bateman, R.J., Xiong, C., Benzinger, T.L., Fagan, A.M., Goate, A., Fox, N.C., Marcus, D.S.,
Alexandrov, P.N., Pogue, A., Bhattacharjee, S., Lukiw, W.J., 2011. Retinal amyloid Cairns, N.J., Xie, X., Blazey, T.M., Holtzman, D.M., Santacruz, A., Buckles, V.,
peptides and complement factor H in transgenic models of Alzheimer’s disease. Oliver, A., Moulder, K., Aisen, P.S., Ghetti, B., Klunk, W.E., McDade, E., Martins, R.
Neuroreport 22 (12), 623–627. N., Masters, C.L., Mayeux, R., Ringman, J.M., Rossor, M.N., Schofield, P.R.,
Alten, F., Motte, J., Ewering, C., Osada, N., Clemens, C.R., Kadas, E.M., Eter, N., Paul, F., Sperling, R.A., Salloway, S., Morris, J.C., Network, D.I.A., 2012. Clinical and
Marziniak, M., 2014. Multimodal retinal vessel analysis in CADASIL patients. PloS biomarker changes in dominantly inherited Alzheimer’s disease. N. Engl. J. Med.
One 9 (11), e112311. 367 (9), 795–804.
Altintaş, O., Işeri, P., Ozkan, B., Cağlar, Y., 2008. Correlation between retinal Bayhan, H.A., Aslan Bayhan, S., Celikbilek, A., Tanık, N., Gürdal, C., 2015. Evaluation of
morphological and functional findings and clinical severity in Parkinson’s disease. the chorioretinal thickness changes in Alzheimer’s disease using spectral-domain
Doc. Ophthalmol. 116 (2), 137–146. optical coherence tomography. Clin. Exp. Ophthalmol. 43 (2), 145–151.
Alves, C., Jorge, L., Canário, N., Santiago, B., Santana, I., Castelhano, J., Ambrósio, A.F., Bayhan, H.A., Aslan Bayhan, S., Tanık, N., Gürdal, C., 2014. The association of spectral-
Bernardes, R., Castelo-Branco, M., 2019. Interplay between macular retinal changes domain optical coherence tomography determined ganglion cell complex parameters
and white matter integrity in early Alzheimer’s disease. J. Alzheimers Dis. 70 (3), and disease severity in Parkinson’s disease. Curr. Eye Res. 39 (11), 1117–1122.
723–732. Beach, T.G., Adler, C.H., Sue, L.I., Vedders, L., Lue, L., White Iii, C.L., Akiyama, H.,
Anderson, D.H., Talaga, K.C., Rivest, A.J., Barron, E., Hageman, G.S., Johnson, L.V., Caviness, J.N., Shill, H.A., Sabbagh, M.N., Walker, D.G., Consortium, A.P.s.D., 2010.
2004. Characterization of beta amyloid assemblies in drusen: the deposits associated Multi-organ distribution of phosphorylated alpha-synuclein histopathology in
with aging and age-related macular degeneration. Exp. Eye Res. 78 (2), 243–256. subjects with Lewy body disorders. Acta Neuropathol. 119 (6), 689–702.
Alzheimer’s Association, 2020. 2020 Alzheimer’s disease facts and figures. Alzheimers Beach, T.G., Carew, J., Serrano, G., Adler, C.H., Shill, H.A., Sue, L.I., Sabbagh, M.N.,
Dement. https://www.alz.org/media/Documents/alzheimers-facts-and-figures.pdf. Akiyama, H., Cuenca, N., Consortium, A.P.s.D., 2014. Phosphorylated α-synuclein-
(Accessed 16 January 2021). immunoreactive retinal neuronal elements in Parkinson’s disease subjects. Neurosci.
Anderson, T.J., MacAskill, M.R., 2013. Eye movements in patients with Lett. 571, 34–38.
neurodegenerative disorders. Nat. Rev. Neurol. 9 (2), 74–85. Bennett, A.G., Rudnicka, A.R., Edgar, D.F., 1994. Improvements on Littmann’s method of
Andrade, C., Beato, J., Monteiro, A., Costa, A., Penas, S., Guimarães, J., Reis, F.F., determining the size of retinal features by fundus photography. Graefes Arch. Clin.
Garrett, C., 2016. Spectral-domain optical coherence tomography as a potential Exp. Ophthalmol. 232 (6), 361–367.
biomarker in Huntington’s disease. Mov. Disord. 31 (3), 377–383. Bennett, J.L., de Seze, J., Lana-Peixoto, M., Palace, J., Waldman, A., Schippling, S.,
Andrew, S.E., Goldberg, Y.P., Kremer, B., Telenius, H., Theilmann, J., Adam, S., Starr, E., Tenembaum, S., Banwell, B., Greenberg, B., Levy, M., Fujihara, K., Chan, K.H.,
Squitieri, F., Lin, B., Kalchman, M.A., 1993. The relationship between trinucleotide Kim, H.J., Asgari, N., Sato, D.K., Saiz, A., Wuerfel, J., Zimmermann, H., Green, A.,
(CAG) repeat length and clinical features of Huntington’s disease. Nat. Genet. 4 (4), Villoslada, P., Paul, F., Gjcf-Icc&Br, 2015. Neuromyelitis optica and multiple
398–403. sclerosis: seeing differences through optical coherence tomography. Mult. Scler. 21
Ang, M., Wong, C.W., Hoang, Q.V., Cheung, G.C.M., Lee, S.Y., Chia, A., Saw, S.M., Ohno- (6), 678–688.
Matsui, K., Schmetterer, L., 2019. Imaging in myopia: potential biomarkers, current Berisha, F., Feke, G.T., Trempe, C.L., McMeel, J.W., Schepens, C.L., 2007. Retinal
challenges and future developments. Br. J. Ophthalmol. 103 (6), 855–862. abnormalities in early Alzheimer’s disease. Invest. Ophthalmol. Vis. Sci. 48 (5),
Antony, B.J., Chen, M., Carass, A., Jedynak, B.M., Al-Louzi, O., Solomon, S.D., Saidha, S., 2285–2289.
Calabresi, P.A., Prince, J.L., 2016. Voxel based morphometry in optical coherence Bettermann, K., Slocomb, J., Shivkumar, V., Quillen, D., Gardner, T.W., Lott, M.E., 2017.
tomography: validation & core findings. Proc. SPIE-Int. Soc. Opt. Eng. 9788. Impaired retinal vasoreactivity: an early marker of stroke risk in diabetes.
Archibald, N.K., Clarke, M.P., Mosimann, U.P., Burn, D.J., 2009. The retina in J. Neuroimaging 27 (1), 78–84.
Parkinson’s disease. Brain 132 (Pt 5), 1128–1145. Biffi, A., Greenberg, S.M., 2011. Cerebral amyloid angiopathy: a systematic review.
Archibald, N.K., Clarke, M.P., Mosimann, U.P., Burn, D.J., 2011. Retinal thickness in J. Clin. Neurol. 7 (1), 1–9.
Parkinson’s disease. Park. Relat. Disord. 17 (6), 431–436. Bittersohl, D., Stemplewitz, B., Keserü, M., Buhmann, C., Richard, G., Hassenstein, A.,
Arevalo-Rodriguez, I., Smailagic, N., Roqué I Figuls, M., Ciapponi, A., Sanchez-Perez, E., 2015 Nov. Detection of retinal changes in idiopathic Parkinson’s disease using high-
Giannakou, A., Pedraza, O.L., Bonfill Cosp, X., Cullum, S., 2015. Mini-Mental State resolution optical coherence tomography and heidelberg retina tomography. Acta
Examination (MMSE) for the detection of Alzheimer’s disease and other dementias in Ophthalmol 93 (7), e578–e584.
people with mild cognitive impairment (MCI). Cochrane Database Syst. Rev. 3, Blanks, J.C., Hinton, D.R., Sadun, A.A., Miller, C.A., 1989. Retinal ganglion cell
CD010783. degeneration in Alzheimer’s disease. Brain Res. 501 (2), 364–372.
Armstrong, G.W., Kim, L.A., Vingopoulos, F., Park, J.Y., Garg, I., Kasetty, M., Blanks, J.C., Schmidt, S.Y., Torigoe, Y., Porrello, K.V., Hinton, D.R., Blanks, R.H., 1996a.
Silverman, R.F., Zeng, R., Douglas, V.P., Lopera, F., Baena, A., Giraldo, M., Retinal pathology in Alzheimer’s disease. II. Regional neuron loss and glial changes
Norton, D., Cronin-Golomb, A., Arboleda-Velasquez, J.F., Quiroz, Y.T., Miller, J.B., in GCL. Neurobiol. Aging 17 (3), 385–395.
2020 Nov. Retinal imaging findings in carriers with PSEN1-associated early-onset Blanks, J.C., Torigoe, Y., Hinton, D.R., Blanks, R.H., 1996b. Retinal pathology in
familial Alzheimer disease before onset of cognitive symptoms. JAMA Ophthalmol. Alzheimer’s disease. I. Ganglion cell loss in foveal/parafoveal retina. Neurobiol.
e204909. https://doi.org/10.1001/jamaophthalmol.2020.4909. Aging 17 (3), 377–384.
Asanad, S., Fantini, M., Sultan, W., Nassisi, M., Felix, C.M., Wu, J., Karanjia, R., Ross- Blanks, J.C., Torigoe, Y., Hinton, D.R., Blanks, R.H., 1991. Retinal degeneration in the
Cisneros, F.N., Sagare, A.P., Zlokovic, B.V., Chui, H.C., Pogoda, J.M., Arakaki, X., macula of patients with Alzheimer’s disease. Ann. N. Y. Acad. Sci. 640, 44–46.
Fonteh, A.N., Sadun, A.A., Harrington, M.G., 2020. Retinal nerve fiber layer Bodis-Wollner, I., Kozlowski, P.B., Glazman, S., Miri, S., 2014. α-synuclein in the inner
thickness predicts CSF amyloid/tau before cognitive decline. PloS One 15 (5), retina in Parkinson disease. Ann. Neurol. 75 (6), 964–966.
e0232785. Bodis-Wollner, I., Yahr, M.D., 1978. Measurements of visual evoked potentials in
Asanad, S., Ross-Cisneros, F.N., Barron, E., Nassisi, M., Sultan, W., Karanjia, R., Sadun, A. Parkinson’s disease. Brain 101 (4), 661–671.
A., 2019a. The retinal choroid as an oculovascular biomarker for Alzheimer’s Borbe, H.O., Fehske, K.J., Müller, W.E., Nover, A., Woller, U., 1982. The demonstration
dementia: a histopathological study in severe disease. Alzheimers Dement. (Amst.) of several neurotransmitter and drug receptors in human retina. Comp. Biochem.
11, 775–783. Physiol., C 72 (1), 117–119.
Asanad, S., Ross-Cisneros, F.N., Nassisi, M., Barron, E., Karanjia, R., Sadun, A.A., 2019b. Boroomand, A., Wong, A., Li, E., Cho, D.S., Ni, B., Bizheva, K., 2013. Multi-penalty
The retina in Alzheimer’s disease: histomorphometric analysis of an ophthalmologic conditional random field approach to super-resolved reconstruction of optical
biomarker. Invest. Ophthalmol. Vis. Sci. 60 (5), 1491–1500. coherence tomography images. Biomed. Optic Express 4 (10), 2032–2050.
Ascaso, F.J., Cruz, N., Modrego, P.J., Lopez-Anton, R., Santabárbara, J., Pascual, L.F., Borrelli, E., Sarraf, D., Freund, K.B., Sadda, S.R., 2018. OCT angiography and evaluation
Lobo, A., Cristóbal, J.A., 2014. Retinal alterations in mild cognitive impairment and of the choroid and choroidal vascular disorders. Prog. Retin. Eye Res. 67, 30–55.
Alzheimer’s disease: an optical coherence tomography study. J. Neurol. 261 (8), Boucart, M., Bubbico, G., Szaffarczyk, S., Defoort, S., Ponchel, A., Waucquier, N.,
1522–1530. Deplanque, D., Deguil, J., Bordet, R., 2015. Donepezil increases contrast sensitivity
Ashimatey, B., D’Orazio, L., Ma, S., Jann, K., Lu, H., Jiang, X., Wang, D., Ringman, J., for the detection of objects in scenes. Behav. Brain Res. 292, 443–447.
Kashani, A., 2020. Lower retinal capillary density in minimal cognitive impairment Bulens, C., Meerwaldt, J.D., Van der Wildt, G.J., Van Deursen, J.B., 1987. Effect of
among older Latinx adults. Alzheimer’s Dementia: Diagn. Assess. Dis. Monit. 12, levodopa treatment on contrast sensitivity in Parkinson’s disease. Ann. Neurol. 22
e12071. (3), 365–369.
Ashimatey, B.S., Green, K.M., Chu, Z., Wang, R.K., Kashani, A.H., 2019. Impaired retinal Bulut, M., Kurtuluş, F., Gözkaya, O., Erol, M.K., Cengiz, A., Akıdan, M., Yaman, A., 2018.
vascular reactivity in diabetic retinopathy as assessed by optical coherence Evaluation of optical coherence tomography angiographic findings in Alzheimer’s
tomography angiography. Invest. Ophthalmol. Vis. Sci. 60 (7), 2468–2473. type dementia. Br. J. Ophthalmol. 102 (2), 233–237.
Baker, C.L., Hess, R.R., Olsen, B.T., Zrenner, E., 1988. Current source density analysis of Bulut, M., Yaman, A., Erol, M.K., Kurtuluş, F., Toslak, D., Doğan, B., Turgut Çoban, D.,
linear and non-linear components of the primate electroretinogram. J. Physiol. 407, Kaya Başar, E., 2016. Choroidal thickness in patients with mild cognitive impairment
155–176. and Alzheimer’s type dementia. J. Ophthalmol. 2016, 7291257.
Baker, M.L., Hand, P.J., Wong, T.Y., Liew, G., Rochtchina, E., Mitchell, P., Lindley, R.I., Cameron, A., Lui, D., Boroomand, A., Glaister, J., Wong, A., Bizheva, K., 2013. Stochastic
Hankey, G.J., Wang, J.J., M.-C. R. S. S. Group, 2010a. Retinopathy and lobar speckle noise compensation in optical coherence tomography using non-stationary
intracerebral hemorrhage: insights into pathogenesis. Arch. Neurol. 67 (10), spline-based speckle noise modelling. Biomed. Optic Express 4 (9), 1769–1785.
1224–1230. Cameron, J.R., Megaw, R.D., Tatham, A.J., McGrory, S., MacGillivray, T.J., Doubal, F.N.,
Baker, M.L., Wang, J.J., Liew, G., Hand, P.J., De Silva, D.A., Lindley, R.I., Mitchell, P., Wardlaw, J.M., Trucco, E., Chandran, S., Dhillon, B., 2017. Lateral thinking -
Wong, M.C., Rochtchina, E., Wong, T.Y., Wardlaw, J.M., Hankey, G.J., M.-C. R. S. S.

30
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

interocular symmetry and asymmetry in neurovascular patterning, in health and Cubo, E., Tedejo, R.P., Rodriguez Mendez, V., López Peña, M.J., Trejo Gabriel Y Galán, J.
disease. Prog. Retin. Eye Res. 59, 131–157. M., 2010. Retina thickness in Parkinson’s disease and essential tremor. Mov. Disord.
Campbell, G., Mahad, D., 2018. Neurodegeneration in progressive multiple sclerosis. 25 (14), 2461–2462.
Cold Spring Harb. Perspect. Med. 8 (10). Cunha, L.P., Lopes, L.C., Costa-Cunha, L.V., Costa, C.F., Pires, L.A., Almeida, A.L.,
Campbell, J.P., Zhang, M., Hwang, T.S., Bailey, S.T., Wilson, D.J., Jia, Y., Huang, D., Monteiro, M.L., 2016. Macular thickness measurements with frequency domain-OCT
2017. Detailed vascular anatomy of the human retina by projection-resolved optical for quantification of retinal neural loss and its correlation with cognitive impairment
coherence tomography angiography. Sci. Rep. 7, 42201. in Alzheimer’s disease. PloS One 11 (4), e0153830.
Cannistraro, R.J., Badi, M., Eidelman, B.H., Dickson, D.W., Middlebrooks, E.H., Curcio, C.A., Drucker, D.N., 1993. Retinal ganglion cells in Alzheimer’s disease and
Meschia, J.F., 2019. CNS small vessel disease: a clinical review. Neurology 92 (24), aging. Ann. Neurol. 33 (3), 248–257.
1146–1156. Das, A.S., Regenhardt, R.W., Vernooij, M.W., Blacker, D., Charidimou, A.,
Carelli, V., La Morgia, C., Valentino, M.L., Barboni, P., Ross-Cisneros, F.N., Sadun, A.A., Viswanathan, A., 2019. Asymptomatic cerebral small vessel disease: insights from
2009. Retinal ganglion cell neurodegeneration in mitochondrial inherited disorders. population-based studies. J. Stroke 21 (2), 121–138.
Biochim. Biophys. Acta 1787 (5), 518–528. Del Brutto, O.H., Mera, R.M., Viteri, E.M., Pólit, J., Ledesma, E.A., Cano, J.A., Plaza, K.J.,
Chan, V.T.T., Sun, Z., Tang, S., Chen, L.J., Wong, A., Tham, C.C., Wong, T.Y., Chen, C., Zambrano, M., Costa, A.F., 2016. Hypertensive retinopathy and cerebral small vessel
Ikram, M.K., Whitson, H.E., Lad, E.M., Mok, V.C.T., Cheung, C.Y., 2019. Spectral- disease in Amerindians living in rural Ecuador: the Atahualpa Project. Int. J. Cardiol.
domain OCT measurements in Alzheimer’s disease: a systematic review and meta- 218, 65–68.
analysis. Ophthalmology 126 (4), 497–510. Delori, F., Greenberg, J.P., Woods, R.L., Fischer, J., Duncker, T., Sparrow, J., Smith, R.T.,
Chen C, B.K., Xin, C., Wen, J., Gupta, D., Zhang, Q., Mudumbai, R.C., Johnstone, M.A., 2011. Quantitative measurements of autofluorescence with the scanning laser
Chen, P.P., Wang, R.K., 2016. Repeatability and reproducibility of optic disc ophthalmoscope. Invest. Ophthalmol. Vis. Sci. 52 (13), 9379–9390.
perfusion measurements using optical coherence tomography-based angiography. Delori, F.C., Goger, D.G., Dorey, C.K., 2001. Age-related accumulation and spatial
Chen, J.J., Flanagan, E.P., Jitprapaikulsan, J., López-Chiriboga, A.S.S., Fryer, J.P., distribution of lipofuscin in RPE of normal subjects. Invest. Ophthalmol. Vis. Sci. 42
Leavitt, J.A., Weinshenker, B.G., McKeon, A., Tillema, J.M., Lennon, V.A., Tobin, W. (8), 1855–1866.
O., Keegan, B.M., Lucchinetti, C.F., Kantarci, O.H., McClelland, C.M., Lee, M.S., den Haan, J., Janssen, S.F., van de Kreeke, J.A., Scheltens, P., Verbraak, F.D.,
Bennett, J.L., Pelak, V.S., Chen, Y., VanStavern, G., Adesina, O.O., Eggenberger, E.R., Bouwman, F.H., 2018a. Retinal thickness correlates with parietal cortical atrophy in
Acierno, M.D., Wingerchuk, D.M., Brazis, P.W., Sagen, J., Pittock, S.J., 2018. Myelin early-onset Alzheimer’s disease and controls. Alzheimers Dement. (Amst.) 10,
oligodendrocyte glycoprotein antibody-positive optic neuritis: clinical 49–55.
characteristics, radiologic clues, and outcome. Am. J. Ophthalmol. 195, 8–15. den Haan, J., Morrema, T.H.J., Verbraak, F.D., de Boer, J.F., Scheltens, P., Rozemuller, A.
Chen, M., Lang, A., Ying, H.S., Calabresi, P.A., Prince, J.L., Carass, A., 2014. Analysis of J., Bergen, A.A.B., Bouwman, F.H., Hoozemans, J.J., 2018b. Amyloid-beta and
macular OCT images using deformable registration. Biomed. Optic Express 5 (7), phosphorylated tau in post-mortem Alzheimer’s disease retinas. Acta Neuropathol.
2196–2214. Commun. 6 (1), 147.
Chen, Z., Huang, D., Izatt, J.A., Wang, R.K., Werner, J.S., Yasuno, Y., 2016. Re: Spaide den Haan, J., Verbraak, F.D., Visser, P.J., Bouwman, F.H., 2017. Retinal thickness in
et al.: volume-rendering optical coherence tomography angiography of macular Alzheimer’s disease: a systematic review and meta-analysis. Alzheimers Dement.
telangiectasia type 2 (Ophthalmology 2015;122:2261-9). Ophthalmology 123 (3), (Amst.) 6, 162–170.
e24. Dorr, A., Sahota, B., Chinta, L.V., Brown, M.E., Lai, A.Y., Ma, K., Hawkes, C.A.,
Cheung, C.Y., Ong, Y.T., Hilal, S., Ikram, M.K., Low, S., Ong, Y.L., McLaurin, J., Stefanovic, B., 2012. Amyloid-β-dependent compromise of
Venketasubramanian, N., Yap, P., Seow, D., Chen, C.L., Wong, T.Y., 2015. Retinal microvascular structure and function in a model of Alzheimer’s disease. Brain 135
ganglion cell analysis using high-definition optical coherence tomography in (Pt 10), 3039–3050.
patients with mild cognitive impairment and Alzheimer’s disease. J. Alzheimers Dis. Doubal, F.N., de Haan, R., MacGillivray, T.J., Cohn-Hokke, P.E., Dhillon, B., Dennis, M.
45 (1), 45–56. S., Wardlaw, J.M., 2010. Retinal arteriolar geometry is associated with cerebral
Cheung, C.Y., Ong, Y.T., Ikram, M.K., Ong, S.Y., Li, X., Hilal, S., Catindig, J.A., white matter hyperintensities on magnetic resonance imaging. Int. J. Stroke 5 (6),
Venketasubramanian, N., Yap, P., Seow, D., Chen, C.P., Wong, T.Y., 2014. 434–439.
Microvascular network alterations in the retina of patients with Alzheimer’s disease. Drexler, W., Fujimoto, J.G., 2008. State-of-the-art retinal optical coherence tomography.
Alzheimers Dement. 10 (2), 135–142. Prog. Retin. Eye Res. 27 (1), 45–88.
Cheung, N., Mosley, T., Islam, A., Kawasaki, R., Sharrett, A.R., Klein, R., Coker, L.H., Duan, A., Bedggood, P.A., Bui, B.V., Metha, A.B., 2016. Evidence of flicker-induced
Knopman, D.S., Shibata, D.K., Catellier, D., Wong, T.Y., 2010. Retinal microvascular functional Hyperaemia in the smallest vessels of the human retinal blood supply.
abnormalities and subclinical magnetic resonance imaging brain infarct: a PloS One 11 (9), e0162621.
prospective study. Brain 133 (Pt 7), 1987–1993. Duan, A., Bedggood, P.A., Metha, A.B., Bui, B.V., 2017. Reactivity in the human retinal
Chibhabha, F., Yang, Y., Ying, K., Jia, F., Zhang, Q., Ullah, S., Liang, Z., Xie, M., Li, F., microvasculature measured during acute gas breathing provocations. Sci. Rep. 7 (1),
2020. Non-invasive optical imaging of retinal Aβ plaques using curcumin loaded 2113.
polymeric micelles in APP. J. Mater. Chem. B 8 (33), 7438–7452. Duan, W., Zheng, Y., Ding, Y., Hou, S., Tang, Y., Xu, Y., Qin, M., Wu, J., Shen, D., Bi, H.,
Chlebiej, M., Gorczynska, I., Rutkowski, A., Kluczewski, J., Grzona, T., Pijewska, E., 2018. A generative model for OCT retinal layer segmentation by groupwise curve
Sikorski, B.L., Szkulmowska, A., Szkulmowski, M., 2019. Quality improvement of Alignment. IEEE Access 6, 25130–25141.
OCT angiograms with elliptical directional filtering. Biomed. Optic Express 10 (2), Dubois, B., Hampel, H., Feldman, H.H., Scheltens, P., Aisen, P., Andrieu, S.,
1013–1031. Bakardjian, H., Benali, H., Bertram, L., Blennow, K., Broich, K., Cavedo, E.,
Cho, K.J., Yu, J., 2015. Retinal nerve fibre layer defect associated with MELAS syndrome. Crutch, S., Dartigues, J.F., Duyckaerts, C., Epelbaum, S., Frisoni, G.B., Gauthier, S.,
Can. J. Ophthalmol. 50 (5), e85–88. Genthon, R., Gouw, A.A., Habert, M.O., Holtzman, D.M., Kivipelto, M., Lista, S.,
Chrysou, A., Jansonius, N.M., van Laar, T., 2019. Retinal layers in Parkinson’s disease: a Molinuevo, J.L., O’Bryant, S.E., Rabinovici, G.D., Rowe, C., Salloway, S.,
meta-analysis of spectral-domain optical coherence tomography studies. Park. Relat. Schneider, L.S., Sperling, R., Teichmann, M., Carrillo, M.C., Cummings, J., Jack, C.
Disord. 64, 40–49. R., P. o. t. M. o. t. I. W. G. I. a. t. A. A. s. A. o. T. P. S. o. Ad” July 23 and U. S. A.
Chu, Z., Lin, J., Gao, C., Xin, C., Zhang, Q., Chen, C.L., Roisman, L., Gregori, G., Washington DC, 2016. Preclinical Alzheimer’s disease: definition, natural history,
Rosenfeld, P.J., Wang, R.K., 2016. Quantitative assessment of the retinal and diagnostic criteria. Alzheimers Dement. 12 (3), 292–323.
microvasculature using optical coherence tomography angiography. J. Biomed. Dysli, C., Fink, R., Wolf, S., Zinkernagel, M.S., 2017a. Fluorescence lifetimes of drusen in
Optic. 21 (6), 66008. age-related macular degeneration. Invest. Ophthalmol. Vis. Sci. 58 (11), 4856–4862.
Cipollini, V., Abdolrahimzadeh, S., Troili, F., De Carolis, A., Calafiore, S., Scuderi, L., Dysli, C., Wolf, S., Berezin, M.Y., Sauer, L., Hammer, M., Zinkernagel, M.S., 2017b.
Giubilei, F., Scuderi, G., 2020 Sep. Neurocognitive assessment and retinal thickness Fluorescence lifetime imaging ophthalmoscopy. Prog. Retin. Eye Res. 60, 120–143.
alterations in Alzheimer disease: is there a correlation? J. Neuro Ophthalmol. 40 (3), Eastline, M., Munk, M.R., Wolf, S., Schaal, K.B., Ebneter, A., Tian, M., Giannakaki-
370–377. Zimmermann, H., Zinkernagel, M.S., 2019. Repeatability of wide-field optical
Cogan, D.G., Kuwabara, T., 1984. Comparison of retinal and cerebral vasculature in coherence tomography angiography in normal retina. Transl. Vis. Sci. Technol. 8 (3),
trypsin digest preparations. Br. J. Ophthalmol. 68 (1), 10–12. 6.
Conti, F.F., Young, J.M., Silva, F.Q., Rodrigues, E.B., Singh, R.P., 2018. Repeatability of Ehle, A.L., Stewart, R.M., Lellelid, N.A., Leventhal, N.A., 1984. Evoked potentials in
split-spectrum amplitude-decorrelation angiography to assess capillary perfusion Huntington’s disease. A comparative and longitudinal study. Arch. Neurol. 41 (4),
density within optical coherence tomography. Ophthalmic Surg. Lasers Imag. Retina 379–382.
49 (9), e9–e19. Einarsdottir, A.B., Hardarson, S.H., Kristjansdottir, J.V., Bragason, D.T., Snaedal, J.,
Cooper, L.S., Wong, T.Y., Klein, R., Sharrett, A.R., Bryan, R.N., Hubbard, L.D., Couper, D. Stefánsson, E., 2016. Retinal oximetry imaging in Alzheimer’s disease. J. Alzheimers
J., Heiss, G., Sorlie, P.D., 2006. Retinal microvascular abnormalities and MRI- Dis. 49 (1), 79–83.
defined subclinical cerebral infarction: the Atherosclerosis Risk in Communities Eladawi, N., Elmogy, M., Helmy, O., Aboelfetouh, A., Riad, A., Sandhu, H., Schaal, S., El-
Study. Stroke 37 (1), 82–86. Baz, A., 2017. Automatic blood vessels segmentation based on different retinal maps
Coppola, G., Di Renzo, A., Ziccardi, L., Martelli, F., Fadda, A., Manni, G., Barboni, P., from OCTA scans. Comput. Biol. Med. 89, 150–161.
Pierelli, F., Sadun, A.A., Parisi, V., 2015. Optical coherence tomography in Elbøl, P., Work, K., 1990. Retinal nerve fiber layer in multiple sclerosis. Acta
Alzheimer’s disease: a meta-analysis. PloS One 10 (8), e0134750. Ophthalmol. 68 (4), 481–486.
Costello, F., Coupland, S., Hodge, W., Lorello, G.R., Koroluk, J., Pan, Y.I., Freedman, M. Ellenberger, C., Petro, D.J., Ziegler, S.B., 1978. The visually evoked potential in
S., Zackon, D.H., Kardon, R.H., 2006. Quantifying axonal loss after optic neuritis Huntington disease. Neurology 28 (1), 95–97.
with optical coherence tomography. Ann. Neurol. 59 (6), 963–969. Esen, E., Sizmaz, S., Demir, T., Demirkiran, M., Unal, I., Demircan, N., 2016. Evaluation
Cowey, A., Alexander, I., Stoerig, P., 2011. Transneuronal retrograde degeneration of of choroidal vascular changes in patients with multiple sclerosis using enhanced
retinal ganglion cells and optic tract in hemianopic monkeys and humans. Brain 134 depth imaging optical coherence tomography. Ophthalmologica 235 (2), 65–71.
(Pt 7), 2149–2157.

31
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Evangelou, N., Konz, D., Esiri, M.M., Smith, S., Palace, J., Matthews, P.M., 2001. Size- Gavaghan, K.A., Peterhans, M., Oliveira-Santos, T., Weber, S., 2011. A portable image
selective neuronal changes in the anterior optic pathways suggest a differential overlay projection device for computer-aided open liver surgery. IEEE Trans.
susceptibility to injury in multiple sclerosis. Brain 124 (Pt 9), 1813–1820. Biomed. Eng. 58 (6), 1855–1864.
Fang, L., Li, S., Nie, Q., Izatt, J.A., Toth, C.A., Farsiu, S., 2012. Sparsity based denoising Gelfand, J.M., Nolan, R., Schwartz, D.M., Graves, J., Green, A.J., 2012. Microcystic
of spectral domain optical coherence tomography images. Biomed. Optic Express 3 macular oedema in multiple sclerosis is associated with disease severity. Brain 135
(5), 927–942. (Pt 6), 1786–1793.
Fang, X.J., Yu, M., Wu, Y., Zhang, Z.H., Wang, W.W., Wang, Z.X., Yuan, Y., 2017. Study Gharbiya, M., Trebbastoni, A., Parisi, F., Manganiello, S., Cruciani, F., D’Antonio, F., De
of enhanced depth imaging optical coherence tomography in cerebral autosomal Vico, U., Imbriano, L., Campanelli, A., De Lena, C., 2014. Choroidal thinning as a
dominant arteriopathy with subcortical infarcts and leukoencephalopathy. Chin. new finding in Alzheimer’s disease: evidence from enhanced depth imaging spectral
Med. J. (Engl.) 130 (9), 1042–1048. domain optical coherence tomography. J. Alzheimers Dis. 40 (4), 907–917.
Federico, A., Di Donato, I., Bianchi, S., Di Palma, C., Taglia, I., Dotti, M.T., 2012. Ghilardi, M.F., Bodis-Wollner, I., Onofrj, M.C., Marx, M.S., Glover, A.A., 1988a. Spatial
Hereditary cerebral small vessel diseases: a review. J. Neurol. Sci. 322 (1–2), 25–30. frequency-dependent abnormalities of the pattern electroretinogram and visual
Feke, G.T., Hyman, B.T., Stern, R.A., Pasquale, L.R., 2015. Retinal blood flow in mild evoked potentials in a parkinsonian monkey model. Brain 111 (Pt 1), 131–149.
cognitive impairment and Alzheimer’s disease. Alzheimers Dement. (Amst.) 1 (2), Ghilardi, M.F., Chung, E., Bodis-Wollner, I., Dvorzniak, M., Glover, A., Onofrj, M., 1988b.
144–151. Systemic 1-methyl,4-phenyl,1-2-3-6-tetrahydropyridine (MPTP) administration
Ferrara, D., Waheed, N.K., Duker, J.S., 2016. Investigating the choriocapillaris and decreases retinal dopamine content in primates. Life Sci. 43 (3), 255–262.
choroidal vasculature with new optical coherence tomography technologies. Prog. Giaschi, D., Lang, A., Regan, D., 1997. Reversible dissociation of sensitivity to dynamic
Retin. Eye Res. 52, 130–155. stimuli in Parkinson’s disease: is magnocellular function essential to reading motion-
Ferrari, L., Huang, S.C., Magnani, G., Ambrosi, A., Comi, G., Leocani, L., 2017. Optical defined letters? Vis. Res. 37 (24), 3531–3534.
coherence tomography reveals retinal neuroaxonal thinning in frontotemporal Gibson, E., Young, M., Sarunic, M.V., Beg, M.F., 2010. Optic nerve head registration via
dementia as in Alzheimer’s disease. J. Alzheimers Dis. 56 (3), 1101–1107. hemispherical surface and volume registration. IEEE Trans. Biomed. Eng. 57 (10),
Feucht, N., Maier, M., Lepennetier, G., Pettenkofer, M., Wetzlmair, C., Daltrozzo, T., 2592–2595.
Scherm, P., Zimmer, C., Hoshi, M.M., Hemmer, B., Korn, T., Knier, B., 2019. Optical Gilmore, G.C., Spinks, R.A., Thomas, C.W., 2006. Age effects in coding tasks:
coherence tomography angiography indicates associations of the retinal vascular componential analysis and test of the sensory deficit hypothesis. Psychol. Aging 21
network and disease activity in multiple sclerosis. Mult. Scler. 25 (2), 224–234. (1), 7–18.
Filippatou, A.G., Mukharesh, L., Saidha, S., Calabresi, P.A., Sotirchos, E.S., 2020a. AQP4- Gobron, C., Erginay, A., Massin, P., Lutz, G., Tessier, N., Vicaut, E., Chabriat, H., 2014.
IgG and MOG-IgG related optic neuritis-prevalence, optical coherence tomography Microvascular retinal abnormalities in acute intracerebral haemorrhage and lacunar
findings, and visual outcomes: a systematic review and meta-analysis. Front. Neurol. infarction. Rev. Neurol. (Paris) 170 (1), 13–18.
11, 540156. González-López, J.J., Rebolleda, G., Leal, M., Oblanca, N., Muñoz-Negrete, F.J., Costa-
Filippatou, A.G., Vasileiou, E.S., He, Y., Fitzgerald, K.C., Kalaitzidis, G., Lambe, J., Frossard, L., Alvarez-Cermeño, J.C., 2014. Comparative diagnostic accuracy of
Mealy, M.A., Levy, M., Liu, Y., Prince, J.L., Mowry, E.M., Saidha, S., Calabresi, P.A., ganglion cell-inner plexiform and retinal nerve fiber layer thickness measures by
Sotirchos, E.S., 2020b. Evidence of subclinical quantitative retinal layer Cirrus and Spectralis optical coherence tomography in relapsing-remitting multiple
abnormalities in AQP4-IgG seropositive NMOSD. Mult. Scler. https://doi.org/ sclerosis. BioMed Res. Int. 2014, 128517.
10.1177/1352458520977771, 1352458520977771, Online ahead of print. Gordon-Lipkin, E., Chodkowski, B., Reich, D.S., Smith, S.A., Pulicken, M., Balcer, L.J.,
Forman, M.S., Farmer, J., Johnson, J.K., Clark, C.M., Arnold, S.E., Coslett, H.B., Frohman, E.M., Cutter, G., Calabresi, P.A., 2007. Retinal nerve fiber layer is
Chatterjee, A., Hurtig, H.I., Karlawish, J.H., Rosen, H.J., Van Deerlin, V., Lee, V.M., associated with brain atrophy in multiple sclerosis. Neurology 69 (16), 1603–1609.
Miller, B.L., Trojanowski, J.Q., Grossman, M., 2006. Frontotemporal dementia: Gorelick, P.B., Pantoni, L., 2013. Advances in vascular cognitive impairment. Stroke 44
clinicopathological correlations. Ann. Neurol. 59 (6), 952–962. (2), 307–308.
Frangi, A.F., Niessen, W.J., Vincken, K.L., Viergever, M.A., 1998. Multiscale Vessel Gorelick, P.B., Scuteri, A., Black, S.E., Decarli, C., Greenberg, S.M., Iadecola, C.,
Enhancement Filtering. Springer Berlin Heidelberg, Berlin, Heidelberg. Launer, L.J., Laurent, S., Lopez, O.L., Nyenhuis, D., Petersen, R.C., Schneider, J.A.,
Frank, R.N., Turczyn, T.J., Das, A., 1990. Pericyte coverage of retinal and cerebral Tzourio, C., Arnett, D.K., Bennett, D.A., Chui, H.C., Higashida, R.T., Lindquist, R.,
capillaries. Invest. Ophthalmol. Vis. Sci. 31 (6), 999–1007. Nilsson, P.M., Roman, G.C., Sellke, F.W., Seshadri, S., C. o. E. a. P. American Heart
Frishman, L.J., 2013. Electrogenesis of the Electroretinogram. Elsevier, New York. Association Stroke Council, Council on Cardiovascular Nursing, C.uncil on
Frost, S., Kanagasingam, Y., Sohrabi, H., Vignarajan, J., Bourgeat, P., Salvado, O., Cardiovascular Radiology and Intervention, Council on Cardiovascular Surgery and
Villemagne, V., Rowe, C.C., Macaulay, S.L., Szoeke, C., Ellis, K.A., Ames, D., Anesthesia, 2011. Vascular contributions to cognitive impairment and dementia: a
Masters, C.L., Rainey-Smith, S., Martins, R.N., AIBL Research Group, 2013. Retinal statement for healthcare professionals from the american heart association/american
vascular biomarkers for early detection and monitoring of Alzheimer’s disease. stroke association. Stroke 42 (9), 2672–2713.
Transl. Psychiatry 3, e233. Goto, K., Miki, A., Yamashita, T., Araki, S., Takizawa, G., Nakagawa, M., Ieki, Y.,
Gabilondo, I., Martínez-Lapiscina, E.H., Fraga-Pumar, E., Ortiz-Perez, S., Torres- Kiryu, J., 2016. Sectoral analysis of the retinal nerve fiber layer thinning and its
Torres, R., Andorra, M., Llufriu, S., Zubizarreta, I., Saiz, A., Sanchez-Dalmau, B., association with visual field loss in homonymous hemianopia caused by post-
Villoslada, P., 2015. Dynamics of retinal injury after acute optic neuritis. Ann. geniculate lesions using spectral-domain optical coherence tomography. Graefes
Neurol. 77 (3), 517–528. Arch. Clin. Exp. Ophthalmol. 254 (4), 745–756.
Garcia-Martin, E., Bambo, M.P., Marques, M.L., Satue, M., Otin, S., Larrosa, J.M., Green, A.J., McQuaid, S., Hauser, S.L., Allen, I.V., Lyness, R., 2010. Ocular pathology in
Polo, V., Pablo, L.E., 2016. Ganglion cell layer measurements correlate with disease multiple sclerosis: retinal atrophy and inflammation irrespective of disease duration.
severity in patients with Alzheimer’s disease. Acta Ophthalmol. 94 (6), e454–459. Brain 133 (Pt 6), 1591–1601.
Garcia-Martin, E., Larrosa, J.M., Polo, V., Satue, M., Marques, M.L., Alarcia, R., Seral, M., Greenberg, S.M., Bacskai, B.J., Hernandez-Guillamon, M., Pruzin, J., Sperling, R., van
Fuertes, I., Otin, S., Pablo, L.E., 2014a. Distribution of retinal layer atrophy in Veluw, S.J., 2020. Cerebral amyloid angiopathy and Alzheimer disease - one peptide,
patients with Parkinson disease and association with disease severity and duration. two pathways. Nat. Rev. Neurol. 16 (1), 30–42.
Am. J. Ophthalmol. 157 (2), 470–478 e472. Gregg, R., McCall, M., Massey, S., 2013. Function and Anatomy of the Mammalian
Garcia-Martin, E., Rodriguez-Mena, D., Satue, M., Almarcegui, C., Dolz, I., Alarcia, R., Retina. Elsevier, New York.
Seral, M., Polo, V., Larrosa, J.M., Pablo, L.E., 2014b. Electrophysiology and optical Group, O.N.S., 2008. Multiple sclerosis risk after optic neuritis: final optic neuritis
coherence tomography to evaluate Parkinson disease severity. Invest. Ophthalmol. treatment trial follow-up. Arch. Neurol. 65 (6), 727–732.
Vis. Sci. 55 (2), 696–705. Gulmez Sevim, D., Unlu, M., Gultekin, M., Karaca, C., 2019. Retinal single-layer analysis
Garcia-Martin, E., Pablo, L.E., Herrero, R., Satue, M., Polo, V., Larrosa, J.M., Martin, J., with optical coherence tomography shows inner retinal layer thinning in
Fernandez, J., 2012a. Diagnostic ability of a linear discriminant function for Huntington’s disease as a potential biomarker. Int. Ophthalmol. 39 (3), 611–621.
spectral-domain optical coherence tomography in patients with multiple sclerosis. Gupta, V.B., Chitranshi, N., den Haan, J., Mirzaei, M., You, Y., Lim, J.K.,
Ophthalmology 119 (8), 1705–1711. Basavarajappa, D., Godinez, A., Di Angelantonio, S., Sachdev, P., Salekdeh, G.H.,
Garcia-Martin, E., Satue, M., Fuertes, I., Otin, S., Alarcia, R., Herrero, R., Bambo, M.P., Bouwman, F., Graham, S., Gupta, V., 2020. Retinal changes in Alzheimer’s disease-
Fernandez, J., Pablo, L.E., 2012b. Ability and reproducibility of Fourier-domain integrated prospects of imaging, functional and molecular advances. Prog. Retin. Eye
optical coherence tomography to detect retinal nerve fiber layer atrophy in Res. 100899.
Parkinson’s disease. Ophthalmology 119 (10), 2161–2167. Guzmán-Vélez, E., Jaimes, S., Aguirre-Acevedo, D.C., Norton, D.J., Papp, K.V.,
Garcia-Martin, E., Pueyo, V., Almarcegui, C., Martin, J., Ara, J.R., Sancho, E., Pablo, L.E., Amariglio, R., Rentz, D., Baena, A., Henao, E., Tirado, V., Muñoz, C., Giraldo, M.,
Dolz, I., Fernandez, J., 2011. Risk factors for progressive axonal degeneration of the Sperling, R.A., Lopera, F., Quiroz, Y.T., 2018. A three-factor structure of cognitive
retinal nerve fibre layer in multiple sclerosis patients. Br. J. Ophthalmol. 95 (11), functioning among unimpaired carriers and non-carriers of autosomal-dominant
1577–1582. Alzheimer’s disease. J. Alzheimers Dis. 65 (1), 107–115.
Garcia-Martin, E., Pueyo, V., Martin, J., Almarcegui, C., Ara, J.R., Dolz, I., Honrubia, F. Haan, J.D., van de Kreeke, J.A., Konijnenberg, E., Kate, M.T., Braber, A.D., Barkhof, F.,
M., Fernandez, F.J., 2010. Progressive changes in the retinal nerve fiber layer in van Berckel, B.N., Teunissen, C.E., Scheltens, P., Visser, P.J., Verbraak, F.D.,
patients with multiple sclerosis. Eur. J. Ophthalmol. 20 (1), 167–173. Bouwman, F.H., 2019a. Retinal thickness as a potential biomarker in patients with
Garcia-Martin, E., Rodriguez-Mena, D., Herrero, R., Almarcegui, C., Dolz, I., Martin, J., amyloid-proven early- and late-onset Alzheimer’s disease. Alzheimers Dement.
Ara, J.R., Larrosa, J.M., Polo, V., Fernández, J., Pablo, L.E., 2013. Neuro- (Amst.) 11, 463–471.
ophthalmologic evaluation, quality of life, and functional disability in patients with Haan, J.D., van de Kreeke, J.A., van Berckel, B.N., Barkhof, F., Teunissen, C.E.,
MS. Neurology 81 (1), 76–83. Scheltens, P., Verbraak, F.D., Bouwman, F.H., 2019b. Is retinal vasculature a
Gatto, E., Parisi, V., Persi, G., Fernandez Rey, E., Cesarini, M., Luis Etcheverry, J., biomarker in amyloid proven Alzheimer’s disease? Alzheimers Dement. (Amst.) 11,
Rivera, P., Squitieri, F., 2018. Optical coherence tomography (OCT) study in 383–391.
Argentinean Huntington’s disease patients. Int. J. Neurosci. 128 (12), 1157–1162. Hadoux, X., Hui, F., Lim, J.K.H., Masters, C.L., Pébay, A., Chevalier, S., Ha, J., Loi, S.,
Fowler, C.J., Rowe, C., Villemagne, V.L., Taylor, E.N., Fluke, C., Soucy, J.P.,

32
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Lesage, F., Sylvestre, J.P., Rosa-Neto, P., Mathotaarachchi, S., Gauthier, S., Ikram, M.K., De Jong, F.J., Van Dijk, E.J., Prins, N.D., Hofman, A., Breteler, M.M., De
Nasreddine, Z.S., Arbour, J.D., Rhéaume, M.A., Beaulieu, S., Dirani, M., Nguyen, C.T. Jong, P.T., 2006. Retinal vessel diameters and cerebral small vessel disease: the
O., Bui, B.V., Williamson, R., Crowston, J.G., van Wijngaarden, P., 2019. Non- Rotterdam Scan Study. Brain 129 (Pt 1), 182–188.
invasive in vivo hyperspectral imaging of the retina for potential biomarker use in Ikuta, F., Zimmerman, H.M., 1976. Distribution of plaques in seventy autopsy cases of
Alzheimer’s disease. Nat. Commun. 10 (1), 4227. multiple sclerosis in the United States. Neurology 26 (6 PT 2), 26–28.
Hajee, M.E., March, W.F., Lazzaro, D.R., Wolintz, A.H., Shrier, E.M., Glazman, S., Bodis- Inzelberg, R., Ramirez, J.A., Nisipeanu, P., Ophir, A., 2004. Retinal nerve fiber layer
Wollner, I.G., 2009. Inner retinal layer thinning in Parkinson disease. Arch. thinning in Parkinson disease. Vis. Res. 44 (24), 2793–2797.
Ophthalmol. 127 (6), 737–741. Irwin, D.J., Cairns, N.J., Grossman, M., McMillan, C.T., Lee, E.B., Van Deerlin, V.M.,
Halliday, A.M., McDonald, W.I., Mushin, J., 1972. Delayed visual evoked response in Lee, V.M., Trojanowski, J.Q., 2015. Frontotemporal lobar degeneration: defining
optic neuritis. Lancet 1 (7758), 982–985. phenotypic diversity through personalized medicine. Acta Neuropathol. 129 (4),
Halliday, A.M., McDonald, W.I., Mushin, J., 1973. Visual evoked response in diagnosis of 469–491.
multiple sclerosis. Br. Med. J. 4 (5893), 661–664. Isas, J.M., Luibl, V., Johnson, L.V., Kayed, R., Wetzel, R., Glabe, C.G., Langen, R.,
Hanff, T.C., Sharrett, A.R., Mosley, T.H., Shibata, D., Knopman, D.S., Klein, R., Klein, B. Chen, J., 2010. Soluble and mature amyloid fibrils in drusen deposits. Invest.
E., Gottesman, R.F., 2014. Retinal microvascular abnormalities predict progression Ophthalmol. Vis. Sci. 51 (3), 1304–1310.
of brain microvascular disease: an atherosclerosis risk in communities magnetic Jack, C.R., Albert, M.S., Knopman, D.S., McKhann, G.M., Sperling, R.A., Carrillo, M.C.,
resonance imaging study. Stroke 45 (4), 1012–1017. Thies, B., Phelps, C.H., 2011. Introduction to the recommendations from the
Haritoglou, C., Rudolph, G., Hoops, J.P., Opherk, C., Kampik, A., Dichgans, M., 2004. National Institute on Aging-Alzheimer’s Association workgroups on diagnostic
Retinal vascular abnormalities in CADASIL. Neurology 62 (7), 1202–1205. guidelines for Alzheimer’s disease. Alzheimers Dement. 7 (3), 257–262.
Harnois, C., Di Paolo, T., 1990. Decreased dopamine in the retinas of patients with Jack, C.R., Bennett, D.A., Blennow, K., Carrillo, M.C., Dunn, B., Haeberlein, S.B.,
Parkinson’s disease. Invest. Ophthalmol. Vis. Sci. 31 (11), 2473–2475. Holtzman, D.M., Jagust, W., Jessen, F., Karlawish, J., Liu, E., Molinuevo, J.L.,
Hasanov, S., Demirkilinc Biler, E., Acarer, A., Akkın, C., Colakoglu, Z., Uretmen, O., Montine, T., Phelps, C., Rankin, K.P., Rowe, C.C., Scheltens, P., Siemers, E.,
2019. Functional and morphological assessment of ocular structures and follow-up Snyder, H.M., Sperling, R., Contributors, 2018. NIA-AA Research Framework:
of patients with early-stage Parkinson’s disease. Int. Ophthalmol. 39 (6), 1255–1262. toward a biological definition of Alzheimer’s disease. Alzheimers Dement. 14 (4),
He, Y., Carass, A., Liu, Y., Jedynak, B.M., Solomon, S.D., Saidha, S., Calabresi, P.A., 535–562.
Prince, J.L., 2019. Deep learning based topology guaranteed surface and MME Jaime, G.R., Kashani, A.H., Saati, S., Martin, G., Chader, G., Humayun, M.S., 2012. Acute
segmentation of multiple sclerosis subjects from retinal OCT. Biomed. Optic Express variations in retinal vascular oxygen content in a rabbit model of retinal venous
10 (10), 5042–5058. occlusion. PloS One 7 (11), e50179.
Helmlinger, D., Yvert, G., Picaud, S., Merienne, K., Sahel, J., Mandel, J.L., Devys, D., Jia, Y., Tan, O., Tokayer, J., Potsaid, B., Wang, Y., Liu, J.J., Kraus, M.F., Subhash, H.,
2002. Progressive retinal degeneration and dysfunction in R6 Huntington’s disease Fujimoto, J.G., Hornegger, J., Huang, D., 2012. Split-spectrum amplitude-
mice. Hum. Mol. Genet. 11 (26), 3351–3359. decorrelation angiography with optical coherence tomography. Optic Express 20 (4),
Henderson, A.D., Bruce, B.B., Newman, N.J., Biousse, V., 2011. Hypertension-related eye 4710–4725.
abnormalities and the risk of stroke. Rev. Neurol. Dis. 8 (1–2), 1–9. Jian, Z., Yu, L., Rao, B., Tromberg, B.J., Chen, Z., 2010. Three-dimensional speckle
Henderson, A.P., Trip, S.A., Schlottmann, P.G., Altmann, D.R., Garway-Heath, D.F., suppression in Optical Coherence Tomography based on the curvelet transform.
Plant, G.T., Miller, D.H., 2010. A preliminary longitudinal study of the retinal nerve Optic Express 18 (2), 1024–1032.
fiber layer in progressive multiple sclerosis. J. Neurol. 257 (7), 1083–1091. Jian, Z., Yu, Z., Yu, L., Rao, B., Chen, Z., Tromberg, B.J., 2009. Speckle attenuation in
Hickman, S.J., Brierley, C.M., Brex, P.A., MacManus, D.G., Scolding, N.J., Compston, D. optical coherence tomography by curvelet shrinkage. Opt. Lett. 34 (10), 1516–1518.
A., Miller, D.H., 2002. Continuing optic nerve atrophy following optic neuritis: a Jiang, H., Delgado, S., Tan, J., Liu, C., Rammohan, K.W., DeBuc, D.C., Lam, B.L.,
serial MRI study. Mult. Scler. 8 (4), 339–342. Feuer, W.J., Wang, J., 2016. Impaired retinal microcirculation in multiple sclerosis.
Hickman, S.J., Toosy, A.T., Jones, S.J., Altmann, D.R., Miszkiel, K.A., MacManus, D.G., Mult. Scler. 22 (14), 1812–1820.
Barker, G.J., Plant, G.T., Thompson, A.J., Miller, D.H., 2004. A serial MRI study Jiang, H., Liu, Y., Wei, Y., Shi, Y., Wright, C.B., Sun, X., Rundek, T., Baumel, B.S.,
following optic nerve mean area in acute optic neuritis. Brain 127 (Pt 11), Landman, J., Wang, J., 2018a. Impaired retinal microcirculation in patients with
2498–2505. Alzheimer’s disease. PloS One 13 (2), e0192154.
Hinton, D.R., Sadun, A.A., Blanks, J.C., Miller, C.A., 1986. Optic-nerve degeneration in Jiang, H., Wei, Y., Shi, Y., Wright, C.B., Sun, X., Gregori, G., Zheng, F., Vanner, E.A.,
Alzheimer’s disease. N. Engl. J. Med. 315 (8), 485–487. Lam, B.L., Rundek, T., Wang, J., 2018b. Altered macular microvasculature in mild
Hiroki, M., Miyashita, K., Yoshida, H., Hirai, S., Fukuyama, H., 2003. Central retinal cognitive impairment and Alzheimer disease. J. Neuro Ophthalmol. 38 (3), 292–298.
artery Doppler flow parameters reflect the severity of cerebral small-vessel disease. Jiménez, B., Ascaso, F.J., Cristóbal, J.A., López del Val, J., 2014. Development of a
Stroke 34 (7), e92–94. prediction formula of Parkinson disease severity by optical coherence tomography.
Ho, C.Y., Troncoso, J.C., Knox, D., Stark, W., Eberhart, C.G., 2014. Beta-amyloid, Mov. Disord. 29 (1), 68–74.
phospho-tau and alpha-synuclein deposits similar to those in the brain are not Jonas, J.B., Wang, Y.X., Wei, W.B., Zhu, L.P., Shao, L., Xu, L., 2016. Cognitive function
identified in the eyes of Alzheimer’s and Parkinson’s disease patients. Brain Pathol. and subfoveal choroidal thickness: the Beijing eye study. Ophthalmology 123 (1),
24 (1), 25–32. 220–222.
Hoh Kam, J., Lenassi, E., Jeffery, G., 2010. Viewing ageing eyes: diverse sites of amyloid Josiassen, R.C., Shagass, C., Mancall, E.L., Roemer, R.A., 1984. Auditory and visual
Beta accumulation in the ageing mouse retina and the up-regulation of macrophages. evoked potentials in Huntington’s disease. Electroencephalogr. Clin. Neurophysiol.
PloS One 5 (10). 57 (2), 113–118.
Holden, B.A., Fricke, T.R., Wilson, D.A., Jong, M., Naidoo, K.S., Sankaridurg, P., Jung, N.Y., Han, J.C., Ong, Y.T., Cheung, C.Y., Chen, C.P., Wong, T.Y., Kim, H.J., Kim, Y.
Wong, T.Y., Naduvilath, T.J., Resnikoff, S., 2016. Global prevalence of myopia and J., Lee, J., Lee, J.S., Jang, Y.K., Kee, C., Lee, K.H., Kim, E.J., Seo, S.W., Na, D.L.,
high myopia and temporal trends from 2000 through 2050. Ophthalmology 123 (5), 2019. Retinal microvasculature changes in amyloid-negative subcortical vascular
1036–1042. cognitive impairment compared to amyloid-positive Alzheimer’s disease. J. Neurol.
Holmen, I.C., Konda, M.S., Pak, J.W., McDaniel, K.W., Blodi, B., Stepien, K.E., Sci. 396, 94–101.
Domalpally, A., 2020 Feb 1. Prevalence and severity of artifacts in optical coherence Kanamori, A., Escano, M.F., Eno, A., Nakamura, M., Maeda, H., Seya, R., Ishibashi, K.,
tomographic angiograms. JAMA Ophthalmol. 138 (2), 119–126. Negi, A., 2003. Evaluation of the effect of aging on retinal nerve fiber layer thickness
Howland, D., Ellederova, Z., Aronin, N., Fernau, D., Gallagher, J., Taylor, A., measured by optical coherence tomography. Ophthalmologica 217 (4), 273–278.
Hennebold, J., Weiss, A.R., Gray-Edwards, H., McBride, J., 2020. Large animal Kardon, R.H., 1998. Drop the Alzheimer’s drop test. Neurology 50 (3), 588–591.
models of Huntington’s disease: what we have learned and where we need to go Kashani, A.H., Brown, K.T., Chang, E., Drenser, K.A., Capone, A., Trese, M.T., 2014a.
next. J. Huntingtons Dis. 9 (3), 201–216. Diversity of retinal vascular anomalies in patients with familial exudative
Hu, J., Gottlieb, C.B., Barajas, D.J., Barnett, C.J., Schoenholz, T., Sadda, S.R., 2015. vitreoretinopathy. Ophthalmology 121 (11), 2220–2227.
Improved repeatability of retinal thickness measurements using line-scan Kashani, A.H., Learned, D., Nudleman, E., Drenser, K.A., Capone, A., Trese, M.T., 2014b.
ophthalmoscope image-based retinal tracking. Ophthalmic Surg. Lasers Imag. Retina High prevalence of peripheral retinal vascular anomalies in family members of
46 (3), 310–314. patients with familial exudative vitreoretinopathy. Ophthalmology 121 (1),
Huang, D., Swanson, E.A., Lin, C.P., Schuman, J.S., Stinson, W.G., Chang, W., Hee, M.R., 262–268.
Flotte, T., Gregory, K., Puliafito, C.A., 1991. Optical coherence tomography. Science Kashani, A.H., Chen, C.L., Gahm, J.K., Zheng, F., Richter, G.M., Rosenfeld, P.J., Shi, Y.,
254 (5035), 1178–1181. Wang, R.K., 2017. Optical coherence tomography angiography: a comprehensive
Huang, J., Li, Y., Xiao, J., Zhang, Q., Xu, G., Wu, G., Liu, T., Luo, W., 2018. Combination review of current methods and clinical applications. Prog. Retin. Eye Res. 60,
of multifocal electroretinogram and spectral-domain OCT can increase diagnostic 66–100.
efficacy of Parkinson’s disease. Park. Dis. 2018, 4163239. Kashani, A.H., Kirkman, E., Martin, G., Humayun, M.S., 2011. Hyperspectral computed
Hutton, J.T., Morris, J.L., Elias, J.W., 1993. Levodopa improves spatial contrast tomographic imaging spectroscopy of vascular oxygen gradients in the rabbit retina
sensitivity in Parkinson’s disease. Arch. Neurol. 50 (7), 721–724. in vivo. PloS One 6 (9), e24482.
Iadecola, C., 2013. The pathobiology of vascular dementia. Neuron 80 (4), 844–866. Kashani, A.H., Lee, S.Y., Moshfeghi, A., Durbin, M.K., Puliafito, C.A., 2015. Optical
Iijima, A., Haida, M., Ishikawa, N., Ueno, A., Minamitani, H., Shinohara, Y., 2003. Re- coherence tomography angiography OF retinal venous occlusion. Retina 35 (11),
evaluation of tropicamide in the pupillary response test for Alzheimer’s disease. 2323–2331.
Neurobiol. Aging 24 (6), 789–796. Kashani, A.H., Zimmer-Galler, I.E., Shah, S.M., Dustin, L., Do, D.V., Eliott, D., Haller, J.
Ikeda, H., Head, G.M., Ellis, C.J., 1994. Electrophysiological signs of retinal dopamine A., Nguyen, Q.D., 2010. Retinal thickness analysis by race, gender, and age using
deficiency in recently diagnosed Parkinson’s disease and a follow up study. Vis. Res. Stratus OCT. Am. J. Ophthalmol. 149 (3), 496–502 e491.
34 (19), 2629–2638. Katz, B., Rimmer, S., 1989. Ophthalmologic manifestations of Alzheimer’s disease. Surv.
Ophthalmol. 34 (1), 31–43.

33
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Kaur, M., Saxena, R., Singh, D., Behari, M., Sharma, P., Menon, V., 2015. Correlation Kushner-Lenhoff, S., Ashimatey, B.S., Kashani, A.H., 2020. Retinal vascular reactivity as
between structural and functional retinal changes in Parkinson disease. J. Neuro assessed by optical coherence tomography angiography. JoVE 157.
Ophthalmol. 35 (3), 254–258. Kutzelnigg, A., Lucchinetti, C.F., Stadelmann, C., Brück, W., Rauschka, H., Bergmann, M.,
Kellner, U., Kellner, S., Weinitz, S., 2010. Fundus autofluorescence (488 NM) and near- Schmidbauer, M., Parisi, J.E., Lassmann, H., 2005. Cortical demyelination and
infrared autofluorescence (787 NM) visualize different retinal pigment epithelium diffuse white matter injury in multiple sclerosis. Brain 128 (Pt 11), 2705–2712.
alterations in patients with age-related macular degeneration. Retina 30 (1), 6–15. Kwa, V.I., van der Sande, J.J., Stam, J., Tijmes, N., Vrooland, J.L., Group, A.V.M., 2002.
Kerrison, J.B., Flynn, T., Green, W.R., 1994. Retinal pathologic changes in multiple Retinal arterial changes correlate with cerebral small-vessel disease. Neurology 59
sclerosis. Retina 14 (5), 445–451. (10), 1536–1540.
Kersten, H.M., Danesh-Meyer, H.V., Kilfoyle, D.H., Roxburgh, R.H., 2015. Optical Kwapong, W.R., Ye, H., Peng, C., Zhuang, X., Wang, J., Shen, M., Lu, F., 2018. Retinal
coherence tomography findings in Huntington’s disease: a potential biomarker of microvascular impairment in the early stages of Parkinson’s disease. Invest.
disease progression. J. Neurol. 262 (11), 2457–2465. Ophthalmol. Vis. Sci. 59 (10), 4115–4122.
Kertesz, A., McMonagle, P., Blair, M., Davidson, W., Munoz, D.G., 2005. The evolution La Morgia, C., Barboni, P., Rizzo, G., Carbonelli, M., Savini, G., Scaglione, C.,
and pathology of frontotemporal dementia. Brain 128 (Pt 9), 1996–2005. Capellari, S., Bonazza, S., Giannoccaro, M.P., Calandra-Buonaura, G., Liguori, R.,
Kesler, A., Vakhapova, V., Korczyn, A.D., Naftaliev, E., Neudorfer, M., 2011. Retinal Cortelli, P., Martinelli, P., Baruzzi, A., Carelli, V., 2013. Loss of temporal retinal
thickness in patients with mild cognitive impairment and Alzheimer’s disease. Clin. nerve fibers in Parkinson disease: a mitochondrial pattern? Eur. J. Neurol. 20 (1),
Neurol. Neurosurg. 113 (7), 523–526. 198–201.
Khansari, M.M., Zhang, J., Qiao, Y., Gahm, J.K., Sarabi, M.S., Kashani, A.H., Shi, Y., La Morgia, C., Di Vito, L., Carelli, V., Carbonelli, M., 2017. Patterns of retinal ganglion
2020. Automated deformation-based analysis of 3D optical coherence tomography in cell damage in neurodegenerative disorders: parvocellular vs magnocellular
diabetic retinopathy. IEEE Trans. Med. Imag. 39 (1), 236–245. degeneration in optical coherence tomography studies. Front. Neurol. 8, 710.
Kim, A.Y., Chu, Z., Shahidzadeh, A., Wang, R.K., Puliafito, C.A., Kashani, A.H., 2016a. La Morgia, C., Ross-Cisneros, F.N., Koronyo, Y., Hannibal, J., Gallassi, R., Cantalupo, G.,
Quantifying microvascular density and morphology in diabetic retinopathy using Sambati, L., Pan, B.X., Tozer, K.R., Barboni, P., Provini, F., Avanzini, P.,
spectral-domain optical coherence tomography angiography. Invest. Ophthalmol. Carbonelli, M., Pelosi, A., Chui, H., Liguori, R., Baruzzi, A., Koronyo-Hamaoui, M.,
Vis. Sci. 57 (9), OCT362–370. Sadun, A.A., Carelli, V., 2016. Melanopsin retinal ganglion cell loss in Alzheimer
Kim, A.Y., Rodger, D.C., Shahidzadeh, A., Chu, Z., Koulisis, N., Burkemper, B., Jiang, X., disease. Ann. Neurol. 79 (1), 90–109.
Pepple, K.L., Wang, R.K., Puliafito, C.A., Rao, N.A., Kashani, A.H., 2016b. Lachowicz, E., Lubiński, W., 2018. The importance of the electrophysiological tests in the
Quantifying retinal microvascular changes in uveitis using spectral domain optical early diagnosis of ganglion cells and/or optic nerve dysfunction coexisting with
coherence tomography angiography (SD-OCTA). Am. J. Ophthalmol. 171, 101–112. pituitary adenoma: an overview. Doc. Ophthalmol. 137 (3), 193–202.
https://doi.org/10.1016/j.ajo.2016.08.035. Epub 2016 Sep 2. Lad, E.M., Mukherjee, D., Stinnett, S.S., Cousins, S.W., Potter, G.G., Burke, J.R.,
Kim, B.J., Grossman, M., Song, D., Saludades, S., Pan, W., Dominguez-Perez, S., Farsiu, S., Whitson, H.E., 2018. Evaluation of inner retinal layers as biomarkers in
Dunaief, J.L., Aleman, T.S., Ying, G.S., Irwin, D.J., 2019a. Persistent and progressive mild cognitive impairment to moderate Alzheimer’s disease. PloS One 13 (2),
outer retina thinning in frontotemporal degeneration. Front. Neurosci. 13, 298. e0192646.
Kim, M.S., Jeong, H.Y., Cho, K.H., Oh, S.W., Byun, S.J., Woo, S.J., Yang, H.K., Hwang, J. Lambe, J., Saidha, S., Bermel, R.A., 2020. Optical coherence tomography and multiple
M., Park, K.H., Kim, C.K., Park, S.J., 2019b. Nonarteritic anterior ischemic optic sclerosis: update on clinical application and role in clinical trials. Mult. Scler. 26 (6),
neuropathy is associated with cerebral small vessel disease. PloS One 14 (11), 624–639.
e0225322. Law, M., Saindane, A.M., Ge, Y., Babb, J.S., Johnson, G., Mannon, L.J., Herbert, J.,
Kim, B.J., Irwin, D.J., Song, D., Daniel, E., Leveque, J.D., Raquib, A.R., Pan, W., Ying, G. Grossman, R.I., 2004. Microvascular abnormality in relapsing-remitting multiple
S., Aleman, T.S., Dunaief, J.L., Grossman, M., 2017. Optical coherence tomography sclerosis: perfusion MR imaging findings in normal-appearing white matter.
identifies outer retina thinning in frontotemporal degeneration. Neurology 89 (15), Radiology 231 (3), 645–652.
1604–1611. Law, M.W.K., Chung, A.C.S., 2008. Three Dimensional Curvilinear Structure Detection
Kim, M., Park, K.H., Kwon, J.W., Jeoung, J.W., Kim, T.W., Kim, D.M., 2011. Retinal Using Optimally Oriented Flux. Springer Berlin Heidelberg, Berlin, Heidelberg.
nerve fiber layer defect and cerebral small vessel disease. Invest. Ophthalmol. Vis. Lee, J.Y., Kim, J.M., Ahn, J., Kim, H.J., Jeon, B.S., Kim, T.W., 2014. Retinal nerve fiber
Sci. 52 (9), 6882–6886. layer thickness and visual hallucinations in Parkinson’s Disease. Mov. Disord. 29 (1),
Kirbas, S., Turkyilmaz, K., Tufekci, A., Durmus, M., 2013. Retinal nerve fiber layer 61–67.
thickness in Parkinson disease. J. Neuro Ophthalmol. 33 (1), 62–65. Lee, M.J., Abraham, A.G., Swenor, B.K., Sharrett, A.R., Ramulu, P.Y., 2018. Application
Klunk, W.E., Engler, H., Nordberg, A., Wang, Y., Blomqvist, G., Holt, D.P., Bergström, M., of optical coherence tomography in the detection and classification of cognitive
Savitcheva, I., Huang, G.F., Estrada, S., Ausén, B., Debnath, M.L., Barletta, J., decline. J. Curr. Glaucoma Pract. 12 (1), 10–18.
Price, J.C., Sandell, J., Lopresti, B.J., Wall, A., Koivisto, P., Antoni, G., Mathis, C.A., Lee, S., Lebed, E., Sarunic, M.V., Beg, M.F., 2015. Exact surface registration of retinal
Långström, B., 2004. Imaging brain amyloid in Alzheimer’s disease with Pittsburgh surfaces from 3-D optical coherence tomography images. IEEE Trans. Biomed. Eng.
Compound-B. Ann. Neurol. 55 (3), 306–319. 62 (2), 609–617.
Knapp, J., VanNasdale, D.A., Ramsey, K., Racine, J., 2018. Retinal dysfunction in a Leger, F., Fernagut, P.O., Canron, M.H., Léoni, S., Vital, C., Tison, F., Bezard, E., Vital, A.,
presymptomatic patient with Huntington’s disease. Doc. Ophthalmol. 136 (3), 2011. Protein aggregation in the aging retina. J. Neuropathol. Exp. Neurol. 70 (1),
213–221. 63–68.
Knier, B., Schmidt, P., Aly, L., Buck, D., Berthele, A., Mühlau, M., Zimmer, C., Leuba, G., Saini, K., 1995. Pathology of subcortical visual centres in relation to cortical
Hemmer, B., Korn, T., 2016. Retinal inner nuclear layer volume reflects response to degeneration in Alzheimer’s disease. Neuropathol. Appl. Neurobiol. 21 (5),
immunotherapy in multiple sclerosis. Brain 139 (11), 2855–2863. 410–422.
Knopman, D.S., Boeve, B.F., Parisi, J.E., Dickson, D.W., Smith, G.E., Ivnik, R.J., Liew, G., Baker, M.L., Wong, T.Y., Hand, P.J., Wang, J.J., Mitchell, P., De Silva, D.A.,
Josephs, K.A., Petersen, R.C., 2005. Antemortem diagnosis of frontotemporal lobar Wong, M.C., Rochtchina, E., Lindley, R.I., Wardlaw, J.M., Hankey, G.J., M.-C. R. S. S.
degeneration. Ann. Neurol. 57 (4), 480–488. Group, 2014. Differing associations of white matter lesions and lacunar infarction
Knopman, D.S., Roberts, R.O., 2011. Estimating the number of persons with with retinal microvascular signs. Int. J. Stroke 9 (7), 921–925.
frontotemporal lobar degeneration in the US population. J. Mol. Neurosci. 45 (3), Lim, J.K.H., Li, Q.X., He, Z., Vingrys, A.J., Chinnery, H.R., Mullen, J., Bui, B.V.,
330–335. Nguyen, C.T.O., 2020. Retinal functional and structural changes in the 5xFAD mouse
Ko, F., Muthy, Z.A., Gallacher, J., Sudlow, C., Rees, G., Yang, Q., Keane, P.A., Petzold, A., model of Alzheimer’s disease. Front. Neurosci. 14, 862.
Khaw, P.T., Reisman, C., Strouthidis, N.G., Foster, P.J., Patel, P.J., U. B. E. V. Lindley, R.I., Wang, J.J., Wong, M.C., Mitchell, P., Liew, G., Hand, P., Wardlaw, J., De
Consortium, 2018. Association of retinal nerve fiber layer thinning with current and Silva, D.A., Baker, M., Rochtchina, E., Chen, C., Hankey, G.J., Chang, H.M., Fung, V.
future cognitive decline: a study using optical coherence tomography. JAMA Neurol. S., Gomes, L., Wong, T.Y., M.-C. R. a. S. S. M. C. Group, 2009. Retinal
75 (10), 1198–1205. microvasculature in acute lacunar stroke: a cross-sectional study. Lancet Neurol. 8
Koronyo, Y., Biggs, D., Barron, E., Boyer, D.S., Pearlman, J.A., Au, W.J., Kile, S.J., (7), 628–634.
Blanco, A., Fuchs, D.T., Ashfaq, A., Frautschy, S., Cole, G.M., Miller, C.A., Hinton, D. Litvan, I., Aarsland, D., Adler, C.H., Goldman, J.G., Kulisevsky, J., Mollenhauer, B.,
R., Verdooner, S.R., Black, K.L., Koronyo-Hamaoui, M., 2017. Retinal amyloid Rodriguez-Oroz, M.C., Tröster, A.I., Weintraub, D., 2011. MDS Task Force on mild
pathology and proof-of-concept imaging trial in Alzheimer’s disease. JCI Insight 2 cognitive impairment in Parkinson’s disease: critical review of PD-MCI. Mov. Disord.
(16). 26 (10), 1814–1824.
Koronyo-Hamaoui, M., Koronyo, Y., Ljubimov, A.V., Miller, C.A., Ko, M.K., Black, K.L., Liu, Y., Carass, A., Filippatou, A., He, Y., Solomon, S.D., Saidha, S., Calabresi, P.A.,
Schwartz, M., Farkas, D.L., 2011. Identification of amyloid plaques in retinas from Prince, J.L., F. 2, IEEE, 2019a. Projection artifact suppression for inner retina IN OCT
Alzheimer’s patients and noninvasive in vivo optical imaging of retinal plaques in a angiography. In: 2019 IEEE 16th International Symposium on Biomedical Imaging
mouse model. Neuroimage 54 (Suppl. 1), S204–S217. (ISBI 2019). IEEE, Venice, Italy, Italy, pp. 592–596.
Koulisis, N., Kim, A.Y., Chu, Z., Shahidzadeh, A., Burkemper, B., Olmos de Koo, L.C., Liu, Y., Delgado, S., Jiang, H., Lin, Y., Hernandez, J., Deng, Y., Gameiro, G.R., Wang, J.,
Moshfeghi, A.A., Ameri, H., Puliafito, C.A., Isozaki, V.L., Wang, R.K., Kashani, A.H., 2019b. Retinal tissue perfusion in patients with multiple sclerosis. Curr. Eye Res. 44
2017. Quantitative microvascular analysis of retinal venous occlusions by spectral (10), 1091–1097.
domain optical coherence tomography angiography. PloS One 12 (4), e0176404. Llanas, S., Linderman, R.E., Chen, F.K., Carroll, J., 2019. Assessing the use of incorrectly
Krasodomska, K., Lubiński, W., Potemkowski, A., Honczarenko, K., 2010. Pattern scaled optical coherence tomography angiography images in peer-reviewed studies:
electroretinogram (PERG) and pattern visual evoked potential (PVEP) in the early a systematic review. JAMA Ophthalmol. https://doi.org/10.1001/
stages of Alzheimer’s disease. Doc. Ophthalmol. 121 (2), 111–121. jamaophthalmol.2019.4821. Online ahead of print.
Kraus, M.F., Potsaid, B., Mayer, M.A., Bock, R., Baumann, B., Liu, J.J., Hornegger, J., Longstreth, W., Larsen, E.K., Klein, R., Wong, T.Y., Sharrett, A.R., Lefkowitz, D.,
Fujimoto, J.G., 2012. Motion correction in optical coherence tomography volumes Manolio, T.A., 2007. Associations between findings on cranial magnetic resonance
on a per A-scan basis using orthogonal scan patterns. Biomed. Optic Express 3 (6), imaging and retinal photography in the elderly: the Cardiovascular Health Study.
1182–1199. Am. J. Epidemiol. 165 (1), 78–84.

34
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Luan, F., Wu, Y., 2013. Application of RPCA in optical coherence tomography for speckle neuropathologic evaluation of Alzheimer’s disease. Alzheimers Dement. 12 (2),
noise reduction. Laser Phys. Lett. 10 (3), 035603. 164–169.
Lynch, S.K., Abràmoff, M.D., 2017. Diabetic retinopathy is a neurodegenerative disorder. Mordant, D.J., Al-Abboud, I., Muyo, G., Gorman, A., Sallam, A., Ritchie, P., Harvey, A.R.,
Vis. Res. 139, 101–107. McNaught, A.I., 2011. Spectral imaging of the retina. Eye 25 (3), 309–320.
López-de-Eguileta, A., Lage, C., López-García, S., Pozueta, A., García-Martínez, M., More, S.S., Beach, J.M., McClelland, C., Mokhtarzadeh, A., Vince, R., 2019. In vivo
Kazimierczak, M., Bravo, M., de Arcocha-Torres, M., Banzo, I., Jimenez-Bonilla, J., assessment of retinal biomarkers by hyperspectral imaging: early detection of
Cerveró, A., Goikoetxea, A., Rodríguez-Rodríguez, E., Sánchez-Juan, P., Casado, A., Alzheimer’s disease. ACS Chem. Neurosci. 10 (11), 4492–4501.
2020. Evaluation of choroidal thickness in prodromal Alzheimer’s disease defined by Moschos, M.M., Chatziralli, I.P., 2018. Evaluation of choroidal and retinal thickness
amyloid PET. PloS One 15 (9), e0239484. changes in Parkinson’s disease using spectral domain optical coherence tomography.
Löffler, K.U., Edward, D.P., Tso, M.O., 1995. Immunoreactivity against tau, amyloid Semin. Ophthalmol. 33 (4), 494–497.
precursor protein, and beta-amyloid in the human retina. Invest. Ophthalmol. Vis. Moschos, M.M., Markopoulos, I., Chatziralli, I., Rouvas, A., Papageorgiou, S.G., Ladas, I.,
Sci. 36 (1), 24–31. Vassilopoulos, D., 2012. Structural and functional impairment of the retina and optic
Mailankody, P., Battu, R., Khanna, A., Lenka, A., Yadav, R., Pal, P.K., 2015. Optical nerve in Alzheimer’s disease. Curr. Alzheimer Res. 9 (7), 782–788.
coherence tomography as a tool to evaluate retinal changes in Parkinson’s disease. Moschos, M.M., Tagaris, G., Markopoulos, I., Margetis, I., Tsapakis, S., Kanakis, M.,
Park. Relat. Disord. 21 (10), 1164–1169. Koutsandrea, C., 2011. Morphologic changes and functional retinal impairment in
Marquié, M., Valero, S., Castilla-Marti, M., Martínez, J., Rodríguez-Gómez, O., patients with Parkinson disease without visual loss. Eur. J. Ophthalmol. 21 (1),
Sanabria, Á., Tartari, J.P., Monté-Rubio, G.C., Sotolongo-Grau, O., Alegret, M., 24–29.
Pérez-Cordón, A., Roberto, N., de Rojas, I., Moreno-Grau, S., Montrreal, L., Moulder, K.L., Snider, B.J., Mills, S.L., Buckles, V.D., Santacruz, A.M., Bateman, R.J.,
Hernández, I., Rosende-Roca, M., Mauleón, A., Vargas, L., Abdelnour, C., Gil, S., Morris, J.C., 2013. Dominantly Inherited Alzheimer Network: facilitating research
Esteban-De Antonio, E., Espinosa, A., Ortega, G., Lomeña, F., Pavia, J., Vivas, A., and clinical trials. Alzheimer’s Res. Ther. 5 (5), 48.
Tejero, M., Gómez-Chiari, M., Simó, R., Ciudin, A., Hernández, C., Orellana, A., Murphy, M.J., Wei, Z., Fatyga, M., Williamson, J., Anscher, M., Wallace, T., Weiss, E.,
Benaque, A., Ruiz, A., Tárraga, L., Boada, M., group, F.s., 2020. Association between 2008. How does CT image noise affect 3D deformable image registration for image-
retinal thickness and β-amyloid brain accumulation in individuals with subjective guided radiotherapy planning? Med. Phys. 35 (3), 1145–1153.
cognitive decline: fundació ACE Healthy Brain Initiative. Alzheimer’s Res. Ther. 12 Murphy, O.C., Kwakyi, O., Iftikhar, M., Zafar, S., Lambe, J., Pellegrini, N., Sotirchos, E.S.,
(1), 37. Gonzalez-Caldito, N., Ogbuokiri, E., Filippatou, A., Risher, H., Cowley, N.,
Martinez-Lapiscina, E.H., Arnow, S., Wilson, J.A., Saidha, S., Preiningerova, J.L., Feldman, S., Fioravante, N., Frohman, E.M., Frohman, T.C., Balcer, L.J., Prince, J.L.,
Oberwahrenbrock, T., Brandt, A.U., Pablo, L.E., Guerrieri, S., Gonzalez, I., Channa, R., Calabresi, P.A., Saidha, S., 2020. Alterations in the retinal vasculature
Outteryck, O., Mueller, A.K., Albrecht, P., Chan, W., Lukas, S., Balk, L.J., Fraser, C., occur in multiple sclerosis and exhibit novel correlations with disability and visual
Frederiksen, J.L., Resto, J., Frohman, T., Cordano, C., Zubizarreta, I., Andorra, M., function measures. Mult. Scler. 26 (7), 815–828.
Sanchez-Dalmau, B., Saiz, A., Bermel, R., Klistorner, A., Petzold, A., Schippling, S., Mutlu, U., Colijn, J.M., Ikram, M.A., Bonnemaijer, P.W.M., Licher, S., Wolters, F.J.,
Costello, F., Aktas, O., Vermersch, P., Oreja-Guevara, C., Comi, G., Leocani, L., Tiemeier, H., Koudstaal, P.J., Klaver, C.C.W., Ikram, M.K., 2018. Association of
Garcia-Martin, E., Paul, F., Havrdova, E., Frohman, E., Balcer, L.J., Green, A.J., retinal neurodegeneration on optical coherence tomography with dementia: a
Calabresi, P.A., Villoslada, P., consortium, I., 2016. Retinal thickness measured with population-based study. JAMA Neurol. 75 (10), 1256–1263.
optical coherence tomography and risk of disability worsening in multiple sclerosis: Müller, A.K., Blasberg, C., Südmeyer, M., Aktas, O., Albrecht, P., 2014. Photoreceptor
a cohort study. Lancet Neurol. 15 (6), 574–584. layer thinning in parkinsonian syndromes. Mov. Disord. 29 (9), 1222–1223.
Marziani, E., Pomati, S., Ramolfo, P., Cigada, M., Giani, A., Mariani, C., Staurenghi, G., Neary, D., Snowden, J., Mann, D., 2005. Frontotemporal dementia. Lancet Neurol. 4
2013. Evaluation of retinal nerve fiber layer and ganglion cell layer thickness in (11), 771–780.
Alzheimer’s disease using spectral-domain optical coherence tomography. Invest. Nelis, P., Kleffner, I., Burg, M.C., Clemens, C.R., Alnawaiseh, M., Motte, J.,
Ophthalmol. Vis. Sci. 54 (9), 5953–5958. Marziniak, M., Eter, N., Alten, F., 2018. OCT-Angiography reveals reduced vessel
Matej, R., Tesar, A., Rusina, R., 2019. Alzheimer’s disease and other neurodegenerative density in the deep retinal plexus of CADASIL patients. Sci. Rep. 8 (1), 8148.
dementias in comorbidity: a clinical and neuropathological overview. Clin. Biochem. Nelson, P.T., Alafuzoff, I., Bigio, E.H., Bouras, C., Braak, H., Cairns, N.J., Castellani, R.J.,
73, 26–31. Crain, B.J., Davies, P., Del Tredici, K., Duyckaerts, C., Frosch, M.P., Haroutunian, V.,
Matsunaga, D., Yi, J., Puliafito, C.A., Kashani, A.H., 2014. OCT angiography in healthy Hof, P.R., Hulette, C.M., Hyman, B.T., Iwatsubo, T., Jellinger, K.A., Jicha, G.A.,
human subjects. Ophthalmic Surg. Lasers Imag. Retina 45 (6), 510–515. Kövari, E., Kukull, W.A., Leverenz, J.B., Love, S., Mackenzie, I.R., Mann, D.M.,
Matthews, W.B., Small, D.G., Small, M., Pountney, E., 1977. Pattern reversal evoked Masliah, E., McKee, A.C., Montine, T.J., Morris, J.C., Schneider, J.A., Sonnen, J.A.,
visual potential in the diagnosis of multiple sclerosis. J. Neurol. Neurosurg. Thal, D.R., Trojanowski, J.Q., Troncoso, J.C., Wisniewski, T., Woltjer, R.L., Beach, T.
Psychiatry 40 (10), 1009–1014. G., 2012. Correlation of Alzheimer disease neuropathologic changes with cognitive
Mauschitz, M.M., Bonnemaijer, P.W.M., Diers, K., Rauscher, F.G., Elze, T., Engel, C., status: a review of the literature. J. Neuropathol. Exp. Neurol. 71 (5), 362–381.
Loeffler, M., Colijn, J.M., Ikram, M.A., Vingerling, J.R., Williams, K.M., Nguyen-Legros, J., 1988. Functional neuroarchitecture of the retina: hypothesis on the
Hammond, C.J., Creuzot-Garcher, C., Bron, A.M., Silva, R., Nunes, S., Delcourt, C., dysfunction of retinal dopaminergic circuitry in Parkinson’s disease. Surg. Radiol.
Cougnard-Grégoire, A., Holz, F.G., Klaver, C.C.W., Breteler, M.M.B., Finger, R.P., Anat. 10 (2), 137–144.
Consortium, E.E.E.E., 2018. Systemic and ocular determinants of peripapillary Nightingale, S., Mitchell, K.W., Howe, J.W., 1986. Visual evoked cortical potentials and
retinal nerve fiber layer thickness measurements in the European eye epidemiology pattern electroretinograms in Parkinson’s disease and control subjects. J. Neurol.
(E3) population. Ophthalmology 125 (10), 1526–1536. Neurosurg. Psychiatry 49 (11), 1280–1287.
Mayer, M.A., Borsdorf, A., Wagner, M., Hornegger, J., Mardin, C.Y., Tornow, R.P., 2012. Nolan-Kenney, R.C., Liu, M., Akhand, O., Calabresi, P.A., Paul, F., Petzold, A., Balk, L.,
Wavelet denoising of multiframe optical coherence tomography data. Biomed. Optic Brandt, A.U., Martínez-Lapiscina, E.H., Saidha, S., Villoslada, P., Al-Hassan, A.A.,
Express 3 (3), 572–589. Behbehani, R., Frohman, E.M., Frohman, T., Havla, J., Hemmer, B., Jiang, H.,
McGrory, S., Ballerini, L., Doubal, F.N., Staals, J., Allerhand, M., Valdes-Hernandez, M.D. Knier, B., Korn, T., Leocani, L., Papadopoulou, A., Pisa, M., Zimmermann, H.,
C., Wang, X., MacGillivray, T., Doney, A.S.F., Dhillon, B., Starr, J.M., Bastin, M.E., Galetta, S.L., Balcer, L.J., I. M. S. V. S. Consortium, 2019. Optimal intereye difference
Trucco, E., Deary, I.J., Wardlaw, J.M., 2019. Retinal microvasculature and cerebral thresholds by optical coherence tomography in multiple sclerosis: an international
small vessel disease in the Lothian nirth cohort 1936 and mild stroke study. Sci. Rep. study. Ann. Neurol. 85 (5), 618–629.
9 (1), 6320. Nowacka, B., Lubiński, W., Honczarenko, K., Potemkowski, A., Safranow, K., 2015.
McKhann, G., Drachman, D., Folstein, M., Katzman, R., Price, D., Stadlan, E.M., 1984. Bioelectrical function and structural assessment of the retina in patients with early
Clinical diagnosis of Alzheimer’s disease: report of the NINCDS-ADRDA work group stages of Parkinson’s disease (PD). Doc. Ophthalmol. 131 (2), 95–104.
under the auspices of department of health and human services task force on O’Bryhim, B.E., Apte, R.S., Kung, N., Coble, D., Van Stavern, G.P., 2018. Association of
Alzheimer’s disease. Neurology 34 (7), 939–944. preclinical Alzheimer disease with optical coherence tomographic angiography
Meier, P.G., Maeder, P., Kardon, R.H., Borruat, F.X., 2015. Homonymous ganglion cell findings. JAMA Ophthalmol. 136 (11), 1242–1248.
layer thinning after isolated occipital lesion: macular OCT demonstrates Obis, J., Garcia-Martin, E., Orduna, E., Vilades, E., Alarcia, R., Rodrigo, M.J., Pablo, L.E.,
transsynaptic retrograde retinal degeneration. J. Neuro Ophthalmol. 35 (2), Polo, V., Larrosa, J.M., Satue, M., 2020. Reproducibility of retinal and choroidal
112–116. measurements using swept-source optical coherence tomography in patients with
Miller, D.H., Newton, M.R., van der Poel, J.C., du Boulay, E.P., Halliday, A.M., Parkinson’s disease. Arq. Bras. Oftalmol. 83 (1), 19–27.
Kendall, B.E., Johnson, G., MacManus, D.G., Moseley, I.F., McDonald, W.I., 1988. Oepen, G., Doerr, M., Thoden, U., 1981. Visual (VEP) and somatosensory (SSEP) evoked
Magnetic resonance imaging of the optic nerve in optic neuritis. Neurology 38 (2), potentials in Huntington’s chorea. Electroencephalogr. Clin. Neurophysiol. 51 (6),
175–179. 666–670.
Miri, S., Shrier, E.M., Glazman, S., Ding, Y., Selesnick, I., Kozlowski, P.B., Bodis- Ohno-Matsui, K., 2011. Parallel findings in age-related macular degeneration and
Wollner, I., 2015. The avascular zone and neuronal remodeling of the fovea in Alzheimer’s disease. Prog. Retin. Eye Res. 30 (4), 217–238.
Parkinson disease. Ann. Clin. Transl. Neurol. 2 (2), 196–201. Oliveira, F.P., Tavares, J.M., 2014. Medical image registration: a review. Comput.
Monet-Leprêtre, M., Bardot, B., Lemaire, B., Domenga, V., Godin, O., Dichgans, M., Methods Biomech. Biomed. Eng. 17 (2), 73–93.
Tournier-Lasserve, E., Cohen-Tannoudji, M., Chabriat, H., Joutel, A., 2009. Distinct Ong, Y.T., De Silva, D.A., Cheung, C.Y., Chang, H.M., Chen, C.P., Wong, M.C., Wong, T.
phenotypic and functional features of CADASIL mutations in the Notch3 ligand Y., Ikram, M.K., 2013. Microvascular structure and network in the retina of patients
binding domain. Brain 132 (Pt 6), 1601–1612. with ischemic stroke. Stroke 44 (8), 2121–2127.
Montine, T.J., Monsell, S.E., Beach, T.G., Bigio, E.H., Bu, Y., Cairns, N.J., Frosch, M., Onofrj, M., Ghilardi, M.F., Basciani, M., Gambi, D., 1986. Visual evoked potentials in
Henriksen, J., Kofler, J., Kukull, W.A., Lee, E.B., Nelson, P.T., Schantz, A.M., parkinsonism and dopamine blockade reveal a stimulus-dependent dopamine
Schneider, J.A., Sonnen, J.A., Trojanowski, J.Q., Vinters, H.V., Zhou, X.H., function in humans. J. Neurol. Neurosurg. Psychiatry 49 (10), 1150–1159.
Hyman, B.T., 2016. Multisite assessment of NIA-AA guidelines for the

35
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Ortuño-Lizarán, I., Beach, T.G., Serrano, G.E., Walker, D.G., Adler, C.H., Cuenca, N., glaucoma, ischemic neuropathy, papilledema, and toxic neuropathy. Arch.
2018. Phosphorylated α-synuclein in the retina is a biomarker of Parkinson’s disease Ophthalmol. 100 (1), 135–146.
pathology severity. Mov. Disord. 33 (8), 1315–1324. Ragauskas, S., Leinonen, H., Puranen, J., Rönkkö, S., Nymark, S., Gurevicius, K.,
Otsu, N., 1979. A threshold selection method from gray-level histograms. IEEE Trans. Lipponen, A., Kontkanen, O., Puoliväli, J., Tanila, H., Kalesnykas, G., 2014. Early
Syst. Man Cybern. 9 (1), 62–66. retinal function deficit without prominent morphological changes in the R6/2 mouse
Pantoni, L., 2010. Cerebral small vessel disease: from pathogenesis and clinical model of Huntington’s disease. PloS One 9 (12), e113317.
characteristics to therapeutic challenges. Lancet Neurol. 9 (7), 689–701. Ravalico, G., Toffoli, G., Pastori, G., Crocè, M., Calderini, S., 1996. Age-related ocular
Panwar, N., Huang, P., Lee, J., Keane, P.A., Chuan, T.S., Richhariya, A., Teoh, S., Lim, T. blood flow changes. Invest. Ophthalmol. Vis. Sci. 37 (13), 2645–2650.
H., Agrawal, R., 2016. Fundus photography in the 21st century–A review of recent Regan, D., Neima, D., 1984. Visual fatigue and visual evoked potentials in multiple
technological advances and their implications for worldwide healthcare. Telemed. J. sclerosis, glaucoma, ocular hypertension and Parkinson’s disease. J. Neurol.
e Health 22 (3), 198–208. Neurosurg. Psychiatry 47 (7), 673–678.
Paquet, C., Boissonnot, M., Roger, F., Dighiero, P., Gil, R., Hugon, J., 2007. Abnormal Rensma, S.P., van Sloten, T.T., Launer, L.J., Stehouwer, C.D.A., 2018. Cerebral small
retinal thickness in patients with mild cognitive impairment and Alzheimer’s vessel disease and risk of incident stroke, dementia and depression, and all-cause
disease. Neurosci. Lett. 420 (2), 97–99. mortality: a systematic review and meta-analysis. Neurosci. Biobehav. Rev. 90,
Parisi, V., Manni, G., Spadaro, M., Colacino, G., Restuccia, R., Marchi, S., Bucci, M.G., 164–173.
Pierelli, F., 1999. Correlation between morphological and functional retinal Reshef, E.R., Miller, J.B., Vavvas, D.G., 2020. Hyperspectral imaging of the retina: a
impairment in multiple sclerosis patients. Invest. Ophthalmol. Vis. Sci. 40 (11), review. Int. Ophthalmol. Clin. 60 (1), 85–96.
2520–2527. Rimmer, S., Iragui, V., Klauber, M.R., Katz, B., 1989. Retinocortical time exhibits spatial
Parisi, V., Restuccia, R., Fattapposta, F., Mina, C., Bucci, M.G., Pierelli, F., 2001. selectivity. Invest. Ophthalmol. Vis. Sci. 30 (9), 2045–2049.
Morphological and functional retinal impairment in Alzheimer’s disease patients. Ringman, J.M., Goate, A., Masters, C.L., Cairns, N.J., Danek, A., Graff-Radford, N.,
Clin. Neurophysiol. 112 (10), 1860–1867. Ghetti, B., Morris, J.C., Network, D.I.A., 2014. Genetic heterogeneity in Alzheimer
Paulus, W., Schwarz, G., Werner, A., Lange, H., Bayer, A., Hofschuster, M., Müller, N., disease and implications for treatment strategies. Curr. Neurol. Neurosci. Rep. 14
Zrenner, E., 1993. Impairment of retinal increment thresholds in Huntington’s (11), 499.
disease. Ann. Neurol. 34 (4), 574–578. Ringman, J.M., Younkin, S.G., Pratico, D., Seltzer, W., Cole, G.M., Geschwind, D.H.,
Pearl, J.R., Heath, L.M., Bergey, D.E., Kelly, J.P., Smith, C., Laurino, M.Y., Weiss, A., Rodriguez-Agudelo, Y., Schaffer, B., Fein, J., Sokolow, S., Rosario, E.R., Gylys, K.H.,
Price, N.D., LaSpada, A., Bird, T.D., Jayadev, S., 2017. Enhanced retinal responses in Varpetian, A., Medina, L.D., Cummings, J.L., 2008. Biochemical markers in persons
Huntington’s disease patients. J. Huntingtons Dis. 6 (3), 237–247. with preclinical familial Alzheimer disease. Neurology 71 (2), 85–92.
Peca, S., McCreary, C.R., Donaldson, E., Kumarpillai, G., Shobha, N., Sanchez, K., Rizzo, P.A., Pierelli, F., Pozzessere, G., Morocutti, C., 1980. Pattern visual evoked
Charlton, A., Steinback, C.D., Beaudin, A.E., Flück, D., Pillay, N., Fick, G.H., potentials in Huntington’s disease. Acta Neurol. 2 (2), 128–131.
Poulin, M.J., Frayne, R., Goodyear, B.G., Smith, E.E., 2013. Neurovascular Rohani, M., Langroodi, A.S., Ghourchian, S., Falavarjani, K.G., SoUdi, R., Shahidi, G.,
decoupling is associated with severity of cerebral amyloid angiopathy. Neurology 81 2013. Retinal nerve changes in patients with tremor dominant and akinetic rigid
(19), 1659–1665. Parkinson’s disease. Neurol. Sci. 34 (5), 689–693.
Peng, C., Trojanowski, J.Q., Lee, V.M., 2020. Protein transmission in neurodegenerative Ross, C.A., Aylward, E.H., Wild, E.J., Langbehn, D.R., Long, J.D., Warner, J.H., Scahill, R.
disease. Nat. Rev. Neurol. 16 (4), 199–212. I., Leavitt, B.R., Stout, J.C., Paulsen, J.S., Reilmann, R., Unschuld, P.G., Wexler, A.,
Peppe, A., Stanzione, P., Pierelli, F., De Angelis, D., Pierantozzi, M., Bernardi, G., 1995. Margolis, R.L., Tabrizi, S.J., 2014. Huntington disease: natural history, biomarkers
Visual alterations in de novo Parkinson’s disease: pattern electroretinogram latencies and prospects for therapeutics. Nat. Rev. Neurol. 10 (4), 204–216.
are more delayed and more reversible by levodopa than are visual evoked potentials. Roth, N.M., Saidha, S., Zimmermann, H., Brandt, A.U., Isensee, J., Benkhellouf-
Neurology 45 (6), 1144–1148. Rutkowska, A., Dornauer, M., Kühn, A.A., Müller, T., Calabresi, P.A., Paul, F., 2014.
Petrasch-Parwez, E., Saft, C., Schlichting, A., Andrich, J., Napirei, M., Arning, L., Photoreceptor layer thinning in idiopathic Parkinson’s disease. Mov. Disord. 29 (9),
Wieczorek, S., Dermietzel, R., Epplen, J.T., 2005. Is the retina affected in Huntington 1163–1170.
disease? Acta Neuropathol. 110 (5), 523–525. Ryman, D.C., Acosta-Baena, N., Aisen, P.S., Bird, T., Danek, A., Fox, N.C., Goate, A.,
Petzold, A., Balcer, L.J., Calabresi, P.A., Costello, F., Frohman, T.C., Frohman, E.M., Frommelt, P., Ghetti, B., Langbaum, J.B., Lopera, F., Martins, R., Masters, C.L.,
Martinez-Lapiscina, E.H., Green, A.J., Kardon, R., Outteryck, O., Paul, F., Mayeux, R.P., McDade, E., Moreno, S., Reiman, E.M., Ringman, J.M., Salloway, S.,
Schippling, S., Vermersch, P., Villoslada, P., Balk, L.J., Imsvisual, E.-E., 2017. Retinal Schofield, P.R., Sperling, R., Tariot, P.N., Xiong, C., Morris, J.C., Bateman, R.J.,
layer segmentation in multiple sclerosis: a systematic review and meta-analysis. Network, D.I.A., 2014. Symptom onset in autosomal dominant Alzheimer disease: a
Lancet Neurol. 16 (10), 797–812. systematic review and meta-analysis. Neurology 83 (3), 253–260.
Petzold, A., de Boer, J.F., Schippling, S., Vermersch, P., Kardon, R., Green, A., Sachdev, P., Kalaria, R., O’Brien, J., Skoog, I., Alladi, S., Black, S.E., Blacker, D.,
Calabresi, P.A., Polman, C., 2010. Optical coherence tomography in multiple Blazer, D.G., Chen, C., Chui, H., Ganguli, M., Jellinger, K., Jeste, D.V., Pasquier, F.,
sclerosis: a systematic review and meta-analysis. Lancet Neurol. 9 (9), 921–932. Paulsen, J., Prins, N., Rockwood, K., Roman, G., Scheltens, P., I. S. f. V. B. a. C.
Pham, T.Q., Kifley, A., Mitchell, P., Wang, J.J., 2006. Relation of age-related macular Disorders, 2014. Diagnostic criteria for vascular cognitive disorders: a VASCOG
degeneration and cognitive impairment in an older population. Gerontology 52 (6), statement. Alzheimer Dis. Assoc. Disord. 28 (3), 206–218.
353–358. Sadda, S.R., Borrelli, E., Fan, W., Ebraheem, A., Marion, K.M., Harrington, M., Kwon, S.,
Pilat, A., McLean, R.J., Proudlock, F.A., Maconachie, G.D., Sheth, V., Rajabally, Y.A., 2019. A pilot study of fluorescence lifetime imaging ophthalmoscopy in preclinical
Gottlob, I., 2016. In vivo morphology of the optic nerve and retina in patients with Alzheimer’s disease. Eye 33 (8), 1271–1279.
Parkinson’s disease. Invest. Ophthalmol. Vis. Sci. 57 (10), 4420–4427. Sadun, A.A., Bassi, C.J., 1990. Optic nerve damage in Alzheimer’s disease.
Pillai, J.A., Bermel, R., Bonner-Jackson, A., Rae-Grant, A., Fernandez, H., Bena, J., Ophthalmology 97 (1), 9–17.
Jones, S.E., Ehlers, J.P., Leverenz, J.B., 2016. Retinal nerve fiber layer thinning in Sadun, A.A., Borchert, M., DeVita, E., Hinton, D.R., Bassi, C.J., 1987. Assessment of
Alzheimer’s disease: a case-control study in comparison to normal aging, Parkinson’s visual impairment in patients with Alzheimer’s disease. Am. J. Ophthalmol. 104 (2),
disease, and non-Alzheimer’s dementia. Am. J. Alzheimers Dis. Other Demen. 31 (5), 113–120.
430–436. Saidha, S., Naismith, R.T., 2019. Optical coherence tomography for diagnosing optic
Pingale, T., Gupta, G.L., 2020. Classic and evolving animal models in Parkinson’s neuritis: are we there yet? Neurology 92 (6), 253–254.
disease. Pharmacol. Biochem. Behav. 199, 173060. Saidha, S., Sotirchos, E.S., Ibrahim, M.A., Crainiceanu, C.M., Gelfand, J.M., Sepah, Y.J.,
Polo, V., Satue, M., Rodrigo, M.J., Otin, S., Alarcia, R., Bambo, M.P., Fuertes, M.I., Ratchford, J.N., Oh, J., Seigo, M.A., Newsome, S.D., Balcer, L.J., Frohman, E.M.,
Larrosa, J.M., Pablo, L.E., Garcia-Martin, E., 2016. Visual dysfunction and its Green, A.J., Nguyen, Q.D., Calabresi, P.A., 2012. Microcystic macular oedema,
correlation with retinal changes in patients with Parkinson’s disease: an thickness of the inner nuclear layer of the retina, and disease characteristics in
observational cross-sectional study. BMJ Open 6 (5), e009658. multiple sclerosis: a retrospective study. Lancet Neurol. 11 (11), 963–972.
Pringsheim, T., Wiltshire, K., Day, L., Dykeman, J., Steeves, T., Jette, N., 2012. The Saidha, S., Sotirchos, E.S., Oh, J., Syc, S.B., Seigo, M.A., Shiee, N., Eckstein, C.,
incidence and prevalence of Huntington’s disease: a systematic review and meta- Durbin, M.K., Oakley, J.D., Meyer, S.A., Frohman, T.C., Newsome, S., Ratchford, J.
analysis. Mov. Disord. 27 (9), 1083–1091. N., Balcer, L.J., Pham, D.L., Crainiceanu, C.M., Frohman, E.M., Reich, D.S.,
Promjunyakul, N.O., Dodge, H.H., Lahna, D., Boespflug, E.L., Kaye, J.A., Rooney, W.D., Calabresi, P.A., 2013. Relationships between retinal axonal and neuronal measures
Silbert, L.C., 2018. Baseline NAWM structural integrity and CBF predict and global central nervous system pathology in multiple sclerosis. JAMA Neurol. 70
periventricular WMH expansion over time. Neurology 90 (24), e2119–e2126. (1), 34–43.
Qin, T., Prins, S., Groeneveld, G.J., Van Westen, G., de Vries, H.E., Wong, Y.C., Salobrar-García, E., de Hoz, R., Ramírez, A.I., López-Cuenca, I., Rojas, P., Vazirani, R.,
Bischoff, L.J.M., de Lange, E.C.M., 2020. Utility of animal models to understand Amarante, C., Yubero, R., Gil, P., Pinazo-Durán, M.D., Salazar, J.J., Ramírez, J.M.,
human Alzheimer’s disease, using the mastermind research approach to avoid 2019. Changes in visual function and retinal structure in the progression of
unnecessary further sacrifices of animals. Int. J. Mol. Sci. 21 (9). Alzheimer’s disease. PloS One 14 (8), e0220535.
Qu, M., Kwapong, W.R., Peng, C., Cao, Y., Lu, F., Shen, M., Han, Z., 2020. Retinal Santangelo, R., Huang, S.C., Bernasconi, M.P., Falautano, M., Comi, G., Magnani, G.,
sublayer defect is independently associated with the severity of hypertensive white Leocani, L., 2020. Neuro-retina might reflect Alzheimer’s disease stage.
matter hyperintensity. Brain Behav. 10 (2), e01521. J. Alzheimers Dis. 77 (4), 1455–1468.
Querques, G., Borrelli, E., Sacconi, R., De Vitis, L., Leocani, L., Santangelo, R., Santos, C.Y., Johnson, L.N., Sinoff, S.E., Festa, E.K., Heindel, W.C., Snyder, P.J., 2018.
Magnani, G., Comi, G., Bandello, F., 2019. Functional and morphological changes of Change in retinal structural anatomy during the preclinical stage of Alzheimer’s
the retinal vessels in Alzheimer’s disease and mild cognitive impairment. Sci. Rep. 9 disease. Alzheimers Dement. (Amst.) 10, 196–209.
(1), 63. Sarabi, M.S., Gahm, J.K., Khansari, M.M., Zhang, J., Kashani, A.H., Shi, Y., 2019. An
Quigley, H.A., 2011. Glaucoma. Lancet 377 (9774), 1367–1377. Automated 3D Analysis Framework for Optical Coherence Tomography
Quigley, H.A., Addicks, E.M., Green, W.R., 1982. Optic nerve damage in human Angiography. bioRxiv, p. 655175.
glaucoma. III. Quantitative correlation of nerve fiber loss and visual field defect in

36
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Sari, E.S., Koc, R., Yazici, A., Sahin, G., Ermis, S.S., 2015. Ganglion cell-inner plexiform Smith, E.E., Vijayappa, M., Lima, F., Delgado, P., Wendell, L., Rosand, J., Greenberg, S.
layer thickness in patients with Parkinson disease and association with disease M., 2008. Impaired visual evoked flow velocity response in cerebral amyloid
severity and duration. J. Neuro Ophthalmol. 35 (2), 117–121. angiopathy. Neurology 71 (18), 1424–1430.
Sartucci, F., Borghetti, D., Bocci, T., Murri, L., Orsini, P., Porciatti, V., Origlia, N., Snyder, H.M., Corriveau, R.A., Craft, S., Faber, J.E., Greenberg, S.M., Knopman, D.,
Domenici, L., 2010. Dysfunction of the magnocellular stream in Alzheimer’s disease Lamb, B.T., Montine, T.J., Nedergaard, M., Schaffer, C.B., Schneider, J.A.,
evaluated by pattern electroretinograms and visual evoked potentials. Brain Res. Wellington, C., Wilcock, D.M., Zipfel, G.J., Zlokovic, B., Bain, L.J., Bosetti, F.,
Bull. 82 (3–4), 169–176. Galis, Z.S., Koroshetz, W., Carrillo, M.C., 2015. Vascular contributions to cognitive
Sato, Y., Nakajima, S., Shiraga, N., Atsumi, H., Yoshida, S., Koller, T., Gerig, G., impairment and dementia including Alzheimer’s disease. Alzheimers Dement. 11
Kikinis, R., 1998. Three-dimensional multi-scale line filter for segmentation and (6), 710–717.
visualization of curvilinear structures in medical images. Med. Image Anal. 2 (2), Sodi, A., Lenzetti, C., Bacherini, D., Finocchio, L., Verdina, T., Borg, I., Cipollini, F.,
143–168. Patwary, F.U., Tanini, I., Zoppetti, C., Rizzo, S., Virgili, G., 2019a. Quantitative
Satue, M., Obis, J., Rodrigo, M.J., Otin, S., Fuertes, M.I., Vilades, E., Gracia, H., Ara, J.R., analysis of conjunctival and retinal vessels in Fabry disease. J. Ophthalmol. 2019,
Alarcia, R., Polo, V., Larrosa, J.M., Pablo, L.E., Garcia-Martin, E., 2016. Optical 4696429. https://doi.org/10.1155/2019/4696429 eCollection 2019.
coherence tomography as a biomarker for diagnosis, progression, and prognosis of Sodi, A., Nicolosi, C., Vicini, G., Lenzetti, C., Virgili, G., Rizzo, S., 2019b. Computer-
neurodegenerative diseases. J. Ophthalmol. 2016, 8503859. assisted retinal vessel diameter evaluation in Fabry disease. Eur. J. Ophthalmol.
Satue, M., Rodrigo, M.J., Obis, J., Vilades, E., Gracia, H., Otin, S., Fuertes, M.I., 1120672119886985.
Alarcia, R., Crespo, J.A., Polo, V., Larrosa, J.M., Pablo, L.E., Garcia-Martin, E., 2017. Song, G., Steelman, Z.A., Finkelstein, S., Yang, Z., Martin, L., Chu, K.K., Farsiu, S.,
Evaluation of progressive visual dysfunction and retinal degeneration in patients Arshavsky, V.Y., Wax, A., 2020. Multimodal coherent imaging of retinal biomarkers
with Parkinson’s disease. Invest. Ophthalmol. Vis. Sci. 58 (2), 1151–1157. of Alzheimer’s disease in a mouse model. Sci. Rep. 10 (1), 7912.
Satue, M., Seral, M., Otin, S., Alarcia, R., Herrero, R., Bambo, M.P., Fuertes, M.I., Spaide, R.F., 2016. Volume-Rendered optical coherence tomography of retinal vein
Pablo, L.E., Garcia-Martin, E., 2014. Retinal thinning and correlation with functional occlusion pilot study. Am. J. Ophthalmol. 165, 133–144.
disability in patients with Parkinson’s disease. Br. J. Ophthalmol. 98 (3), 350–355. Spaide, R.F., Fujimoto, J.G., Waheed, N.K., 2015a. Image artifacts IN optical coherence
Schiess, N., Calabresi, P.A., 2016. Multiple sclerosis. Semin. Neurol. 36 (4), 350–356. tomography angiography. Retina 35 (11), 1–18.
Schneider, J.A., Aggarwal, N.T., Barnes, L., Boyle, P., Bennett, D.A., 2009a. The Spaide, R.F., Fujimoto, J.G., Waheed, N.K., 2015b. Image artifacts IN optical coherence
neuropathology of older persons with and without dementia from community versus tomography angiography. Retina 35 (11), 2163–2180.
clinic cohorts. J. Alzheimers Dis. 18 (3), 691–701. Spaide, R.F., Klancnik, J.M., Cooney, M.J., 2015c. Retinal vascular layers imaged by
Schneider, J.A., Arvanitakis, Z., Leurgans, S.E., Bennett, D.A., 2009b. The fluorescein angiography and optical coherence tomography angiography. JAMA
neuropathology of probable Alzheimer disease and mild cognitive impairment. Ann. Ophthalmol. 133 (1), 45–50.
Neurol. 66 (2), 200–208. Spaide, R.F., Fujimoto, J.G., Waheed, N.K., Sadda, S.R., Staurenghi, G., 2018. Optical
Schneider, J.A., Arvanitakis, Z., Bang, W., Bennett, D.A., 2007. Mixed brain pathologies coherence tomography angiography. Prog. Retin. Eye Res. 64, 1–55.
account for most dementia cases in community-dwelling older persons. Neurology Spaide, R.F., Suzuki, M., Yannuzzi, L.A., Matet, A., Behar-Cohen, F., 2017. Volume-
69 (24), 2197–2204. rendered angiographic and structural optical coherence tomography angiography of
Schneider, J.A., Wilson, R.S., Bienias, J.L., Evans, D.A., Bennett, D.A., 2004. Cerebral macular telangiectasia type 2. Retina 37 (3), 424–435.
infarctions and the likelihood of dementia from Alzheimer disease pathology. Spain, R.I., Liu, L., Zhang, X., Jia, Y., Tan, O., Bourdette, D., Huang, D., 2018. Optical
Neurology 62 (7), 1148–1155. coherence tomography angiography enhances the detection of optic nerve damage in
Schneider, M., Müller, H.P., Lauda, F., Tumani, H., Ludolph, A.C., Kassubek, J., multiple sclerosis. Br. J. Ophthalmol. 102 (4), 520–524.
Pinkhardt, E.H., 2014. Retinal single-layer analysis in Parkinsonian syndromes: an Sparrow, J.R., Duncker, T., Schuerch, K., Paavo, M., de Carvalho, J.R.L., 2020. Lessons
optical coherence tomography study. J. Neural. Transm. 121 (1), 41–47. learned from quantitative fundus autofluorescence. Prog. Retin. Eye Res. 74,
Schultz, N., Byman, E., Wennström, M., Bank, N.B., 2018. Levels of retinal IAPP are 100774.
altered in Alzheimer’s disease patients and correlate with vascular changes and Spencer, R.J., Wendell, C.R., Giggey, P.P., Katzel, L.I., Lefkowitz, D.M., Siegel, E.L.,
hippocampal IAPP levels. Neurobiol. Aging 69, 94–101. Waldstein, S.R., 2013. Psychometric limitations of the mini-mental state
Schwaber, E.J., Thompson, A.C., Smilnak, G., Stinnett, S.S., Whitson, H.E., Lad, E.M., examination among nondemented older adults: an evaluation of neurocognitive and
2020. Co-prevalence of Alzheimer’s disease and age-related macular degeneration magnetic resonance imaging correlates. Exp. Aging Res. 39 (4), 382–397.
established by histopathologic diagnosis. J. Alzheimers Dis. 76 (1), 207–215. Steel, D.H., Waldock, A., 1998. Measurement of the retinal nerve fibre layer with
Scinto, L.F., Daffner, K.R., Dressler, D., Ransil, B.I., Rentz, D., Weintraub, S., scanning laser polarimetry in patients with previous demyelinating optic neuritis.
Mesulam, M., Potter, H., 1994. A potential noninvasive neurobiological test for J. Neurol. Neurosurg. Psychiatry 64 (4), 505–509.
Alzheimer’s disease. Science 266 (5187), 1051–1054. Steen, C., D’haeseleer, M., Hoogduin, J.M., Fierens, Y., Cambron, M., Mostert, J.P.,
Sen, S., Saxena, R., Tripathi, M., Vibha, D., Dhiman, R., 2020a. Neurodegeneration in Heersema, D.J., Koch, M.W., De Keyser, J., 2013. Cerebral white matter blood flow
Alzheimer’s disease and glaucoma: overlaps and missing links. Eye 34 (9), and energy metabolism in multiple sclerosis. Mult. Scler. 19 (10), 1282–1289.
1546–1553. Stefánsson, E., Olafsdottir, O.B., Einarsdottir, A.B., Eliasdottir, T.S., Eysteinsson, T.,
Sen, S., Saxena, R., Vibha, D., Tripathi, M., Sharma, P., Phuljhele, S., Tandon, R., Vehmeijer, W., Vandewalle, E., Bek, T., Hardarson, S.H., 2017. Retinal oximetry
Kumar, P., 2020b. Detection of structural and electrical disturbances in macula and discovers novel biomarkers in retinal and brain diseases. Invest. Ophthalmol. Vis.
optic nerve in Alzheimer’s patients and their correlation with disease severity. Sci. 58 (6), BIO227–BIO233.
Semin. Ophthalmol. 35 (2), 116–125. Stemplewitz, B., Keserü, M., Bittersohl, D., Buhmann, C., Skevas, C., Richard, G.,
Sengillo, J.D., Winkler, E.A., Walker, C.T., Sullivan, J.S., Johnson, M., Zlokovic, B.V., Hassenstein, A., 2015 Dec. Scanning laser polarimetry and spectral domain optical
2013. Deficiency in mural vascular cells coincides with blood-brain barrier coherence tomography for the detection of retinal changes in Parkinson’s disease.
disruption in Alzheimer’s disease. Brain Pathol. 23 (3), 303–310. Acta Ophthalmol. 93 (8), e672–e677.
Sepulcre, J., Murie-Fernandez, M., Salinas-Alaman, A., García-Layana, A., Bejarano, B., Sultan, H., Kellogg, C., El-Annan, J., 2017. Retinal detachment and microangiopathy in
Villoslada, P., 2007. Diagnostic accuracy of retinal abnormalities in predicting mitochondrial encephalomyopathy, lactic acidosis, and stroke-like episodes
disease activity in MS. Neurology 68 (18), 1488–1494. syndrome. Can. J. Ophthalmol. 52 (6), e208–e211.
Shao, Y., Jiang, H., Wei, Y., Shi, Y., Shi, C., Wright, C.B., Sun, X., Vanner, E.A., Sung, M.S., Choi, S.M., Kim, J., Ha, J.Y., Kim, B.C., Heo, H., Park, S.W., 2019. Inner
Rodriguez, A.D., Lam, B.L., Rundek, T., Baumel, B.S., Gameiro, G.R., Dong, C., retinal thinning as a biomarker for cognitive impairment in de novo Parkinson’s
Wang, J., 2018. Visualization of focal thinning of the ganglion cell-inner plexiform disease. Sci. Rep. 9 (1), 11832.
layer in patients with mild cognitive impairment and Alzheimer’s disease. Szaruga, M., Veugelen, S., Benurwar, M., Lismont, S., Sepulveda-Falla, D., Lleo, A.,
J. Alzheimers Dis. 64 (4), 1261–1273. Ryan, N.S., Lashley, T., Fox, N.C., Murayama, S., Gijsen, H., De Strooper, B., Chávez-
Sharafi, S.M., Sylvestre, J.P., Chevrefils, C., Soucy, J.P., Beaulieu, S., Pascoal, T.A., Gutiérrez, L., 2015. Qualitative changes in human γ-secretase underlie familial
Arbour, J.D., Rhéaume, M.A., Robillard, A., Chayer, C., Rosa-Neto, P., Alzheimer’s disease. J. Exp. Med. 212 (12), 2003–2013.
Mathotaarachchi, S.S., Nasreddine, Z.S., Gauthier, S., Lesage, F., 2019. Vascular Szegedi, S., Dal-Bianco, P., Stögmann, E., Traub-Weidinger, T., Rainer, M., Masching, A.,
retinal biomarkers improves the detection of the likely cerebral amyloid status from Schmidl, D., Werkmeister, R.M., Chua, J., Schmetterer, L., Garhöfer, G., 2020 Nov.
hyperspectral retinal images. Alzheimers Dement. (N. Y.) 5, 610–617. Anatomical and functional changes in the retina in patients with Alzheimer’s disease
Sherman, T.F., 1981. On connecting large vessels to small. The meaning of Murray’s law. and mild cognitive impairment. Acta Ophthalmol. 98 (7), e914–e921.
J. Gen. Physiol. 78 (4), 431–453. Sánchez, D., Castilla-Marti, M., Marquié, M., Valero, S., Moreno-Grau, S., Rodríguez-
Shi, Y., Gahm, J., Kashani, A.H., 2017. Curvelet-based vessel enhancement for 3D OCT Gómez, O., Piferrer, A., Martínez, G., Martínez, J., Rojas, I., Hernández, I.,
angiography. Invest. Ophthalmol. Vis. Sci. 58 (8), 648-648. Abdelnour, C., Rosende-Roca, M., Vargas, L., Mauleón, A., Gil, S., Alegret, M.,
Shoda, C., Kitagawa, Y., Shimada, H., Yuzawa, M., Tateno, A., Okubo, Y., 2018. Ortega, G., Espinosa, A., Pérez-Cordón, A., Sanabria, Á., Roberto, N., Ciudin, A.,
Relationship of area of soft drusen in retina with cerebral amyloid-β accumulation Simó, R., Hernández, C., Tárraga, L., Boada, M., Ruiz, A., 2020. Evaluation of
and blood amyloid-β level in the elderly. J. Alzheimers Dis. 62 (1), 239–245. macular thickness and volume tested by optical coherence tomography as
Shrier, E.M., Adam, C.R., Spund, B., Glazman, S., Bodis-Wollner, I., 2012. Interocular biomarkers for Alzheimer’s disease in a memory clinic. Sci. Rep. 10 (1), 1580.
asymmetry of foveal thickness in Parkinson disease. J. Ophthalmol. 2012, 728457. Tagliati, M., Bodis-Wollner, I., Kovanecz, I., Stanzione, P., 1994. Spatial frequency tuning
Sidiqi, A., Wahl, D., Lee, S., Ma, D., To, E., Cui, J., Beg, M.F., Sarunic, M., Matsubara, J. of the monkey pattern ERG depends on D2 receptor-linked action of dopamine. Vis.
A., 2020. Retinal fluorescence imaging with curcumin in an Alzheimer mouse model. Res. 34 (16), 2051–2057.
Front. Neurosci. 14, 713. Tagliati, M., Bodis-Wollner, I., Yahr, M.D., 1996. The pattern electroretinogram in
Singer, M., Ashimatey, B.S., Zhou, X., Chu, Z., Wang, R., Kashani, A.H., 2020. Impaired Parkinson’s disease reveals lack of retinal spatial tuning. Electroencephalogr. Clin.
layer specific retinal vascular reactivity among diabetic subjects. PloS One 15 (9), Neurophysiol. 100 (1), 1–11.
e0233871.

37
A.H. Kashani et al. Progress in Retinal and Eye Research 83 (2021) 100938

Tatton, W.G., Kwan, M.M., Verrier, M.C., Seniuk, N.A., Theriault, E., 1990. MPTP Walter, S.D., Ishikawa, H., Galetta, K.M., Sakai, R.E., Feller, D.J., Henderson, S.B.,
produces reversible disappearance of tyrosine hydroxylase-containing retinal Wilson, J.A., Maguire, M.G., Galetta, S.L., Frohman, E., Calabresi, P.A., Schuman, J.
amacrine cells. Brain Res. 527 (1), 21–31. S., Balcer, L.J., 2012. Ganglion cell loss in relation to visual disability in multiple
Teng, C.C., Shapiro, L.G., Kalet, I.J., 2010. Head and neck lymph node region delineation sclerosis. Ophthalmology 119 (6), 1250–1257.
with image registration. Biomed. Eng. Online 9, 30. Wang, B., Camino, A., Pi, S., Guo, Y., Wang, J., Huang, D., Hwang, T.S., Jia, Y., 2019.
Tham, Y.C., Li, X., Wong, T.Y., Quigley, H.A., Aung, T., Cheng, C.Y., 2014. Global Three-dimensional structural and angiographic evaluation of foveal ischemia in
prevalence of glaucoma and projections of glaucoma burden through 2040: a diabetic retinopathy: method and validation. Biomed. Optic Express 10 (7),
systematic review and meta-analysis. Ophthalmology 121 (11), 2081–2090. 3522–3532.
The Huntington’s Disease Collaborative Research Group, 1993. A novel gene containing Wang, D., Li, Y., Wang, C., Xu, L., You, Q.S., Wang, Y.X., Zhao, L., Wei, W.B., Zhao, X.,
a trinucleotide repeat that is expanded and unstable on Huntington’s disease Jonas, J.B., 2014. Localized retinal nerve fiber layer defects and stroke. Stroke 45
chromosomes. Cell 72 (6), 971–983. (6), 1651–1656.
Thomson, K.L., Yeo, J.M., Waddell, B., Cameron, J.R., Pal, S., 2015. A systematic review Wardlaw, J.M., Smith, E.E., Biessels, G.J., Cordonnier, C., Fazekas, F., Frayne, R.,
and meta-analysis of retinal nerve fiber layer change in dementia, using optical Lindley, R.I., O’Brien, J.T., Barkhof, F., Benavente, O.R., Black, S.E., Brayne, C.,
coherence tomography. Alzheimers Dement. (Amst.) 1 (2), 136–143. Breteler, M., Chabriat, H., Decarli, C., de Leeuw, F.E., Doubal, F., Duering, M.,
Toga, A.W., Thompson, P.M., 2001. The role of image registration in brain mapping. Fox, N.C., Greenberg, S., Hachinski, V., Kilimann, I., Mok, V., Oostenbrugge, R.,
Image Vis Comput. 19 (1–2), 3–24. Pantoni, L., Speck, O., Stephan, B.C., Teipel, S., Viswanathan, A., Werring, D.,
Toledo, J., Sepulcre, J., Salinas-Alaman, A., García-Layana, A., Murie-Fernandez, M., Chen, C., Smith, C., van Buchem, M., Norrving, B., Gorelick, P.B., Dichgans, M.,
Bejarano, B., Villoslada, P., 2008. Retinal nerve fiber layer atrophy is associated with 2013. Neuroimaging standards for research into small vessel disease and its
physical and cognitive disability in multiple sclerosis. Mult. Scler. 14 (7), 906–912. contribution to ageing and neurodegeneration. Lancet Neurol. 12 (8), 822–838.
Toussaint, D., Périer, O., Verstappen, A., Bervoets, S., 1983. Clinicopathological study of Wei, W., Xia, Z., Gao, H., Gong, J., Yan, L., Huang, Y., Chen, F., Zhang, W., 2016.
the visual pathways, eyes, and cerebral hemispheres in 32 cases of disseminated Correlation of retinopathy with leukoaraiosis in patients with anterior circulation
sclerosis. J. Clin. Neuro Ophthalmol. 3 (3), 211–220. infarcts. J. Clin. Neurosci. 33, 105–110.
Trebbastoni, A., Marcelli, M., Mallone, F., D’Antonio, F., Imbriano, L., Campanelli, A., de Weil, R.S., Schrag, A.E., Warren, J.D., Crutch, S.J., Lees, A.J., Morris, H.R., 2016. Visual
Lena, C., Gharbiya, M., 2017. Attenuation of choroidal thickness in patients with dysfunction in Parkinson’s disease. Brain 139 (11), 2827–2843.
Alzheimer disease: evidence from an Italian prospective study. Alzheimer Dis. Assoc. Weinreb, R.N., Aung, T., Medeiros, F.A., 2014. The pathophysiology and treatment of
Disord. 31 (2), 128–134. glaucoma: a review. J. Am. Med. Assoc. 311 (18), 1901–1911.
Trost, A., Lange, S., Schroedl, F., Bruckner, D., Motloch, K.A., Bogner, B., Kaser- Williams, E.A., McGuone, D., Frosch, M.P., Hyman, B.T., Laver, N., Stemmer-
Eichberger, A., Strohmaier, C., Runge, C., Aigner, L., Rivera, F.J., Reitsamer, H.A., Rachamimov, A., 2017. Absence of Alzheimer disease neuropathologic changes in
2016. Brain and retinal pericytes: origin, function and role. Front. Cell. Neurosci. 10, eyes of subjects with Alzheimer disease. J. Neuropathol. Exp. Neurol. 76 (5),
20. 376–383.
Tsai, C.S., Ritch, R., Schwartz, B., Lee, S.S., Miller, N.R., Chi, T., Hsieh, F.Y., 1991. Optic Williams, M.A., McGowan, A.J., Cardwell, C.R., Cheung, C.Y., Craig, D., Passmore, P.,
nerve head and nerve fiber layer in Alzheimer’s disease. Arch. Ophthalmol. 109 (2), Silvestri, G., Maxwell, A.P., McKay, G.J., 2015. Retinal microvascular network
199–204. attenuation in Alzheimer’s disease. Alzheimers Dement. (Amst.) 1 (2), 229–235.
Tsironi, E.E., Dastiridou, A., Katsanos, A., Dardiotis, E., Veliki, S., Patramani, G., Witkovsky, P., 2004. Dopamine and retinal function. Doc. Ophthalmol. 108 (1), 17–40.
Zacharaki, F., Ralli, S., Hadjigeorgiou, G.M., 2012. Perimetric and retinal nerve fiber Wong, A., Mishra, A., Bizheva, K., Clausi, D.A., 2010. General Bayesian estimation for
layer findings in patients with Parkinson’s disease. BMC Ophthalmol. 12, 54. speckle noise reduction in optical coherence tomography retinal imagery. Optic
Tsoi, K.K., Chan, J.Y., Hirai, H.W., Wong, S.Y., Kwok, T.C., 2015. Cognitive tests to detect Express 18 (8), 8338–8352.
dementia: a systematic review and meta-analysis. JAMA Intern. Med. 175 (9), Wong, T.Y., Wong, T., Mitchell, P., 2007. The eye in hypertension. Lancet 369 (9559),
1450–1458. 425–435.
Tuo, Q.Z., Zou, J.J., Lei, P., 2020 Oct 27. Rodent models of vascular cognitive Wu, H.Q., Wu, H., Shi, L.L., Yu, L.Y., Wang, L.Y., Chen, Y.L., Geng, J.S., Shi, J., Jiang, K.,
impairment. J. Mol. Neurosci. https://doi.org/10.1007/s12031-020-01733-2. Dong, J.C., 2017. The association between retinal vasculature changes and stroke: a
Online ahead of print. literature review and Meta-analysis. Int. J. Ophthalmol. 10 (1), 109–114.
Törnquist, P., Alm, A., 1986. Carrier-mediated transport of amino acids through the Xu, D., Vaswani, N., Huang, Y., Kang, J.U., 2012. Modified compressive sensing optical
blood-retinal and the blood-brain barriers. Graefes Arch. Clin. Exp. Ophthalmol. 224 coherence tomography with noise reduction. Opt. Lett. 37 (20), 4209–4211.
(1), 21–25. Yannuzzi, L.A., Ober, M.D., Slakter, J.S., Spaide, R.F., Fisher, Y.L., Flower, R.W.,
Ucak, T., Alagoz, A., Cakir, B., Celik, E., Bozkurt, E., Alagoz, G., 2016. Analysis of the Rosen, R., 2004. Ophthalmic fundus imaging: today and beyond. Am. J. Ophthalmol.
retinal nerve fiber and ganglion cell - inner plexiform layer by optical coherence 137 (3), 511–524.
tomography in Parkinson’s patients. Park. Relat. Disord. 31, 59–64. Yoon, S.P., Grewal, D.S., Thompson, A.C., Polascik, B.W., Dunn, C., Burke, J.R.,
Uchida, A., Pillai, J.A., Bermel, R., Bonner-Jackson, A., Rae-Grant, A., Fernandez, H., Fekrat, S., 2019. Retinal microvascular and neurodegenerative changes in
Bena, J., Jones, S.E., Leverenz, J.B., Srivastava, S.K., Ehlers, J.P., 2018. Outer retinal Alzheimer’s disease and mild cognitive impairment compared with control
assessment using spectral-domain optical coherence tomography in patients with participants. Ophthalmol. Retina 3 (6), 489–499.
Alzheimer’s and Parkinson’s disease. Invest. Ophthalmol. Vis. Sci. 59 (7), Yousefi, S., Liu, T., Wang, R.K., 2015. Segmentation and quantification of blood vessels
2768–2777. for OCT-based micro-angiograms using hybrid shape/intensity compounding.
Uhlmann, R.F., Larson, E.B., Koepsell, T.D., Rees, T.S., Duckert, L.G., 1991. Visual Microvasc. Res. 97, 37–46.
impairment and cognitive dysfunction in Alzheimer’s disease. J. Gen. Intern. Med. 6 Yu, J.G., Feng, Y.F., Xiang, Y., Huang, J.H., Savini, G., Parisi, V., Yang, W.J., Fu, X.A.,
(2), 126–132. 2014. Retinal nerve fiber layer thickness changes in Parkinson disease: a meta-
Ukalovic, K., Cao, S., Lee, S., Tang, Q., Beg, M.F., Sarunic, M.V., Hsiung, G.R., analysis. PloS One 9 (1), e85718.
Mackenzie, I.R., Hirsch-Reinshagen, V., Cui, J.Z., Matsubara, J.A., 2018. Drusen in Zabel, P., Kaluzny, J.J., Wilkosc-Debczynska, M., Gebska-Toloczko, M., Suwala, K.,
the peripheral retina of the Alzheimer’s eye. Curr. Alzheimer Res. 15 (8), 743–750. Zabel, K., Zaron, A., Kucharski, R., Araszkiewicz, A., 2019. Comparison of retinal
Ulusoy, M.O., Horasanlı, B., Işık-Ulusoy, S., 2020. Optical coherence tomography microvasculature in patients with Alzheimer’s disease and primary open-Angle
angiography findings of multiple sclerosis with or without optic neuritis. Neurol. glaucoma by optical coherence tomography angiography. Invest. Ophthalmol. Vis.
Res. 42 (4), 319–326. Sci. 60 (10), 3447–3455.
van de Kreeke, J.A., Nguyen, H.T., den Haan, J., Konijnenberg, E., Tomassen, J., den Zamir, M., 1976a. Optimality principles in arterial branching. J. Theor. Biol. 62 (1),
Braber, A., Ten Kate, M., Collij, L., Yaqub, M., van Berckel, B., Lammertsma, A.A., 227–251.
Boomsma, D.I., Tan, H.S., Verbraak, F.D., Visser, P.J., 2019. Retinal layer thickness Zamir, M., 1976b. The role of shear forces in arterial branching. J. Gen. Physiol. 67 (2),
in preclinical Alzheimer’s disease. Acta Ophthalmol. 97 (8), 798–804. 213–222.
Veys, L., Vandenabeele, M., Ortuño-Lizarán, I., Baekelandt, V., Cuenca, N., Moons, L., De Zaveri, M.S., Conger, A., Salter, A., Frohman, T.C., Galetta, S.L., Markowitz, C.E.,
Groef, L., 2019. Retinal α-synuclein deposits in Parkinson’s disease patients and Jacobs, D.A., Cutter, G.R., Ying, G.S., Maguire, M.G., Calabresi, P.A., Balcer, L.J.,
animal models. Acta Neuropathol. 137 (3), 379–395. Frohman, E.M., 2008. Retinal imaging by laser polarimetry and optical coherence
Villemagne, V.L., Burnham, S., Bourgeat, P., Brown, B., Ellis, K.A., Salvado, O., tomography evidence of axonal degeneration in multiple sclerosis. Arch. Neurol. 65
Szoeke, C., Macaulay, S.L., Martins, R., Maruff, P., Ames, D., Rowe, C.C., Masters, C. (7), 924–928.
L., A. I. B. a. L. A. R. Group, 2013. Amyloid β deposition, neurodegeneration, and Zhang, J., Qiao, Y., Sarabi, M.S., Khansari, M.M., Gahm, J.K., Kashani, A.H., Shi, Y., 2019
cognitive decline in sporadic Alzheimer’s disease: a prospective cohort study. Lancet May. 3D shape modeling and analysis of retinal microvasculature in OCT-
Neurol. 12 (4), 357–367. angiography images. IEEE Trans. Med. Imag. 39 (5), 1335–1346.
Visser, F., Vermeer, K.A., Ghafaryasl, B., Vlaar, A.M.M., Apostolov, V., van Hellenberg Zhang, X., Saaddine, J.B., Chou, C.F., Cotch, M.F., Cheng, Y.J., Geiss, L.S., Gregg, E.W.,
Hubar, J., Weinstein, H.C., de Boer, J.F., Berendse, H.W., 2018. In vivo exploration Albright, A.L., Klein, B.E., Klein, R., 2010. Prevalence of diabetic retinopathy in the
of retinal nerve fiber layer morphology in Parkinson’s disease patients. J. Neural. United States, 2005-2008. J. Am. Med. Assoc. 304 (6), 649–656.
Transm. 125 (6), 931–936.

38

You might also like