Quantitative optical coherence tomography …Optical coherence tomography (OCT) angiography (3 × 3...

8
Quantitative optical coherence tomography angiography of vascular abnormalities in the living human eye Yali Jia a , Steven T. Bailey a , Thomas S. Hwang a , Scott M. McClintic a , Simon S. Gao a , Mark E. Pennesi a , Christina J. Flaxel a , Andreas K. Lauer a , David J. Wilson a , Joachim Hornegger b , James G. Fujimoto c , and David Huang a,1 a Casey Eye Institute, Oregon Health & Science University, Portland, OR 97239; b Pattern Recognition Lab and School of Advanced Optical Technologies, University Erlangen-Nuremberg, D-91058 Erlangen, Germany; and c Department of Electrical Engineering and Computer Science and Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, MA 02139 Edited by Morton Goldberg, The Johns Hopkins Wilmer Eye Institute, Baltimore, MD, and accepted by the Editorial Board March 26, 2015 (received for review January 6, 2015) Retinal vascular diseases are important causes of vision loss. A detailed evaluation of the vascular abnormalities facilitates diagno- sis and treatment in these diseases. Optical coherence tomography (OCT) angiography using the highly efficient split-spectrum ampli- tude decorrelation angiography algorithm offers an alternative to conventional dye-based retinal angiography. OCT angiography has several advantages, including 3D visualization of retinal and cho- roidal circulations (including the choriocapillaris) and avoidance of dye injection-related complications. Results from six illustrative cases are reported. In diabetic retinopathy, OCT angiography can detect neovascularization and quantify ischemia. In age-related macular degeneration, choroidal neovascularization can be ob- served without the obscuration of details caused by dye leakage in conventional angiography. Choriocapillaris dysfunction can be detected in the nonneovascular form of the disease, furthering our understanding of pathogenesis. In choroideremia, OCTs ability to show choroidal and retinal vascular dysfunction separately may be valuable in predicting progression and assessing treatment re- sponse. OCT angiography shows promise as a noninvasive alter- native to dye-based angiography for highly detailed, in vivo, 3D, quantitative evaluation of retinal vascular abnormalities. optical coherence tomography angiography | ophthalmic imaging | ocular circulation O ptical coherence tomography (OCT) has become the most commonly used imaging modality in ophthalmology. It pro- vides cross-sectional and 3D imaging of the retina and optic nerve head with micrometer-scale depth resolution. Structural OCT enhances the clinicians ability to detect and monitor fluid exudation associated with retinal vascular diseases. Whereas anatomical alterations that impact vision are readily visible, structural OCT has a limited ability to image the retinal or choroidal vasculatures. Furthermore, it is unable to directly de- tect capillary dropout or pathologic new vessel growth (neo- vascularization) that are the major vascular changes associated with two of the leading causes of blindness, age-related macular degeneration (AMD) and diabetic retinopathy (1). To visualize these changes, traditional i.v. contrast dye-based angiography techniques are currently used. Fluorescein dye is primarily used to visualize the retinal vas- culature. A separate dye, indocyanine green (ICG), is necessary to evaluate the choroidal vasculature. Both fluorescein angiog- raphy (FA) and ICG angiography require i.v. injection, which is time consuming, and which can cause nausea, vomiting, and, rarely, anaphylaxis (2). Dye leakage or staining provides infor- mation regarding vascular incompetence (e.g., from abnormal capillary growth), but it also obscures the image and blurs the boundaries of neovascularization. Additionally, conventional an- giography is 2D, which makes it difficult to distinguish vascular abnormalities within different layers. Therefore, it is desirable to develop a no-injection, dye-free method for 3D visualization of ocular circulation. In recent years, several OCT angiography methods have been developed to detect changes in the OCT signal caused by flowing red blood cells in blood vessels. Initially, Doppler OCT angiog- raphy methods were investigated for the visualization and mea- surement of blood flow (38). Because Doppler OCT is only sensitive to motion parallel to the OCT probe beam, it is limited in its ability to image retinal and choroidal circulations, which are predominantly perpendicular to the OCT beam. More recent approaches, based on detecting variation in the speckle pattern over time, are sensitive to both transverse and axial flow. Several types of speckle-based techniques have been described, including amplitude-based (911), phase-based (12), or a combination of both amplitude and phase (13) variance methods. We developed an amplitude-based method called split-spectrum amplitude-decorrelation angiography (SSADA). The SSADA al- gorithm detects motion in the blood vessel lumen by measuring the variation in reflected OCT signal amplitude between consecutive cross-sectional scans. The novelty of SSADA lies in how the OCT Significance Retinal vascular diseases are a leading cause of blindness. Optical coherence tomography (OCT) has become the standard imaging modality for evaluating fluid accumulation in these diseases and for guiding treatment. However, fluorescein an- giography (FA) is still required for initial evaluation of retinal ischemia and choroidal neovascularization, which are not visi- ble in conventional structural OCT. The limitations of FA include poor penetration of fluorescence through blood and pigment, inability to determine the depth of the pathology due to its two-dimensional nature, and some uncommon but potentially severe complications. As a noninvasive three-dimensional al- ternative, OCT angiography may be used in routine screening and monitoring to provide new information for clinical diag- nosis and management. Author contributions: Y.J., S.T.B., T.S.H., M.E.P., C.J.F., A.K.L., D.J.W., and D.H. designed re- search; J.H. and J.G.F. contributed new reagents/analytic tools; Y.J., S.T.B., T.S.H., S.M.M., S.S.G., M.E.P., C.J.F., A.K.L., D.J.W., and D.H. analyzed data; and Y.J., S.T.B., T.S.H., S.M.M., S.S.G., M.E.P., and D.H. wrote the paper. Conflict of interest statement: Oregon Health & Science University (OHSU), Y.J., J.G.F., and D.H. have a significant financial interest in Optovue, Inc., a company that may have a commercial interest in the results of this research and technology. These potential con- flicts of interest have been reviewed and managed by OHSU. J.G.F. and D.H. receive royalties on an optical coherence tomography patent licensed by the Massachusetts In- stitute of Technology (MIT) to Carl Zeiss Meditec. J.G.F. and J.H. receive royalties from intellectual property owned by MIT and licensed to Optovue, Inc. Other authors do not have financial interest in the subject of this article. This article is a PNAS Direct Submission. M.G. is a guest editor invited by the Editorial Board. 1 To whom correspondence should be addressed. Email: [email protected]. www.pnas.org/cgi/doi/10.1073/pnas.1500185112 PNAS | Published online April 20, 2015 | E2395E2402 MEDICAL SCIENCES ENGINEERING PNAS PLUS Downloaded by guest on August 18, 2020

Transcript of Quantitative optical coherence tomography …Optical coherence tomography (OCT) angiography (3 × 3...

Page 1: Quantitative optical coherence tomography …Optical coherence tomography (OCT) angiography (3 × 3 mm) of a healthy human eye, acquired using a 70-kHz spectral OCT system with an

Quantitative optical coherence tomographyangiography of vascular abnormalities in the livinghuman eyeYali Jiaa, Steven T. Baileya, Thomas S. Hwanga, Scott M. McClintica, Simon S. Gaoa, Mark E. Pennesia, Christina J. Flaxela,Andreas K. Lauera, David J. Wilsona, Joachim Horneggerb, James G. Fujimotoc, and David Huanga,1

aCasey Eye Institute, Oregon Health & Science University, Portland, OR 97239; bPattern Recognition Lab and School of Advanced Optical Technologies,University Erlangen-Nuremberg, D-91058 Erlangen, Germany; and cDepartment of Electrical Engineering and Computer Science and Research Laboratory ofElectronics, Massachusetts Institute of Technology, Cambridge, MA 02139

Edited by Morton Goldberg, The Johns Hopkins Wilmer Eye Institute, Baltimore, MD, and accepted by the Editorial Board March 26, 2015 (received for reviewJanuary 6, 2015)

Retinal vascular diseases are important causes of vision loss. Adetailed evaluation of the vascular abnormalities facilitates diagno-sis and treatment in these diseases. Optical coherence tomography(OCT) angiography using the highly efficient split-spectrum ampli-tude decorrelation angiography algorithm offers an alternative toconventional dye-based retinal angiography. OCT angiography hasseveral advantages, including 3D visualization of retinal and cho-roidal circulations (including the choriocapillaris) and avoidance ofdye injection-related complications. Results from six illustrativecases are reported. In diabetic retinopathy, OCT angiography candetect neovascularization and quantify ischemia. In age-relatedmacular degeneration, choroidal neovascularization can be ob-served without the obscuration of details caused by dye leakagein conventional angiography. Choriocapillaris dysfunction can bedetected in the nonneovascular form of the disease, furthering ourunderstanding of pathogenesis. In choroideremia, OCT’s ability toshow choroidal and retinal vascular dysfunction separately may bevaluable in predicting progression and assessing treatment re-sponse. OCT angiography shows promise as a noninvasive alter-native to dye-based angiography for highly detailed, in vivo, 3D,quantitative evaluation of retinal vascular abnormalities.

optical coherence tomography angiography | ophthalmic imaging |ocular circulation

Optical coherence tomography (OCT) has become the mostcommonly used imaging modality in ophthalmology. It pro-

vides cross-sectional and 3D imaging of the retina and opticnerve head with micrometer-scale depth resolution. StructuralOCT enhances the clinician’s ability to detect and monitor fluidexudation associated with retinal vascular diseases. Whereasanatomical alterations that impact vision are readily visible,structural OCT has a limited ability to image the retinal orchoroidal vasculatures. Furthermore, it is unable to directly de-tect capillary dropout or pathologic new vessel growth (neo-vascularization) that are the major vascular changes associatedwith two of the leading causes of blindness, age-related maculardegeneration (AMD) and diabetic retinopathy (1). To visualizethese changes, traditional i.v. contrast dye-based angiographytechniques are currently used.Fluorescein dye is primarily used to visualize the retinal vas-

culature. A separate dye, indocyanine green (ICG), is necessaryto evaluate the choroidal vasculature. Both fluorescein angiog-raphy (FA) and ICG angiography require i.v. injection, which istime consuming, and which can cause nausea, vomiting, and,rarely, anaphylaxis (2). Dye leakage or staining provides infor-mation regarding vascular incompetence (e.g., from abnormalcapillary growth), but it also obscures the image and blurs theboundaries of neovascularization. Additionally, conventional an-giography is 2D, which makes it difficult to distinguish vascularabnormalities within different layers. Therefore, it is desirable to

develop a no-injection, dye-free method for 3D visualization ofocular circulation.In recent years, several OCT angiography methods have been

developed to detect changes in the OCT signal caused by flowingred blood cells in blood vessels. Initially, Doppler OCT angiog-raphy methods were investigated for the visualization and mea-surement of blood flow (3–8). Because Doppler OCT is onlysensitive to motion parallel to the OCT probe beam, it is limitedin its ability to image retinal and choroidal circulations, whichare predominantly perpendicular to the OCT beam. More recentapproaches, based on detecting variation in the speckle patternover time, are sensitive to both transverse and axial flow. Severaltypes of speckle-based techniques have been described, includingamplitude-based (9–11), phase-based (12), or a combination ofboth amplitude and phase (13) variance methods.We developed an amplitude-based method called split-spectrum

amplitude-decorrelation angiography (SSADA). The SSADA al-gorithm detects motion in the blood vessel lumen by measuring thevariation in reflected OCT signal amplitude between consecutivecross-sectional scans. The novelty of SSADA lies in how the OCT

Significance

Retinal vascular diseases are a leading cause of blindness.Optical coherence tomography (OCT) has become the standardimaging modality for evaluating fluid accumulation in thesediseases and for guiding treatment. However, fluorescein an-giography (FA) is still required for initial evaluation of retinalischemia and choroidal neovascularization, which are not visi-ble in conventional structural OCT. The limitations of FA includepoor penetration of fluorescence through blood and pigment,inability to determine the depth of the pathology due to itstwo-dimensional nature, and some uncommon but potentiallysevere complications. As a noninvasive three-dimensional al-ternative, OCT angiography may be used in routine screeningand monitoring to provide new information for clinical diag-nosis and management.

Author contributions: Y.J., S.T.B., T.S.H., M.E.P., C.J.F., A.K.L., D.J.W., and D.H. designed re-search; J.H. and J.G.F. contributed new reagents/analytic tools; Y.J., S.T.B., T.S.H., S.M.M., S.S.G.,M.E.P., C.J.F., A.K.L., D.J.W., and D.H. analyzed data; and Y.J., S.T.B., T.S.H., S.M.M., S.S.G., M.E.P.,and D.H. wrote the paper.

Conflict of interest statement: Oregon Health & Science University (OHSU), Y.J., J.G.F., andD.H. have a significant financial interest in Optovue, Inc., a company that may have acommercial interest in the results of this research and technology. These potential con-flicts of interest have been reviewed and managed by OHSU. J.G.F. and D.H. receiveroyalties on an optical coherence tomography patent licensed by the Massachusetts In-stitute of Technology (MIT) to Carl Zeiss Meditec. J.G.F. and J.H. receive royalties fromintellectual property owned by MIT and licensed to Optovue, Inc. Other authors do nothave financial interest in the subject of this article.

This article is a PNAS Direct Submission. M.G. is a guest editor invited by the EditorialBoard.1To whom correspondence should be addressed. Email: [email protected].

www.pnas.org/cgi/doi/10.1073/pnas.1500185112 PNAS | Published online April 20, 2015 | E2395–E2402

MED

ICALSC

IENCE

SEN

GINEE

RING

PNASPL

US

Dow

nloa

ded

by g

uest

on

Aug

ust 1

8, 2

020

Page 2: Quantitative optical coherence tomography …Optical coherence tomography (OCT) angiography (3 × 3 mm) of a healthy human eye, acquired using a 70-kHz spectral OCT system with an

signal is processed to enhance flow detection and reject axial bulkmotion noise. Compared with the full-spectrum amplitude method,SSADA using fourfold spectral splits improved the signal-to-noiseratio (SNR) by a factor of two, which is equivalent to reducing thescan time by a factor of four (14). More recent SSADA imple-mentations use even more than a fourfold split to further enhancethe SNR of flow detection. This highly efficient algorithm gener-ates high-quality angiograms of both the retina and choroid. Theangiograms have capillary-level detail and can be obtained withcurrently available commercial OCT systems.This article uses six illustrative cases and highlights the various

types of vascular pathologies that can be detected and measuredusing SSADA and an OCT angiography system of visualization.Additionally, we describe techniques designed to help cliniciansrapidly interpret OCT angiograms and to easily identify patho-logical vascular features. These techniques include (i) separationof the 3D angiogram into individual vascular beds via segmenta-tion algorithms, (ii) presentation of en face OCT angiograms,analogous to traditional angiography, (iii) creation of cross-sec-tional structural OCT images with superimposed OCT angiogramsto help correlate anatomical alterations with vascular abnormali-ties, and (iv) quantification of neovascularization and capillarydropout in both the retinal and choroidal circulation.

ResultsRetinal and Choroidal Microcirculation in Normal Subjects. The ret-ina and choroid are two distinct vascular beds. The 3D nature ofOCT angiography allows separate visualizations of these twocirculations from the same volumetric scan. In healthy eyes,retinal circulation is located between the internal limiting mem-brane (ILM) and the outer plexiform layer (OPL), whereas thechoroidal circulation is beneath Bruch’s membrane (BM). Thevitreous (anterior to the ILM) and outer retina (between OPLand Bruch’s membrane) are avascular in normal eyes. One wayto present OCT angiography is to use different colors, eachrepresenting different vascular beds, superimposed on gray-scale, cross-sectional, structural OCT images (Fig. 1A). Using

this technique, both blood flow and retinal structural informationare presented on a single image.En face presentation of OCT angiography is comparable to

the traditional view of dye-based angiography and is necessaryfor clinicians to identify vascular patterns. Segmented en faceOCT angiograms can be displayed as individual retinal andchoroidal circulations (Fig. 1 C and E). In healthy eyes, the vit-reous and outer retinal layers are colored black (Fig. 1 B and D),representing the absence of flow.In vivo imaging of the choroid is limited with current imaging

modalities (15). High-quality OCT angiograms of the chorioca-pillaris can be obtained by segmenting a thin slice (10 μm) of theinner choroid below Bruch’s membrane (Fig. 1E). The larger anddeeper choroidal vessels are more difficult to visualize withSSADA-based OCT angiography because choriocapillaris flowcauses a flow projection artifact in the deeper layers. Selectiveremoval of this artifact is impossible, given the near confluentnature of the choriocapillaris. Another reason is that high flow inthe larger choroidal vessels reduces the OCT signal due to in-terference fringe washout. However, visualization of the larger,deeper choroidal vessels is possible using an inverse reflectancedisplay scale (Fig. 1F). This en face choroidal image is achieved bytaking advantage of the OCT signal void produced by interferencefringe washout in regions of very high flow velocity (16).

Retinal Neovascularization and Capillary Dropout in Diabetic Retinopathy.Diabetic retinopathy is a microangiopathy characterized bycapillary occlusion, hyperpermeability, and neovasculariza-tion (17). These pathophysiologic changes cause proliferativediabetic retinopathy and macular edema, which are respon-sible for most of the vision loss associated with this disease(18). Assessment of the microvascular changes with FA hasbeen validated as a way to classify disease severity and predictprogression (19).Like FA, OCT angiography can visualize areas of low capillary

perfusion or dropout. Fig. 2 compares en face retinal OCT an-giograms from a normal subject to those from a patient withnonproliferative diabetic retinopathy (NPDR). The central 0.6-mm

Fig. 1. Optical coherence tomography (OCT) angiography (3 × 3 mm) of a healthy human eye, acquired using a 70-kHz spectral OCT system with an 840-nmcenter wavelength. (A) Cross-sectional composite OCT angiogram. Depth layer segmentation lines are shown in green. BM, Bruch’s membrane; ILM, internallimiting membrane; OPL, outer plexiform layer. Flow signals are color coded by depth: purple, anterior to the OPL; red, posterior to BM. (B) En face OCTangiogram above the ILM shows the normal, avascular vitreous. (C) En face OCT angiogram between the ILM and OPL shows the normal retinal vasculature.(D) En face OCT angiogram between the OPL and BM shows the normal, avascular outer retina. (E) En face OCT angiogram of the inner 10 μm of the choroidshows dense, relatively even flow throughout the central macula (3 × 3 mm). (F) En face OCT structural image with an inverse gray scale shows the deeperchoroid with medium- and large-sized vessels.

E2396 | www.pnas.org/cgi/doi/10.1073/pnas.1500185112 Jia et al.

Dow

nloa

ded

by g

uest

on

Aug

ust 1

8, 2

020

Page 3: Quantitative optical coherence tomography …Optical coherence tomography (OCT) angiography (3 × 3 mm) of a healthy human eye, acquired using a 70-kHz spectral OCT system with an

circle encompasses the normal foveal avascular zone (FAZ), andareas of capillary loss are represented by dark areas elsewhere inthe macula. No dark areas are present outside of the FAZ in thenormal subject (Fig. 2 B1, C1, and D1), whereas the patient withNPDR has enlargement of the normal FAZ and extensive loss ofthe macular capillary bed (Fig. 2B2). Nonperfusion areas on FAcorrespond to those on OCT angiography (Fig. 2 B2 and C2).SSADA also enables the quantitative evaluation of local cir-

culation by determining the flow index and vessel density in areasof interest, such as the parafoveal and perifoveal areas thatcorrespond respectively to the 3-mm and 6-mm early treatmentdiabetic retinopathy study (ETDRS) macular fields (as demon-strated in Fig. 2) (19). Nonperfusion maps may be created (Fig. 2D1 and D2) that allows for the area of nonperfusion to be cal-culated and compared between sequential angiograms.The development of retinal neovascularization (RNV) sig-

nifies progression to the proliferative phase of the disease.Recognition of this change is important because it may guidepanretinal photocoagulation and other treatments to reduce therisk of vision loss due to RNV (20). Because OCT angiogramscan be segmented by anatomic planes such as the ILM, it can beparticularly effective in distinguishing between intraretinal mi-crovascular abnormalities, which occur in the same plane as theretinal blood vessels, and early RNV, which develops anterior tothe retinal vessels and may extend into the vitreous cavity (Fig.3). The extent and viability of RNV can also be quantified onOCT angiography by vessel area and flow index. Thus, comparedwith FA, OCT angiography has the advantages of 3D localizationand quantification.

Choroidal Neovascularization in Age-Related Macular Degeneration.Choroidal neovascularization (CNV), the hallmark pathologic

feature of neovascular AMD, consists of abnormal blood vesselsthat grow from the choriocapillaris and penetrate throughBruch’s membrane into the subretinal pigment epithelium (sub-RPE) space and subretinal space. Subsequent exudation andhemorrhage damage retinal tissue, resulting in vision loss (21).FA is the gold standard for CNV diagnosis (22); however, it islimited by its 2D nature. In addition, blocked fluorescence fromthe RPE or hemorrhage (if present) reduces visibility of theCNV beneath the RPE, as well as visualization of the choroid(23). OCT angiography has the capability to generate 3D an-giograms of the retina, choroid, and CNV that is otherwise ob-scured in FA by RPE blockage or hemorrhage.In an example of neovascular AMD (Fig. 4A), the late FA

image (Fig. 4B) shows a stippled hyperfluoresence leakage pat-tern, indicating the presence of an occult CNV. The outer retinais devoid of blood flow in healthy eyes, and the flow detected inthis area is associated with the presence of CNV. Color-codedcomposite en face OCT angiography allows for 3D representa-tion of retinal flow, outer retinal flow, and choroidal flow on asingle 2D image. In this case, the CNV is highlighted in yellow,and the extent and microvascular structure (Fig. 4C) is betterdefined compared with traditional FA image (Fig. 4B). Thecross-sectional structural OCT image with a color-coded OCTangiogram overlay (Fig. 4D) demonstrates the depth of theCNV, in this case beneath the RPE, as well as the presence offluid exudation and disruption of outer retinal anatomy.This case illustrates the capability of OCT angiography to

assess the morphology, extent, and depth of CNV in AMD. Wehave demonstrated that OCT angiography can be used to classifyCNV as type I (between the RPE and Bruch’s membrane, Fig.4D), type II (above the RPE), type III (in the inner retina), or acombined type (24). OCT angiography can furthermore provide

Fig. 2. Quantification of inner retinal blood flows in normal control (A1–D1) and nonproliferative diabetic retinopathy (NPDR) with macular edema (A2–D2).OCT angiography was acquired using a 70-kHz spectral OCT system with a center wavelength of 840 nm. White dashed circle indicates normal foveal avascularzone (FAZ, 0.6 mm diameter white dashed circle). Area between white and blue dashed circles indicates parafoveal zone. Area between blue and greendashed circles indicates perifoveal zone. First column (A1 and A2) indicates fundus photo (A1) and fluorescein angiography (A2). Second column (B1 and B2)indicates en face 3 × 3 mm OCT angiograms. Enlargement of the FAZ is present in the parafoveal region for NPDR (B2). Parafoveal and perifoveal retinal flowindexes (vessel densities) are shown on en face 6 × 6 mm OCT angiograms (C1 and C2). Nonperfusion areas (blue) are shown of 6 × 6 mm OCT angiograms (D1and D2). The normal subject had a FAZ of 0.30 mm2, whereas the NPDR case showed an enlarged FAZ and scattered areas of macular nonperfusion totaling7.07 mm2. Decorrelation (shown by the scale bar) as calculated by the SSADA algorithm measures fluctuation in the backscattered OCT signal amplitude(intensity). It is linear to blood flow velocity up to a saturation limit after which it reaches a maximum value. The linear range is approximately within thelimits of capillary flow velocity.

Jia et al. PNAS | Published online April 20, 2015 | E2397

MED

ICALSC

IENCE

SEN

GINEE

RING

PNASPL

US

Dow

nloa

ded

by g

uest

on

Aug

ust 1

8, 2

020

Page 4: Quantitative optical coherence tomography …Optical coherence tomography (OCT) angiography (3 × 3 mm) of a healthy human eye, acquired using a 70-kHz spectral OCT system with an

quantitative data regarding CNV flow and area. These capabil-ities may prove to be valuable in the assessment of disease se-verity and monitoring of the effectiveness of treatment. It islikely that, in many cases, CNV growth through Bruch’s mem-brane occurs before the onset of exudation and visual symptoms.Whereas visual acuity at presentation is strongly predictive of theposttreatment outcome (25), OCT angiography may enable thedetection of CNV before the development of symptoms or de-tectable changes with structural OCT or FA. Because it is a safe,noninvasive, and rapid imaging technique, at-risk patients maybenefit from OCT angiography screening.

Choriocapillaris Loss in Age-Related Macular Degeneration. Thechoriocapillaris has been implicated in the progression of AMD.In advanced nonneovascular AMD, geographic atrophy (GA) isassociated with loss of photoreceptors, RPE, and the chorioca-pillaris. Whether the RPE or choriocapillaris alterations are theprimary event in pathogenesis has been a matter of debate (26).Histologic studies have described choriocapillaris dysfunction inthe intermediate stage of AMD, before development of GA orchoroidal neovascularization (27, 28). Whereas efforts have beenmade to assess the choriocapillaris in vivo (29), its small size,high density, and high permeability have made it difficult forconventional imaging modalities, including FA and ICG, toprovide meaningful assessment. By segmenting a layer extending10 μm posterior to Bruch’s membrane, OCT angiography pro-vides qualitative and quantitative evaluation of the chorioca-pillaris, which may be valuable in understanding its role in AMD.OCT angiography has utility in elucidating the state of the

choriocapillaris in GA (Fig. 5). In a patient with perifoveal GA,

the fundus photograph (Fig. 5A) and autofluorescence image(Fig. 5B) show the affected region. The drusen–RPE complexthickness map (the axial distance from the apex of the drusenand the RPE layer to BM, Fig. 5C) shows an area of RPE losscorresponding to the clinically evident GA. The choriocapillarisflow is absent in most of the area correlating to clinical GA (Fig. 5E and F). Prior histopathological findings (27) showed that viablechoriocapillaris can exist in areas of GA, but highly constricted.OCT angiography is based on flow detection; therefore, con-stricted choriocapillaris with little or no flow may appear absent.In some areas near the border of the GA with RPE loss, intactchoriocapillaris flow is present (Fig. 5 G and H). In this area ofpreserved choriocapillaris and slightly beyond it, the outer nuclearlayer is also preserved; however, in most of the GA area, the outernuclear layer, photoreceptors, and RPE are absent.

Choriocapillaris Loss in Choroideremia. Pathology of the chorioca-pillaris has been implicated in many disease processes apart fromAMD, including those with genetic, inflammatory, and infectiousetiologies. Choroideremia is an X-linked recessive chorioretinaldystrophy associated with mutation of the CHM gene. This geneencodes the Rab escort protein 1 (REP-1) and is characterized bysignificant atrophy of the RPE and choriocapillaris with photore-ceptor loss (30). Affected patients typically experience nightblindness in their first or second decade, followed by constriction ofthe peripheral visual fields until central vision is lost. Patients withchoroideremia often demonstrate retinal vessels of normal caliberuntil late in the disease, suggesting a relatively greater loss ofchoroidal blood flow (30). Spectral-domain OCT studies of

Fig. 3. Proliferative diabetic retinopathy (PDR) case imaged using a 100-kHzswept-source OCT system with a center wavelength of 1,050 nm. (A) Con-focal scanning laser ophthalmoscope (cSLO) shows retinal neovascularizationat the optic disk (NVD) and attenuated retinal vessels. (B) FA showing NVDand peripapillary capillary dropout. The green squares in A and B outline the3 × 3 mm area shown on the OCT angiogram below. (C) En face OCT angi-ography showing NVD and areas of capillary dropout that correspond to FA(NV is shown in light red gold; normal retinal vessels are in purple). The areaof NVD was 0.47 mm2. The vitreous flow index was 0.022. (D) Cross-sectionalcomposite OCT angiogram showing NVD above the inner limiting mem-brane (red/gold).

Fig. 4. Type I CNV case imaged by a 100-kHz swept-source OCT system witha center wavelength of 1,050 nm. The CNV is identified by OCT angiography(3 × 3 mm), but it is ill defined by fluorescein angiography (FA). (A) Fundusphotograph. The black square outlines the areas shown on the angiograms.(B) Late stage fluorescein angiograph showing an occult CNV. (C) Compositeen face color-coded OCT angiogram with CNV flow highlighted in yellow.The CNV area was 0.96 mm2 and the outer retina flow index was 0.012. Theyellow dashed line indicates the position of the OCT cross-section. (D) Cross-sectional color OCT angiogram. Both composite en face (C) and cross-sec-tional color OCT angiograms (D) show inner retinal flow in purple, outerretinal flow (CNV) in yellow, and choroidal flow in red. The CNV is pre-dominantly under the RPE.

E2398 | www.pnas.org/cgi/doi/10.1073/pnas.1500185112 Jia et al.

Dow

nloa

ded

by g

uest

on

Aug

ust 1

8, 2

020

Page 5: Quantitative optical coherence tomography …Optical coherence tomography (OCT) angiography (3 × 3 mm) of a healthy human eye, acquired using a 70-kHz spectral OCT system with an

patients with choroideremia show photoreceptor rosettes, sug-gesting loss of RPE before that of photoreceptors. However, thetemporal relationship of RPE loss versus choriocapillaris lossis unknown.OCT angiography with SSADA may aid in understanding the

disease pathogenesis and inform the debate on whether de-generation occurs first in the choriocapillaris, RPE, or photo-receptors. En face OCT angiography has the ability to map bothretinal and choroidal perfusion down to the capillary level. In thethree subjects with choroideremia, OCT angiography (Fig. 6)showed that the choriocapillaris nonperfusion area was moreextensive than the retinal nonperfusion in all cases. The area ofRPE loss was even more extensive than choriocapillaris non-perfusion. These patterns suggest that RPE loss might be theprimary event, with subsequent choroidal perfusion loss follow-ing more closely than retinal perfusion loss. With the recentinitiation of gene therapy trials (31), the ability to quantify andmap both choroidal and retinal circulations may prove valuablein assessing disease severity and response to treatment.

DiscussionThe cases presented in this article show the potential clinicalapplications of OCT angiography for noninvasive vascular im-aging in the eye. Although clinical studies using OCT angiogra-phy have investigated polypoidal choroidal vasculopathy (32),macular degeneration (33, 34), diabetic retinopathy (35), andmacular telangiectasia type 2 (36, 37), they were descriptive innature. In this study, a comprehensive OCT angiography systemwas demonstrated that includes scanning, flow detection, seg-mentation, display, and quantification. There are several im-portant contributions in this report. First, OCT angiography withSSADA is able to capture a large 6 × 6 mm view of the maculawith adequate resolution using a commercially available OCTsystem. Second, the multicolor display system shows multiple

circulations in the same image panel so that the location of pa-thologies can be located in relation to the retinal vasculaturewith minimal interference from flow projection artifacts. Third,quantification of RNV and quantification of the area of capillarydropout in the retinal circulation and choriocapillaris are alsoshown in vivo for the first time to our knowledge.OCT angiography differs from FA, in which ocular pathology

is typically highlighted by leakage and staining. On OCT angi-ography, ocular pathologies are identified by the abnormalpresence of flow in layers that usually lack blood vessels or theabsence of flow in normally vascular layers. Because dye leakageand staining do not occur in OCT angiography, the boundaries,and therefore areas, of capillary dropout and neovascularizationcan be measured without these complications. OCT angiographyhas several other compelling characteristics that make it a prom-ising modality for clinical use. OCT angiography using SSADAcan be acquired in a few seconds, a dramatic improvement com-pared with several minutes for FA. Its 3D nature allows separateevaluation of abnormalities in retinal and choroidal circulations.Additionally, quantitative information, such as vessel density, flowindex, and nonperfusion areas, can now be obtained. OCT angi-ography might be useful in assessing the response of CNV toantiangiogenic treatment by quantifying the size and flow index ofCNV before and after therapy. Furthermore, the scan pattern andSSADA processing can be implemented on spectral-domain orswept-source OCT systems without any special hardware modifi-cation, provided that imaging speeds are sufficient; i.e., ≥70 kHz isrequired to provide 6 × 6 mm angiograms with capillary detailswithin a reasonable scan time (≤3 s).OCT angiography has several limitations and requirements.

First, projection artifacts are observed on the cross-sectionalangiograms. These artifacts are due to fluctuating shadows castby flowing blood in large inner retinal vessels that cause variationin the OCT signal from deeper layers. The projection artifact

Fig. 5. Geographic atrophy case imaged by a 100-kHz swept-source OCT system with a center wavelength of 1,050 nm. (A) Fundus photograph showing thearea of geographic atrophy (GA) adjacent to the foveal center. (B) An autofluorescence image sharply outlines the area of absent RPE and is surrounded by ahalo of hyperautofluorescence. The green squares in A and B outline the area shown in C–F. (C) A drusen–RPE complex thickness map shows the area of RPEthinning (the dark area nasally). (D) An en face OCT structural image on an inverse gray scale of the deeper choroid reveals medium- and large-sized vessels.(E) An en face OCT angiogram (3 × 3 mm) of the choriocapillaris shows dramatically decreased, but not absent, choriocapillaris flow in the area of GA. (F) Thelight blue color represents the choriocapillaris nonperfusion area (2.75 mm2). (G and H) Cross-sectional composite OCT angiograms show the absence ofchoriocapillaris flow in most of the area of GA, but the flow at the edge of the atrophy is spared (shown by the green arrows in the magnified views).

Jia et al. PNAS | Published online April 20, 2015 | E2399

MED

ICALSC

IENCE

SEN

GINEE

RING

PNASPL

US

Dow

nloa

ded

by g

uest

on

Aug

ust 1

8, 2

020

Page 6: Quantitative optical coherence tomography …Optical coherence tomography (OCT) angiography (3 × 3 mm) of a healthy human eye, acquired using a 70-kHz spectral OCT system with an

from the retinal circulation can be easily removed from the enface views of deeper layers because the retinal vasculature isrelatively sparse. However, the choriocapillaris is nearly conflu-ent, and its shadow often obscures OCT angiography visualiza-tion of deeper choroidal vessels. Second, the flow signal fadeoutin large vessels with very fast blood flow can induce fringewashout of OCT signals (16). This means that central retinalvessels in the disk and large vessels in the deep choroid cannot bevisualized using SSADA. However, we can use an inverted signalscale to take advantage of fringe washout to visualize deepchoroidal vessels. Third, decorrelation (flow signal) is linearlyrelated to velocity over a limited range. A higher decorrelation

value thus implies higher velocity flow. This range is dependenton the time scale of the SSADA measurement. With a 70-kHzspectral OCT system and 200+ A-scans per cross-sectionalB-scan, SSADA should be sensitive to the normal capillary flowspeeds, which have been estimated at between 0.4 and 3 mm/s(38, 39). If blood flow is extremely slow in pathological condi-tions, the decorrelation values may be below noise background,in which case flow would not be detected by OCT angiography.In larger vessels with higher velocities, the SSADA signal rea-ches a maximum value (saturates). This suggests that the angi-ography-based flow index, at current scan speeds, is a mixedmeasure of vascular density of both large and small vessels plus a

Fig. 6. Choroideremia case imaged by a 100-kHz swept-source OCT system with a center wavelength of 1,050 nm. The large-field en face OCT angiograms(∼3 × 8.5 mm) were produced by stitching together three 3 × 3 mm scans. (A) OCT angiography of inner retinal blood flow. (B) OCT angiography withquantification of inner retinal blood flow demonstrating patchy areas of nonperfusion (blue) in the extrafoveal macula. The total nonperfusion area of theinner retina was 7.65 mm2. (C) OCT angiography of the choroidal blood flow, including choriocapillaris and deep choroid. It should be noted that OCTangiography is able to image deeper choroidal vessels in this case due to the relative absence of the overlying choriocapillaris and RPE. (D) OCT angiographyof the choroidal blood flow with quantification of the choriocapillaris nonperfusion area (purple), which was 12.11 mm2 (47.5% of image area). (E) Anautofluorescence image outlines the area of existing RPE (hyperautofluorescent area).

E2400 | www.pnas.org/cgi/doi/10.1073/pnas.1500185112 Jia et al.

Dow

nloa

ded

by g

uest

on

Aug

ust 1

8, 2

020

Page 7: Quantitative optical coherence tomography …Optical coherence tomography (OCT) angiography (3 × 3 mm) of a healthy human eye, acquired using a 70-kHz spectral OCT system with an

component of capillary flow velocity. Fourth, unlike FA, OCTangiography cannot visualize dye leakage. However, the visuali-zation of intraretinal and subretinal fluid accumulation onstructural OCT may provide information associated with fluidleakage (40), whereas CNV activity may be evaluated by assess-ing the presence or characteristics of its perfusion (24). Fifth, thescan area is relatively small (3 × 3–6 × 6 mm). Larger area an-giograms can be achieved, but it would require higher speedOCT systems that are not yet commercially available (41). Lastly,because OCT angiography identifies depth resolve pathology,practical clinical applications require automated image process-ing software that accurately delineates tissue layers.In summary, OCT angiography can provide safe and repeatable

(42) high-resolution images of retinal and choroidal circulations, inaddition to structural OCT information. Its ability to noninvasivelyvisualize neovascularization and nonperfusion unaffected by leak-age and staining offers advantages over dye-based angiographymethods. The ability to quantify perfusion and separately visualizeretinal and choroidal vasculature makes it uniquely suited for avariety of research and clinical applications.

MethodsHuman Subjects Imaging. Study subjects were enrolled after informed consentin accordance with an Institutional Review Board/Ethics Committee approvedprotocol at Oregon Health & Science University and at the MassachusettsInstitute of Technology in compliance with the Declaration of Helsinki.Healthy participants or patients with a diagnosis of retinal disease (diabeticretinopathy/AMD/choroideremia) were selected from the Retina Division ofthe Casey Eye Institute for their clear media and ability to fixate. In total, 15healthy subjects, 14 with diabetic retinopathy, 26 with AMD, and 3 withchoroideremia were enrolled. To better demonstrate the potential clinicalapplication of this OCT methodology, six cases with characteristic patho-logical and clinical features were selected for this article.

Color fundus and optic disk photographs were acquired with Zeiss funduscameras (FF3 for Figs. 2A1 and 4A and FF450 for Fig. 5; Carl Zeiss Meditec)and the Optos 200Tx confocal scanning laser ophthalmoscope (cSLO) (Fig.3A; Optos PLC). For fluorescein angiography, 10% sodium fluorescein inwater (500 mg/5 mL) was injected i.v. using a 23- or 25-gauge needle, fol-lowed by a flush of normal saline. Fluorescein angiography was performedusing either the Optos 200Tx cSLO (Figs. 2A2 and 3B) or the Spectralis HRA+OCT cSLO (Fig. 4B; Heidelberg Engineering). For both cSLO devices, a 488-nmwavelength laser excited the fluorescein, and a barrier filter at 500 nmseparated the excitation and emission light. The fluorescein autofluores-cence image in Figs. 5 and 6 was also acquired using the Spectralis HRA+OCTcSLO. These procedures, imaging systems, and contrast dyes are approved bythe Food and Drug Administration.

OCT Systems. TwoOCT systems were used in this study. The first was a custom-built swept-source OCT instrument (14, 24, 42, 43). The device operated at anaxial scan rate of 100 kHz using a swept-source cavity laser operating at∼1,050 nm with a sweep range of 100 nm. The instrument has a 5.3-μm axialresolution and 18-μm lateral resolution with an imaging range of 2.9 mm intissue. The ocular light exposure was 1.9 mW, which was within the Amer-ican National Standards Institute safety limit (44). The second OCT systemwas a commercial spectral domain OCT instrument (RTVue-XR; Optovue)(36). The center wavelength was ∼840 nm with a full-width half-maximumbandwidth of 45 nm and an axial scan rate of 70 kHz.

OCT Imaging. A 3 × 3 or 6 × 6 mm scanning area was used for OCT angi-ography. In the fast transverse scanning direction, 200 axial scans weresampled to obtain a single B-scan. Multiple repeated B-scans (eight for theswept-source OCT and five for the spectral OCT), were captured at a fixedposition before proceeding to the next sampling location. A total of 200locations along a 3- or 6-mm distance in the slow transverse direction weresampled to form a 3D data cube. For the swept-source system, with a B-scanframe rate of 455 frames per second, the 1,600 B-scans in each scan wereacquired in ∼3.5 s. Four volumetric raster scans, including two horizontalpriority fast transverse (x-fast) scans and two vertical priority fast transverse(y-fast) scans, were obtained consecutively in one session. For the spectral

domain OCT system, with a B-scan frame rate of 320 frames per second, the1,000 B-scans in each scan were acquired in ∼3.1 s. Two volumetric rasterscans including one x-fast scan and one y-fast scan were obtained.

SSADA Processing. The SSADA algorithm was used to distinguish blood flowfrom static tissue as described in detail in a previous publication (14). Bycalculating the decorrelation of the signal amplitude from consecutiveB-scans, contrast between static and nonstatic tissue is created that enablesvisualization of blood flow. Decorrelation is a mathematical function thatquantifies variation without being affected by the average signal strength,as long as the signal is strong enough to predominate over optical and/orelectronic noise. Specifically, the algorithm splits the OCT image into dif-ferent spectral bands, thus increasing the number of usable image frames. Inthe optimized SSADA technique, the OCT signal is first split into 11 spectralbands to obtain 11 low axial resolution images instead of a single imageframe with high axial resolution. Each new frame has a lower axial resolu-tion that is less susceptible to axial eye motion caused by retrobulbar pul-sation. This lower resolution also translates to a wider coherence gate overwhich reflected signal from a moving particle such as a blood cell can in-terfere with adjacent structures, thereby increasing speckle contrast. Inaddition, each spectral band contains a different speckle pattern and in-dependent information on flow. When amplitude decorrelation imagesfrom multiple spectral bands are combined, the flow signal is increased. Byenhancing the flow signal and suppressing bulk motion noise, SSADA im-proves the signal-to-noise ratio of flow detection by at least a factor of two(14). Motion artifacts were further corrected and the flow signal increasedby applying an image registration algorithm that registered orthogonalraster-scanned volumes (45).

OCT Angiogram Visualization and Quantification. To enhance visualization, the3D angiogram was separately projected as en face views in five layers: vitreous,inner retina, outer retina, choriocapillaris, and deep choroid (Fig. 1). The sep-arated en face OCT angiography images were presented in a sepia color scale.A composite view was created, where each layer was assigned a different color.The color-coded angiogram can be superimposed on a gray-scale, cross-sec-tional, structural OCT image to demonstrate blood flow and structural in-formation simultaneously. Flow projection artifacts are a common problem forexisting OCT angiography techniques (3–13). To better distinguish and interpretthe blood flow within different layers, a negative filter was used to maskprojection artifacts from the larger caliber retinal vessels (24).

To quantify the blood flow within the regions of interest, the flow index,vessel density, and neovascularization areawere determined from the en facemaximum projection angiogram. The flow index was calculated as the av-erage decorrelation value (which is correlated with flow velocity) in theselected region, and the vessel density was calculated as the percentage areaoccupied by vessels and microvasculature in the selected region. For scans ofthemacula, flow index and vessel density can be routinely determined for theparafovea and/or perifovea. The parafovea is defined to be an annular regionwith an inner diameter of 0.6 mm and outer diameter of 2.5 mm centered onthe FAZ. The perifovea is defined to be the annular region extending from theedge of the parafovea to an outer diameter of 5.5mm. For a 3 × 3mm scan, onlythe parafovea values can be determined. Examples from macular angiogramsare shown in Fig. 2 C1 and C2. The vitreous or outer retinal flow index can beused to indicate the RNV or CNV flow within the scanned area (3 × 3 or 6 × 6mm) of the vitreous or outer retina. The retinal or choroidal neovascularizationarea was the area occupied by vessels in the vitreous or outer retina. To identifyand quantify the capillary dropout of the inner retina or choriocapillaris, thenonperfusion map was created by identifying decorrelation values lower than aset cutoff point, typically 2.33 SD below the mean according to the normaldistribution. Morphologic operations were then used to remove areas below acertain size to reduce noise. The remaining areas were then summed andconverted from pixels to metric units.

ACKNOWLEDGMENTS. This work was supported by National Institutes ofHealth Grants R01-EY023285, R01-EY024544, DP3 DK104397, R01-EY11289,K08-EY021186, T32-EY23211, and P30-EY010572; Clinical and TranslationalScience Award Grant UL1TR000128; unrestricted Grant and Career De-velopment Award CD-NMT-0914-0659-OHSU from Research to PreventBlindness; enhanced Career Development Award AFOSR FA9550-10-1-0551from Foundation Fighting Blindness; and German Research FoundationGrants DFG-HO-1791/11-1 and DFG-GSC80-SAOT.

1. Congdon N, et al.; Eye Diseases Prevalence Research Group (2004) Causes and prevalence of

visual impairment among adults in the United States. Arch Ophthalmol 122(4):477–485.

2. López-Sáez MP, et al. (1998) Fluorescein-induced allergic reaction. Ann Allergy

Asthma Immunol 81(5 Pt 1):428–430.

Jia et al. PNAS | Published online April 20, 2015 | E2401

MED

ICALSC

IENCE

SEN

GINEE

RING

PNASPL

US

Dow

nloa

ded

by g

uest

on

Aug

ust 1

8, 2

020

Page 8: Quantitative optical coherence tomography …Optical coherence tomography (OCT) angiography (3 × 3 mm) of a healthy human eye, acquired using a 70-kHz spectral OCT system with an

3. Wang RK, et al. (2007) Three dimensional optical angiography. Opt Express 15(7):4083–4097.

4. Grulkowski I, et al. (2009) Scanning protocols dedicated to smart velocity ranging inspectral OCT. Opt Express 17(26):23736–23754.

5. Yu L, Chen Z (2010) Doppler variance imaging for three-dimensional retina andchoroid angiography. J Biomed Opt 15(1):016029.

6. Makita S, Jaillon F, Yamanari M, Miura M, Yasuno Y (2011) Comprehensive in vivomicro-vascular imaging of the human eye by dual-beam-scan Doppler optical co-herence angiography. Opt Express 19(2):1271–1283.

7. Zotter S, et al. (2011) Visualization of microvasculature by dual-beam phase-resolvedDoppler optical coherence tomography. Opt Express 19(2):1217–1227.

8. Braaf B, Vermeer KA, Vienola KV, de Boer JF (2012) Angiography of the retina andthe choroid with phase-resolved OCT using interval-optimized backstitched B-scans.Opt Express 20(18):20516–20534.

9. Mariampillai A, et al. (2008) Speckle variance detection of microvasculature usingswept-source optical coherence tomography. Opt Lett 33(13):1530–1532.

10. Motaghiannezam R, Fraser S (2012) Logarithmic intensity and speckle-based motioncontrast methods for human retinal vasculature visualization using swept sourceoptical coherence tomography. Biomed Opt Express 3(3):503–521.

11. Enfield J, Jonathan E, Leahy M (2011) In vivo imaging of the microcirculation of thevolar forearm using correlation mapping optical coherence tomography (cmOCT).Biomed Opt Express 2(5):1184–1193.

12. Fingler J, Zawadzki RJ, Werner JS, Schwartz D, Fraser SE (2009) Volumetric micro-vascular imaging of human retina using optical coherence tomography with a novelmotion contrast technique. Opt Express 17(24):22190–22200.

13. Liu G, Lin AJ, Tromberg BJ, Chen Z (2012) A comparison of Doppler optical coherencetomography methods. Biomed Opt Express 3(10):2669–2680.

14. Jia Y, et al. (2012) Split-spectrum amplitude-decorrelation angiography with opticalcoherence tomography. Opt Express 20(4):4710–4725.

15. Mrejen S, Spaide RF (2013) Optical coherence tomography: Imaging of the choroidand beyond. Surv Ophthalmol 58(5):387–429.

16. Hendargo HC, McNabb RP, Dhalla A-H, Shepherd N, Izatt JA (2011) Doppler velocitydetection limitations in spectrometer-based versus swept-source optical coherencetomography. Biomed Opt Express 2(8):2175–2188.

17. Frank RN (2004) Diabetic retinopathy. N Engl J Med 350(1):48–58.18. Antonetti DA, Klein R, Gardner TW (2012) Diabetic retinopathy. N Engl J Med 366(13):

1227–1239.19. Group ETDRSR; Early Treatment Diabetic Retinopathy Study Research Group (1991)

Classification of diabetic retinopathy from fluorescein angiograms. ETDRS reportnumber 11. Ophthalmology 98(5, Suppl):807–822.

20. Group DRSR; The Diabetic Retinopathy Study Research Group (1981) Photocoagula-tion treatment of proliferative diabetic retinopathy. Clinical application of DiabeticRetinopathy Study (DRS) findings, DRS Report Number 8. Ophthalmology 88(7):583–600.

21. Ambati J, Ambati BK, Yoo SH, Ianchulev S, Adamis AP (2003) Age-related maculardegeneration: Etiology, pathogenesis, and therapeutic strategies. Surv Ophthalmol48(3):257–293.

22. Hee MR, et al. (1996) Optical coherence tomography of age-related macular de-generation and choroidal neovascularization. Ophthalmology 103(8):1260–1270.

23. Gass J (1997) Stereoscopic Atlas Of Macular Diseases: Diagnosis and Management(Mosby, St. Louis), Ed 4, pp 24–26.

24. Jia Y, et al. (2014) Quantitative optical coherence tomography angiography of cho-roidal neovascularization in age-related macular degeneration. Ophthalmology121(7):1435–1444.

25. Ying GS, et al.; Comparison of Age-related Macular Degeneration Treatments TrialsResearch Group (2013) Baseline predictors for one-year visual outcomes with ranibi-zumab or bevacizumab for neovascular age-related macular degeneration. Oph-thalmology 120(1):122–129.

26. Bhutto I, Lutty G (2012) Understanding age-related macular degeneration (AMD):Relationships between the photoreceptor/retinal pigment epithelium/Bruch’s mem-brane/choriocapillaris complex. Mol Aspects Med 33(4):295–317.

27. McLeod DS, et al. (2009) Relationship between RPE and choriocapillaris in age-relatedmacular degeneration. Invest Ophthalmol Vis Sci 50(10):4982–4991.

28. Mullins RF, Johnson MN, Faidley EA, Skeie JM, Huang J (2011) Choriocapillaris vasculardropout related to density of drusen in human eyes with early age-related maculardegeneration. Invest Ophthalmol Vis Sci 52(3):1606–1612.

29. Grunwald JE, Metelitsina TI, Dupont JC, Ying G-S, Maguire MG (2005) Reduced fo-veolar choroidal blood flow in eyes with increasing AMD severity. Invest OphthalmolVis Sci 46(3):1033–1038.

30. Coussa RG, Traboulsi EI (2012) Choroideremia: A review of general findings andpathogenesis. Ophthalmic Genet 33(2):57–65.

31. MacLaren RE, et al. (2014) Retinal gene therapy in patients with choroideremia: Initialfindings from a phase 1/2 clinical trial. Lancet 383(9923):1129–1137.

32. Miura M, Makita S, Iwasaki T, Yasuno Y (2011) Three-dimensional visualization ofocular vascular pathology by optical coherence angiography in vivo. Invest Oph-thalmol Vis Sci 52(5):2689–2695.

33. Kim DY, et al. (2013) Optical imaging of the chorioretinal vasculature in the livinghuman eye. Proc Natl Acad Sci USA 110(35):14354–14359.

34. Moult E, et al. (2014) Ultrahigh-speed swept-source OCT angiography in exudativeAMD. Ophthalmic Surg Lasers Imaging Retina 45(6):496–505.

35. Schwartz DM, et al. (2014) Phase-variance optical coherence tomography: A tech-nique for noninvasive angiography. Ophthalmology 121(1):180–187.

36. Spaide RF, Klancnik JM, Jr, Cooney MJ (2014) Retinal vascular layers in macular tel-angiectasia type 2 imaged by optical coherence tomographic angiography. JAMAOphthalmol 133(1):66–73.

37. Thorell M, et al. (2014) Swept-source OCT angiography of macular telangiectasia type2. Ophthalmic Surg Lasers Imaging Retina 45(5):369–380.

38. Riva CE, Petrig B (1980) Blue field entoptic phenomenon and blood velocity in theretinal capillaries. J Opt Soc Am 70(10):1234–1238.

39. Tam J, Tiruveedhula P, Roorda A (2011) Characterization of single-file flow throughhuman retinal parafoveal capillaries using an adaptive optics scanning laser oph-thalmoscope. Biomed Opt Express 2(4):781–793.

40. Drexler W, Fujimoto JG (2008) State-of-the-art retinal optical coherence tomography.Prog Retin Eye Res 27(1):45–88.

41. Blatter C, et al. (2012) Ultrahigh-speed non-invasive widefield angiography. J BiomedOpt 17(7):070505.

42. Jia Y, et al. (2014) Optical coherence tomography angiography of optic disc perfusionin glaucoma. Ophthalmology 121(7):1322–1332.

43. Choi W, et al. (2013) Choriocapillaris and choroidal microvasculature imaging withultrahigh speed OCT angiography. PLoS ONE 8(12):e81499.

44. Anonymous (2007) American National Standard for Safe Use of Lasers, ANSI Z136Series (American National Standards Institute, New York).

45. Kraus MF, et al. (2012) Motion correction in optical coherence tomography volumes on aper A-scan basis using orthogonal scan patterns. Biomed Opt Express 3(6):1182–1199.

E2402 | www.pnas.org/cgi/doi/10.1073/pnas.1500185112 Jia et al.

Dow

nloa

ded

by g

uest

on

Aug

ust 1

8, 2

020