AMD Book

AMD Book

English

AMD BOOK

www.amdbook.org

The most updated clinical and scientific information about the treatment of AMD: Age-related macular degeneration

Coordination
Silva, Rufino, MD, PhD

Department of Ophthalmology
Coimbra University Hospital. Faculty of Medicine. Coimbra. Portugal

Invited Professor of Ophthalmology. University of Coimbra.
Coimbra, Portugal.

Revision
Bandello, Francesco, MD, FEBO
Professor and Chairman
Department of Ophthalmology
University Vita-Salute
Scientific Institute San Raffaele,
Milan, Italy

Introduction
Cunha-Vaz, José Guilherme, MD, PhD
Emeritus Professor of Ophthalmology, University of Coimbra, Portugal
President of the Association for Innovation and Biomedical Research on Light and Image
(AIBILI), Coimbra, Portugal

 ©  GER GROUP - All rights reserved.

 

Authors

Authors

English

Ágoas, Victor, MD
Instituto Oftalmológico Gama Pinto, Lisbon, Portugal 

Alvarez-Vidal, Aurora
Ophthalmology Department. Clínica Universidad de Navarra. Pamplona. Spain
 
Biarnés, Marc, OD, MPH
Institut de la Màcula i de la Retina, Centro Médico Teknon, Barcelona. Spain
 
Bonet-Farriol, Elvira, MD

Ophthalmology Department. Clínica Universidad de Navarra. Pamplona. Spain

Brandão, Elisete, MD
Hospital S. João, Porto, Portugal

Cunha-Vaz, José Guilherme, MD, PhD
Emeritus Professor of Ophthalmology, University of Coimbra, Portugal President of the Association for Innovation and Biomedical Research on Light and Image (AIBILI), Coimbra, Portugal
 
Carneiro, Ângela, MD, PhD
Department of Ophthamology, Hospital São João, Faculty of Medicine of University of Porto. Porto, Portugal
Consultant of Ophthalmology, Retinal Specialist Invited Assistant of the Faculty of Medicine of University of Porto

Cachulo, Maria Luz, MD
Coimbra University Hospital - Coimbra, Portugal

Ciuffo, Gianfranco
Ophthalmology Department. Clínica Universidad de Navarra. Pamplona. Spain
 
Delcourt, Cecile, PhD
Inserm U897 (National Institute for Health and Medical Research), University Victor Segalen Bordeaux 2, Bordeaux, France
Researcher, in charge of the “Nutrition and eye diseases” axis.
 
De Nova Fernandez-Yañez, Elisa, MD

Ophthalmology Department. Clínica Universidad de Navarra. Pamplona. Spain

Ferreira, Natália, MD
Hospital Santo António-CHP - Porto, Portugal

Fernández, Maribel, MD, PhD
Instituto Oftalmológico Gómez-Ulla- Complexo Hospitalario Universitario de Santiago de Compostela. University of Santiago de Compostela. Spain.
 
Figueira, João P., MD
Department of Ophthamology Coimbra University Hospital
Assistant of Pathophysiology, Faculty of Medicine, Coimbra
 
Filomena Costa e Silva, MD
Prof. Dr. Fernando Fonseca Hospital, Amadora, Lisbon, Portugal
 
Flores, Rita, MD
Department of Ophthalmology Lisbon Hospital Center, Lisbon

Fonseca, Ana Fernandes
Instituto Oftalmológico Gama Pinto, Lisbon, Portugal 
ALM, Lisbon, Portugal
 
Garcia-Layana, Alfredo, MD, PhD
Ophthalmology Department. Clínica Universidad de Navarra. Pamplona. Spain
 
Gómez-Ulla, Francisco, MD, PhD
Medical Director of Instituto Oftalmológico Gómez-Ulla.University of Santiago de Compostela. Spain.
 
Gonzalez, Felipe
Optometrist. Unit of Optometry. Instituto Oftalmológico Gómez-Ulla
 
Guitana, Mário João Caeiro, MD
Service Head, Lisbon Ophtalmological Centre, Lisbon, Portugal

Gil, María,  MD
Instituto Oftalmológico Gómez-Ulla. Santiago de Compostela. Spain.
Department of Ophthalmology, Complejo Hospitalario Universitario de Santiago, Santiago de Compostela. Spain.

Henriques, José, MD
Instituto Oftalmológico Gama Pinto, Lisbon, Portugal
 
Introini, Ugo, MD
Department of Ophthalmology, University Vita-Salute Scientific Institute San Raffaele, Milano, Italy
Chief of Macula Service

Koh, Adrian, MD, FRCOphth, FRCS (Edin), MMed (Ophthalmology)
Founding Partner & Senior Consultant, Eye & Retina Surgeons, Singapore
 
Luis Arias, MD
Hospital Universitari de Bellvitge - University of Barcelona. Spain

Martinho, Rui, MD , FEBO
Ophthalmology Department of Hospital da Boavista (Hospitais Privados de Portugal) Porto, Portugal
Coordinator of the Ophthalmology Department of Hospital da Boavista (HPP) Porto, Portugal
 
Meireles, Angelina, MD
Department of Ophthalmology. Centro Hospitalar do Porto - Hospital Santo António, Porto
 
Monés, Jordi , MD
Institut de la Màcula i de la Retina - Centro Médico Teknon - Barcelona. Spain
 
Montero, Javier A MD, PhD
Pio del Rio Hortega Hospital, University of Valladolid. Valladolid. Spain
Alicante Institute of Ophthalmology, VISSUM, Vitreo-Retina Unit. Alicante. Spain

Marques, João Pedro, MD, MSc
Department of Ophthalmology, Centro Hospitalar e Universitário de Coimbra (CHUC), Coimbra, Portugal
Association for Innovation and Biomedical Research on Light and Image (AIBILI), Coimbra, Portugal
Faculty of Medicine, University of Coimbra (FMUC), Coimbra, Portugal
 
Nascimento, João C., MD
Retina Surgery Medical Chief. Retina Department - Vitreo Retinal Surgery Center Gama Pinto Ophthalmology Institut, Lisbon and Lisbon Retina Institute, Lisbon, Portugal

Picoto, Maria, MD
Ophthalmology Department, Hospital Beatriz Ângelo, Lisbon, Portugal

Pessoa, Bernardete, MD
Hospital Santo António-CHP - Porto, Portugal
 
Quintão, Teresa Luísa, MD
Hospital Medical career conultant - Ophtalmology. Gama Pinto Ophthalmology Institut, Lisbon, Portugal
 
Rebika, Hayette, MD
Ophthalmology, Clermont Ferrand Hospital, Clermont Ferrand, France
 
Rosa, Paulo Jorge Caldeira, MD
Graduate assistant and Medical Retina responsible. Gama Pinto Ophthalmology Institut, Lisbon, Portugal
 
Ruescas, Virginia Bautista, MD
Pio del Rio Hortega Hospital, University of Valladolid. Valladolid. Spain
 
Ruiz-Moreno, José Mª, MD, PhD
Clinic/hospital: Department of Ophthamology. Vissum Alicante & CHUA. Spain. Charge: Professor at the University Castilla La Mancha. Spain

Ruiz-Medrano, Jorge,  MD
Fellow of surgical and medical retina. Jules Gonin Eye Hospital. Fondation Asile des Aveugles. Lausanne. Switzerland.

Silva, Rufino, MD, PhD
Coimbra University Hospital. Faculty of Medicine. Coimbra. Portugal
Professor of Ophthalmology. University of Coimbra. Portugal 

Sabater Gozalvo, Alfonso, MD
Ophthalmology Department. Clínica Universidad de Navarra. Pamplona. Spain
 
Salinas-Alaman, Angel, MD, PhD
Ophthalmology Department. Clínica Universidad de Navarra. Pamplona. Spain

Teixeira, Susana, MD
Prof. Dr. Fernando Fonseca Hospital, Amadora, Lisbon, Portugal
​​​​​​​Vitreo-Retinal Consultant. Hospital Fernando da Fonseca - Medical and Surgical. Retina Department and Pediatric Retina Section, Portugal
 
Veludo, Maria João, MD
Department of Ophthalmology. Lisbon Hospital Center, Lisbon, Portugal

Vaz, Fernanda, MD
Ophthalmology Department, Centro Hospitalar de Lisboa Ocidental, Lisbon, Portugal
 
Zarranz-Ventura Javier, MD
Ophthalmology Department. Clínica Universidad de Navarra. Pamplona. Spain

Introduction

Introduction

English

Author:

José Cunha-Vaz, MD, PhD
Association for Innovation and Biomedical Research on Light and Image (AIBILI), Coimbra, Portugal

 

Age-related macular degeneration (AMD) is now one of the major causes of central vision loss.

It involves the macular area and when it progresses and destroys the central fovea, quality of life is seriously compromised.

The ability to read, drive, recognize faces or watch television is impaired or lost.

It is a disease associated with aging and progressive tissue degeneration.

Most of these senior citizens had anticipated the opportunity to enjoy life at leisure doing their preferred activities and find profound limitations and are deeply disappointed in their expectations.

This was the dismal state of affairs until a few years ago when the introduction of new therapeutic agents changed dramatically the expected outcome.

The most important development has been the clinical demonstration that agents inhibiting vascular endothelial growth factor and, therefore, the formation and development of new vessels not only preserve visual acuity but also improve visual function.

This find was a true revolution in ophthalmology.

The retina like the central nervous system, particularly in old age, was not considered capable of regeneration and, therefore, new therapies were expected to offer only stabilization of the disease.

When laser photocoagulation was shown to be effective for the treatment of diabetic retinopathy this was a breakthrough but it offered only stabilization of the disease and maintenance of visual acuity present at time of treatment.

Now, new available treatments offer real improvement in visual acuity, implicating improved visual function and some degree of recovery of the retinal neuronal network.

AMD is, therefore, an active scientific area with new information becoming available almost everyday on the pathophysiology, clinical phenotypes, markers of disease progression, new treatments and novel treatment regimens.

It is a complex multifactorial disease where aging associates with genetic factors and inflammatory responses to local cell injury.

This means that the treatment of AMD must also address the various factors involved in disease progression and, most likely, will involve a combination of therapies after clear identification of the different AMD phenotypes.

This book reviews the new concepts on AMD and is, therefore, timely.

It is a concerted effort of Portuguese ophthalmologists which together with a few other European experts in the field cover the subject paying particular attention to daily practice management of AMD.

I am particularly happy to see this collaborative effort coming from my country, Portugal.

It shows that Portuguese ophthalmology is at the forefront of European and International ophthalmology.

It shows also that there is in Portugal a real spirit of collaboration between colleagues working together to improve the vision of their patients.

Epidemiology of AMD

Epidemiology of AMD

English

Author:

Cécile Delcourt, MD, PhD
Inserm, U897, Bordeaux, France Université Bordeaux 2, Bordeaux, France

 

1. Introduction

Although age-related macular degeneration (AMD) is the third cause of blindness worldwide, and the first in industrialized countries(1), epidemiological data on this disease remain scarce and partial. The very first studies were published in the 1980’s. Since then, a number of population-based studies have been conducted, first in Caucasian populations of the United States and other industrialized countries (Australia, Europe). Studies have more recently been extended to other ethnical subgroups of industrialized countries (African Americans, Latinos) and to other parts of the world (India, China, and Japan).

 

2. Prevalence of late AMD in Caucasians from industrialized countries

The prevalence of late AMD in the main epidemiological studies performed in Caucasians of industrialized countries are presented in Table 1.

Author

Study Years study conducted

Country Number of subjects, age

Prevalence of late AMD(%)

United States

 

 

 

Friedman, Baltimore Eye Study(28)

1985-1988, USA

N=2518, age >= 40 years

1.23

Klein, Beaver Dam Eye Study(29)

1988-1990, USA

N=4752, age >= 40 years

1.64

Bressler, Salisbury Eye Evaluation Project(30)

1993-1995, USA

N=1773, age >= 65 years

Exudative AMD : 1.7 Geographic atrophy: 1

Klein, Atherosclerosis Risk in Communities Study(31)

1993-1995, USA

N=8984 , age 48-72 years (one eye only)

0.

Klein, Cardiovascular Health Study(32)

1997-1998, USA

N=1998, age 69-97 years (one eye only)

1.3

Klein, Multi-Ethnic Study of Atherosclerosis(10)

2000-2002, US

N=2315, age 45-84 years

0.6

Australia

 

 

 

Mitchell, Blue Mountains Eye Study(33)

1992-1994, Australia

N=3632, age >= 50

2.06

Van Newkirk, Melbourne VIP(34)

1992-1996, Australia

N=4345, age >= 40 years

0.68

Europe

 

 

 

Vingerling, Rotterdam Study(35)

1990-1993, Netherlands

N=6774, age >= 55 years

1.65

Vingerling, Rotterdam Study(35)

1995-1997, France

N=2196, age >= 60 years

1.9

Jonasson, Reykjavik Eye Study(5)

1996, Iceland

N=1022, age >= 50 years

3.5

Topouzis, Thessaloniki Eye Study(37)

2000, Greece

N=1022, age >= 50 years

2.5

Bjornsson, Oslo Macular Study(6)

2002, Norway

N=459, age > 50 years

2.8

Augood, Eureye(4)

2001-2002,

N=4753, age >= 65 years

3.3

 

7 european countries

 

 

Table 1. Prevalence of late AMD in Caucasians from industrialized countries

 

The included studies were restricted to those having classified AMD from retinal photographs and used the international classification(2), in order to make comparisons between studies easier. In these studies, late AMD is defined by the presence of neovascular AMD and/or geographic atrophy. In the United States, the prevalence of late AMD ranged from 0.2% to 1.6% according to studies. Since the prevalence of late AMD increases sharply with age, most of the differences between studies were due to differences in the age distribution. For instance, the lowest prevalence rate (0.2%) was observed for the Atherosclerosis Risk in Communities (ARIC) Study. In this study, the age range of participants was 48-72 years, thus excluding the oldest subjects, in which the prevalence of late AMD is the highest. This study also performed photographs in only one eye of each participant, which may have led to undetected unilateral cases and thus underestimation of the prevalence of late AMD. This is also the case for the Cardiovascular Health Study, which also found relatively low prevalence (1.3% in subjects aged 69 to 97 years). Prevalence rates observed in Caucasians from Australia and Europe appear relatively similar to those observed in the United States (Table 1), when taking into account differing age distributions in the different studies. In European studies, the prevalence of late AMD ranged from 1.65% to 3.5%.

In order to better compare the prevalence rates among studies, age-specific prevalence rates need to be estimated. This has been done in a meta-analysis performed in 2004 by Friedman et al(3). The authors concluded that prevalence rates were not different among Caucasian populations of industrialized countries, including the United States, Australia and the Netherlands. Similarly, in the EUREYE Study, which included 7 European countries (Norway, Estonia, Ireland, France, Italy, Greece, Spain), no significant differences were found among the participating countries(4). Therefore, the prevalence of late AMD appears to be similar in the Caucasian populations from the United States, Australia and European countries, despite major geographical and lifestyle differences. Thus, the meta-analysis from Friedman et al. probably constitutes the most reliable available estimate of prevalence rates for these countries, since it bears on studies gathering altogether more than 25000 subjects(3). In this meta-analysis, the prevalence rates increase sharply with age (Figure 1), from less than 0.5% in subjects 50 to 60 years, to 12% and 16% in men and women aged 80 years or more, respectively. While men tend to have higher prevalence rates than women in the younger age groups (less than 80 years), late AMD is more frequent in women than in men in the oldest age group (80 years or more). This may be due to the higher mortality in ageing men, which leads to a higher selection of the oldest subjects. Finally, in the late AMD groups, two clinical entities exist: neovascular AMD and geographic atrophy. In most studies, neovascular cases represent about 55% of all late AMD cases(3). Nordic European countries (in particular Iceland and Norway) may be an exception, since the few studies performed in these countries suggest a higher rate of geographic atrophy(5,6).

 

epidemiology of amd_figure1.jpg

Figure 1. Prevalence of late AMD according to age and gender in Caucasians from industrialized countries (adapted from the meta-analysis by Friedman et al.(3))

 

3. Prevalence of late AMD in other ethnic groups

As shown in Table 2, the prevalence of late AMD is much lower in African-Americans than in Caucasians.

Author

Study Years study conducted

Country Number of subjects, age

Prevalence of late AMD(%)

African-Americans

 

 

 

Friedman, Baltimore Eye Study(28)

1985-1988, USA

N=1843, age >= 40 years

0.22

Bressler, Salisbury Eye Evaluation Project(30)

1993-1995, USA

N=666, age >= 65 years

Exudative AMD : 1.1

Geographic atrophy : 0.3

Klein, Atherosclerosis Risk in Communities

1993-1995, USA

N= 2548 , age 48-72 years (one eye only)

0.04

Study(31)

 

 

 

Klein, Cardiovascular Health Study(32)

1997-1998, USA

N=363, age 69-97 years (one eye only)

0.3

Hispanics

 

 

 

San Luis Vally(17)

1983, USA

N=3995, age 43-74 years

0.09

Proyecto VER(15)

1997-1999, USA

N=2776, age >= 50 years

0.5

Los Angeles Latino Eye Study(16)

2000-2003, USA

N=5875, age >= 40 years

0.43

Klein,Multi-Ethnic Study of Atherosclerosis(10)

2000-2002, USA

N=1280, age 45-84 years

0.2

Chinese

 

 

 

Klein, Multi-Ethnic Study of Atherosclerosis(10)

Klein, Multi-Ethnic Study of Atherosclerosis(10)

N=699, age 45-84 years

1.0

Table 2. Prevalence of late AMD in other ethnic groups from the United States

Most of the cited studies were multi-ethnic and performed direct comparisons between African-Americans and Caucasians. Statistical power was generally low for these comparisons, but these studies globally suggest that the prevalence of late AMD is at least 2-fold lower in African-Americans. In the Barbados Eye Study, bearing on subjects of African origin from the Barbados Island, the prevalence rate of late AMD (0.57%) was similar to those observed in African-Americans (Table 3)(7).

Author

Study Years study conducted

Country Number of subjects, age

Prevalence of late AMD(%)

Barbados

 

 

 

Schachat, Barbados Eye Study(7)

11987-1992, Barbados

N=3444, age 40-84 years

0.57

Japan

 

 

 

Oshima, Hisayama Study(8)

1998, Japan

N=889, age >= 50 years

0.89

Kawasaki, Funagata Study(9)

200-2002, Japan

N=1625, age >= 35 years

0.5

 

 

N=1037, age >= 55 years

0.8

India

 

 

 

Krishnaiah, Andrah Pradesh Study(38)

1996-2000, India

N=3723, age >= 40 years

1.9

Nirmalan, Aravind Eye Study(18)

1995-1997, India

N=4197, age >= 40 years

0.6

Krishnan, INDEYE(19)

2005-2007, India

N=4266, age >= 60 years

1.2

China

 

 

 

Li, Beijing Eye Study(11)

2001, China

N=4376, age >= 40 years

0.2

Taïwan

 

 

 

Chen, Shihpay Study(12)

1999-2000, Taïwan

N=1105, age >= 65 years

1.9

Singapore

 

 

 

Kawasaki, Singapore Malay Eye Study(13)

2004-2006, Singapore

N=3265, age 40-80 years

0.7

Greenland

 

 

 

Andersen, Greenland Inuit Eye Study(14)

2002-2003, Greenland

N=642, age >= 60 years

9.1

Table 3. Prevalence of late AMD in other countries

 

Therefore, people of African origin appear to be at much lower risk of late AMD than Caucasians, although no data are available from African countries. American Hispanics also appear to be at lower risk for late AMD than Caucasian, with prevalence rates ranging from 0.09% to 0.5% among studies (Table 2). The reasons for the lower prevalence rates in African-American and Hispanics are unclear. No data are available on minorities of other industrialized countries, in particular in Europe. This constitutes a limitation to the estimation of the prevalence of AMD in European countries, which have different types of ethnic groups than in the United States (populations originating from North Africa, Middle-east countries, India…). Given the major differences observed in ethnic groups in the United States, such epidemiological data in European minorities are warranted.

Other epidemiological studies have mainly been conducted in Asia (Japan, India and China). In contrast to what was originally thought, late AMD is not rare in these Asian populations, as shown in Table 3. In two Japanese studies(8,9), the prevalence rates in Japanese men were similar to those observed in Caucasian men from industrialized countries, while late AMD appeared rare in Japanese women. This gender-effect, which is not observed in industrialized countries, may be due to gender-related smoking habits in Japanese. Indeed, smoking is a major risk factor for AMD, in all studied populations. In these two studies, smoking was very frequent in Japanese men, while rare in Japanese women.

In three epidemiological studies performed in India, the prevalence rates of late AMD ranged from 0.6% to 1.9% (Table 3). Below the age of 80 years, the prevalence rates were rather similar to those observed in industrialized countries. Above 80 years, the figures are mostly unreliable because of the very high prevalence of unoperated cataracts in this population, leading to a very high percentage of ungradable retinal photographs. The global prevalence rates were therefore probably underestimated.

Three studies have been performed in subjects of Chinese origin: one in the United States (Table 2)(10), one in mainland Beijing(11)and one in Taïwan(12). While prevalence rates in Chinese from the United States and Taïwan seemed similar to those in Caucasians, the prevalence rates in Chinese from Mainland China were much lower.
The reasons for these differences are unclear, and may be related to gene-environment interactions, since Chinese from the United States and Taïwan tend to live a more industrialized lifestyle. More epidemiological data are needed to confirm and understand the reasons for these differences.

Finally, in a study from Singapore, the prevalence of late AMD was similar to that in Caucasians(13). Globally, the prevalence of late AMD appears high in Asians, with the only exception of Chinese from Mainland China, which remains to be confirmed.

A study in Inuits from Greenland shows strikingly high rates of late AMD, with particularly high rates of neovascular AMD(14). This is in contrast from the observations of European Nordic countries, which found higher rates of geographic atrophy(5,6) and also in contrast with original observations in this population, which also found higher rates of geographic atrophy in Inuits. The reasons for these differences are unclear.

 

4. Prevalence of early AMD

Late AMD is preceded by early, usually asymptomatic, retinal abnormalities. The long-term cohort studies have helped to characterize the lesions that are the most predictive of incident late AMD. While large soft drusen (>125 microns) and pigmentary abnormalities clearly are the hallmarks of early AMD, there is no consensus on the precise definition of early AMD, and several classification systems coexist. This makes it difficult to assess and compare the prevalence of early AMD among geographical areas and ethnic groups.

Despite these difficulties, it appears that early AMD is frequent in the elderly. For instance, in the meta-analysis by Friedman et al, the prevalence of large drusen increased from about 1.5% in Caucasians aged 40-49 years, to more than 25% in those aged 80 years or more(3). Similar rates were observed in the EUREYE Study, performed in Europe(4).

Consistently with what was observed for late AMD, early AMD appears less frequent in African-Americans than in Caucasian(3,10). By contrast, while late AMD is also less frequent in Hispanic-Americans, the prevalence of early abnormalities appears similar, or even higher than in Caucasians(10,15-17). The reasons for these differences are unclear, and suggest that different factors may be implicated in the etiology of early and late AMD. In most studies performed in Asians, the prevalence of early AMD appears similar to that observed in Caucasians, consistently to what was found for late AMD(8-10,12,13,18,19).

 

5. Prevalence of late AMD in Portugal

To our knowledge, there are no published data on the prevalence of AMD in Portugal. What can therefore be extrapolated from available data in other countries? As explained above, the prevalence of late AMD appears to be similar in Caucasian populations from industrialized countries. Its best estimate therefore derives from the meta-analysis by Friedman et al(3). We, thus, applied the age and gender-specific prevalence rates of the meta-analysis to the demographic data from Portugal, leading to an estimation of about 84 000 cases of late AMD in Portugal (Table 4).

Population of Portugal

AMD prevalence (%) *

Number with AMD

Men

 

 

 

50-54

309 484

0,34

1 052

55-59

268 899

0,41

1 102

60-64

256 179

0,63

1 614

65-69

244 230

1,08

2 638

70-74

196 615

1,98

3 893

75-79

143 439

3,97

5 695

80+

123 934

11,90

14 748

Total men

1 542 780

1,99

30 742

 

 

 

 

Women

 

 

 

50-54

333 032

0,20

666

55-59

302 553

0,22

666

60-64

294 737

0,35

1 032

65-69

293 935

0,70

2 058

70-74

257 347

1,52

3 912

75-79

204 627

3,44

7 039

80+

229 366

16,39

37 593

Total women

1 915 597

2,76

52 965

Total

3 458 377

2,42

83 707

Table 4. Prevalence of late AMD in Portugal according to age and gender

 

A major limitation to this approach is that it does not take into account the prevalence rates of AMD in non Caucasians living in Portugal. As shown above, the prevalence of late AMD appear lower in subjects of African origin. However, no data are available on the prevalence of AMD in subjects of African, or other ethnic origins (for instance North Africa and Asia) in European populations. It appears difficult to assess the prevalence of AMD while taking into account the multi-ethnic structure of the Portuguese population. However, these ethnic subgroups represent small minorities in Portugal, so that this may only marginally affect the global prevalence rates.

According to these estimations, about two thirds of cases are women, and two thirds of cases are aged 80 years or more. Since it represents about 55% of late cases, neovascular AMD probably affects about 46 000 subjects.

 

>> References

References - Epidemiology of AMD

References - Epidemiology of AMD

English

1. Resnikoff S, Pascolini D, Etya’ale D, Kocur I, Pararajasegaram R, Pokharel GP, Mariotti SP. Global data on visual impairment in the year 2002. Bull World Health Organ 2004; 82 (11): 844-51.

2. Bird AC, Bressler NM, Bressler SB, Chisholm IH, Coscas G, Davis MD, de Jong PT, Klaver CC, Klein BE, Klein R, et al. An international classification and grading system for age-related maculopathy and age-related macular degeneration. The International ARM Epidemiological Study Group. Surv Ophthalmol. 1995; 39 (5): 367-74.

3. Friedman DS, O’Colmain BJ, Muñoz B, Tomany SC, McCarty C, de Jong PT, Nemesure B, Mitchell P, Kempen J; Eye Diseases Prevalence Research Group. Prevalence of age-related macular degeneration in the United States. Arch Ophthalmol 2004; 122 (4): 564-72.

4. Augood CA, Vingerling JR, de Jong PT, Chakravarthy U, Seland J, Soubrane G, Tomazzoli L, Topouzis F, Bentham G, Rahu M, Vioque J, Young IS, Fletcher AE. Prevalence of age-related maculopathy in older Europeans: the European Eye Study (EUREYE). Arch Ophthalmol 2006; 124 (4): 529-35.

5. Jonasson F, Arnarsson A, Sasaki H, Peto T, Sasaki K, Bird AC. The prevalence of age-related maculopathy in iceland: Reykjavik eye study. Arch Ophthalmol 2003; 121 (3):379-85.

6. Björnsson OM, Syrdalen P, Bird AC, Peto T, Kinge B. The prevalence of age-related maculopathy (ARM) in an urban Norwegian population: the Oslo Macular study. Acta Ophthalmol Scand 2006; 84 (5): 636-41.

7. Schachat AP, Hyman L, Leske MC, Connell AM, Wu SY. Features of age-related macular degeneration in a black population. The Barbados Eye Study Group. Arch Ophthalmol 1995; 113 (6): 728-35.

8. Oshima Y, Ishibashi T, Murata T, Tahara Y, Kiyohara Y, Kubota T. Prevalence of age related maculopathy in a representative Japanese population: the Hisayama study. Br J Ophthalmol 2001; 85 (10): 1153-7

9. Kawasaki R, Wang JJ, Ji GJ, Taylor B, Oizumi T, Daimon M, Kato T, Kawata S, Kayama T, Tano Y, Mitchell P, Yamashita H, Wong TY. Prevalence and risk factors for age-related macular degeneration in an adult Japanese population: the Funagata study. Ophthalmology 2008; 115 (8): 1376-81, 1381.e1-2.

10. Klein R, Klein BE, Knudtson MD, Wong TY, Cotch MF, Liu K, Burke G, Saad MF, Jacobs DR Jr. Prevalence of age-related macular degeneration in 4 racial/ethnic groups in the multi-ethnic study of atherosclerosis. Ophthalmology 2006; 113 (3): 373-80.

11. Li Y, Xu L, Wang YX, You QS, Yang H, Jonas JB. Prevalence of age-related maculopathy in the adult population in China: the Beijing eye study. Am J Ophthalmol 2008; 146 (2): 329.

12. Chen SJ, Cheng CY, Peng KL, Li AF, Hsu WM, Liu JH, Chou P. Prevalence and associated risk factors of age-related macular degeneration in an elderly Chinese population in Taiwan: the Shihpai Eye Study. Invest Ophthalmol Vis Sci 2008; 49 (7): 3126-33.

13. Kawasaki R, Wang JJ, Aung T, Tan DT, Mitchell P, Sandar M, Saw SM, Wong TY; Singapore Malay Eye Study Group. Prevalence of age-related macular degeneration in a Malay population: the Singapore Malay Eye Study. Ophthalmology 2008; 115 (10): 1735-41. 14. Andersen MV, Rosenberg T, la Cour M, Kiilgaard JF, Prause JU, Alsbirk PH, Borch-Johnsen K, Peto T, Carstensen B, Bird AC. Prevalence of age-related maculopathy and age-related macular degeneration among the inuit in Greenland. The Greenland Inuit Eye Study. Ophthalmology 2008; 115 (4): 700-707.e1. 15. Muñoz B, Klein R, Rodriguez J, Snyder R, West SK. Prevalence of age-related macular degeneration in a population-based sample of Hispanic people in Arizona: Proyecto VER. Arch Ophthalmol 2005; 123 (11): 1575-80.

16. Varma R, Fraser-Bell S, Tan S, Klein R, Azen SP; Los Angeles Latino Eye Study Group. Prevalence of age-related macular degeneration in Latinos: the Los Angeles Latino eye study. Ophthalmology 2004; 111 (7): 1288-97.

17. Cruickshanks KJ, Hamman RF, Klein R, Nondahl DM, Shetterly SM. The prevalence of age-related maculopathy by geographic region and ethnicity. The Colorado-Wisconsin Study of Age-Related Maculopathy. Arch Ophthalmol 1997; 115 (2): 242-50.

18. Nirmalan PK, Katz J, Robin AL, Tielsch JM, Namperumalsamy P, Kim R, Narendran V, Ramakrishnan R, Krishnadas R, Thulasiraj RD, Suan E. Prevalence of vitreoretinal disorders in a rural population of southern India: the Aravind Comprehensive Eye Study. Arch Ophthalmol 2004; 122 (4): 581-6.

19. Krishnan T, Ravindran RD, Murthy GVS, Vashist P, Fitzpatrick KE, Thulasiraj RD, John N, Maraini G, Camparini M, Chakravarthy U, Fletcher AE. Prevalence of early and late Age-Related Macular Degeneration in India: The INDEYE Study. Invest Ophthalmol Vis Sci 2010; 51: 701-707

20. Klein R, Klein BE, Knudtson MD, Meuer SM, Swift M, Gangnon RE. Fifteen-year cumulative incidence of age-related macular degeneration: the Beaver Dam Eye Study. Ophthalmology 2007; 114 (2): 253-62.

21. Wang JJ, Rochtchina E, Lee AJ, Chia EM, Smith W, Cumming RG, Mitchell P. Ten-year incidence and progression of age-related maculopathy: the blue Mountains Eye Study. Ophthalmology 2007; 114 (1): 92-8.

22. Mukesh BN, Dimitrov PN, Leikin S, Wang JJ, Mitchell P, McCarty CA, Taylor HR. Five-year incidence of age-related maculopathy: the Visual Impairment Project. Ophthalmology 2004; 111 (6): 1176-82.

23. Jonasson F, Arnarsson A, Peto T, Sasaki H, Sasaki K, Bird AC. 5-year incidence of age-related maculopathy in the Reykjavik Eye Study. Ophthalmology 2005; 112 (1): 132-8.

24. Delcourt C, Lacroux A, Carrière I; POLA Study Group. The three-year incidence of age-related macular degeneration: the “Pathologies Oculaires Liées à l’Age” (POLA) prospective study. Am J Ophthalmol 2005; 140 (5): 924-6.

25. van Leeuwen R, Klaver CC, Vingerling JR, Hofman A, de Jong PT. The risk and natural course of age-related maculopathy: follow-up at 6 1/2 years in the Rotterdam study. Arch Ophthalmol 2003; 121 (4): 519-26.

26. Leske MC, Wu SY, Hennis A, Nemesure B, Yang L, Hyman L, Schachat AP; Barbados Eye Studies Group. Nine-year incidence of age-related macular degeneration in the Barbados Eye Studies. Ophthalmology 2006; 113 (1): 29-35.

27. Miyazaki M, Kiyohara Y, Yoshida A, Iida M, Nose Y, Ishibashi T. The 5-year incidence and risk factors for age-related maculopathy in a general Japanese population: the Hisayama study. Invest Ophthalmol Vis Sci 2005; 46 (6): 1907-10.

28. Friedman DS, Katz J, Bressler NM, Rahmani B, Tielsch JM. Racial differences in the prevalence of age-related macular degeneration: the Baltimore Eye Survey. Ophthalmology 1999; 106 (6): 1049-55.

29. Klein R, Klein BE, Linton KL. Prevalence of age-related maculopathy. The Beaver Dam Eye Study. Ophthalmology 1992; 99 (6): 933-43.

30. Bressler SB, Muñoz B, Solomon SD, West SK; Salisbury Eye Evaluation (SEE) Study Team. Racial differences in the prevalence of age-related macular degeneration: the Salisbury Eye Evaluation (SEE) Project. Arch Ophthalmol 2008; 126 (2): 241-5.

31. Klein R, Clegg L, Cooper LS, Hubbard LD, Klein BE, King WN, Folsom AR. Prevalence of age-related maculopathy in the Atherosclerosis Risk in Communities Study. Arch Ophthalmol 1999; 117 (9): 1203-10.

32. Klein R, Klein BE, Marino EK, Kuller LH, Furberg C, Burke GL, Hubbard LD. Early age-related maculopathy in the cardiovascular health study. Ophthalmology 2003; 110 (1): 25-33.

33. Mitchell P, Smith W, Attebo K, Wang JJ. Prevalence of age-related maculopathy in Australia. The Blue Mountains Eye Study. Ophthalmology 1995; 102 (10): 1450-60..

34. VanNewkirk MR, Nanjan MB, Wang JJ, Mitchell P, Taylor HR, McCarty CA. The prevalence of age-related maculopathy: the visual impairment project. Ophthalmology 2000; 107 (8): 1593-600.

35. Vingerling JR, Dielemans I, Hofman A, Grobbee DE, Hijmering M, Kramer CF, de Jong PT. The prevalence of age-related maculopathy in the Rotterdam Study. Ophthalmology 1995; 102 (2): 205-10.

36. Delcourt C, Diaz JL, Ponton-Sanchez A, Papoz L. Smoking and age-related macular degeneration. The POLA Study. Pathologies Oculaires Liées à l’Age. Arch Ophthalmol 1998; 116 (8): 1031-5.

37. Topouzis F, Coleman AL, Harris A, Anastasopoulos E, Yu F, Koskosas A, Pappas T, Mavroudis L, Wilson MR. Prevalence of age-related macular degeneration in Greece: the Thessaloniki Eye Study. Am J Ophthalmol 2006; 142 (6): 1076-9. 5.

38. Krishnaiah S, Das T, Nirmalan PK, Nutheti R, Shamanna BR, Rao GN, Thomas R. Risk factors for age-related macular degeneration: findings from the Andhra Pradesh eye disease study in South India. Invest Ophthalmol Vis Sci 2005; 46 (12): 4442-9.

Modifiable risk factors for AMD

Modifiable risk factors for AMD

English

Author:

Cécile Delcourt, MD, PhD
Inserm, U897, Bordeaux, France Université Bordeaux 2, Bordeaux, France

 

1. Introduction

The epidemiological studies conducted in the past 25 years have helped identifying major modifiable risk factors for AMD. In particular, smoking and nutrition appear ever more important in determining the occurrence of AMD, and may, in the future, lead to prevention strategies.

 

2. Smoking

Smoking is the best characterized risk factor for AMD(1). The initial observations performed in Caucasian populations from Western countries(2), are now being confirmed in other ethnic groups, such as African-Americans(3), Latino-American(4), or Asian populations(5-7). In most studies, the risk for late AMD was multiplied by 2.5 to 4.5 in current smokers. In addition, the dose-response relationship was explored in some studies(6,8-12). Most of these studies found that the risk for AMD increased with increasing number of cigarettes smoked per day, and, even more, with number of pack-years smoked, which is an indicator of cumulative smoking over the lifetime (mean number of packs smoked/day x duration of smoking (years). Moreover, the risk for AMD appeared to decrease with time from cessation of smoking.  Former smokers generally demonstrated a lower risk for AMD than current smokers. Several studies have shown that the risk for AMD in subjects having ceased smoking for more than 20 years was similar to the risk in never smokers(8-10,13). One study suggested that passive smoking is also associated with an increased risk for AMD(9), while this association did not reach statistical significance in another study(14). Finally, smoking appeared to be related to similar risks for both types of late AMD (geographic atrophy and neovascular AMD)(9,13,15-16). By contrast, associations with early AMD were weaker in the vast majority of published studies, and often not statistically significant(8-9,13,15,17-18)

Overall, the strength of the association (about 3-fold increased risk in current smokers), its consistency across different populations, the observation of a clear dose-response relationship in most studies, and the decrease of the risk with stopping smoking are all strong arguments in favour of a causal role of tobacco smoking in the aetiology of late AMD. 

The exact mechanisms by which smoking increases the risk for AMD are unclear, and probably multiple, including oxidative stress, inflammation and decreased macular pigment. Finally, recent studies gave important insights on the joint effects of smoking and genetic polymorphisms, showing that the risk for AMD is particularly high in smokers bearing at-risk polymorphisms in the CFH or LOC387715 genes(19-21).

Other vascular risk factors, such as systemic hypertension, obesity, diabetes, plasma lipids or alcohol drinking may be associated with an increased risk of AMD, but epidemiological studies have been inconsistent in this field(22). At the time being, they remain putative, but not clearly identified risk factors for AMD. 

 

Nutritional factors

More recently, epidemiological studies have focused on the potential association of AMD with nutritional factors. Mainly three types of nutritional factors have been investigated for their potential protection against eye ageing: antioxidants (mainly vitamins C and E, zinc), the carotenoids lutein and zeaxanthin and omega 3 polyunsaturated fatty acids (PUFA). 

The retina is particularly susceptible to oxidative stress because of the high level of in-site reactive oxygen species production, due in particular to light exposure and high metabolic activity(23).

Epidemiological studies are mostly in favour of a protective role of antioxidants for AMD(24).

Moreover, the Age-Related Eye Diseases Study (AREDS), a randomized clinical trial performed in the United States and including on almost 5000 subjects supplemented for five years, showed a significant 25% reduction of the incidence of late AMD with supplementation in antioxidants and zinc, by comparison with placebo(25). In this field, data from the United States should be extrapolated to European populations with caution. Indeed, vitamin supplements are widely used in the American population, while this is rarely the case in Europe. For instance, two thirds of the AREDS participants used vitamin supplements, in addition to the supplementation tested in the study(25). Plasma vitamin C concentration at baseline in the AREDS (before the initiation of the study supplementation) was 62 micromol/l(25), whereas it was 31.6 micromol/l in men and 40.5 micromol/l in women of the Pathologies Oculaires Liées à l’Age (POLA) Study, performed in the South of France(26). Similarly, in the EUREYE Study, plasma vitamin C concentrations ranged from 35.5 micromol/l to 48.4 micromol/l in seven European countries(27). Therefore, antioxidant intake is much lower in European populations than in the United States, with part of European populations being at risk of clinical deficiency in these vitamins. Two European studies suggested that the benefit to be expected from increased antioxidant intake may be more important in our populations with low antioxidant intake. Indeed, in the French POLA Study, we observed an 80% decreased risk for late AMD in the subjects with higher plasma vitamin E, by comparison to those with lower concentrations(28), a much stronger effect than the 25% reduction in risk observed in the AREDS Study. Moreover, we observed a 25% reduction in risk for early ARM, whereas the AREDS Study showed no benefit of antioxidant supplements for early ARM. Similarly, results from the Rotterdam Study showed a decreased risk for early ARM in subjects with high dietary intake of vitamin E or zinc, by comparison with those with low intake(29). A European supplementation study would be needed to better assess the benefit of antioxidant supplementation in European populations.

A more recent research domain evaluated the role of two carotenoids, lutein and zeaxanthin, in the protection of the retina and the lens. These carotenoids accumulate in the macula, where they are known as the macular pigment(30). Besides their antioxidant properties, they probably act as a filter against the phototoxic effects of blue light(30). To date, five epidemiological studies have assessed the associations of the risk of AMD with plasma concentration of lutein and zeaxanthin(31-35). As shown in Fig. 1, all five studies showed a decreased risk for AMD in subjects with high plasma concentrations of lutein and zeaxanthin, although the association was statistically significant only in 2 studies(32-35). With regard to dietary intake, four prospective population-based studies were published(29,36-38).

Figure 1. Association of the risk of AMD with plasma levels of lutein and zeaxanthin in cross-sectional and case-control studies (odds-ratios with 95% confidence interval)
OR below 1 suggest a protective role and OR greater than 1 suggest a deleterious role. References of the cited studies: Beaver Dam(31); EDCC(32); NHANES III(33); Gale et al(34); POLA(35)

 

These studies assessed the risk for developing AMD (in subjects initially free of AMD), according to their dietary intake of lutein and zeaxanthin. As shown in Fig. 2, the results for these dietary studies are less clear than for those on plasma measurements.

Figure 2. Associations of the risk for AMD with dietary lutein and zeaxanthin, in published epidemiological prospective studies. References of the cited studies: Beaver Dam(36); Health Professionals(37); Rotterdam(29); Blue Mountains(38).

 

Only one study found a significantly reduced risk for AMD in subjects with high dietary intake of lutein and zeaxanthin(38). However, dietary assessment methods rely on the subjects’ memory and perceptions and face the difficulties of the extreme day-to-day variability of human diet, the bias in reporting due to social standards and nutritional recommendations and the estimations of nutritional contents of food items. Biomarkers have the advantages of being objective, and of taking into account individual variations in bioavailability and metabolism. For instance, smoking and obesity are known to decrease the bioavailability of carotenoids(39-40). Despite normal dietary intake in lutein and zeaxanthin, subjects may be at higher risk for AMD because of decreased bioavailability, associated with lower plasma concentrations of these components. However, currently available studies including plasma measurements are cross-sectional or case-control studies, where the plasma measurements were performed in subjects already affected by the disease. The stronger associations found in these studies may therefore be explained by reverse causality (i.e. plasma carotenoids were lower because of change of dietary habits in subjects with AMD, for instance).

Globally, the few available epidemiological studies suggest a protective role of lutein and zeaxanthin in AMD, but other studies are needed in this field, in particular larger, prospective studies including dietary and plasma measurements.

Finally, omega 3 PUFA include a precursor (alpha-linolenic acid (ALA)), and three long-chain derivatives (eicosapentaenoic acid (EPA), docosapentaenoic acid (DPA) and docosahexaenoic acid (DHA)). ALA is an essential nutrient, since humans cannot synthesize it de novo and therefore rely on diet as its sole source. Synthesis of the long-chain derivatives is very limited in humans(41), who must therefore rely on their dietary supply, mainly by fish and seafood. Long-chain omega-3 PUFA have important structural and protective functions in the retina. DHA is a major component of the photoreceptors(42). They also have protective functions, including the systemic anti-inflammatory, anti-angiogenic and anti-apoptotic functions and specific actions in the retina such as increase in lysosomial acid lipase, leading to increased lipid degradation in the retinal pigment epithelium(42). As shown in Fig. 3 and Fig. 4, seven case-control(43-44) or cross-sectional studies(45-49) and five prospective studies(50-54) assessed the associations of dietary intake of long-chain omega 3 PUFA or fish with AMD. In spite of differences in populations, methods and types of studies, all studies showed a reduced risk for AMD in subjects with high dietary intake of long chain omega 3 PUFA or fish, although some of these associations did not reach statistical significance.

Figure 3. Association of the risk for AMD with dietary long chain omega 3 or fish (cross-sectional and case-control studies). References of the cited studies: EDCC(43); Twin Study(45); AREDS(44); Beaver Dam(46); NHANES(47); POLA(48); EUREYE(49)

 

Figure 4. Association of the risk for AMD with dietary long chain omega 3 or fish (prospective studies). References of the cited studies: Blue Mountains(51); Nurses and Physicians’Health Study(50); Reykjavik Study(52); Seddon et al.(53); AREDS(54)

 

Most of these studies were grouped in a meta-analysis performed in 2008, which concluded at a 38% reduction in risk for AMD in subjects with high dietary long-chain omega 3 PUFA(55). By contrast, two recent studies found an increased risk for ARM in subjects with high omega 3 PUFA(56-57). However, in these studies, only total omega 3 PUFA intake was studied, including ALA and long-chain omega 3 PUFA, while most studies found a reduced risk for ARM only with high long chain omega 3 PUFA, in accordance with the scientific rationale. As stated in one of these studies(57), the main source for ALA is vegetable oil, which is also the main source of omega 6 PUFA, and was found to increase the risk for ARM in some studies(43,53). Future studies need to separate the precursor from the long chain derivatives.
Globally, available epidemiological studies strongly suggest a reduced risk for AMD in subjects with high consumption of long chain omega 3 fatty acids and fish. 
In addition, some studies have suggested that the risk for AMD may be decreased in subjects with high intake of vitamins B(56-63). Consistently, supplementation with vitamins B reduced the incidence of AMD in a randomized interventional study(64). A role for vitamin D in the aetiology of AMD has also been suggested(65)
 

4. Light exposure and cataract extraction

Sunlight exposure has been investigated as a potential risk factor for AMD. Light exposure may have deleterious effect on the eye, in particular through the production of reactive oxygen species(66). Only blue light reaches the retina, since ultraviolet radiations are absorbed by the cornea and the lens. Intense blue light exposure has been shown to induce retinal damage(66), and the macular pigment, which absorbs blue light, is thought to protect the macula against photo-toxic damage(30). However, results have been inconsistent in the few studies that have investigated the associations of the light exposure with AMD in humans(27,67-74). Globally, the available studies suggest that the effect of sunlight exposure in the aetiology of AMD is at most modest. Interestingly, a recent study evidenced an association of the risk for AMD with blue light exposure, only in those subjects with low plasma antioxidants and zeaxanthin(27). This suggests that light exposure may increase the risk for AMD only when defences against the produced reactive oxygen species are not appropriate. Sunlight exposure therefore does not appear to be a major determinant of AMD, but may be a risk factor in susceptible individuals. Appropriate nutritional intake in antioxidants and macular pigment may be particularly important in subjects highly exposed to light. These data will need to be confirmed in future studies.
Besides, cataract surgery was associated with a major increase in AMD incidence in a few studies(7,75-81), although not all(82-83). For instance, in a pooled analysis of two major population-based studies (Beaver Dam and Blue Mountains), eyes which had undergone lens extraction had a 5.7-fold increased risk of developing late AMD(79). The reasons for increased risk of AMD in aphakic and pseudophakic eyes are unknown, but may include increased light exposure. Indeed, the lens naturally absorbs ultraviolet light, and, with the lens yellowing observed with ageing and cataract, also part of blue light. In this context, use of blue light filters in the implanted artificial lenses have been proposed(84), and are currently widely used, although their potential effect on the reduction of incidence of AMD has not been evaluated. 
Because cataract surgery is the most frequent surgical procedure in most industrialized countries, the potential increased risk of AMD in operated eye needs further study, in order to better characterize it, to determine its causes and to identify strategies to limit this potentially deleterious effect.

In conclusion, AMD is emerging as a disease resulting from major genetic susceptibility, the effect of which is modulated by lifestyle. Among lifestyle factors, smoking is the best characterized risk factor, now considered as causal, while the role of nutrition is increasingly identified. The respective roles of antioxidants, macular pigment and omega 3 fatty acids, together with other potential nutritional factors such as vitamins B and D, will be better understood in the future. In this field, several new epidemiological studies are being conducted, among which, in France, the Alienor Study(85). This population-based cohort study aims at assessing the association of AMD with nutritional, vascular and genetic risk factors. It bears on almost 1000 subjects, recruited from an existing cohort study on brain ageing (the 3C Study). The main nutritional factors studied are lutein and zeaxanthin, antioxidants (vitamins C, E, zinc) and omega 3 fatty acids, and are measured both in the diet, in plasma and, for lutein and zeaxanthin, on the retina. This study will add to the existing literature in this field, which is still relatively scarce and partially inconsistent. Moreover, several large controlled interventional trials, including the AREDS2 Study, will help demonstrating their causal role in the aetiology of AMD. Finally, the role of light exposure (in particular blue light) does not seem to be major determinant in this disease, but may be important in subgroups of the population (subjects with low antioxidant and macular pigment intake, subjects undergoing cataract surgery).

 

>> References

References - Modifiable risk factors for AMD

References - Modifiable risk factors for AMD

English

1. Jager RD, Mieler WF, Miller JW. Age-related macular degeneration. N Engl J Med 2008; 358 (24): 2606-2617.

2. Thornton J, Edwards R, Mitchell P, Harrison RA, Buchan I, Kelly SP. Smoking and age-related macular degeneration: a review of association. Eye 2005; 19 (9): 935-944.

3. Chang MA, Bressler SB, Munoz B, West SK. Racial differences and other risk factors for incidence and progression of age-related macular degeneration: Salisbury Eye Evaluation (SEE) Project. Invest Ophthalmol Vis Sci 2008; 49 (6): 2395-2402.

4. Fraser-Bell S, Wu J, Klein R, Azen SP, Varma R. Smoking, alcohol intake, estrogen use, and age-related macular degeneration in Latinos: the Los Angeles Latino Eye Study. Am J Ophthalmol 2006; 141 (1): 79-87.

5. Kawasaki R, Wang JJ, Ji GJ, Taylor B, Oizumi T, Daimon M, Kato T, Kawata S, Kayama T, Tano Y, Mitchell P, Yamashita H, Wong TY. Prevalence and risk factors for age-related macular degeneration in an adult Japanese population: the Funagata study. Ophthalmology 2008; 115 (8): 1376-81, 1381.

6. Cackett P, Wong TY, Aung T, Saw SM, Tay WT, Rochtchina E, Mitchell P, Wang JJ. Smoking, cardiovascular risk factors, and age-related macular degeneration in Asians: the Singapore Malay Eye Study. Am J Ophthalmol 2008; 146 (6): 960-967.

7. Krishnaiah S, Das T, Nirmalan PK, Nutheti R, Shamanna BR, Rao GN, Thomas R. Risk factors for age-related macular degeneration: findings from the Andhra Pradesh eye disease study in South India. Invest Ophthalmol Vis Sci 2005; 46 (12): 4442-4449.

8. Delcourt C, Diaz JL, Ponton-Sanchez A, Papoz L. Smoking and age-related macular degeneration. The POLA Study. Pathologies Oculaires Liées a l’Age. Arch Ophthalmol 1998; 116 (8): 1031-1035.

9. Khan JC, Thurlby DA, Shahid H, Clayton DG, Yates JR, Bradley M, Moore AT, Bird AC. Smoking and age related macular degeneration: the number of pack years of cigarette smoking is a major determinant of risk for both geographic atrophy and choroidal neovascularisation. Br J Ophthalmol 2006; 90 (1): 75-80.

10. Vingerling JR, Hofman A, Grobbee DE, de Jong PT. Age-related macular degeneration and smoking. The Rotterdam Study. Arch Ophthalmol 1996; 114 (10): 1193-1196.

11. Seddon JM, Willett WC, Speizer FE, Hankinson SE. A prospective study of cigarette smoking and age-related macular degeneration in women. JAMA 1996; 276 (14): 1141-1146.

12. Klein R, Klein BE, Moss SE. Relation of smoking to the incidence of age-related maculopathy. The Beaver Dam Eye Study. Am J Epidemiol 1998; 147 (2): 103-110.

13. Chakravarthy U, Augood C, Bentham GC, de Jong PT, Rahu M, Seland J, Soubrane G, Tomazzoli L, Topouzis F, Vingerling JR, Vioque J, Young IS, Fletcher AE. Cigarette smoking and age-related macular degeneration in the EUREYE Study. Ophthalmology 2007; 114 (6): 1157-1163.

14. Smith W, Mitchell P, Leeder SR. Smoking and age-related maculopathy. The Blue Mountains Eye Study. Arch Ophthalmol 1996; 114 (12): 1518-1523.

15. Tan JS, Mitchell P, Kifley A, Flood V, Smith W, Wang JJ. Smoking and the long-term incidence of age-related macular degeneration: the Blue Mountains Eye Study. Arch Ophthalmol 2007; 125 (8): 1089-1095.

16. Tomany SC, Wang JJ, van Leeuwen R, Klein R, Mitchell P, Vingerling JR, Klein BE, Smith W, de Jong PT. Risk factors for incident age-related macular degeneration: pooled findings from 3 continents. Ophthalmology 2004; 111 (7): 1280-1287.

17. Mitchell P, Wang JJ, Smith W, Leeder SR. Smoking and the 5-year incidence of age-related maculopathy: the Blue Mountains Eye Study. Arch Ophthalmol 2002; 120 (10): 1357-1363.

18. Klein R, Klein BE, Tomany SC, Moss SE. Ten-year incidence of age-related maculopathy and smoking and drinking: the Beaver Dam Eye Study. Am J Epidemiol 2002; 156 (7): 589-598.

19. Francis PJ, George S, Schultz DW, Rosner B, Hamon S, Ott J, Weleber RG, Klein ML, Seddon JM. The LOC387715 gene, smoking, body mass index, environmental associations with advanced age-related macular degeneration. Hum Hered 2007; 63 (3-4): 212-218.

20. Seddon JM, George S, Rosner B, Klein ML. CFH gene variant, Y402H, and smoking, body mass index, environmental associations with advanced age-related macular degeneration. Hum Hered 2006; 61 (3): 157-165.

21. Schmidt S, Hauser MA, Scott WK, Postel EA, Agarwal A, Gallins P, Wong F, Chen YS, Spencer K, Schnetz-Boutaud N, Haines JL, Pericak-Vance MA. Cigarette smoking strongly modifies the association of LOC387715 and age-related macular degeneration. Am J Hum Genet 2006; 78 (5): 852-864.

22. Klein R. Overview of progress in the epidemiology of age-related macular degeneration. Ophthalmic Epidemiol 2007; 14 (4): 184-187.

23. Beatty S, Koh H, Phil M, Henson D, Boulton M. The role of oxidative stress in the pathogenesis of age-related macular degeneration. Surv Ophthalmol 2000; 45 (2): 115-134.

24. Chiu CJ, Taylor A. Nutritional antioxidants and age-related cataract and maculopathy. Exp Eye Res 2007; 84 (2): 229-245.

25. A randomized, placebo-controlled, clinical trial of high-dose supplementation with vitamins C and E, beta carotene, and zinc for age-related macular degeneration and vision loss: AREDS report no. 8. Arch Ophthalmol 2001; 119 (10): 1417-1436. Erratum in: 2008; 126 (9): 1251.

26. Birlouez-Aragon I, Delcourt C, Tessier F, Papoz L. Associations of age, smoking habits and diabetes with plasma vitamin C of elderly of the POLA study. Int J Vitam Nutr Res 2001; 71 (1): 53-59.

27. Fletcher AE, Bentham GC, Agnew M, Young IS, Augood C, Chakravarthy U, de Jong PT, Rahu M, Seland J, Soubrane G, Tomazzoli L, Topouzis F, Vingerling JR, Vioque J. Sunlight exposure, antioxidants, and age-related macular degeneration. Arch Ophthalmol 2008; 126 (10): 1396-1403.

28. Delcourt C, Cristol JP, Tessier F, Leger CL, Descomps B, Papoz L. Age-related macular degeneration and antioxidant status in the POLA study. POLA Study Group. Pathologies Oculaires Liees a l’Age. Arch Ophthalmol 1999; 117 (10): 1384-1390.

29. van Leeuwen R, Boekhoorn S, Vingerling JR, Witteman JC, Klaver CC, Hofman A, de Jong PT. Dietary intake of antioxidants and risk of age-related macular degeneration. JAMA 2005; 294 (24): 3101-3107.

30. Whitehead AJ, Mares JA, Danis RP. Macular pigment: a review of current knowledge. Arch Ophthalmol 2006; 124 (7): 1038-1045. AMD.

31. Mares-Perlman JA, Brady WE, Klein R, Klein BE, Bowen P, Stacewicz-Sapuntzakis M, Palta M. Serum antioxidants and age-related macular degeneration in a population-based case-control study. Arch Ophthalmol 1995; 113 (12): 1518-1523.

32. Antioxidant status and neovascular age-related macular degeneration. Eye Disease Case-Control Study Group. Arch Ophthalmol 1993; 111 (1): 104-109. Erratum in: Arch Ophthalmol 1993; 111 (11): 1499. Arch Ophthalmol 1993; 111 (9): 1228. Arch Ophthalmol 1993; 111 (10): 1366.

33. Mares-Perlman JA, Fisher AI, Klein R, Palta M, Block G, Millen AE, Wright JD. Lutein and zeaxanthin in the diet and serum and their relation to age-related maculopathy in the third national health and nutrition examination survey. Am J Epidemiol 2001; 153 (5): 424-432.

34. Gale CR, Hall NF, Phillips DI, Martyn CN. Lutein and zeaxanthin status and risk of age-related macular degeneration. Invest Ophthalmol Vis Sci 2003; 44 (6): 2461-2465.

35. Delcourt C, Carriere I, Delage M, Barberger-Gateau P, Schalch W. Plasma lutein and zeaxanthin and other carotenoids as modifiable risk factors for age-related maculopathy and cataract: the POLA Study. Invest Ophthalmol Vis Sci 2006; 47 (6): 2329-2335.

36. VandenLangenberg GM, Mares-Perlman JA, Klein R, Klein BE, Brady WE, Palta M. Associations between antioxidant and zinc intake and the 5-year incidence of early age-related maculopathy in the Beaver Dam Eye Study. Am J Epidemiol 1998; 148 (2): 204-214.

37. Cho E, Hankinson SE, Rosner B, Willett WC, Colditz GA. Prospective study of lutein/zeaxanthin intake and risk of age-related macular degeneration. Am J Clin Nutr 2008; 87 (6): 1837-1843.

38. Tan JS, Wang JJ, Flood V, Rochtchina E, Smith W, Mitchell P. Dietary antioxidants and the long-term incidence of age-related macular degeneration: the Blue Mountains Eye Study. Ophthalmology 2008; 115 (2): 334-341.

39. Wang L, Gaziano JM, Norkus EP, Buring JE, Sesso HD. Associations of plasma carotenoids with risk factors and biomarkers related to cardiovascular disease in middle-aged and older women. Am J Clin Nutr 2008; 88 (3): 747-754.

40. Gruber M, Chappell R, Millen A, LaRowe T, Moeller SM, Iannaccone A, Kritchevsky SB, Mares J. Correlates of serum lutein + zeaxanthin: findings from the Third National Health and Nutrition Examination Survey. J Nutr 2004; 134 (9): 2387-2394.

41. Arterburn LM, Hall EB, Oken H. Distribution, interconversion, and dose response of n-3 fatty acids in humans. Am J Clin Nutr 2006; 83 (6 Suppl): 1467S-1476S.

42. SanGiovanni JP, Chew EY. The role of omega-3 long-chain polyunsaturated fatty acids in health and disease of the retina. Prog Retin Eye Res 2005; 24 (1): 87-138.

43. Seddon JM, Rosner B, Sperduto RD, Yannuzzi L, Haller JA, Blair NP, Willett W. Dietary fat and risk for advanced age-related macular degeneration. Arch Ophthalmol 2001; 119 (8): 1191-1199.

44. SanGiovanni JP, Chew EY, Clemons TE, Davis MD, Ferris FL, III, Gensler GR, Kurinij N, Lindblad AS, Milton RC, Seddon JM, Sperduto RD. The relationship of dietary lipid intake and age-related macular degeneration in a case-control study: AREDS Report No. 20. Arch Ophthalmol 2007; 125 (5): 671-679.

45. Seddon JM, George S, Rosner B. Cigarette smoking, fish consumption, omega-3 fatty acid intake, and associations with age-related macular degeneration: the US Twin Study of Age-Related Macular Degeneration. Arch Ophthalmol 2006; 124 (7): 995-1001.

46. Mares-Perlman JA, Brady WE, Klein R, VandenLangenberg GM, Klein BE, Palta M. Dietary fat and age-related maculopathy. Arch Ophthalmol 1995; 113 (6): 743-748.

47. Heuberger RA, Mares-Perlman JA, Klein R, Klein BE, Millen AE, Palta M. Relationship of dietary fat to age-related maculopathy in the Third National Health and Nutrition Examination Survey. Arch Ophthalmol 2001; 119 (12): 1833-1838.

48. Delcourt C, Carriere I, Cristol JP, Lacroux A, Gerber M. Dietary fat and the risk of age-related maculopathy: the POLANUT study. Eur J Clin Nutr 2007; 61 (11): 1341-1344.

49. Augood C, Chakravarthy U, Young I, Vioque J, de Jong PT, Bentham G, Rahu M, Seland J, Soubrane G, Tomazzoli L, Topouzis F, Vingerling JR, Fletcher AE. Oily fish consumption, dietary docosahexaenoic acid and eicosapentaenoic acid intakes, and associations with neovascular age-related macular degeneration. Am J Clin Nutr 2008; 88 (2): 398-406.

50. Cho E, Hung S, Willett WC, Spiegelman D, Rimm EB, Seddon JM, Colditz GA, Hankinson SE. Prospective study of dietary fat and the risk of age-related macular degeneration. Am J Clin Nutr 2001; 73 (2): 209-218.

51. Chua B, Flood V, Rochtchina E, Wang JJ, Smith W, Mitchell P. Dietary fatty acids and the 5-year incidence of age-related maculopathy. Arch Ophthalmol 2006; 124 (7): 981-986.

52. Arnarsson A, Sverrisson T, Stefansson E, Sigurdsson H, Sasaki H, Sasaki K, Jonasson F. Risk factors for five-year incident age-related macular degeneration: the Reykjavik Eye Study. Am J Ophthalmol 2006; 142 (3): 419-428.

53. Seddon JM, Cote J, Rosner B. Progression of age-related macular degeneration: association with dietary fat, transunsaturated fat, nuts, and fish intake. Arch Ophthalmol 2003; 121 (12): 1728-1737. Erratum in: Arch Ophthalmol 2004; 122 (3): 426.

54. Sangiovanni JP, Agrón E, Meleth AD, Reed GF, Sperduto RD, Clemons TE, Chew EY; Age-Related Eye Disease Study Research Group. {omega}-3 Long-chain polyunsaturated fatty acid intake and 12-y incidence of neovascular age-related macular degeneration and central geographic atrophy: AREDS report 30, a prospective cohort study from the Age-Related Eye Disease Study. Am J Clin Nutr 2009; 90 (6): 1601-1607.

55. Chong EW, Kreis AJ, Wong TY, Simpson JA, Guymer RH. Dietary omega-3 fatty acid and fish intake in the primary prevention of age-related macular degeneration: a systematic review and meta-analysis. Arch Ophthalmol 2008; 126 (6): 826-833.

56. Robman L, Vu H, Hodge A, Tikellis G, Dimitrov P, McCarty C, Guymer R. Dietary lutein, zeaxanthin, and fats and the progression of age-related macular degeneration. Can J Ophthalmol 2007; 42 (5): 720-726.

57. Parekh N, Voland RP, Moeller SM, Blodi BA, Ritenbaugh C, Chappell RJ, Wallace RB, Mares JA. Association between dietary fat intake and age-related macular degeneration in the Carotenoids in Age-Related Eye Disease Study (CAREDS): an ancillary study of the Women’s Health Initiative. Arch Ophthalmol 2009; 127 (11): 1483-1493.

58. Rochtchina E, Wang JJ, Flood VM, Mitchell P. Elevated serum homocysteine, low serum vitamin B12, folate, and age-related macular degeneration: the Blue Mountains Eye Study. Am J Ophthalmol 2007; 143 (2): 344-346.

59. Kamburoglu G, Gumus K, Kadayifcilar S, Eldem B. Plasma homocysteine, vitamin B12 and folate levels in age-related macular degeneration. Graefes Arch Clin Exp Ophthalmol 2006; 244 (5): 565-569.

60. Nowak M, Swietochowska E, Wielkoszynski T, Marek B, Kos-Kudla B, Szapska B, Kajdaniuk D, Glogowska-Szelag J, Sieminska L, Ostrowska Z, Koziol H, Klimek J. Homocysteine, vitamin B12, and folic acid in age-related macular degeneration. Eur J Ophthalmol 2005; 15 (6): 764-767.

61. Axer-Siegel R, Bourla D, Ehrlich R, Dotan G, Benjamini Y, Gavendo S, Weinberger D, Sela BA. Association of neovascular age-related macular degeneration and hyperhomocysteinemia. Am J Ophthalmol 2004; 137 (1): 84-89.

62. Vine AK, Stader J, Branham K, Musch DC, Swaroop A. Biomarkers of cardiovascular disease as risk factors for age-related macular degeneration. Ophthalmology 2005; 112 (12): 2076-2080. 63. Seddon JM, Gensler G, Klein ML, Milton RC. Evaluation of plasma homocysteine and risk of age-related macular degeneration. Am J Ophthalmol 2006; 141 (1): 201-203.

64. Christen WG, Glynn RJ, Chew EY, Albert CM, Manson JE. Folic acid, pyridoxine, and cyanocobalamin combination treatment and age-related macular degeneration in women: the Women’s Antioxidant and Folic Acid Cardiovascular Study. Arch Intern Med 2009; 169 (4): 335-341.

65. Parekh N, Chappell RJ, Millen AE, Albert DM, Mares JA. Association between vitamin D and age-related macular degeneration in the Third National Health and Nutrition Examination Survey, 1988 through 1994. Arch Ophthalmol 2007; 125 (5): 661-669.

66. Algvere PV, Marshall J, Seregard S. Age-related maculopathy and the impact of blue light hazard. Acta Ophthalmol Scand 2006;

84 (1): 4-15. 67. Taylor HR, West S, Muñoz B, Rosenthal FS, Bressler SB, Bressler NM. The long-term effects of visible light on the eye. Arch Ophthalmol 1992; 110 (1): 99-104. 68. Cruickshanks KJ, Klein R, Klein BE. Sunlight and age-related macular degeneration. The Beaver Dam Eye Study. Arch Ophthalmol 1993; 111 (4): 514-518.

69. Tomany SC, Cruickshanks KJ, Klein R, Klein BE, Knudtson MD. Sunlight and the 10-year incidence of age-related maculopathy: the Beaver Dam Eye Study. Arch Ophthalmol 2004; 122 (5): 750-757.

70. [No authors listed]. Risk factors for neovascular age-related macular degeneration. The Eye Disease Case-Control Study Group. Arch Ophthalmol 1992; 110 (12): 1701-1708.

71. McCarty CA, Mukesh BN, Fu CL, Mitchell P, Wang JJ, Taylor HR72. Risk factors for age-related maculopathy: the Visual Impairment Project. Arch Ophthalmol 2001; 119 (10): 1455-1462.

72. Delcourt C, Carriere I, Ponton-Sanchez A, Fourrey S, Lacroux A, Papoz L. Light exposure and the risk of age-related macular degeneration: the Pathologies Oculaires Liées a l’Age (POLA) study. Arch Ophthalmol 2001; 119 (10): 1463-1468.

73. Darzins P, Mitchell P, Heller RF. Sun exposure and age-related macular degeneration. An Australian case-control study. Ophthalmology 1997; 104 (5): 770-776.

74. Pham TQ, Rochtchina E, Mitchell P, Smith W, Wang JJ. Sunlight-related factors and the 10-year incidence of age-related maculopathy. Ophthalmic Epidemiol 2009; 16 (2): 136-141.

75. Pollack A, Marcovich A, Bukelman A, Oliver M. Age-related macular degeneration after extracapsular cataract extraction with intraocular lens implantation. Ophthalmology 1996; 103 (10): 1546-1554.

76. Ho L, Boekhoorn SS, Liana, van Duijn CM, Uitterlinden AG, Hofman A, de Jong PT, Stijnen T, Vingerling JR. Cataract surgery and the risk of aging macula disorder: the rotterdam study. Invest Ophthalmol Vis Sci 2008; 49 (11): 4795-800.

77. Kaiserman I, Kaiserman N, Elhayany A, Vinker S. Cataract surgery is associated with a higher rate of photodynamic therapy for age-related macular degeneration. Ophthalmology. 2007; 114 (2): 278-282.

78. Cugati S, Mitchell P, Rochtchina E, Tan AG, Smith W, Wang JJ. Cataract surgery and the 10-year incidence of age-related maculopathy: the Blue Mountains Eye Study. Ophthalmology 2006; 113 (11): 2020-2025.

79. Wang JJ, Klein R, Smith W, Klein BE, Tomany S, Mitchell P. Cataract surgery and the 5-year incidence of late-stage age-related maculopathy: pooled findings from the Beaver Dam and Blue Mountains eye studies. Ophthalmology 2003; 110 (10): 1960-1967.

80. Klein R, Klein BE, Wong TY, Tomany SC, Cruickshanks KJ. The association of cataract and cataract surgery with the long-term incidence of age-related maculopathy: the Beaver Dam eye study. Arch Ophthalmol 2002; 120 (11): 1551-1558.

81. Pham TQ, Cugati S, Rochtchina E, Mitchell P, Maloof A, Wang JJ. Early age-related maculopathy in eyes after cataract surgery. Eye 2007; 21 (4): 512-517.

82. Sutter FK, Menghini M, Barthelmes D, Fleischhauer JC, Kurz-Levin MM, Bosch MM, Helbig H. Is pseudophakia a risk factor for neovascular age-related macular degeneration? Invest Ophthalmol Vis Sci 2007;48 (4): 1472-1475.

83. Chew EY, Sperduto RD, Milton RC, Clemons TE, Gensler GR, Bressler SB, Klein R, Klein BE, Ferris FL, III. Risk of advanced age-related macular degeneration after cataract surgery in the Age-Related Eye Disease Study: AREDS report 25. Ophthalmology 2009; 116 (2): 297-303.

84. Braunstein RE, Sparrow JR. A blue-blocking intraocular lens should be used in cataract surgery. Arch Ophthalmol 2005; 123(4): 547-549.

85. Delcourt C, Korobelnik JF, Barberger Gateau P, et al. Nutrition and age-related eye diseases : the ALIENOR (Antioxydants, Lipides Essentiels, Nutrition et maladies Oculaires) Study. Journal Nutr Health Aging. 2010; (in press).

Pathogenic Mechanisms

Pathogenic Mechanisms

English

Authors:

Ângela Carneiro, MD, PhD
Faculty of Medicine of University of Porto, Hospital S. João, Porto, Portugal.

Luis Mendonça, MD

 

Updated/reviewed by the authors, July 2017.

The outer retina

Age-related changes that predispose to age-related macular degeneration (AMD) occur in the outer retina, more specifically the region that includes the photoreceptors, the retinal pigment epithelium (RPE), Bruch

Retinal anatomy is highly organized and vascular and avascular compartments are strictly segregated in the retina(1). The blood-retinal barriers, inner and outer, are fundamental for the integrity of structure and optimization of function in neuro-sensorial retina(2).

The outer blood-retinal barrier is formed, among its various components, by the RPE tight junctions. The intercellular cohesiveness of the RPE is not easily disrupted. Tight junctions appear as a necklace of strands that encircle each cell, binding each cell to its neighbors in a monolayer that separate the outer layer of the neural retina from the choriocapillaris(3). Choriocapillaris is a great vascular network of fenestrated capillaries with high blood flow, fundamental for the metabolism of outer retina. This outer blood-retinal barrier retards transepithelial diffusion through the paracellular spaces(3).

The RPE is a polarized epithelium that consist of a continuous pavement-like monolayer of cuboidal shaped cells that in macular area are tall, narrow and highly uniform in size and shape(4). Interdigitation of the apical processes of the RPE with the cone and rod outer segments provides only a tenuous adhesion of the RPE to the sensory retina(5).

RPE cells have at least ten known functions, but regeneration of visual pigments, transport of fluids and ions between photoreceptors and choriocapillaris, formation and maintenance of the interphotoreceptor matrix and Bruch’s membrane, phagocytosis of outer segments of photoreceptors, and supplying trophic factors such as VEGF-A should be emphasized(6,7).

Bruch’s membrane is a thin, acellular and well-delineated membrane with five layers. From internal to external these layers are: the basement membrane of the RPE, the inner collagenous zone, the elastic tissue layer, the outer collagenous zone and the basement membrane of the choriocapillaris. It is composed of elements from both, the retina and choroid, but is an integral part of choroid(8). Its inner surface is smooth, whereas its outer surface is composed of a series of collagenous protrusions that extend externally to form the pillars separating and supporting the choriocapillaris(5). Due to its specific location and properties, this tissue is thought to be a vital limiting layer for metabolic transport between the RPE cells and the choriocapillaris(9).

The choriocapillaris consists of a continuous layer of fenestrated endothelial cells surrounded by a basement membrane. In the macula the choriocapillaris is arranged in a lobular pattern of highly concentrated interconnecting capillaries supplied by a central arteriole and drained by circumferential venules(5,10). The fenestrations, 60-80 nm in diameter, are abundant and seem to play an important role in permitting the passage of glucose and vitamin A to the RPE and retina. The choriocapillaris supplies oxygen and nutrients to Bruch´s membrane and the outer third of the retina, except in the macula, where it supplies the entire retina(8). The peculiar structure of the choroidal vascular tree in the macula provides this area with the highest rate of blood flow of any tissue in the body(5)

 

Aging changes in outer retina and early AMD

With aging, the neurosensory retina was shown to develop thickening of the internal limiting membrane, diminution of neural elements with age-related loss of rods before cones, gliosis in the peripheral retina, and diminution of capillaries around the fovea, while the lumina of the choriocapillaris and the choroidal thickness become reduced by half(11-13).

With the advancement of age, both the thickness and complexity of Bruch’s membrane increase primarily due to extracellular matrix remodeling and accumulation of inclusions in this region(11). Bruch’s membrane calcifies and doubles in thickness between the ages of 10 and 90 years(11). There is a linear thickening due to deposits of collagen, lipids and debris. After the 30’s its lipid concentration increases during life and consequently the fluid permeability and nutrient transport across the membrane decreases(14). In normal conditions Bruch’s membrane acts as an intercellular matrix regulating survival of adjacent RPE and choriocapillaris cells. Its diminished function results in apoptosis of these cells from incorrect cell adhesion(14). On the other hand extracellular deposits around Bruch’s membrane instigate chronic inflammation, invasion by dendritic cells and release of inflammatory cytokines, angiogenic factors and immune complexes(15,16).

The RPE is a monolayer of regularly arranged hexagonal cells that spans the retina from the margins of the optic disc anteriorly to the ora serrata. The number of RPE cells diminishes with age. Macular RPE cells become wither, flatter and increase in height with advancing age(4,17). In each RPE cell there is a progressive accumulation of lipofuscin during life and in people over 80 years of age, the debris can occupy more than one fifth of the total volume of an RPE cell(18,19).  RPE cells have a brown color in young eyes but with age, they become increasingly more golden colored, owing to the accumulation of lipofuscin pigment granules(20). Lipofuscin in the RPE is the source of fundus autofluorescence. The major component of lipofuscin is N-retinylidene-N-retinylethanol-amine (A2E), a retinoid product of the visual cycle(21). The A2E produced interferes with the function of RPE cells, leading to its apoptosis and subsequent geographic atrophy(22). Age-related changes also include a decrease in the number of melanin granules, loss of basal digitations and irregularity in shape. The RPE cells become separated from their basal membrane by membranous debris and abnormal secretory products and subsequently occurs deposition of collagen and fibronectin and latter formation of basal laminar deposits(23).

Basal laminar deposits are composed of basement membrane protein and long-spacing collagen located between the RPE plasma and basement membranes(24). Basal laminar deposits are considered the precursors of AMD and can appear around the age of 40 years(25).

Basal linear deposits consist of granular, vesicular or membranous lipid-rich material located external to the basement membrane of the RPE, in the inner collagenous layer of Bruch’s membrane, and represent a specific marker of AMD(26).

These two types of deposits can only be shown on pathological specimens and not by clinical evaluation(22).

The combination of the deposits with secondary changes in the RPE results in the formation of drusen. Drusen are localized deposits of extracellular material lying between the basement membrane of the RPE and the inner collagen layer of Bruch’s membrane(23,27). Drusen often have a core of glycoproteins but they also contain fragments of RPE cells, crystallins, apolipoproteins B and E, and proteins related to inflammation such as amyloid P and β, C5 and C5b-9 complement complex(28-31)

Drusen change in size, shape, color, distribution and consistency with the passing years(32).

Small drusen are defined as being less than 63 μm in diameter(33). The presence of small, hard drusen alone is not sufficient to diagnose early AMD. These deposits are ubiquitous and the new development of small drusen in an adult eye without prior evidence of hard drusen is not age dependent(34).

Hard drusen are discrete nodules or deposits composed of hyaline-like material. During fluorescein angiography hard drusen behave as pin-point window defects(35).

Soft drusen are larger and associated with pigment epithelium detachment and diffuse abnormal Bruch’s membrane alterations(36,37). Soft drusen have a tendency to cluster and merge with one another demonstrating confluence(35). During fluorescein angiography soft drusen hyperfluoresce early and either fade or stain in the late phase(5).

Drusen can be visible in ophthalmoscopy when their diameter exceeds 25 μm as dots ranging in color from white to yellow(6). However soft drusen are clinically identified whenever there is sufficient RPE hypopigmentation or atrophy overlying diffuse Bruch’s membrane thickening, or, when there are focal detachments within this material. These findings suggest that the clinical identification of soft drusen identifies an eye with diffuse changes at the RPE-Bruch’s membrane complex(35). When they become larger (>125 μm), and greater the area that they cover, the risk of late AMD becomes higher(38).

The RPE degeneration and nongeographic atrophy of the RPE are characterized by pigment mottling and stippled hypopigmentation with thinning of the neurosensory retina(39). Histopathology shows mottled areas of RPE hypopigmentation or atrophy overlying diffuse basal linear and basal laminar deposits(36). Incidence and prevalence rates of RPE depigmentation are age dependent(34).

Focal hyperpigmentation of the RPE, clinically evident as pigment clumping at the level of the outer retina or sub-retinal space, increases the risk of progression to the late phases of the disease(34,36,40).

Aging is known to be associated with increased oxidative damage and the retina is a fertile environment for reactive oxygen species. Apart from the presence of two retinal blood supplies that generate an highly oxygenated environment, the exposure to high levels of cumulative irradiation, high levels of photosensitizers, large amounts of polyunsaturated fatty acids, readily oxidizable lipid, protein and carbohydrate substrates, and the huge proteolytic burden in the RPE contribute to this particular predisposition to oxidative stress(41,42). This state of accumulation of toxic elements is said to ultimately tip the balance of the ocular immune privilege towards immune activation and inflammation. The end result of an innappropriate activation of diverse immune pathways, including classical and alternative complement pathways, is an immune-mediated retinal damage and/or an impaired immune-mediated retinal maintenance with RPE damage(13). 

With all of the above in mind, Zarbin(43) has summarized his review on AMD pathogenesis in five sequential steps as follows: (1) AMD involves aging changes plus additional pathological changes (ie, AMD is not just an aging change); (2) in aging and AMD, oxidative stress causes RPE and, possibly, choriocapillaris injury; (3) in AMD (and perhaps in aging), RPE and, possibly, choriocapillaris injury results in a chronic inflammatory response within the Bruch membrane and the choroid; (4) in AMD, RPE and, possibly, choriocapillaris injury and inflammation lead to formation of an abnormal extracellular matrix, which causes altered diffusion of nutrients to the retina and RPE, possibly precipitating further RPE and retinal damage; and (5) the abnormal extracellular matrix results in altered RPE-choriocapillaris behavior leading ultimately to atrophy of the retina, RPE, and choriocapillaris and/or choroidal new vessel growth. In this sequence of events, patient's susceptibility to AMD would be determined by both the environment and his genetic profile(43).

 

Late AMD

The primary clinical characteristic of late dry AMD is the appearance of geographic atrophy of RPE. On microscopy, geographic atrophy is seen as abnormal RPE cells with hypotrophy, atrophy, hypertrophy, hypopigmentation, hyperpigmentation, migration, loss of photoreceptors, attenuation of Bruch’s membrane and choriocapillaris degeneration(44,45). Geographic atrophy is clinically characterized by roughly oval areas of hypopigmentation that allows for increased visualization of the underlying choroidal vessels and is the consequence of RPE cell loss. Loss of RPE cells leads to gradual degeneration of photoreceptors and thinning of the retina that may extend to the outer plexiform and inner nuclear layers(6, 29). Compensatory RPE cell proliferation leads to hyperpigmentary changes frequently observed at the periphery of the hypopigmented areas(6). The atrophy of RPE is usually more severe than the loss of choriocapillaris but the choriocapillaris seem to be highly constricted in areas of complete RPE cell loss(44).

In neovascular AMD early choroidal neovascularization occurs under the RPE(46)and eventually breaks through(47), leading to accumulation of lipid-rich fluid under the RPE or neuroretina. In haemorrhagic forms blood breaks through the RPE into the subretinal space and sometimes through the retina and into the vitreous(35).

The pattern of growth of CNV often simulates that of a sea fan with radial arterioles and venules supplying and draining a circumferential dilated capillary sinus(5).

As neovascularization of the sub-RPE space occurs, initially the blood flow through the neovascular network is sluggish and there is little or no exudation. This is a period of occult neovascularization and the overlying RPE and neuroretina may be minimally affected(5). With an increase of blood flow through the network the endothelium decompensates and exudation extends into the subpigment epithelial space creating in some cases RPE detachments. The exudation may also extend through the RPE and detach the overlying retina.

In type II CNV the new vessels extend from the choroid through defects in Bruch´s membrane enters the space between the photoreceptors and RPE cells and growth laterally in the subretinal space(5). This is usually accompanied by varying amounts of subretinal exudates and/or blood.

Macrophages have been documented both morphologically and functionally in neovascular AMD(48,49). Activated macrophages and microglia may secret cytokines and chemokines that promote cellular damage and angiogenesis(50).

Involution of CNV eventually occurs and is associated with varying degrees of subretinal scar tissue, reactive hyperplasia of the RPE and/or atrophy, and can partially or totally replace the neuroretina(21). The outer nuclear layer can be severely attenuated with a reduction of photoreceptor length of almost 70%(51). Often anastomosis between the retinal circulation and the underlying choroidal circulation develops within these old disciforme scars(32,52).

Other factors like complement factor H, that downregulates the alternative complement pathway(53), HtrA1 – a secretory protein and an inhibitor of transforming  growth factor β (TGF-β)(54)- and ARMS2(55) play a role in development of AMD. However its specific role and relevance in development and progression to neovascular and atrophic forms of AMD are discussed in others chapters of this book.

 

Key points

  • Age-related changes that predispose to AMD occur in the outer retina, more specifically the region that includes the photoreceptors, the RPE, Bruch’s membrane and the choriocapillaris.
  • The deposition of insoluble material, the calcification and increase in thickness of Bruch’s membrane, and a less fenestrated and thinner choriocapillaris leads to photoreceptors/RPE hypoxia. The number of RPE cells reduces with age and in each cell there is a progressive accumulation of lipofuscin during life. Basal laminar and basal linear deposits, drusen, RPE degeneration and atrophy develop and finally, geographic atrophy of RPE and choroidal neovascularization occur.
  • Choroidal neovascularization is a relative self-limited disease with development, growth and involutional stages leading to the formation of a scar and destruction of macular structure and function.

 

>> References

References - Pathogenic Mechanisms

References - Pathogenic Mechanisms

English
  1. Gariano RF, Gardner TW. Retinal angiogenesis in development and disease. Nature 2005;438:960-6.
  2. Cunha-Vaz JG. The blood-retinal barriers system. Basic concepts and clinical evaluation. Exp Eye Res 2004;78:715-21.
  3. Rizzolo LJ. Development and role of tight junctions in the retinal pigment epithelium. Int Rev Cytol 2007;258:195-234.
  4. Friedman E, Ts'o MO. The retinal pigment epithelium. II. Histologic changes associated with age. Arch Ophthalmol 1968;79:315-20.
  5. Gass JDM. Stereoscopic atlas of macular diseases: diagnosis and treatment, 4th ed. St. Louis: Mosby, 1997.
  6. de Jong PT. Age-related macular degeneration. N Engl J Med 2006;355:1474-85.
  7. Saint-Geniez M, Maharaj AS, Walshe TE, et al. Endogenous VEGF is required for visual function: evidence for a survival role on müller cells and photoreceptors. PLoS ONE 2008;3:e3554.
  8. Jaeger EA, Anderson DR, Glaser JS, et al. Duane's ophthalmology on CD-ROM, 2006 ed. Philadelphia: Lippincott Williams & Wilkins, 2006:1 computer optical disc.
  9. Huang JD, Presley JB, Chimento MF, Curcio CA, Johnson M. Age-related changes in human macular Bruch's membrane as seen by quick-freeze/deep-etch. Exp Eye Res 2007;85:202-18.
  10. Hayreh SS. Segmental nature of the choroidal vasculature. Br J Ophthalmol 1975;59:631-48.
  11. Ramrattan RS, van der Schaft TL, Mooy CM, de Bruijn WC, Mulder PG, de Jong PT. Morphometric analysis of Bruch's membrane, the choriocapillaris, and the choroid in aging. Invest Ophthalmol Vis Sci 1994;35:2857-64.
  12. Grossniklaus HE, Nickerson JM, Edelhauser HF, Bergman LAMK, Berglin L. Anatomic alterations in aging and age-related diseases of the eye. Invest Ophthalmol Vis Sci 2013;54:ORSF23-7.
  13. Ambati J, Atkinson JP, Gelfand BD. Immunology of age-related macular degeneration. Nat Rev Immunol 2013;13(6): 438–451.
  14. Starita C, Hussain AA, Pagliarini S, Marshall J. Hydrodynamics of ageing Bruch's membrane: implications for macular disease. Exp Eye Res 1996;62:565-72.
  15. Gilmore AP. Anoikis. Cell Death Differ 2005;12(Suppl 2):1473-7.
  16. Penfold PL, Liew SC, Madigan MC, Provis JM. Modulation of major histocompatibility complex class II expression in retinas with age-related macular degeneration. Invest Ophthalmol Vis Sci 1997;38:2125-33.
  17. Guymer RH, Bird AC, Hageman GS. Cytoarchitecture of choroidal capillary endothelial cells. Invest Ophthalmol Vis Sci 2004;45:1660-6.
  18. Feeney-Burns L, Burns RP, Gao CL. Age-related macular changes in humans over 90 years old. Am J Ophthalmol 1990;109:265-78.
  19. Feeney-Burns L, Hilderbrand ES, Eldridge S. Aging human RPE: morphometric analysis of macular, equatorial, and peripheral cells. Invest Ophthalmol Vis Sci 1984;25:195-200.
  20. Feeney L. Lipofuscin and melanin of human retinal pigment epithelium. Fluorescence, enzyme cytochemical, and ultrastructural studies. Invest Ophthalmol Vis Sci 1978;17:583-600.
  21. Sparrow JR, Boulton M. RPE lipofuscin and its role in retinal pathobiology. Exp Eye Res 2005;80:595-606.
  22. Coleman HR, Chan CC, Ferris FL, 3rd, Chew EY. Age-related macular degeneration. Lancet 2008;372:1835-45.
  23. Tabandeh H, Dubovy S, Green WR. Bilateral midperipheral large drusen and retinal pigment epithelial detachments associated with multifocal areas of choroidal neovascularization: a histopathologic study. Retina 2006;26:1063-9.
  24. Sarks S, Cherepanoff S, Killingsworth M, Sarks J. Relationship of Basal laminar deposit and membranous debris to the clinical presentation of early age-related macular degeneration. Invest Ophthalmol Vis Sci 2007;48:968-77.
  25. Van der Schaft TL, Mooy CM, de Bruijn WC, Oron FG, Mulder PG, de Jong PT. Histologic features of the early stages of age-related macular degeneration. A statistical analysis. Ophthalmology 1992;99:278-86.
  26. Curcio CA, Millican CL. Basal linear deposit and large drusen are specific for early age-related maculopathy. Arch Ophthalmol 1999;117:329-39.
  27. Ding X, Patel M, Chan CC. Molecular pathology of age-related macular degeneration. Prog Retin Eye Res 2009;28:1-18.
  28. Hageman GS, Luthert PJ, Victor Chong NH, Johnson LV, Anderson DH, Mullins RF. An integrated hypothesis that considers drusen as biomarkers of immune-mediated processes at the RPE-Bruch's membrane interface in aging and age-related macular degeneration. Prog Retin Eye Res 2001;20:705-32.
  29. Crabb JW, Miyagi M, Gu X, et al. Drusen proteome analysis: an approach to the etiology of age-related macular degeneration. Proc Natl Acad Sci U S A 2002;99:14682-7.
  30. Umeda S, Suzuki MT, Okamoto H, et al. Molecular composition of drusen and possible involvement of anti-retinal autoimmunity in two different forms of macular degeneration in cynomolgus monkey (Macaca fascicularis). FASEB J 2005;19:1683-5.
  31. Wang J, Ohno-Matsui K, Yoshida T, et al. Amyloid-beta up-regulates complement factor B in retinal pigment epithelial cells through cytokines released from recruited macrophages/microglia: Another mechanism of complement activation in age-related macular degeneration. J Cell Physiol 2009;220:119-28.
  32. Gass JD. Pathogenesis of disciform detachment of the neuroepithelium. Am J Ophthalmol 1967;63:Suppl:1-139.
  33. Bird AC, Bressler NM, Bressler SB, et al. An international classification and grading system for age-related maculopathy and age-related macular degeneration. The International ARM Epidemiological Study Group. Surv Ophthalmol 1995;39:367-74.
  34. Klein R, Klein BE, Jensen SC, Meuer SM. The five-year incidence and progression of age-related maculopathy: the Beaver Dam Eye Study. Ophthalmology 1997;104:7-21.
  35. Yannuzzi LA, Green WR, Guyer DR. The retina atlas. [St. Louis, MO]: Mosby, 1998.
  36. Bressler NM, Silva JC, Bressler SB, Fine SL, Green WR. Clinicopathologic correlation of drusen and retinal pigment epithelial abnormalities in age-related macular degeneration. Retina 1994;14:130-42.
  37. Sarks SH, Arnold JJ, Killingsworth MC, Sarks JP. Early drusen formation in the normal and aging eye and their relation to age related maculopathy: a clinicopathological study. Br J Ophthalmol 1999;83:358-68.
  38. Gass JD. Drusen and disciform macular detachment and degeneration. Arch Ophthalmol 1973;90:206-17.
  39. Bressler NM, Bressler SB, West SK, Fine SL, Taylor HR. The grading and prevalence of macular degeneration in Chesapeake Bay watermen. Arch Ophthalmol 1989;107:847-52.
  40. Bressler SB, Maguire MG, Bressler NM, Fine SL. Relationship of drusen and abnormalities of the retinal pigment epithelium to the prognosis of neovascular macular degeneration. The Macular Photocoagulation Study Group. Arch Ophthalmol 1990;108:1442-7.
  41. Whitcomb EA, Shang F, Taylor A. Common cell biologic and biochemical changes in aging and age-related diseases of the eye: toward new therapeutic approaches to age-related ocular diseases. Invest Ophthalmol Vis Sci 2013;54:ORSF31–ORSF36.
  42. Bowes Rickman C, Farsiu S, Toth CA, Klingeborn M. Dry age- related macular degeneration: mecha- nisms, therapeutic targets, and imaging. Invest Ophthalmol Vis Sci 2013;54:ORSF68–ORSF80.
  43. Zarbin M. Current Concepts in the Pathogenesis of Age-Related Macular Degeneration.Arch Ophthalmol 2004;122(4):598-614.
  44. McLeod DS, Taomoto M, Otsuji T, Green WR, Sunness JS, Lutty GA. Quantifying changes in RPE and choroidal vasculature in eyes with age-related macular degeneration. Invest Ophthalmol Vis Sci 2002;43:1986-93.
  45. Lutty G, Grunwald J, Majji AB, Uyama M, Yoneya S. Changes in choriocapillaris and retinal pigment epithelium in age-related macular degeneration. Mol Vis 1999;5:35.
  46. Sarks JP, Sarks SH, Killingsworth MC. Morphology of early choroidal neovascularisation in age-related macular degeneration: correlation with activity. Eye 1997;11(Pt 4):515-22.
  47. Grossniklaus HE, Green WR. Choroidal neovascularization. Am J Ophthalmol 2004;137:496-503.
  48. Penfold PL, Madigan MC, Gillies MC, Provis JM. Immunological and aetiological aspects of macular degeneration. Prog Retin Eye Res 2001;20:385-414.
  49. Anderson DH, Mullins RF, Hageman GS, Johnson LV. A role for local inflammation in the formation of drusen in the aging eye. Am J Ophthalmol 2002;134:411-31.
  50. Chen J, Connor KM, Smith LE. Overstaying their welcome: defective CX3CR1 microglia eyed in macular degeneration. J Clin Invest 2007;117:2758-62.
  51. Kim SY, Sadda S, Pearlman J, et al. Morphometric analysis of the macula in eyes with disciform age-related macular degeneration. Retina 2002;22:471-7.
  52. Jalkh AE, Avila MP, Trempe CL, McMeel JW, Schepens CL. Choroidal neovascularization in fellow eyes of patients with advanced senile macular degeneration. Role of laser photocoagulation. Arch Ophthalmol 1983;101:1194-7.
  53. Klein RJ, Zeiss C, Chew EY, et al. Complement factor H polymorphism in age-related macular degeneration. Science 2005;308:385-9.
  54. Yang Z, Camp NJ, Sun H, et al. A variant of the HTRA1 gene increases susceptibility to age-related macular degeneration. Science 2006;314:992-3.
  55. Allikmets R, Dean M. Bringing age-related macular degeneration into focus. Nat Genet 2008;40:820-1.

Genetics of AMD

Genetics of AMD

English

Author:

Elisete Brandão, MD
Hospital S. João, Porto, Portugal

 

Introduction:

Age-related macular degeneration (AMD) is a complex disease with demographic and environmental risk factors (age, diet, and smoking) but also genetic risk factors.

In fact, instead of having a single contributory gene, there are multiple genes of variable effects that seem to be involved turning the issue of genetics of AMD a complex one: AMD involves environmental factors and varying susceptibilities to these external factors based upon diferent genetic backgrounds(1).

The genetic component of the disease has been suspected from family, twin and sibling studies. According to several family studies, patients with a family history of AMD have an increased risk for developing AMD(2,3).

The concordance for the presence of the disease is greater among homozigous twins than among heterozygous twins.

In 2008, Luo et al.(4) estimated the magnitude of familial risks in a population–based cross-sectional and case–control study.

Recurrence risk in relatives indicate increased relative risks in siblings (2.95), first cousins (1.29), second cousins (1.13) and parents (5.66) of affected elements.

Many linkage and association studies have showed that chromosomes 1q (1q25-31) and 10q (10q26) had genes involved in this pathology(5-7).

It was the completion of the Human Genoma Project, 5 years ago, resulting in the knowledge of the sequencing of the human genome that allowed improved DNA sequencing and mapping technologies and consequently, identification of Single Nucleotide Polymorphism (SNPs).

There are several types of genetic sequences variations (polymorphisms) in the human genome: repeated polymorphisms, insertions and deletions. However the majority of the DNA sequence variations in the human genome is in the form of SNPs which are persistent single changes, substitutions or variants of a single base in at least a population and with a frequency of more than 1% referred as alleles and representing altered forms of a gene: different alleles may produce variations in inherited characteristics(8,9).

These variants are also important because they serve as genetic markers and in this way they can help in determining those which confer increased or decreased risk of several diseases including AMD.

Dissection of the genetic background of AMD has undergone tremendous progress in the last 2 years. We know, now, some polymorphisms which modulate AMD risk.

 

Complement system factors and AMD

Complement factor H (CFH)

 

CFH is a negative regulator of alternative pathway of the complement system which means that in normal conditions, it inhibits the alternative pathway complement system.

It is encoded by a gene localized in 1q23-32 and its dysfunction may lead to excessive inflammation and tissue damage(10).

Complement activity is very important for the imune responses against pathogens and dying cells but, over-activation can result in complement-mediated damage to nearby healthy tissue cells.

It is now accepted that CFH gene is an important susceptibility gene, harbouring variants and haplotypes (short DNA sequences containing alleles) associated with increased and reduced risk of AMD.

Six CFH gene variants have been reported in AMD association studies as major genetic factors for developing AMD in Caucasians(11-15): rs1061170, (CFH Y402H); rs3753394; rs800292; rs1061147; rs380390; rs1329428.

However in the Chinese and Japanese populations only three of these CFH SNPs (rs1329498, rs800292 and rs3753394) were associated with risk of AMD(16,17).

So it is possible that CFH could play a central role in AMD pathogenesis and that multiple SNPs that alter CFH function might contribute to the development of AMD.

Their importance varies among the race of the population.

In the variant (polymorphism) CFH Y402H of the CFH gene, there is a substitution on the nucleotide in exon 9 (1277) where thymine (T) is changed for cytosine (C) (rs1061170) which is the allele risk.

This change leads to the substitution of the aminoacid in the position 402 in the protein, from tyrosine (Y) to histidine (H).

Homozygote CC or heterozygote TC can account for 50% of AMD cases.

The risk attributable for a disease is the rate of disease among individuals with a given characteristic minus the rate of the disease among indivuals without that characteristic.

The population attributable risk (PAR) in individuals with this polymorphism for developing AMD is 43% to 50%(11,12,18).

When compared with those with no risk allele TT, one copy of the Tyr402His polimorphysm (heterozygous for the risk allele TC), increases the risk of AMD by a factor of 2.2 to 4.6 (these individuals are at least twice and half more likely to develop AMD) and two copies of the risk allele (homozygous for the risk allele CC) increases the risk by a factor of 3.3 to 7.4 in Whites(19).

In addition to the common risk haplotype carrying the C allele of CFH Y402H, haplotype analysis of CFH has revealed two common protective haplotypes: homozygous deletions CFHR1 or CFHR3.

The gene cluster of CFH includes other “CFH-related genes“: CFHR1, CFHR2, CFHR3, CFHR4 and CFHR5. This means that the CFH gene resides within the region of complement activation (RCA), which includes also five “CFH–related” genes.

While the function of the CFH related genes is largely unknown, the high degree of sequence similarity between these genes and the suggestion that they arose out of duplication events with CFH, suggest an overlapping function of the CFH-related genes in immune system function and /or regulation.

There is a common and widespread (commonly found in many different populations in the world) deletion within the RCA locus that encompasses the CFHR1 and CFHR3 genes.

However the frequency of homozygous CFHR1 or CFHR3 deletion shows considerable variation between ethnic groups and occurs in 17.3% of African populations, 15.9% of African-American, 6.8% of Hispanic, 4.7% of Caucasian and 2.2% of Chinese cohorts(20).

This is in agreement with AMD less frequency among African-Americans compared with Caucasians and Chinese populations.

CFH1 and CFH3 protein may compete with CFH for C3 binding and therefore interfere with normal regulation of the complement system.

Those individuals, who are homozygous for the CFHR1/CFHR3 deletions and, therefore do not express the respective proteins, are highly protected from developing AMD(20).

 

Complement factor B (BF), complement component 2 (C2)

Complement factor B (BF) is involved in the activation of the complement alternative pathway and complement component 2 (C2) is involved in the activation of the classical pathway of the complement and both have adjacent genes located 500 base pair apart on chromosome 6p21.3 within the major histocompatibility complex class III region(21).

Haplotypes in BF and C2 have been linked to AMD. In particular, the L9H in BF and the E318D in C2 and also the R32Q in BF and a variant in intron 10 of C2 have been showed to be protective for AMD by Gold et al.(21).

They hypothesized that the significance of the haplotypes is due largely to the BF variants, which are in strong linkage disequilibrium with C2.

BF is a complement activating factor and studies have demonstrated that at least one of the two variants associated with AMD (R32Q BF) leads to an impairment in the complement activation function of BF.

This means that the absence of these variants C2/BF can predispose patients to AMD(21).

Thus, much like impaired CFH-mediated complement inhibition confers AMD risk, decreased complement activation by BF might serve to protect against AMD risk.

 

Complement component 3 (C3)

C3 is the central element of the complement cascade and a candidate gene to be involved in AMD, since its cleavage product, C3a, not only was found in drusen but also was proved to induce vascular endothelial growth factor expression and promote choroidal neovascularization in both in vitro and in vivo(22-24).

The variants R102G and P314L of the C3 gene significantly increase the risk of early and all subtypes of AMD and this risk seems to be independent of CFH Y402H, LOC387715 A69S and smoking(25).

 

LOC387715/ARMS2 and HTRA1

In 2003 Majewski et al. suggested that chromosome 10q26 might contain an AMD gene(7).

Later this finding has been replicated by other genome-wide linkage studies(26) and supported by a genome–scan meta-analysis(5).

This locus contains three tightly linked genes: PLEKHA1, LOC387715/ARMS2 (age related maculopathy suceptibility gene 2) and HTRA1, a secreted heat shock serine protease.

In 2005 Jakobsdottir et al.(27) found that the strongest association was over LOC387715/ARMS2 and HTRA1, which share an extensive linkage disequilibrium (LD) block harbouring the high risk haplotype.

There has been more dispute than agreement between the studies in what concerns this locus.

All initial genetic studies, about ten years ago, lacked statistical power (because small samples were used), used cumbersome genotyping technologies and poorly defined cohort.

In recent years, there are some publications of preliminary and unconfirmed genetic associations of the genes in this locus to AMD(28).

 

ARMS2 (LOC387715) SNP and AMD

The association of ARMS2 gene and AMD has been now replicated in various independent studies especially the advanced form of the disease, that means to say, the “wet” or with choroidal neovascularization and the “dry” or geographic atrophy form of the disease(29-32).

The risk confering polymorphism consists in a change in the 69 position aminoacid alanine (A) to serine (S).

According to Ross(31), heterozygosis at the ARMS2/LOC387715 ( A69A/A69S) is associated with odds ratio (OR) of 1.69-3.0 for advanced AMD while homozygozity for the risk conferring allele (A69S/A69S) results in a OR of 2.20-12.1.

The frequency of the risk allele is higher in patients with advanced AMD than in those with early or intermediate AMD(27,33).

Later two studies, based on semiquantitavive expression data of allele associated differences in HTRA1 mRNA or protein levels, suggested a diferent variant (rs11200638) in the same LD block, in the promoter of HTRA1 gene as the functional variant(34,35).

 

HTRA1 (high temperature required factor A-1) SNP and AMD

HTRA1 gene is located on chromosome 10q26.3, extremely close to the locus of the ARMS2 gene (10q26.13) and because of its role in extracellular matrix homeostasis (its extracellular protease activity may favour neovascularization) and in cellular growth or survival (it is an inhibitor of TGF-ß family member(36) and it could play a critical role in controlling TGF-ß dependent neuronal survival(37) it seems a possible functional candidate gene.

Four significant SNPs have been reported in the promoter and the first exon of HTRA1: G625A (rs11200638); T487C (rs2672598); C102T, A34A (rs1049331); G108T, G36G (rs2293870). However the most well documented, statistically significant AMD associated SNP is rs11200638 (G625A) in the promotor region.

Caucasians, Chinese and Japanese heterozygous for the risk allele (G/A) have a high OR of 1.60-2.61 and Caucasians, Chinese and Japanese homozygous for the risk allele (A/A), 6.56-10.0(34,35,38-40).

According to Tam et al.(40), there is an increase in population attributable risk (about 5.5 fold increase) by the joint effect of smoking and HTRA1 allele.

This means that smokers homozygous for the risk allele had a substantially higher risk of developing wet AMD than non smokers with the risk allele.

However Deangelis et al(41), in 2008 reported no interaction between this SNP and smoking.

In what concerns the studies which relate HRTA1 promoter polymorphisms to risk factors for developing AMD, three problems arise according to Allikmets and Dean(28).

The variant encoding the A69S (rs10490924) in ARMS2 and the rs11200638 variant in HTRA1 are almost in complete LD, so it is impossible to assign causality on the basis of allele frequency alone.

10q26 locus doesn’t harbour a wet AMD gene as the authors claimed but a late AMD gene as showned by Weber and colleagues in 2005(30).

All subsequent studies have failed to replicate the functional data(32,42).

This basically means that, as there is strong linkage disequilibrium (LD) across ARMS2-HTRA1 region, genetic association studies alone are insufficient to distinguish between the two candidates.

It is also necessary not only the characterization of the extent of the variants associated to the disease but also the analysis of their possible functional relevance in the disease process(42). Doing this, Fritsche et al.(42) claimed that the functional variant in this locus is the deletion-insertion polymorphism variant 372-815delins54 in the ARMS2 gene.

The deletion removes the polyadenilation signal sequence at position 395-400 exclusively used for the addition of a poly A tract 19 bp downstream.

The insertion introduces a 64 bp AU-rich element, known for its properties to control mRNA decay in many transcripts that encode a wide variety of proteins(43,45).

They demonstrated that it is a major risk factor for AMD: individuals carrying a single copy of the risk allele deletion-insertion in ARMS2 gene have a 2.8–fold increased risk compared with an 8.1–fold increased risk in homozigous individuals.

Their work, also revealed that in homozygous for the deletion-insertion variant, expression of ARMS2 is absent.

They localized the ARMS2 protein within the photoreceptor layer namely, to the mitochondria–enriched ellipsoid region of the inner segments and in accordance; they proposed a functional role of ARMS2 in mitochondrial homeostasis.

According to Fritsche et al., this suggests, that this polimorphysm is the sought-after functional variant with relevance in AMD etiology in 10q26 locus.

However, as Fritsche et al. recognize, it is ultimately required formal exclusion of functional consequences for the remaining polymorphisms on the risk haplotype namely, A69S in ARMS2 gene and HTRA1 promoter variant.

The A69S and the InDel are in 100% LD and on the same haplotype and so the effects are not independent to each other(46).

The work of Fritsche and colleagues does not eliminate all other possibilities(28), nobody disputes the role of complement genes in AMD in spite of the functional consequences of the disease associated variants being not known for CFH, CFB/C2 and/or C3.

 

Apolipoprotein E gene (ApoE)

The ApoE gene, located on chromosome 19q13.2, is polymorphic and has three isoforms wich are common, E2, E3 and E4, coded by different alleles: the ancestral E3 and the SNPs, E2 and E4(47).

Most studies favour a protective role for the ApoE4 SNP and a slight risk-conferring role for ApoE2(47-51).

However other studies do not(52-55).

 

SNP genotype and therapeutic responses

Genotype and response to antioxidative/zinc therapeutics

One of the first works in this area was that of Michael Klein et al.(56).

These authors correlated the CFH and LOC387715 A69S genotypes with the therapeutic responses to supplementation with antioxidants and zinc.

They concluded that, in homozygous individuals for the non-risk phenotype (Y402Y/Y402Y), 34% of those treated with placebo progressed to advanced AMD, compared to 11% of those treated with antioxidants and zinc: a reduction of approximately 68%.

In homozygous individuals for the risk allele (Y402H/Y402H), 44% of those treated with placebo progressed to advanced AMD, compared with 39% of those treated with antioxidants plus zinc: a reduction of only 11%.

A similar interaction was observed in the groups taking zinc versus those not taking zinc: intake had a more protective effect in patients with non-risk alleles compared to patients with risk alleles.

These results suggest that the zinc plus antioxidative treatment seems to have less impact on those with the high-risk CFH variant.

These authors found no association between LOC387715/ARMS2 A69S and the response to AREDS treatment.

 

Genotype and response to intravitreous bevacizumab

Brantley et al(57) investigated 86 wet AMD patients for the association between CFH and LOC387715/ARMS2 genotypes and the response to treatment with bevacizumab.

For the CFH genotypes results show that only 10.5% of patients homozigous for the risk - conferring allele Y402H/Y402H genotype demonstrated improved vision with treatment compared with 53.7% of patients homozigous for the non-risk allele Y402Y/Y402Y and the heterozigous for the non-risk allele Y402Y/Y402H variants.

They found that the CFH variants are associated to the responses to bevacizumab but that LOC387715/ARMS2 are not.

 

Genotype and response to photodynamic therapy

Goverdhan et al(58) and Brantley et al(59) studied the association of the CFH and LOC387715 genotypes with the response to PDT.

They found no statistical association with the LOC387715 genotype but a statistical association with CFH genotype: risk allele genotypes have better results than non-risk allele genotypes.

However more studies are warranted before any definitive conclusions.

 

Conclusions

1. Genetics variants at two chromosomal loci, 1q31 and 10q26, confer major disease risks, together accounting for more than 50% of AMD pathology(11-14, 27,30).

At present SNPs are the best available markers of AMD risk: SNPs in complementt factor H and ARMS2/HTRA1 capture a substancial fraction of AMD risk and permit the identification of individuals at high risk of developing AMD.

Genetic markers can successefully identify individuals whose lifetime risk of age-related macular degeneration ranges from 1% to greater than 50%.

2. Understanding the genetic basis of AMD has important implications for the ophthalmologists as it allows the identification of the biochemical pathway for a large proportion of AMD patients, raises the possibility to perform pre-symptomatic diagnostic testing of risk genotypes and to stratify the response to therapy based on genetic risks and supports the development of new therapies being the inhibition of complement a potential one among others that are already being tested.

 

>> References

References - Genetics of AMD

References - Genetics of AMD

English

1. Guymer R. The genetics of age-related macular degeneration. Clin Exp Optom. 2001; 84: 182–9.

2. Seddon JM, Francis PJ, George S, Schultz DW, Rosner B, Klein ML. Association of CFH Y402H and LOC387715 A69S with progression of age-related macular degeneration. JAMA 2007; 297 (16): 1793-800.

3. Smith W, Mitchell P. Family history and age-related maculopathy: the Blue Mountains Eye Study. Aust N Z J Ophthalmol. 1998; 26 (3): 203-6.

4. Luo L, Harmon J, Yang X, Chen H, Patel S, Mineau G, Yang Z, Constantine R, Buehler J, Kaminoh Y, Ma X, Wong TY, Zhang M, Zhang K. Familial aggregation of age-related macular degeneration in the Utah population. Vision Res 2008; 48 (3): 494-500.

5. Fisher SA, Abecasis GR, Yashar BM, Zareparsi S, Swaroop A, Iyengar SK, Klein BE, Klein R, Lee KE, Majewski J, Schultz DW, Klein ML, Seddon JM, Santangelo SL, Weeks DE, Conley YP, Mah TS, Schmidt S, Haines JL, Pericak-Vance MA, Gorin MB, Schulz HL, Pardi F, Lewis CM, Weber BH. Meta-analysis of genome scans of age-related macular degeneration. Hum Mol Genet 2005; 14 (15): 2257-64.

6. Klein ML, Schultz DW, Edwards A, Matise TC, Rust K, Berselli CB, Trzupek K, Weleber RG, Ott J, Wirtz MK, Acott TS. Age-related macular degeneration. Clinical features in a large family and linkage to chromosome 1q. Arch Ophthalmol 1998; 116 (8): 1082-8.

7. Majewski J, Schultz DW, Weleber RG, Schain MB, Edwards AO, Matise TC, Acott TS, Ott J, Klein ML. Age-related macular degeneration--a genome scan in extended families. Am J Hum Genet 2003; 73 (3): 540-50.

8. Donoso LA, Kim D, Frost A, Callahan A, Hageman G. The role of inflammation in the pathogenesis of age-related macular degeneration. Surv Ophthalmol 2006; 51 (2): 137-52.

9. Tuo J, Bojanowski CM, Chan CC. Genetic factors of age-related macular degeneration. Prog Retin Eye Res 2004; 23 (2): 229-49.

10. Johnson PT, Betts KE, Radeke MJ, Hageman GS, Anderson DH, Johnson LV. Individuals homozygous for the age-related macular degeneration risk-conferring variant of complement factor H have elevated levels of CRP in the choroid. Proc Natl Acad Sci U S A 2006; 103 (46): 17456-61.

11. Edwards AO, Ritter R, Abel KJ, Manning A, Panhuysen C, Farrer LA. Complement factor H polymorphism and age-related macular degeneration. Science 2005; 308 (5720): 362-4.

12. Hageman GS, Anderson DH, Johnson LV, Hancox LS, Taiber AJ, Hardisty LI, Hageman JL, Stockman HA, Borchardt JD, Gehrs KM, Smith RJ, Silvestri G, Russell SR, Klaver CC, Barbazetto I, Chang S, Yannuzzi LA, Barile GR, Merriam JC, Smith RT, Olsh AK, Bergeron J, Zernant J, Merriam JE, Gold B, Dean M, Allikmets R. A common haplotype in the complement regulatory gene factor H (HF1/CFH) predisposes individuals to age-related macular degeneration. Proc Natl Acad Sci U S A 2005; 102 (20): 7227-32.

13. Haines JL, Hauser MA, Schmidt S, Scott WK, Olson LM, Gallins P, Spencer KL, Kwan SY, Noureddine M, Gilbert JR, Schnetz-Boutaud N, Agarwal A, Postel EA, Pericak-Vance MA. Complement factor H variant increases the risk of age-related macular degeneration. Science 2005; 308 (5720): 419-21.

14. Klein RJ, Zeiss C, Chew EY, Tsai JY, Sackler RS, Haynes C, Henning AK, SanGiovanni JP, Mane SM, Mayne ST, Bracken MB, Ferris FL, Ott J, Barnstable C, Hoh J. Complement factor H polymorphism in age-related macular degeneration. Science 2005; 308 (5720): 385-9.

15. Tuo J, Ning B, Bojanowski CM, Lin ZN, Ross RJ, Reed GF, Shen D, Jiao X, Zhou M, Chew EY, Kadlubar FF, Chan CC. Synergic effect of polymorphisms in ERCC6 5’ flanking region and complement factor H on age-related macular degeneration predisposition. Proc Natl Acad Sci U S A 2006; 103 (24): 9256-61.

16. Chen LJ, Liu DT, Tam PO, Chan WM, Liu K, Chong KK, Lam DS, Pang CP. Association of complement factor H polymorphisms with exudative age-related macular degeneration. Mol Vis 2006; 12: 1536-42.

17. Okamoto H, Umeda S, Obazawa M, Minami M, Noda T, Mizota A, Honda M, Tanaka M, Koyama R, Takagi I, Sakamoto Y, Saito Y, Miyake Y, Iwata T. Complement factor H polymorphisms in Japanese population with age-related macular degeneration. Mol Vis 2006; 12: 156-8.

18. Zareparsi S, Branham KE, Li M, Shah S, Klein RJ, Ott J, Hoh J, Abecasis GR, Swaroop A. Strong association of the Y402H variant in complement factor H at 1q32 with susceptibility to age-related macular degeneration. Am J Hum Genet 2005; 77 (1): 149-53.

19. Thakkinstian A, Han P, McEvoy M, Smith W, Hoh J, Magnusson K, Zhang K, Attia J. Systematic review and meta-analysis of the association between complement factor H Y402H polymorphisms and age-related macular degeneration. Hum Mol Genet 2006;15 (18): 2784-90.

20. Hageman GS, Hancox LS, Taiber AJ, Gehrs KM, Anderson DH, Johnson LV, Radeke MJ, Kavanagh D, Richards A, Atkinson J, Meri S, Bergeron J, Zernant J, Merriam J, Gold B, Allikmets R, Dean M; AMD Clinical Study Group. Extended haplotypes in the complement factor H (CFH) and CFH-related (CFHR) family of genes protect against age-related macular degeneration: characterization, ethnic distribution and evolutionary implications. Ann Med 2006; 38 (8): 592-604.

21. Gold B, Merriam JE, Zernant J, Hancox LS, Taiber AJ, Gehrs K, Cramer K, Neel J, Bergeron J, Barile GR, Smith RT; AMD Genetics Clinical Study Group, Hageman GS, Dean M, Allikmets R. Variation in factor B (BF) and complement component 2 (C2) genes is associated with age-related macular degeneration. Nat Genet 2006; 38 (4): 458-62.

22. Hageman GS, Luthert PJ, Victor Chong NH, Johnson LV, Anderson DH, Mullins RF. An integrated hypothesis that considers drusen as biomarkers of immune-mediated processes at the RPE-Bruch’s membrane interface in aging and age-related macular degeneration. Prog Retin Eye Res 2001; 20 (6): 705-32.

23. Johnson LV, Leitner WP, Staples MK, Anderson DH. Complement activation and inflammatory processes in Drusen formation and age related macular degeneration. Exp Eye Res 2001; 73 (6): 887-96.

24. Nozaki M, Raisler BJ, Sakurai E, Sarma JV, Barnum SR, Lambris JD, Chen Y, Zhang K, Ambati BK, Baffi JZ, Ambati J. Drusen complement components C3a and C5a promote choroidal neovascularization. Proc Natl Acad Sci U S A 2006; 103 (7): 2328-33.

25. Despriet DD, van Duijn CM, Oostra BA, Uitterlinden AG, Hofman A, Wright AF, ten Brink JB, Bakker A, de Jong PT, Vingerling JR, Bergen AA, Klaver CC. Complement component C3 and risk of age-related macular degeneration. Ophthalmology 2009; 116 (3):474-480.e2.

26. Weeks DE, Conley YP, Tsai HJ, Mah TS, Schmidt S, Postel EA, Agarwal A, Haines JL, Pericak-Vance MA, Rosenfeld PJ, Paul TO, Eller AW, Morse LS, Dailey JP, Ferrell RE, Gorin MB. Age-related maculopathy: a genomewide scan with continued evidence of susceptibility loci within the 1q31, 10q26, and 17q25 regions. Am J Hum Genet 2004; 75 (2): 174-89.

27. Jakobsdottir J, Conley YP, Weeks DE, Mah TS, Ferrell RE, Gorin MB. Susceptibility genes for age-related maculopathy on chromosome 10q26. Am J Hum Genet 2005; 77 (3): 389-407.

28. Allikmets R, Dean M. Bringing age-related macular degeneration into focus. Nat Genet 2008; 40 (7): 820-1.

29. Conley YP, Jakobsdottir J, Mah T, Weeks DE, Klein R, Kuller L, Ferrell RE, Gorin MB. CFH, ELOVL4, PLEKHA1 and LOC387715 genes and susceptibility to age-related maculopathy: AREDS and CHS cohorts and meta-analyses. Hum Mol Genet 2006; 15 (21): 3206-18.

30. Rivera A, Fisher SA, Fritsche LG, Keilhauer CN, Lichtner P, Meitinger T, Weber BH. Hypothetical LOC387715 is a second major susceptibility gene for age-related macular degeneration, contributing independently of complement factor H to disease risk. Hum Mol Genet 2005; 14 (21): 3227-36.

31. Ross RJ, Bojanowski CM, Wang JJ, Chew EY, Rochtchina E, Ferris FL 3rd, Mitchell P, Chan CC, Tuo J. The LOC387715 polymorphism and age-related macular degeneration: replication in three case-control samples. Invest Ophthalmol Vis Sci 2007; 48 (3): 1128-32.

32. Kanda A, Chen W, Othman M, Branham KE, Brooks M, Khanna R, He S, Lyons R, Abecasis GR, Swaroop A. A variant of mitochondrial protein LOC387715/ARMS2, not HTRA1, is strongly associated with age-related macular degeneration. Proc Natl Acad Sci U S A. 2007; 104 (41): 16227-32.

33. Wang JJ, Ross RJ, Tuo J, Burlutsky G, Tan AG, Chan CC, Favaloro EJ, Williams A, Mitchell P. The LOC387715 polymorphism, inflammatory markers, smoking, and age-related macular degeneration. A population-based case-control study. Ophthalmology 2008; 115 (4): 693-9.

34. Yang Z, Camp NJ, Sun H, Tong Z, Gibbs D, Cameron DJ, Chen H, Zhao Y, Pearson E, Li X, Chien J, Dewan A, Harmon J, Bernstein PS, Shridhar V, Zabriskie NA, Hoh J, Howes K, Zhang K. A variant of the HTRA1 gene increases susceptibility to age-related macular degeneration. Science 2006; 314 (5801): 992-3.

35. Dewan A, Liu M, Hartman S, Zhang SS, Liu DT, Zhao C, Tam PO, Chan WM, Lam DS, Snyder M, Barnstable C, Pang CP, Hoh J. HTRA1 promoter polymorphism in wet age-related macular degeneration. Science 2006; 314 (5801): 989-92.

36. Oka C, Tsujimoto R, Kajikawa M, Koshiba-Takeuchi K, Ina J, Yano M, Tsuchiya A, Ueta Y, Soma A, Kanda H, Matsumoto M, Kawaichi M. HtrA1 serine protease inhibits signaling mediated by Tgfbeta family proteins. Development 2004; 131 (5): 1041-53.

37. Launay S, Maubert E, Lebeurrier N, Tennstaedt A, Campioni M, Docagne F, Gabriel C, Dauphinot L, Potier MC, Ehrmann M, Baldi A, Vivien D. HtrA1-dependent proteolysis of TGF-beta controls both neuronal maturation and developmental survival. Cell Death Differ 2008; 15 (9): 1408-16.

38. Cameron DJ, Yang Z, Gibbs D, Chen H, Kaminoh Y, Jorgensen A, Zeng J, Luo L, Brinton E, Brinton G, Brand JM, Bernstein PS, Zabriskie NA, Tang S, Constantine R, Tong Z, Zhang K. HTRA1 variant confers similar risks to geographic atrophy and neovascular age-related macular degeneration. Cell Cycle 2007; 6 (9): 1122-5.

39. Mori K, Horie-Inoue K, Kohda M, Kawasaki I, Gehlbach PL, Awata T, Yoneya S, Okazaki Y, Inoue S. Association of the HTRA1 gene variant with age-related macular degeneration in the Japanese population. J Hum Genet 2007; 52(7): 636-41.

40. Tam PO, Ng TK, Liu DT, Chan WM, Chiang SW, Chen LJ, DeWan A, Hoh J, Lam DS, Pang CP. HTRA1 variants in exudative age-related macular degeneration and interactions with smoking and CFH. Invest Ophthalmol Vis Sci 2008; 49 (6): 2357-65.

41. Deangelis MM, Ji F, Adams S, Morrison MA, Harring AJ, Sweeney MO, Capone A Jr, Miller JW, Dryja TP, Ott J, Kim IK. Alleles in the HtrA serine peptidase 1 gene alter the risk of neovascular age-related macular degeneration. Ophthalmology 2008; 115 (7): 1209-1215.e7.

42. Fritsche LG, Loenhardt T, Janssen A, Fisher SA, Rivera A, Keilhauer CN, Weber BH. Age-related macular degeneration is associated with an unstable ARMS2 (LOC387715) mRNA. Nat Genet 2008; 40 (7): 892-6.

43. Barreau C, Paillard L, Osborne HB. AU-rich elements and associated factors: are there unifying principles? Nucleic Acids Res 2006; 33 (22): 7138-50.

44. Garneau NL, Wilusz J, Wilusz CJ. The highways and byways of mRNA decay. Nat Rev Mol Cell Biol 2007; 8 (2): 113-26.

45. Khabar KS. The AU-rich transcriptome: more than interferons and cytokines, and its role in disease. J Interferon Cytokine Res 2005; 25 (1): 1-10.

46. Ding X, Patel M, Chan CC. Molecular pathology of age-related macular degeneration. Prog Retin Eye Res 2009; 28 (1): 1-18.

47. Schmidt S, Klaver C, Saunders A, Postel E, De La PM, Agarwal A, Small K, Udar N, Ong J, Chalukya M, Nesburn A, Kenney C, Domurath R, Hogan M, Mah T et al. A pooled case-control study of the apolipoprotein E (APOE) gene in age-related maculopathy. Ophthalmic Genet 2002; 23 (4): 209-223.

48. Baird PN, Richardson AJ, Robman LD, Dimitrov PN, Tikellis G, McCarty CA, Guymer RH. Apolipoprotein (APOE) gene is associated with progression of age-related macular degeneration (AMD). Hum Mutat 2006; 27 (4): 337-342.

49. Kovacs KA, Pamer Z, Kovacs A, Fekete S, Miseta A, Kovacs B, Kovacs GL. Association of apolipoprotein E polymorphism with age-related macular degeneration and Alzheimer’s disease in south-western Hungary. Ideggyogy Sz 2007; 60 (3-4): 169-172.

50. Souied EH, Benlian P, Amouyel P, Feingold J, Lagarde JP, Munnich A, Kaplan J, Coscas G, Soubrane G. The epsilon4 allele of the apolipoprotein E gene as a potential protective factor for exudative age-related macular degeneration. Am J Ophthalmol 1998; 125 (3): 353-359.

51. Zareparsi S, Reddick AC, Branham KE, Moore KB, Jessup L, Thoms S, Smith-Wheelock M, Yashar BM, Swaroop A. Association of apolipoprotein E alleles with susceptibility to age-related macular degeneration in a large cohort from a single center. Invest Ophthalmol Vis Sci 2004; 45 (5): 1306-1310.

52. Kaur I, Hussain A, Hussain N, Das T, Pathangay A, Mathai A, Hussain A, Nutheti R, Nirmalan PK, Chakrabarti S. Analysis of CFH, TLR4, and APOE polymorphism in India suggests the Tyr402His variant of CFH to be a global marker for age-related macular degeneration. Invest Ophthalmol Vis Sci 2006; 47 (9): 3729-3735.

53. Schmidt S, Haines JL, Postel EA, Agarwal A, Kwan SY, Gilbert JR, Pericak-Vance MA, Scott WK. Joint effects of smoking history and APOE genotypes in age-related macular degeneration. Mol Vis 2005; 11: 941-949.

54. Schultz DW, Klein ML, Humpert A, Majewski J, Schain M, Weleber RG, Ott J, Acott TS. Lack of an association of apolipoprotein E gene polymorphisms with familial age-related macular degeneration. Arch Ophthalmol 2003; 121 (5): 679-683.

55. Wong TY, Shankar A, Klein R, Bray MS, Couper DJ, Klein BE, Sharrett AR, Folsom AR. Apolipoprotein E gene and early age-related maculopathy: the Atherosclerosis Risk in Communities Study. Ophthalmology 2006; 113 (2): 255-259.

56. Klein ML, Francis PJ, Rosner B, Reynolds R, Hamon SC, Schultz DW, Ott J, Seddon JM. CFH and LOC387715/ARMS2 genotypes and treatment with antioxidants and zinc for age-related macular degeneration. Ophthalmology 2008; 115 (6): 1019-1025.

57. Brantley MA, Jr., Fang AM, King JM, Tewari A, Kymes SM, Shiels A. Association of complement factor H and LOC387715 genotypes with response of exudative age-related macular degeneration to intravitreal bevacizumab. Ophthalmology 2007; 114 (12): 2168-2173.

58. Goverdhan SV, Hannan S, Newsom RB, Luff AJ, Griffiths H, Lotery AJ. An analysis of the CFH Y402H genotype in AMD patients and controls from the UK, and response to PDT treatment. Eye (Lond) 2008; 22 (6): 849-854.

59. Brantley MA, Jr., Edelstein SL, King JM, Plotzke MR, Apte RS, Kymes SM, Shiels A. Association of complement factor H and LOC387715 genotypes with response of exudative age-related macular degeneration to photodynamic therapy. Eye (Lond) 2009; 23 (3): 626-631.

Angiogenesis

Angiogenesis

English

Author:

Ângela Carneiro, MD, PhD

Manuel Falcão, MD, PhD

Faculty of Medicine of University of Porto, Hospital S. João, Porto, Portugal.

 

Updated/reviewed by the authors, July 2017.

Definition

Blood vessels develop and grow by three different basic mechanisms: vasculogenesis in which vessels form by concatenation of vascular precursor cells into solid cords that then lumenize; angiogenesis that is the growth of new blood vessels from pre-existing ones; and intussusception in which new blood vessels form by the proliferation of endothelial cells that form a pre-existing vessel into the vessel lumen, originating two blood vessels that split into two opposite sides(1).

Angiogenesis, the growth of new vessels from pre-existing ones by sprouting of endothelial cells into a previously avascular tissue, is an essential process both in embryonic development and in adulthood(1,2). It is a complex multistep process involving extracellular matrix degradation and proliferation, survival, migration and anastomosis of endothelial cells(2).

The release of extracellular matrix proteases leads to the degradation of the basal membrane of blood vessels. Endothelial cells change shape, proliferate, invade stroma and form tubular structures that coalesce. This requires the coordinated action of a variety of anti and pro-angiogenic factors and cell-adhesion molecules in endothelial cells. Angiogenesis is of paramount importance as it promotes tissue repair; however, in certain conditions it may cause tissue damage. If not tightly regulated, the angiogenic process is frequently imbalanced, and associated with several pathological situations(1,3).

 

Angiogenic mediators and modulation of their expression

The angiogenic process requires the activation of a series of receptors by numerous ligands including Placental Growth Factor (PIGF), Fibroblast Growth Factors (FGFs), Angiopoietin-1 and -2 (Ang-1 and -2), Platelet-derived Growth Factor (PDGF), Hepatocyte Growth Factor (HGF), Connective Tissue Growth Factor (CTGF) and Transforming Growth Factors (TGF-α e TGF-β), among many others(1, 3-9).

However, there is a consensus that Vascular Endothelial Growth Factor (VEGF) is the most important angiogenic factor and represents the crucial rate-limiting step during angiogenesis(3,10,11).

VEGF-A is the prototype member of a gene family that also includes placental growth factor (PlGF), VEGF-B, VEGF-C, VEGF-D, and the orf-virus-encoded VEGF-E(11). Alternative exon splicing results in the generation of four main VEGF isoforms, which have respectively 121, 165, 189, and 206 amino acids (VEGF121, VEGF165, VEGF189, VEGF206). Less frequent splice variants have also been reported, including VEGF145, VEGF183, VEGF162, and VEGF165b(8,11).

VEGF mediates its biological functions at the endothelial level by binding two highly related tyrosine kinase receptors (RTKs), VEGFR-1 and VEGFR-2. It is generally agreed that VEGFR-2 is the major mediator of the mitogenic, angiogenic and permeability-enhancing effects of VEGF-A(3). VEGFR-1 binds to both VEGF-A and Placenta Growth Factor and fails to mediate a strong mitogenic signal in endothelial cells. It is now generally agreed that VEGFR-1 plays a role in the modulation of VEGF activity(10). VEGFR-3 is mainly a VEGF-C receptor but it plays important roles in lymphangiogenesis and angiogenesis(12).

The mediators of the angiogenic process can be modulated by some molecules and microenvironmental conditions. VEGF is upregulated by cyclo.oxygenase (COX-2)(13). Inflammatory cells within a hypoxic environment release huge amounts of factors that exert effects on endothelial cells and degrade the extracellular matrix(2).

Angiogenesis can also be suppressed by inhibitory molecules, such as interferon-α, thrombospondin-1, angiostatin, endostatin or pigment epithelial-derived factor (PEDF)(14-18).

It is the balance of stimulators and inhibitors that tightly controls the normally quiescent capillary vasculature. When this balance is upset angiogenesis develops(19).

Cell-cell and cell-matrix interactions may also play an important role in angiogenesis. Special focus has been recently given the Rac1 GTPase(20). Recently, microRNA regulation of gene expression has been implicated in the control of pathologic ocular angiogenesis(21).

The angiopoietin (ANG)–TIE signalling pathway has been identified as the second vascular tissue-specific receptor Tyr kinase system. The ANG–TIE pathway is required for lymphatic and blood vessel development and is important for the development of mature blood vessels that originate from the VEGF induced endothelial sprouting. This pathway controls vascular permeability, inflammation and pathological angiogenic responses in adult tissues(22). Maturing of blood vessels include pericyte and coating of endothelial cell walls.

PDGF is mainly believed to be an important mitogen for connective tissue. PDGF promotes migration and proliferation of endothelial cells as well as an increased recruitment of pericytes. These findings suggest that PDGF is not only important in formation of new blood vessels but it is also very important for their maturation and stabilization(23).

Evidence suggests that processes of inflammation and angiogenesis are connected. Newly formed blood vessels enable the recruitment of inflammatory cells, which release a variety of proangiogenic cytokines and growth factors that will perpetuate angiogenesis(24).

 

Angiogenesis during development of retinal vasculature

During embryogenesis retinal vascularization begins in the most superficial (or inner) retinal layers at the optic nerve head, and radiates outwards from this central point. It reaches the retinal periphery just before birth(25).

The migration of large numbers of vascular precursor cells (VPCs) from the optic disc is the first event in human retinal vascularization, and it is apparent before 12 gestational weeks(26). They proliferate and differentiate to form a primordial vascular bed centered on the optic disc. Thus, vasculogenesis is responsible for the formation of the primordial vessels of the inner (superficial) plexus in the central human retina(27). Formation of retinal vessels via vasculogenesis seems to be independent of metabolic demand and hypoxia-induced VEGF expression(28).

Angiogenesis is responsible for the formation of the remaining retinal vessels, including increasing vascular density in the central retina, vessel formation in the inner plexus of the peripheral retina, and formation of the outer plexus and the radial peri-papillary capillaries(28). Formation of the outer plexus begins around the incipient fovea between 25 and 26 weeks of gestation, coincident with signals that indicate a functional visual pathway and photoreceptor activity(27). The timing and topography of angiogenesis in the human retina supports the “physiological hypoxia” model of retinal vascular formation, in which angiogenesis is induced by a transient but physiological level of hypoxia as a result of the increased metabolic activity of retinal neurons as they differentiate and become functional(29).

 

Angiogenesis in retina and choroidal pathologies

Retinal anatomy is highly organized and vascular and avascular compartments are strictly segregated in the retina(1). The blood-retinal barriers, inner and outer, are fundamental for the integrity of structure and optimization of function of the neuro-sensorial retina(30).

Pathological retinal and choroidal angiogenesis generates chaotically orientated and physiologically deficient vessels that do not conform to neuronal histology, which can lead to vision-threatening oedema, exudation and haemorrhage(1).

Angiogenesis is a key aspect in many ocular pathologies that are leading causes of blindness in the world, such as neovascular age-related macular degeneration (AMD), diabetic retinopathy, retinopathy of prematurity, central retinal vein occlusion and other diseases associated with ischemia and neovascularization(31).

Although angiogenesis is a highly complex and coordinated process requiring multiple receptors and ligands in endothelial cells, VEGF is a hypoxia-inducible cytokine that appears to be a pivotal element required for the process in a variety of normal and pathological circumstances(3, 10). VEGF is a surrogate angiogenic marker, since it acts not only as a mitogen, but also as a survival factor for endothelial cells(2). Furthermore, it is also involved in the stimulation of the invasive and migration capacity of endothelial cells and in the enhancement of vascular permeability(10). Bone-marrow derived cells have also been described in choroidal neovascular lesion. Their importance in the pathologic process is still under debate(32).

 

Angiogenesis and Age-related Macular Degeneration

The diagnosis of AMD is based on fundoscopic signs observed on the macula, irrespective of visual acuity(33). The stages of AMD are categorized as early, in which visual symptoms are inconspicuous, (moderate) and late, usually associated with severe loss of vision(34). Early AMD is characterized by the presence of drusen and/or hyperpigmentations or small hypopigmentations(33). Late AMD has “dry” and “wet” forms. However, in the same patient we can find the dry form in one eye and the wet in the other eye, or both forms in the same eye. Moreover with time we can see the conversion of wet in dry or dry becoming wet(35).

Age-related changes that predispose to AMD occur in the outer retina, more specifically the region that includes the photoreceptors, the retinal pigment epithelium (RPE), Bruch’s membrane and the choriocapillaris.

The aging-dependent alterations in the outer retina have been already discussed in another chapter. AMD-related visual loss is a complex process starting by the deposition of debris in the outer retina(36). The deposition of insoluble material, the calcification and increase in thickness of Bruch’s membrane, and a less fenestrated and thinner choriocapillaris leads to photoreceptors/RPE hypoxia resulting in a stimulus for VEGF release(35, 37-39).

All the aging changes in outer retina compromise the nutrition of photoreceptors and RPE and create a favourable environment for the development of choroidal neovascularization (CNV). However other factors – genetic and environmental(40, 41) – are also important, but its role in the development of CNV is discussed in other chapters of this book.

In general terms there are two basic CNV growth patterns, based on the anatomical position of the abnormal vessels with respect to the RPE monolayer, which are related to Gass’s classification of CNV(42, 43). In type 1 CNV, the neovascular complex is located in the plane between the RPE and Bruch’s membrane and in type 2 neovascularization the vessels have penetrated the RPE layer to proliferate in the subneurosensory space(42).

In the type 1 growth pattern, after breaking through Bruch’s membrane, the CNV extend laterally under the RPE in a horizontal fashion that is facilitated by the natural cleavage plane between basal laminar deposits and a lipid rich Bruch’s membrane. This growth pattern recapitulates the choriocapillaris and can provide some nutrients and oxygen to an ischemic RPE/outer retina(43, 44)

The type 2 growth pattern occurs usually with one or few ingrowth sites with vascular leakage under the RPE/outer retina that usually lead to acute visual simptoms(43).

Yannuzzi proposed a type 3 neovascularization, for retinal angiomatous proliferation (RAP), indicating proliferating vessels within or below the retina itself(45). This mixed neovascularization, with a presumed dual origin, may have intraretinal neovascularization driven by angiogenic cytokines from Müller cells, endothelial cells, pericytes, and retinal glial cells, and CNV driven by cytokines from the RPE(45).  There is hypothetically neovascularization extending anteriorly from the choroid in conjunction with retinal neovascularization progressing posteriorly, with both circulations eventually anastomosing. 

The reason that growth patterns vary according to disease and individuals may be related to genetic predispositions, environmental mechanisms, variations in composition and anatomy of Bruch’s membrane, cytokine distribution, or other causes(43).

During the dynamic process of development of CNV there is a balance of angiogenesis promoters and inhibitors. In the initiation stage the RPE and photoreceptors produce VEGF(46). There is also production by RPE of Interleukin-8 (IL-8) and Monocyte Chemoattractant Protein-1 (MCP), witch attract monocytes from the choriocapillaris along the outer surface of Bruch’s membrane(47). The macrophages tend to concentrate around sites of vascular ingrowth through the Bruch’s membrane and express Tumor Necrosis Factor-α (TNF-α) and Interleukin-1 (IL-1), which up-regulate complement factor-B, activates the complement alternative pathway in the subretinal space, and stimulates RPE cells to produce more VEGF(47,48).

After initiation, CNV grows to a certain size and progresses through the tissue planes by the action of Matrix Metalloproteinases (MMP) produced by endothelial cells and macrophages(49). During this stage of active growth, Angiopoietins (Ang-1 and 2) are expressed, FGFs are produced by RPE and endothelial cells, and TGF-β is produced by the RPE(50-52). 

CNV stabilizes during the active stage due to a steady state established between MMP and tissue inhibitors of metalloproteinases, Ang-1 and 2, PEDF and VEGF, PDGF and VEGF, plasminogen and fibrin, and others(43, 53).

At some point the balance shifts toward antiangiogenic, antiproteolytic and antimigratory activity resulting in the involutional stage of CNV. When this occurs the angiogenic/proteolitic/migratory cytokine production decreases with a shift toward TGF-β and tissue inhibitors of metalloproteinases production by the RPE(43). In this involutional stage the CNV may become collagenized and form a disciform scar. Subretinal fibrosis is the hallmark of this stage of the disease and is the result of the activity of inflammatory and endothelial cells and fibroblasts(54).

 

Key points

  • Angiogenesis, the growth of new vessels from pre-existing ones, is a complex multistep process involving the activation of series of receptors by numerous ligands.
  • Vascular Endothelial Growth Factor (VEGF) is the most important angiogenic factor and represents the crucial rate-limiting step during angiogenesis.
  • There are three basic CNV growth patterns, based on the anatomical position of the abnormal vessels with respect to the RPE monolayer and to the proliferation beginning within or below the retina itself: occult (type 1), classic (type 2) and RAP (type 2) lesions.
  • Interleukin-8 (IL-8), Monocyte Chemoattractant Protein-1 (MCP), Tumor Necrosis Factor-α (TNF-α), Interleukin-1 (IL-1), Matrix Metalloproteinases (MMP), Angiopoietins (Ang-1 and 2), Fibroblast Growth Factors (FGFs) and Transforming Growth Factor β (TGF-β), among others, play an active role during the stage of active new vessel growth.

 

>> References

References - Angiogenesis

References - Angiogenesis

English
  1. Gariano RF, Gardner TW. Retinal angiogenesis in development and disease. Nature 2005;438(7070):960-6.
     
  2. Costa C, Incio J, Soares R. Angiogenesis and chronic inflammation: cause or consequence? Angiogenesis
    2007;10(3):149-66.
     
  3. Ferrara N, Kerbel RS. Angiogenesis as a therapeutic target. Nature 2005;438(7070):967-74.
     
  4. Yancopoulos GD, Davis S, Gale NW, et al. Vascular-specific growth factors and blood vessel formation. Nature 2000;407(6801):242-8.
     
  5. Presta M, Dell'Era P, Mitola S, et al. Fibroblast growth factor/fibroblast growth factor receptor system in angiogenesis. Cytokine Growth Factor Rev 2005;16(2):159-78.
     
  6. Miyazono K, Usuki K, Heldin CH. Platelet-derived endothelial cell growth factor. Prog Growth Factor Res 1991;3(3):207-17.
     
  7. Luttun A, Tjwa M, Carmeliet P. Placental growth factor (PlGF) and its receptor Flt-1 (VEGFR-1): novel therapeutic targets for angiogenic disorders. Ann N Y Acad Sci 2002;979:80-93.
     
  8. Ferrara N, Gerber HP, LeCouter J. The biology of VEGF and its receptors. Nat Med 2003;9(6):669-76.
     
  9. Ohnishi T, Daikuhara Y. Hepatocyte growth factor/scatter factor in development, inflammation and carcinogenesis: its expression and role in oral tissues. Arch Oral Biol 2003;48(12):797-804.
     
  10. Ferrara N, Damico L, Shams N, et al. Development of ranibizumab, an anti-vascular endothelial growth factor antigen binding fragment, as therapy for neovascular age-related macular degeneration. Retina 2006;26(8):859-70.
     
  11. Ferrara N, Mass RD, Campa C, et al. Targeting VEGF-A to treat cancer and age-related macular degeneration. Annu Rev Med 2007;58:491-504.
     
  12. Saharinen P, Eklund L, Pulkki K, et al. VEGF and angiopoietin signaling in tumor angiogenesis and metastasis. Trends Mol Med 2011;17(7):347-62.
     
  13. Costa C, Soares R, Reis-Filho JS, et al. Cyclo-oxygenase 2 expression is associated with angiogenesis and lymph node metastasis in human breast cancer. J Clin Pathol 2002;55(6):429-34.
     
  14. Ezekowitz RA, Mulliken JB, Folkman J. Interferon alfa-2a therapy for life-threatening hemangiomas of infancy. N Engl J Med 1992;326(22):1456-63.
     
  15. Tolsma SS, Volpert OV, Good DJ, et al. Peptides derived from two separate domains of the matrix protein thrombospondin-1 have anti-angiogenic activity. J Cell Biol 1993;122(2):497-511.
     
  16. O'Reilly MS, Boehm T, Shing Y, et al. Endostatin: an endogenous inhibitor of angiogenesis and tumor growth. Cell 1997;88(2):277-85.
     
  17. O'Reilly MS, Holmgren L, Shing Y, et al. Angiostatin: a novel angiogenesis inhibitor that mediates the suppression of metastases by a Lewis lung carcinoma. Cell 1994;79(2):315-28.
     
  18. Dawson DW, Volpert OV, Gillis P, et al. Pigment epithelium-derived factor: a potent inhibitor of angiogenesis. Science 1999;285(5425):245-8.
     
  19. Dorrell M, Uusitalo-Jarvinen H, Aguilar E, et al. Ocular neovascularization: basic mechanisms and therapeutic advances. Surv Ophthalmol 2007;52 Suppl 1:S3-19.
     
  20. Wang H, Hartnett ME. Regulation of signaling events involved in the pathophysiology of neovascular AMD. Mol Vis 2016;22:189-202.
     
  21. Berber P, Grassmann F, Kiel C, et al. An Eye on Age-Related Macular Degeneration: The Role of MicroRNAs in Disease Pathology. Mol Diagn Ther 2017;21(1):31-43.
     
  22. Augustin HG, Koh GY, Thurston G, et al. Control of vascular morphogenesis and homeostasis through the angiopoietin-Tie system. Nat Rev Mol Cell Biol 2009;10(3):165-77.
     
  23. Sadiq MA, Hanout M, Sarwar S, et al. Platelet-Derived Growth Factor Inhibitors: A Potential Therapeutic Approach for Ocular Neovascularization. Dev Ophthalmol 2016;55:310-6.
     
  24. Kim YW, West XZ, Byzova TV. Inflammation and oxidative stress in angiogenesis and vascular disease. J Mol Med (Berl) 2013;91(3):323-8.
     
  25. Gariano RF. Cellular mechanisms in retinal vascular development. Prog Retin Eye Res 2003;22(3):295-306.
     
  26. Chan-Ling T, McLeod DS, Hughes S, et al. Astrocyte-endothelial cell relationships during human retinal vascular development. Invest Ophthalmol Vis Sci 2004;45(6):2020-32.
     
  27. Lutty GA, Chan-Ling T, Phelps DL, et al. Proceedings of the Third International Symposium on Retinopathy of Prematurity: an update on ROP from the lab to the nursery (November 2003, Anaheim, California). Mol Vis 2006;12:532-80.
     
  28. Hughes S, Yang H, Chan-Ling T. Vascularization of the human fetal retina: roles of vasculogenesis and angiogenesis. Invest Ophthalmol Vis Sci 2000;41(5):1217-28.
     
  29. Chan-Ling T, Gock B, Stone J. The effect of oxygen on vasoformative cell division. Evidence that 'physiological hypoxia' is the stimulus for normal retinal vasculogenesis. Invest Ophthalmol Vis Sci 1995;36(7):1201-14.
     
  30. Cunha-Vaz JG. The blood-retinal barriers system. Basic concepts and clinical evaluation. Exp Eye Res 2004;78(3):715-21.
     
  31. Eichler W, Yafai Y, Wiedemann P, et al. Antineovascular agents in the treatment of eye diseases. Curr Pharm Des 2006;12(21):2645-60.
     
  32. Gao F, Hou H, Liang H, et al. Bone marrow-derived cells in ocular neovascularization: contribution and mechanisms. Angiogenesis 2016;19(2):107-18.
     
  33. Bird AC, Bressler NM, Bressler SB, et al. An international classification and grading system for age-related maculopathy and age-related macular degeneration. The International ARM Epidemiological Study Group. Surv Ophthalmol 1995;39(5):367-74.
     
  34. Hogg RE, Chakravarthy U. Visual function and dysfunction in early and late age-related maculopathy. Prog Retin Eye Res 2006;25(3):249-76.
     
  35. de Jong PT. Age-related macular degeneration. N Engl J Med 2006;355(14):1474-85.
     
  36. Andreoli CM, Miller JW. Anti-vascular endothelial growth factor therapy for ocular neovascular disease. Curr Opin Ophthalmol 2007;18(6):502-8.
     
  37. Shibuya M. Structure and function of VEGF/VEGF-receptor system involved in angiogenesis. Cell Struct Funct 2001;26(1):25-35.
     
  38. Ng EW, Adamis AP. Targeting angiogenesis, the underlying disorder in neovascular age-related macular degeneration. Can J Ophthalmol 2005;40(3):352-68.
     
  39. Semenza GL. Vasculogenesis, angiogenesis, and arteriogenesis: mechanisms of blood vessel formation and remodeling. J Cell Biochem 2007;102(4):840-7.
     
  40. Thornton J, Edwards R, Mitchell P, et al. Smoking and age-related macular degeneration: a review of association. Eye 2005;19(9):935-44.
     
  41. Seddon JM, Francis PJ, George S, et al. Association of CFH Y402H and LOC387715 A69S with progression of age-related macular degeneration. JAMA 2007;297(16):1793-800.
     
  42. Gass JDM. Stereoscopic atlas of macular diseases : diagnosis and treatment. 4th ed. St. Louis; Mosby 1997.
     
  43. Grossniklaus HE, Green WR. Choroidal neovascularization. Am J Ophthalmol 2004;137(3):496-503.
     
  44. Grossniklaus HE, Gass JD. Clinicopathologic correlations of surgically excised type 1 and type 2 submacular choroidal neovascular membranes. Am J Ophthalmol 1998;126(1):59-69.
     
  45. Yannuzzi LA, Freund KB, Takahashi BS. Review of retinal angiomatous proliferation or type 3 neovascularization. Retina 2008;28(3):375-84.
     
  46. Lopez PF, Sippy BD, Lambert HM, et al. Transdifferentiated retinal pigment epithelial cells are immunoreactive for vascular endothelial growth factor in surgically excised age-related macular degeneration-related choroidal neovascular membranes. Invest Ophthalmol Vis Sci 1996;37(5):855-68.
     
  47. Wang J, Ohno-Matsui K, Yoshida T, et al. Amyloid-beta up-regulates complement factor B in retinal pigment epithelial cells through cytokines released from recruited macrophages/microglia: Another mechanism of complement activation in age-related macular degeneration. J Cell Physiol 2009;220(1):119-28.
     
  48. Oh H, Takagi H, Takagi C, et al. The potential angiogenic role of macrophages in the formation of choroidal neovascular membranes. Invest Ophthalmol Vis Sci 1999;40(9):1891-8.
     
  49. Steen B, Sejersen S, Berglin L, et al. Matrix metalloproteinases and metalloproteinase inhibitors in choroidal neovascular membranes. Invest Ophthalmol Vis Sci 1998;39(11):2194-200.
     
  50. Otani A, Takagi H, Oh H, et al. Expressions of angiopoietins and Tie2 in human choroidal neovascular membranes. Invest Ophthalmol Vis Sci 1999;40(9):1912-20.
     
  51. Amin R, Puklin JE, Frank RN. Growth factor localization in choroidal neovascular membranes of age-related macular degeneration. Invest Ophthalmol Vis Sci 1994;35(8):3178-88.
     
  52. Ogata N, Yamamoto C, Miyashiro M, et al. Expression of transforming growth factor-beta mRNA in experimental choroidal neovascularization. Curr Eye Res 1997;16(1):9-18.
     
  53. Hangai M, Murata T, Miyawaki N, et al. Angiopoietin-1 upregulation by vascular endothelial growth factor in human retinal pigment epithelial cells. Invest Ophthalmol Vis Sci 2001;42(7):1617-25.
     
  54. Ishikawa K, Kannan R, Hinton DR. Molecular mechanisms of subretinal fibrosis in age-related macular degeneration. Exp Eye Res 2016;142:19-25.

Development and Progression of AMD

Development and Progression of AMD

English

Author:

Maria Luz Cachulo, MD PhD; José Costa, MD Msc
Coimbra University Hospital - Coimbra, Portugal

 

Updated/reviewed by the authors, July 2017.

Introduction

Age-related macular degeneration (AMD) is the leading cause of blindness in older population in industrialized nations(1,2). It is characterized by ageing changes at the photoreceptors, retinal pigment epithelium (RPE), Bruch’s membrane and choroid(3,4).

The clinical hallmark of the early stage of AMD is drusen, which are yellowish deposits at the level of the RPE. Focal RPE hyperpigmentation and atrophy can also be seen. The advanced stage of non-neovascular AMD is Geographic Atrophy (GA) where areas of atrophy become confluent and cause visual loss. Neovascular AMD (NV-AMD) is also an advanced manifestation of AMD and is characterized by choroidal neovascularization (CNV).

This review summarizes the various biomarkers of AMD and analyses whether or not they may, one day, be exploited to determine risks of disease onset, measure progression of disease or even assess the effects of treatment of AMD.

 

Environmental Risk Factors

See chapter “Modifiable Risk Factors for AMD”.

 

Genetic biomarkers

Genetic predisposition for AMD has been demonstrated by familial aggregation studies and twin studies(5,6). Using genome linkage scan and association studies, multiple potentially causative genes and/or single nucleotide polymorphisms (SNPs) have been identified(7-9).

The most commonly implicated genetic risk loci are located in chromosome 1q25-31, which are related to several complement system regulators, such as complement factor H (CFH) and CFH-related proteins (CFHR1-5)(10). CFH, in particular, is a key regulator of the complement system of innate immunity: it maintains the homeostasis of the complement system and protects bystander host cells and tissues from damage by complement activation(11). At least 6 genetic variants related to AMD development in Caucasians have been identified(7,12-14), while only three of these have been described in Chinese and Japanese populations(15,16).

The most documented risk-conferring single-nucleotide polymorphism results is a tyrosine-to-histidine substitution in 1q26 at position 402 of the CFH protein. This CFHY402H polymorphism increases AMD risk 2 to 4-fold in heterozygotes and 3 to 7-fold in homozygotes(7,17). It causes a significant loss of binding of factor H to heparin, C-reactive protein and to peroxidised lipids on dead cells combined with reduced anti-inflammatory functions by factor H, leading to increased local complement activation(18).

The CFHR genes comprise five plasma proteins (CFHR1-5) that bind to the central complement component C3b. These are located at 1q32 within the RCA (Regulation of Complement Activation) gene cluster(10). A homozygous deletion of CFHR1 and CFHR3 is a relatively common finding in several populations and appears to confer a lower risk of AMD(19). A possible explanation for the protective effect of these deletions might be that CFHR1 and CFHR3 compete with CFH for binding to C3b, thus a deficiency of these proteins increases the efficacy of factor H and its regulatory effect on complement activation(10). The prevalence of this deletion differs between ethnic groups, occurring in 17.3% of Africans, 15.9% of Afro-Americans, 6.8% of Hispanics, 4.7% of Caucasians and 2.2% of Chinese. This finding is consistent with the AMD prevalence rates, which are lower in African-Americans than in Caucasians and Chinese(12).

For an in-depth review of genetic biomarkers of AMD, please refer to the chapter “Genetics of AMD”.

 

Inflammatory biomarkers

Besides components of the complement system(20), several immunological molecules and inflammatory mediators have been identified at the site of AMD lesions(21). In fact, chronic inflammation appears to be a causative factor for the development of AMD by causing endothelial dysfunction and facilitating interactions between modified lipoproteins, monocyte-derived macrophages, T-cells and normal cellular elements of the retinal vasculature(22). Activated macrophages and microglia may cause cellular damage, Bruch’s membrane degradation and angiogenesis by secreting chemokines and cytokines(22). In fact, using electron microscopy or immunohistochemistry methods, macrophages can be found in the area of geographic atrophy phagocytising pigment debris(23) and around CNV in wet AMD(24).

In order to grasp the role of inflammation on the development and progression of AMD, several markers of systemic inflammation have been extensively studied. C-reactive protein (CRP), an acute phase serum protein, is a surrogate for interleukin-6 (IL-6)(25). Classically seen as an inflammatory marker, CRP is now regarded as an independent risk for both cardiovascular and peripheral arterial disease(26). It directly upregulates endothelial cell adhesion molecules and promotes the release of chemoatractant chemokines, which have a negative effect on the retinal microvasculature(22,27).

A meta-analysis of data from more than 41.000 patients found a two-fold increase of late AMD risk in patients with CRP higher than 3 mg/L when compared to those with serum levels <1mg/L(28). The effect of CRP levels on AMD progression might be due to chronic inflammation leading to oxidative damage, endothelial dysfunction, drusen development and the degeneration of Bruch’s membrane(29). Additionally, CRP may also have a direct role in AMD development through its ability to induce complement activation(30).

IL-6 is a marker for systemic inflammation, such as acute pancreatitis and chronic arthritis. It has been implicated in angiogenesis, along with VEGF, in several in vitro and cancer studies(31). Seddon and colleagues found a correlation between the level of IL-6 and chances of AMD progression(32). Animal studies have shown that CNV induction by laser stimulated IL-6 expression in the RPE–choroid complex, and that blockade of IL-6 led to a significant suppression of CNV(33). Interestingly, high aqueous IL-6 levels might to predict resistance to intravitreal bevacizumab in patients with wet AMD(34).

Fibrinogen is also an established biomarker of acute and chronic inflammation(35). Lip and colleagues found elevated levels of plasma fibrinogen in AMD cases compared with controls(36), and a case–control analysis from the large Blue Mountains Eye Study in Australia detected significantly elevated plasma fibrinogen levels in AMD patients compared with controls (p < 0.05)(37). In another study using patients recruited from the Muenster Aging and Retina Study population in Münster (Germany), plasma fibrinogen levels were found to be elevated as the degree of AMD severity increased(38). In fact, fibrinogen plasma levels appear to be particularly increased in patients with exudative AMD(39).

 

Multimodal imaging evaluation of AMD progression

1) Fundoscopy and Colour Fundus Photographs

Despite the increasingly sophisticated imaging techniques available, fundoscopy and colour fundus photographs (CFP) remain invaluable tools for the assessment of the severity and risk of progression of patients with AMD.

Drusen size, area and location are important predictors of AMD progression(40,41): large soft drusens with ill-defined borders and extensive drusen area are the most important risk factors for the progression to advanced AMD identifiable on CFP(42,43). Yet, earlier macular changes can provide valuable information about the risk of AMD progression. This has been extensively demonstrated in population-based observational studies and provides the basis for several AMD severity classifications.

The Age-Related Eye Disease Study (AREDS) collected stereoscopic CFP of more than 3000 AREDS over 5 years of follow-up and identified large drusen and macular pigmentary abnormalities as features that, cumulatively, increase the risk of progression to advanced AMD(43). A simplified risk scale was developed based on these findings (Table 1). It considers the presence of at least 1 large drusen (diameter greater than or equal to that of a large vein at the disc margin) and the presence of any pigment abnormality as 1 point each, and sums their presence across both eyes when both are free of advanced AMD. One point is assigned to patients who have no large drusen in either eye but intermediate-sized drusen (diameter ≥ one half that of a large drusen) in both eyes. The 5-year risk of advanced AMD using this scale increases as more risk factors are present: 0 factors, 0.5%; 1 factor, 3%; 2 factors, 12%; 3 factors, 25%; and 4 factors, 50%. This simplified scale is particularly useful in clinical practice because it allows fast risk stratification and is useful when discussing with patients their risk of progression to vision-threatening AMD.

Table 1 - Risk Factors for AMD (from AREDS)

 

In recent years, several studies have studied the significance of reticular pseudodrusen (RPD). They were first described in 1990 as a yellowish interlacing network most commonly located on the superotemporal macula(44). RPD are difficult to identify in clinical fundoscopic examination and are poorly visualized with regular CFP, so they are best seen with multimodal imaging(45). In two population-based studies, the overall prevalence of RPD was 0.7%-1.95% for the general population(45,46). In AMD patients, the prevalence of RPD is significantly higher: they are present in up to 58% of patients(47,48). This finding is particularly relevant due to the fact that RPD are associated with greater risk of progression to advanced AMD: the presence of RPD at baseline in fellow eyes of patients with unilateral neovascularization is a significant predictor of progression to advanced AMD (OR =2.5) over a 5-year period, particularly the wet form(48). Similarly, depending on the follow-up, up to 14% of patients with AMD ultimately develop GA(47-49). Unfortunately, RPD presence remains an underreported and underresearched retinal phenotype, but the advent of new imaging modalities will certainly improve our knowledge about this phenotype(48).

 

2) Fundus Autofluorescence

Fundus autofluorescence (FAF) is a non-invasive method that supplies additional information to that obtained using CFP and fluorescence angiography(50). It provides an indirect evaluation of RPE function, an important component of the physiopathology of AMD(51). In GA, the atrophic areas appear as hypoautofluorescent patches surrounded by non-atrophic retina of variable autofluorescence(52). The FAF features of these borders of GA lesions appear to be particularly important to predict the progression of pre-existing GA, as variation in GA growth rates are dependent on the specific phenotype of FAF at baseline(53-55).

Additionally, FAF has a great sensitivity for the detection of RPD: areas covered with RPD usually show a reticular FAF pattern with small areas of decreased FAF surrounded by normal FAF(55).

For further data on FAF, please refer to the chapter “Fundus autofluorescence in age-related macular degeneration”.

 

3) Optical Coherence Tomography

Due to the quantitative and highly reproducible data that it provides, optical coherence tomography (OCT) has quickly become an invaluable tool to evaluate AMD progression and treatment response.

Spectral domain OCT (SD-OCT), in particular, enables automated measurement of drusen area and volume(41), important predictors of AMD progression(56), and has been validated as a method of following GA development and progression(57). In addition, RPD can be easily seen on SD-OCT B-scans as triangular shaped structures located above the RPE line(58).

Detailed information about the role of OCT in the management of AMD can be found on the chapter “Optical Coherence Tomography in Age-related Macular Degeneration”.

 

Conclusion

AMD is a complex disease caused by a combination of genetic predisposition and environmental factors and the interplay between these remains a mystery. A better understanding of the involved pathophysiologic process or the identification of biomarkers for the conversion would enhance our ability to diagnose and treat the wet AMD, to develop better therapies and, eventually, to prevent vision loss associated with the disease.

 

>> References

References - Development and Progression of AMD

References - Development and Progression of AMD

English
  1. Miller JW. Age-related macular degeneration revisited--piecing the puzzle: the LXIX Edward Jackson memorial lecture. Am J Ophthalmol 2013;155:1-35 e13.
     
  2. Wong WL, Su X, Li X, et al. Global prevalence of age-related macular degeneration and disease burden projection for 2020 and 2040: a systematic review and meta-analysis. Lancet Glob Health 2014;2:e106-16.
     
  3. Green WR, Enger C. Age-related macular degeneration histopathologic studies: the 1992 Lorenz E. Zimmerman Lecture. 1992. Retina 2005;25:1519-35.
     
  4. Bressler NM, Silva JC, Bressler SB, Fine SL, Green WR. Clinicopathologic correlation of drusen and retinal pigment epithelial abnormalities in age-related macular degeneration. 1994. Retina 2005;25:130-42.
     
  5. Seddon JM, Francis PJ, George S, Schultz DW, Rosner B, Klein ML. Association of CFH Y402H and LOC387715 A69S with progression of age-related macular degeneration. JAMA 2007;297:1793-800.
     
  6. Schwartz SG, Hampton BM, Kovach JL, Brantley MA, Jr. Genetics and age-related macular degeneration: a practical review for the clinician. Clin Ophthalmol 2016;10:1229-35.
     
  7. Klein RJ, Zeiss C, Chew EY, et al. Complement factor H polymorphism in age-related macular degeneration. Science 2005;308:385-9.
     
  8. Fisher SA, Abecasis GR, Yashar BM, et al. Meta-analysis of genome scans of age-related macular degeneration. Hum Mol Genet 2005;14:2257-64.
     
  9. Majewski J, Schultz DW, Weleber RG, et al. Age-related macular degeneration--a genome scan in extended families. Am J Hum Genet 2003;73:540-50.
     
  10. Skerka C, Chen Q, Fremeaux-Bacchi V, Roumenina LT. Complement factor H related proteins (CFHRs). Mol Immunol 2013;56:170-80.
     
  11. Rodriguez de Cordoba S, Esparza-Gordillo J, Goicoechea de Jorge E, Lopez-Trascasa M, Sanchez-Corral P. The human complement factor H: functional roles, genetic variations and disease associations. Mol Immunol 2004;41:355-67.
     
  12. Hageman GS, Anderson DH, Johnson LV, et al. A common haplotype in the complement regulatory gene factor H (HF1/CFH) predisposes individuals to age-related macular degeneration. Proc Natl Acad Sci U S A 2005;102:7227-32.
     
  13. Haines JL, Hauser MA, Schmidt S, et al. Complement factor H variant increases the risk of age-related macular degeneration. Science 2005;308:419-21.
     
  14. Tuo J, Ning B, Bojanowski CM, et al. Synergic effect of polymorphisms in ERCC6 5' flanking region and complement factor H on age-related macular degeneration predisposition. Proc Natl Acad Sci U S A 2006;103:9256-61.
     
  15. Okamoto H, Umeda S, Obazawa M, et al. Complement factor H polymorphisms in Japanese population with age-related macular degeneration. Mol Vis 2006;12:156-8.
     
  16. Chen LJ, Liu DT, Tam PO, et al. Association of complement factor H polymorphisms with exudative age-related macular degeneration. Mol Vis 2006;12:1536-42.
     
  17. Wegscheider BJ, Weger M, Renner W, et al. Association of complement factor H Y402H gene polymorphism with different subtypes of exudative age-related macular degeneration. Ophthalmology 2007;114:738-42.
     
  18. Weismann D, Hartvigsen K, Lauer N, et al. Complement factor H binds malondialdehyde epitopes and protects from oxidative stress. Nature 2011;478:76-81.
     
  19. Hughes AE, Orr N, Esfandiary H, Diaz-Torres M, Goodship T, Chakravarthy U. A common CFH haplotype, with deletion of CFHR1 and CFHR3, is associated with lower risk of age-related macular degeneration. Nat Genet 2006;38:1173-7.
     
  20. Wang L, Clark ME, Crossman DK, et al. Abundant lipid and protein components of drusen. PLoS One 2010;5:e10329.
     
  21. Hageman GS, Luthert PJ, Victor Chong NH, Johnson LV, Anderson DH, Mullins RF. An integrated hypothesis that considers drusen as biomarkers of immune-mediated processes at the RPE-Bruch's membrane interface in aging and age-related macular degeneration. Prog Retin Eye Res 2001;20:705-32.
     
  22. Colak E, Majkic-Singh N, Zoric L, Radosavljevic A, Kosanovic-Jakovic N. The role of CRP and inflammation in the pathogenesis of age-related macular degeneration. Biochem Med (Zagreb) 2012;22:39-48.
     
  23. Roth F, Bindewald A, Holz FG. Keypathophysiologic pathways in age-related macular disease. Graefes Arch Clin Exp Ophthalmol 2004;242:710-6.
     
  24. Grossniklaus HE, Cingle KA, Yoon YD, Ketkar N, L'Hernault N, Brown S. Correlation of histologic 2-dimensional reconstruction and confocal scanning laser microscopic imaging of choroidal neovascularization in eyes with age-related maculopathy. Arch Ophthalmol 2000;118:625-9.
     
  25. Ridker PM. From C-Reactive Protein to Interleukin-6 to Interleukin-1: Moving Upstream To Identify Novel Targets for Atheroprotection. Circ Res 2016;118:145-56.
     
  26. Lange LA, Carlson CS, Hindorff LA, et al. Association of polymorphisms in the CRP gene with circulating C-reactive protein levels and cardiovascular events. JAMA 2006;296:2703-11.
     
  27. Yim-Lui Cheung C, Wong TY, Lamoureux EL, et al. C-reactive protein and retinal microvascular caliber in a multiethnic asian population. Am J Epidemiol 2010;171:206-13.
     
  28. Hong T, Tan AG, Mitchell P, Wang JJ. A review and meta-analysis of the association between C-reactive protein and age-related macular degeneration. Surv Ophthalmol 2011;56:184-94.
     
  29. Schaumberg DA, Christen WG, Kozlowski P, Miller DT, Ridker PM, Zee RY. A prospective assessment of the Y402H variant in complement factor H, genetic variants in C-reactive protein, and risk of age-related macular degeneration. Invest Ophthalmol Vis Sci 2006;47:2336-40.
     
  30. Scholl HP, Charbel Issa P, Walier M, et al. Systemic complement activation in age-related macular degeneration. PLoS One 2008;3:e2593.
     
  31. Tzeng HE, Tsai CH, Chang ZL, et al. Interleukin-6 induces vascular endothelial growth factor expression and promotes angiogenesis through apoptosis signal-regulating kinase 1 in human osteosarcoma. Biochem Pharmacol 2013;85:531-40.
     
  32. Seddon JM, George S, Rosner B, Rifai N. Progression of age-related macular degeneration: prospective assessment of C-reactive protein, interleukin 6, and other cardiovascular biomarkers. Arch Ophthalmol 2005;123:774-82.
     
  33. zumi-Nagai K, Nagai N, Ozawa Y, et al. Interleukin-6 receptor-mediated activation of signal transducer and activator of transcription-3 (STAT3) promotes choroidal neovascularization. Am J Pathol 2007;170:2149-58.
     
  34. Chalam KV, Grover S, Sambhav K, Balaiya S, Murthy RK. Aqueous interleukin-6 levels are superior to vascular endothelial growth factor in predicting therapeutic response to bevacizumab in age-related macular degeneration. J Ophthalmol 2014;2014:502174.
     
  35. Ridker PM. Clinical application of C-reactive protein for cardiovascular disease detection and prevention. Circulation 2003;107:363-9.
     
  36. Lip PL, Blann AD, Hope-Ross M, Gibson JM, Lip GY. Age-related macular degeneration is associated with increased vascular endothelial growth factor, hemorheology and endothelial dysfunction. Ophthalmology 2001;108:705-10.
     
  37. Smith W, Mitchell P, Leeder SR, Wang JJ. Plasma fibrinogen levels, other cardiovascular risk factors, and age-related maculopathy: the Blue Mountains Eye Study. Arch Ophthalmol 1998;116:583-7.
     
  38. Klein R, Klein BE, Knudtson MD, Wong TY, Shankar A, Tsai MY. Systemic markers of inflammation, endothelial dysfunction, and age-related maculopathy. Am J Ophthalmol 2005;140:35-44.
     
  39. Colak E, Ignjatovic S, Radosavljevic A, Zoric L. The association of enzymatic and non-enzymatic antioxidant defense parameters with inflammatory markers in patients with exudative form of age-related macular degeneration. J Clin Biochem Nutr 2017;60:100-7.
     
  40. Joachim N, Mitchell P, Kifley A, Rochtchina E, Hong T, Wang JJ. Incidence and progression of geographic atrophy: observations from a population-based cohort. Ophthalmology 2013;120:2042-50.
     
  41. Nathoo NA, Or C, Young M, et al. Optical coherence tomography-based measurement of drusen load predicts development of advanced age-related macular degeneration. Am J Ophthalmol 2014;158:757-61 e1.
     
  42. Bressler NM, Bressler SB, Seddon JM, Gragoudas ES, Jacobson LP. Drusen characteristics in patients with exudative versus non-exudative age-related macular degeneration. Retina 1988;8:109-14.
     
  43. Ferris FL, Davis MD, Clemons TE, et al. A simplified severity scale for age-related macular degeneration: AREDS Report No. 18. Arch Ophthalmol 2005;123:1570-4.
     
  44. Mimoun G, Soubrane G, Coscas G. [Macular drusen]. J Fr Ophtalmol 1990;13:511-30.
     
  45. Klein R, Meuer SM, Knudtson MD, Iyengar SK, Klein BE. The epidemiology of retinal reticular drusen. Am J Ophthalmol 2008;145:317-26.
     
  46. Joachim N, Mitchell P, Rochtchina E, Tan AG, Wang JJ. Incidence and progression of reticular drusen in age-related macular degeneration: findings from an older Australian cohort. Ophthalmology 2014;121:917-25.
     
  47. Finger RP, Wu Z, Luu CD, et al. Reticular pseudodrusen: a risk factor for geographic atrophy in fellow eyes of individuals with unilateral choroidal neovascularization. Ophthalmology 2014;121:1252-6.
     
  48. Gil JQ, Marques JP, Hogg R, et al. Clinical features and long-term progression of reticular pseudodrusen in age-related macular degeneration: findings from a multicenter cohort. Eye (Lond) 2017;31:364-71.
     
  49. Pumariega NM, Smith RT, Sohrab MA, Letien V, Souied EH. A prospective study of reticular macular disease. Ophthalmology 2011;118:1619-25.
     
  50. Solbach U, Keilhauer C, Knabben H, Wolf S. Imaging of retinal autofluorescence in patients with age-related macular degeneration. Retina 1997;17:385-9.
     
  51. Boulton M, Dayhaw-Barker P. The role of the retinal pigment epithelium: topographical variation and ageing changes. Eye (Lond) 2001;15:384-9.
     
  52. von Ruckmann A, Fitzke FW, Bird AC. Fundus autofluorescence in age-related macular disease imaged with a laser scanning ophthalmoscope. Invest Ophthalmol Vis Sci 1997;38:478-86.
     
  53. Schmitz-Valckenberg S, Bindewald-Wittich A, Dolar-Szczasny J, et al. Correlation between the area of increased autofluorescence surrounding geographic atrophy and disease progression in patients with AMD. Invest Ophthalmol Vis Sci 2006;47:2648-54.
     
  54. Jeong YJ, Hong IH, Chung JK, Kim KL, Kim HK, Park SP. Predictors for the progression of geographic atrophy in patients with age-related macular degeneration: fundus autofluorescence study with modified fundus camera. Eye (Lond) 2014;28:209-18.
     
  55. Holz FG, Bindewald-Wittich A, Fleckenstein M, et al. Progression of geographic atrophy and impact of fundus autofluorescence patterns in age-related macular degeneration. Am J Ophthalmol 2007;143:463-72.
     
  56. Complications of Age-related Macular Degeneration Prevention Trial Research G. Risk factors for choroidal neovascularization and geographic atrophy in the complications of age-related macular degeneration prevention trial. Ophthalmology 2008;115:1474-9, 9 e1-6.
     
  57. Yehoshua Z, Rosenfeld PJ, Gregori G, et al. Progression of geographic atrophy in age-related macular degeneration imaged with spectral domain optical coherence tomography. Ophthalmology 2011;118:679-86.
     
  58. Zweifel SA, Spaide RF, Curcio CA, Malek G, Imamura Y. Reticular pseudodrusen are subretinal drusenoid deposits. Ophthalmology 2010;117:303-12 e1.

Fluorescein Angiography

Fluorescein Angiography

English

Authors:

Luis Arias, MD
Hospital Universitari de Bellvitge - University of Barcelona. Spain

 

Jordi Monés, MD
Institut de la Màcula i de la Retina - Centro Médico Teknon - Barcelona. Spain

 

Last revision: October 2011 by Luis Arias

Introduction

Fluorescein angiography (FA) was introduced in ophthalmology by Novotny and Alvis in the sixties of the last century.

They took serial fundus photographs after intravenous injection of sodium fluorescein to study the retinal and choroidal circulation(1).

Initially, they used this technique in diabetic and hypertensive patients and after, the technique was used in age-related macular degeneration (AMD).

Although the clinical diagnosis of AMD can be established based on patient’s history and fundus examination, FA is the most important ancillary test for classifying the disease in its different subtypes, especially in its wet form.

Nowadays, optical coherence tomography (OCT) is being more used than FA for monitoring the response to treatment, although FA is still very useful in some cases. Sodium fluorescein is a small molecule, with a molecular weight of 376.27 daltons, and it is highly soluble in water.

It is stimulated by light in the range of 465-490 nm and then it enters into a higher energy state.

The molecule emits longer wavelength fluorescence, between 520 and 530 nm, as it decays to a lower energy state.

In clinical application, two filters are used.

An excitation filter allows the passage of blue light, which stimulates the fluorescein in the eye, which emits yellow-green light.

In addition, a barrier filter is used to block some reflected blue light, allowing only the yellow-green light to pass through.

This resultant fluorescence is recorded by a camera as an image(2).

Sodium fluorescein diffuses through the fenestrated vessels of the choriocapillaris, but does not cross the internal and the external blood-retinal barriers.

Thus, any condition that compromises these barriers, obstructs blood flow, or changes the normal pigmentation of the retinal pigment epithelium (RPE) can cause abnormalities on FA.

The dye is removed from the vascular compartment by the kidney.

It is relatively inert, making intravenous injection safe and severe adverse reactions rare. Nevertheless, the patient should be properly informed of the potential risks of FA injection(3).

 

Interpretation

Several FA patterns can be observed in AMD patients.

They can be classified at those leading to decreased fluorescence (hypofluorescence) or increased fluorescence (hyperfluorescence).

Hypofluorescence either represents blocked fluorescence or a vascular filling defect (Table 1).

Hypofluorescence
 
Blocked
  • Intraretinal or subretinal hemorrhage/exudate
  • Sub-RPE hemorrhage
  • Pigment proliferation
  • Pigment epithelial clumping (RPE rip)
Vascular filling defect
  • Choroidal vascular atrophy
  • Retinal capillary occlusion

Table 1 - Causes of hypofluorescence.

 

Hyperfluorescence can be the result of loss of the normal barrier to background choroidal fluorescence known as transmitted fluorescence.

A second reason for hyperfluorescence can arise from extravascular accumulation of the dye or from leakage from abnormal vessels (Table 2)(4).

Hyperfluorescence
 
Transmitted fluorescence
  • RPE atrophy
  • RPE rip
  • Hard, basal laminar drusen
Extravascular fluorescence
  • Serous pigment epithelial detachment
  • Soft drusen
  • Disciform scar
  • Loculated fluid
  • Cystoid macular edema
Abnormal vessels
  • Choroidal neovascularization
  • Retinal angiomatous proliferation

Table 2 - Causes of hypofluorescence.

 

By convention, leakage of fluorescein into a space is referred to as pooling, while leakage into a tissue is called staining(5).

 

Angiographic patterns in AMD

Drusen and RPE abnormalities

The majority of patients with AMD have drusen and RPE abnormalities with no significant visual loss.

FA is not usually indicated in these cases unless we suspect the presence of choroidal neovascularization (CNV).

Several types of drusen can be identified. Hard drusen are small (<63 µm), round, well-defined, yellowish deposits that correspond to accumulation of hyaline material in the inner and outer collagenous zones of Bruch’s membrane.

On FA, they appear hyperfluorescent as transmission defects due to overlying RPE thinning(6).

On occasion there may be a myriad of small drusen, termed cuticular or basal laminar drusen, which appear as a “starry sky” on FA (Figure 1).

Figure 1 - Cuticular drusen with the typical pattern of “starry sky”

 

Soft drusen are larger (>63 µm) with poorly defined borders and they tend to coalesce and become confluent.

Their angiographic appearance depends on the thinning of the overlying RPE, the histochemical composition and the age of the patient.

They are hyperfluorescent with phospholipid accumulation and in younger patients(7).

Soft drusen represent localized detachments of the RPE.

It is very usual to find both hard and soft drusen in the same eye of a patient (Figure 2).

Figure 2 - Coexistence of hard and soft drusen in the same eye of a patient with AMD.

 

The confluence of soft drusen can produce a drusenoid pigment epithelial detachment (PED), which shows hyperfluorescence and dye pooling without leakage beyond its margin with typical areas of focal hyperpigmentation (Figure 3).

Figure 3 - Drusenoid pigment epithelial detachment.

 

In addition to drusen we can find RPE abnormalities, namely hyperpigmentation.

Focal hyperpigmentation is a risk factor for the development of choroidal neovascularization (CNV) and angiographically appears as a blocked fluorescence (Figure 4).

Figure 4 - RPE abnormalities with focal hyperpigmentation.

 

Histopathologically it is characterized by focal RPE hypertrophy and pigment migration into the subretinal space.

It also displays focal hyperautofluorescence suggesting that these cells contain lipofuscin(8).

 

Atrophic AMD

Atrophy can occur in sharply defined areas of severe atrophy, known as geographic atrophy (GA), or in less well-defined, more granular regions of less severe atrophy, known as non-GA.

Both forms share the feature of RPE loss, more extensive and with associated atrophy of the overlying retina and underlying choriocapillaris in GA.

The angiographic appearance depends on the remaining pigment within the RPE and choriocapillaris vessels.

Non-GA shows mottled early hyperfluorescence, which fades late consistent with window defect (Figure 5).

Figure 5 - Non-geographic atrophy.

 

GA typically shows late well-defined hyperfluorescence from staining of the exposed deep choroid and sclera(9).

In these cases, visual acuity depends on the foveal involvement (Figure 6).

Figure 6 - Geographic atrophy.

 

In advanced cases, larger choroidal vessels show a sclerotic appearance (Figure 7).

Figure 7 - Severe atrophic AMD with sclerotic appearance of larger choroidal vessels.

 

Classic CNV

Classic CNV is characterized by well-demarcated hyperfluorescence in early phases on FA and late leakage that obscures the boundaries of the lesion (Figure. 8).

Figure 8 - Classic CNV.

 

As defined by Donald Gass, classic CNV lies between the neurosensory retina and the RPE (type II CNV)(10).

Angiographic classic CNV appears as a lacy or bicycle-wheel pattern.

Depending on its location, it can be classified as extrafoveal (>200 microns from the foveal center) (Figure 9), juxtafoveal (1-199 microns from the foveal center) (Figure 10) or subfoveal (involving the foveal center) (Figure 11).

Figure 9 - Extrafoveal classic CNV.

 

Figure 10 - Juxtafoveal classic CNV.

 

Figure 11 - Subfoveal classic CNV.

 

Sometimes, a feeder vessel can be localized (Figure 12).

Another typical feature is the presence of a hyperpigmented rim, hypofluorescent on FA, surrounding the CNV (Figure 12).

Figure 12 - Classic CNV with feeder vessel.

 

On occasion, classic CNV can be associated to loculated fluid (Figure 13).

Figure 13 - Classic CNV with loculated fluid.

 

In loculated fluid, dye pooling is well-demarcated in a confined space of a localized sensory retinal detachment or within intraretinal cystic spaces.

It was a common finding in patients with new subfoveal CNV in the Macular Photocoagulation Study (MPS) and may confuse the treating physician as to the boundary of the lesion(11).

Depending on their sizes, classic CNV can be classified as small (Figure 9-11) or medium (Figure 14) or large (Fig. 15).

Figure 14 - Medium size classic CNV.

 

Figure 15 - Large classic CNV.

 

Importantly, larger classic CNV are associated to a poorer visual prognosis since they represent long-term duration of the pathological disorder.

Classic CNV is an emergency and it requires early treatment to halt the progression of the disease.

Without treatment, CNV tend to enlarge and irreversible fibrosis appears (Figure 16).

Figure 16 - Old classic CNV with fibrosis.

 

In the last decade of the last century, the advent of photodynamic therapy (PDT) with verteporfin promoted a classification of the lesions depending on the percentage of classic CNV.

Thus, predominantly classic lesions were defined as having 50% or more of the total lesion size comprised of classic CNV (Figure 17).

Figure 17 - Predominantly classic CNV.

 

On the other hand, minimally classic lesions were characterized by classic CNV occupying less than 50% of the total lesion size (Figure 18)(12).

Figure 18 - Minimally classic CNV.

 

The best results with PDT in wet AMD patients were obtained in the treatment of predominantly classic lesions.

Nowadays, in the antiangiogenic therapy era, this classification has lost popularity among ophthalmologists since lesion composition does not seem to be as relevant as it was with PDT.

Lesion components associated with neovascular AMD that can obscure the boundaries of CNV include changes that block fluorescence, such as blood, fibrous tissue, RPE hyperplasia, or RPE redundancy (from an RPE tear).

Likewise, CNV can be obscured by greater fluorescence from staining or pooling.

 

Occult CNV

Occult CNV has been categorized as fibrovascular PED or late leakage of undetermined source(13).

Fibrovascular PED (type I occult CNV) is defined as an irregular elevation of the RPE associated with stippled hyperfluorescence apparent 1 to 2 minutes after fluorescein injection and ill-defined staining or leakage in the late frames (Figure 19-20).

Figure 19 - Fibrovascular PED (type I occult CNV).

 

Figure 20 - Fibrovascular plaque

 

It differs from classic CNV in that the early hyperfluorescence is not as bright and the boundaries usually are indeterminate.

Late leakage of undetermined source (type II occult CNV) lacks a discernible, well-demarcated area of leakage in the early angiographic frames.

Speckled hyperfluorescence with no visible source becomes apparent 2 to 5 minutes after dye injection (Figure 21).

Figure 21 - Late leakage of undetermined source (type II occult CNV).

 

Serous PED

Although serous PEDs can occur in the context of non-neovascular AMD, most of them are related to CNV.

On fundus biomicroscopy, a serous PED appears as a round or oval translucent elevation of the RPE.

On FA, it is characterized by progressive and uniform hyperfluorescence from early frames with intense pooling of fluorescein in late phases (Figure 22).

Figure 22 - Serous PED.

 

PEDs with a notch usually have occult CNV in the notch (Figure 23).

Figure 23 - Serous PED associated to occult CNV.

 

The association of occult CNV and a serous PED is frequently termed “vascularized PED”(14, 15).

On occasion, a serous PED is associated to classic CNV (Figure 24).

Figure 24 - Serous PED associated to classic CNV.

 

Retinal angiomatous proliferation

Retinal angiomatous proliferation (RAP) has been described and classified by Yannuzzi et al.(16).

In RAP, the vasogenic process originates in the retina and begins as intraretinal neovascularization (stage I), which progresses to subretinal neovascularization (stage II) and finally to CNV (stage III).

In some cases it is possible to find a retinal-retinal anastomosis.

Angiographically, early lesions (stage I) show a focal area of intraretinal hyperfluorescence with indistinct borders corresponding to the intraretinal neovascularization and surrounding intraretinal edema (Figure 25).

Figure 25 - RAP (stage I).

 

Sometimes, these early lesions can mimic the appearance of a classic CNV. Later stages of RAP are often classified as minimally classic or occult CNV.

In stage II, it is very characteristic to find a serous PED with occult CNV associated to overlying cystoid macular edema (Figure 26).

Figure 26 - RAP (stage II).

 

Indocyanine green (ICG) angiography is often more useful than FA for the diagnosis and evaluation of RAP lesions.

 

Polypoidal Choroidal Vasculopathy

In polypoidal choroidal vasculopathy (PCV), the primary abnormality involves the choroidal circulation, and the characteristic lesion is an inner choroidal vascular network of vessels ending in an aneurismal bulge.

Clinically, PCV is associated with multiple, recurrent, serosanguineous detachments of the RPE and neurosensory retina secondary to leakage and bleeding from the choroidal vascular lesion(17)(Figure 27).

Figure 27 - Polypoidal choroidal vasculopathy.

 

Although FA can sometimes confirm the diagnosis of PCV, ICG angiography is the choice for imaging this entity.

 

RPE tears

Although RPE tears can occur spontaneously, it is not uncommon for them to occur after treatment with thermal laser, PDT or antiangiogenic therapy.

RPE tears are commonly related to PEDs, although they have been described in classic lesions too(18).

The detached monolayer of RPE scrolls toward the CNV, leaving a denuded area of choroid exposed.

On FA, the denuded area becomes hyperfluorescent and the scrolled RPE is dark and blocks the underlying fluorescence (Figure 28).

Figure 28 - RPE tear.

 

Hemorrhagic AMD

FA is not very useful in hemorrhagic forms of macular degeneration since blood blocks the underlying fluorescence (Figure 29).

Figure 29 - Hemorrhagic AMD.

 

ICG angiography can detect the presence of occult CNV.

 

Disciform scar

A disciform scar is the end-stage manifestation of untreated CNV, namely formed by fibroblasts and inflammatory cells.

Angiographically, it typically shows late staining (Figure 30).

Figure 30 - Disciform scar.

 

FA for monitoring AMD treatment

In the era of PDT with verteporfin, FA was the gold standard for monitoring the response to treatment(19).

Nowadays, with antiangiogenic therapy, OCT scanning has replaced FA for this purpose since it is highly effective to detect lesion activity and it is a non-invasive procedure(20).

However, in some cases FA is still very useful in the evaluation of treated patients.

 

>> References

References - Fluorescein Angiography

References - Fluorescein Angiography

English
  1. Novotny HR, Alvis DL. A method of photographing fluorescence in circulating blood in the human retina. Circulation 1961; 24: 82-86.
     
  2. Wolfe DR. Fluorescein angiography: basic science and engineering. Ophthalmology 1986; 93: 1617-1620.
     
  3. Yannuzzi LA, Rohrer KT, Tindel LJ, et al. Fluorescein angiography complication survey. Ophthalmology 1986; 93: 611-617.
     
  4. Jumper JM, Fu AD, McDonald HR, Johnson RN, Ai E. Fluorescein angiography. In : Alfaro DV, Liggett PE, Mieler WF, Quiroz-Mercado H, Jager RD, Tano Y, eds. Age-related macular degeneration. Lippincott Williams&Wilkins; Philadelphia 2006: 86-100.
     
  5. Spaide RF. Fundus angiography. In: Holz FG, Pauleikhoff D, Spaide RF, Bird AC, eds. Age-related macular degeneration. Springer; Heidelberg 2004: 87-107.
     
  6. Bressler NM, Bressler SB, Fine SL. Age-related macular degeneration. Surv Ophthalmol 1988; 32: 375-413.
     
  7. Pauleikhoff D, Zuels S, Sheraidah GS, et al. Correlation between biochemical composition and fluorescein binding of deposits in Bruch’s membrane. Ophthalmology 1992; 99: 1548-1553.
     
  8. Spaide RF. Fundus autofluorescence and age-related macular degeneration. Ophthalmology 2003; 110: 392-399.
     
  9. Pieramici DJ, Bressler SB. Fluorescein angiography. In: Berger JW, Fine SL, Maguire MG, eds. Age-related macular degeneration. Mosby; St. Louis: 219-236.
     
  10. Gass JDM. Pathogenesis of disciform detachment of the neuroepithelium. Am J Ophthalmol 1967; 63: 567-659.
     
  11. Bressler NM, Bressler SB, Alexander J, et al. Loculated fluid. A previously undescribed fluorescein angiographic finding in choroidal neovascularization associated with macular degeneration. Macular Photocoagulation Study Reading Center. Arch Ophthalmol 1991; 109: 211-215.
     
  12. Barbazetto I, Burdan A, Bressler NM, et al. Photodynamic therapy of subfoveal choroidal neovascularization with verteporfin: fluorescein angiographic guidelines for evaluation and treatment – TAP and VIP report No. 2. Arch Ophthalmol 2003; 121: 1253-1268.
     
  13. Macular Photocoagulation Study Group. Subfoveal neovascular lesions in age-related macular degeneration. Guidelines for evaluation and treatment in the macular photocoagulation study. Arch Ophthalmol 1991; 109: 1242-1257.
     
  14. Gass JDM. Serous retinal pigment detachment with a notch. Retina 1984; 4: 205-220.
     
  15. Coscas G, Koenig F, Soubrane G. The pretear characteristics of pigment epithelial detachments. A study of 40 eyes. Arch Ophthalmol 1990; 108: 1687-1693.
     
  16. Yannuzzi LA, Negrao S, Iida T, et al. Retinal angiomatous proliferation in age-related macular degeneration. Retina 2001; 21: 416-434.
     
  17. Yannuzzi LA, Sorenson J, Spaide RF, et al. Idiopathic polypoidal choroidal vasculopathy. Retina 1990 ; 10 : 1-8.
     
  18. Arias L, Caminal J, Rubio M, et al. Retinal pigment epithelial tears after intravitreal bevacizumab injection for predominantly classic choroidal neovascularization. Eur J Ophthalmol 2007; 17: 992-995.
     
  19. Verteporfin Roundtable Participants. Guidelines for using verteporfin (Visudyne) in photodynamic therapy for choroidal neovascularization due to age-related macular degeneration and other causes: update. Retina 2005; 25: 119-134.
     
  20. Fung AE, Lalwani GA, Rosenfeld PJ, et al. An optical coherence tomography-guided, variable dosing regimen with intravitreal ranibizumab (Lucentis) for neovascular age-related macular degeneration. Am J Ophthalmol 2007; 143: 566-83.

Diagnostic usefulness of indocyanine green angiography (ICGA) in age-related macular degeneration (AMD)

Diagnostic usefulness of indocyanine green angiography (ICGA) in age-related macular degeneration (AMD)

English

Authors:

Maribel Fernández, MD, PhD1,3,4

María Gil, MD1,3

Felipe Gonzalez2

Francisco Gómez-Ulla, MD, PhD1,3,4

1 Instituto Oftalmológico Gómez-Ulla. Santiago de Compostela. Spain.

2 Optometrist. Unit of Optometry. Instituto Oftalmológico Gómez-Ulla. Santiago de Compostela. Spain.

3 Department of Ophthalmology, Complejo Hospitalario Universitario de Santiago, Santiago de Compostela. Spain.

4 University of Santiago de Compostela. Santiago de Compostela. Spain.

 

Updated/reviewed by the authors, July 2017.

Age-related macular degeneration (AMD), is a leading cause of severe central vision loss and legal blindness among people over 55 years of age in the industrialized countries (1-4), and thus represents a major worldwide sociosanitary problem.

Exudative (neovascular) AMD is a less common form of the disease, but is responsible for approximately 80-90% of all cases of severe vision loss (5-7). Choroidal neovascularization (CNV) in its different forms of presentation is the main cause of complications and vision loss in these patients.

The classification of CNV based on its angiographic characteristics was regarded as very important in the past, when treatment fundamentally consisted on thermal laser ablation or photodynamic therapy (PDT). However, at present, the development of spectral domain optical coherence tomography (SD-OCT) and the use of intravitreal injections of antiangiogenic drugs have led to a genuine revolution in the diagnosis, follow-up and prognosis of these patients.

Anti-vascular endothelial growth factor (VEGF) therapy is the treatment of choice in patients with exudative AMD, though there are cases where other treatments may be used alone or in combination in order to increase efficiency.

OCT has displaced other diagnostic techniques used for patient control in the frequent follow-up visits of the disease. Today, SD-OCT and noninvasive fundus autofluorescence imaging can provide additional anatomical information beyond the fluorescein angiography (FA), including visualization of the layers of the retina and choroid involved with the neovascular membrane. Moreover, swept source OCT (SS-OCT) is a new OCT technology, which have enabled a even faster acquisition and a reduction in the signal-to-noise ratio. The use of OCT has increased dramatically during the past decade, whereas the use of FA had declined considerably. 

In addition, the more recent introduction of OCT angiography (OCTA) and the increasing experience in interpreting OCT images has reduced even further the need for FA and its associated side effects.

FA remains however, in our opinion, a necessary technique at the time of diagnosis, since it affords prognostic information, and is useful for re-evaluating patients with a poor response to treatment. In fact, the American Academy of Ophthalmology AMD Preferred Practice Pattern advocates the use of intravenous fundus FA when a patient complains of a new unexplained blurred vision and it is also indicated in the following situations: to detect the presence and determine the extent, type, size, and location of CNV; and, to detect persistent or recurrent CNV following treatment(9)

Nowadays, the retina and imaging community is excited to see if improvements in OCT technology and the introduction of the OCTA could eliminate the FA(9).

Indocyanine green angiography (ICGA) is useful for the study of occult CNV (OCNV), and particularly in identifying the characteristic patterns of idiopathic polypoidal choroidal vasculopathy (IPCV) and retinal angiomatous proliferation (RAP). It is important to correctly diagnose these disorders, since they respond differently to treatment. The main advantage of ICGA is that it allows specific analysis of choroidal circulation and its interactions with the retina. However, interpretation of the ICGA images is complex, and in some cases requires contrasting with the data obtained by means of other techniques such as FA or OCT, and with correct biomicroscopic assessment of the macula.

The present chapter analyses the information that we can obtain from ICGA in application to the diagnosis and follow-up of AMD, and defines the current most precise indications of the technique, and its usefulness in daily clinical practice.

Thus, an analysis is made of the characteristic ICGA images found in patients with AMD, the precise indications of the technique, and the situations in which ICGA proves essential for establishing a correct diagnosis to ensure adequate treatment of this complex disease.

 

INDOCYANINE GREEN ANGIOGRAPHY (ICGA)

Indocyanine green (ICG) is a tricarbocyanine dye with both hydrophilic and lipophilic properties(10). The empirical formula of ICG is C43H47N2NaO6S2. The molecular weight is 775 Daltons. When dissolved in saline solution, ICG tends to form polymers if the solution concentration is high and monomers if the concentration is low. For this reason the dye is dissolved in water for intravenous injections.

Peak absorption of the monomers and polymers is observed at 785 and 690 nm, respectively. Binding to proteins in turn elevates the absorption spectrum to 810 nm, since polymer formation decreases. The emission spectrum in aqueous solution is 810-820 nm, though in blood it shifts 10 nm towards longer wavelengths. The ICG molecule can be considered active at a wavelength of between 790 and 830 nm.

Following intravenous administration, ICG binds to blood lipoproteins (fundamentally phospholipids) in a proportion that can reach 98%(11-13). The emitted fluorescence intensity increases as a result of such binding. Elimination takes place in the liver, with secretion in bile.

The adverse reactions are similar to those observed with fluorescein, but are less frequent(10). In 1994, explorations with ICG documented a 0.15%, 0.2% and 0.05% incidence of mild, moderate and severe reactions, respectively(10). The dye should not be used in patients with shellfish or iodine allergies, liver disease, and end-stage renal disease. No studies in animal models have been made regarding use of the dye in pregnancy. However, current practice patterns regarding the use of ICGA in pregnant woman may be unnecessarily restrictive(14).

Exploration is carried out in the same way as in FA, though with the corresponding excitatory light and barrier filter. The contrast is injected and images are obtained documenting the fluorescence passing through the ocular fundus during the filling phases, and posteriorly after 1, 3, 5, 10, 20 and 30 minutes.

The ICG molecule has biophysical properties that make it useful for visualizing the choroidal circulation. The dye is excited by light close to infrared, increasing its penetration through the retinal pigment epithelium (RPE), melanin, xanthophyll pigment and thin blood layers. ICG circulates almost entirely bound to plasma proteins, and gradually diffuses across the choriocapillaris – thus allowing better visualization of its circulation within the choroidal vessels. These properties make it possible to visualize CNV beneath blood layers, exudates or RPE detachments (PEDs)(15-16) by differentiating (in the latter case) between the hypofluorescent serous component and the hyperfluorescent vascular component (Figure 1).

Figure 1. (A) The ICG molecule has biophysical properties that make it useful for visualizing the choroidal circulation. (B, C and D) These properties make it possible to visualize choroidal neovessels beneath blood layers, exudates or RPE detachments (PEDs) by differentiating (in the latter case) between the hypofluorescent serous component and the hyperfluorescent vascular component. With FA, both the serous and the vascular component appear hyperfluorescent, and the limits of CNV may be masked as a result. A and B. Visualization of the choroidal circulation under thin blood layers. (C) Hot spots visible in an area of exudation. (D) PED with small CNV at the margin of the lesion. With FA, both the serous and the vascular component appear hyperfluorescent, while with ICGA only the vascular component of PED appears hyperfluorescent.

 

The most important contribution of ICGA to the study of CNV in AMD is its capacity to detect choroidal new vessels which in some cases remain occult when using FA(15,16). Following the development of ICGA in the early 1970s(17,18), two decades went by before scanning laser ophthalmoscopy (SLO) and digital angiography made it possible to improve ICGA visualization(19-21)and registry, and routine application of the technique in clinical practice.

In the 1990s, different studies(22) showed that ICGA allows the visualization of OCNV, the most frequent type of CNV in the context of neovascular AMD. The capacity of ICGA to locate new vessels particularly in those cases with associated PEDs made it possible to increase the number of cases of CNV amenable to laser photocoagulation therapy – the latter being the only treatment option available at that time for such patients.

Until the end 1990s, the visualization and delimitation of neovascular membranes was regarded as very important, due to the inclusion of extra- and juxta-foveal locations as an indication for laser photocoagulation(22). In this context, the use of ICGA was of great help due to its capacity to delimit OCNV. In 1992, Yannuzzi et al.(23) showed that late frames of ICGA allowed visualization of the entire choroidal neovascular membrane(22). Before this time, ICGA had already been found to be useful in detecting CNV in cases of block caused by media opacity(24,25).

ICGA served to guide photocoagulation of OCNV without classic CNV, particularly when this OCNV was located inside or at the margin of PED(26,27). The images obtained with this technique in such cases of vascularized PEDs allowed the individualization of a new form of CNV known as chorioretinal anastomosis, later termed retinal angiomatous proliferation (RAP) or type 3 neovascularization(28-30). Scanning laser ophthalmoscopy (SLO), more than vid

In his description, Yannuzzi (58) defended this retinal onset of neovascularization, while Gass (38) suggested that the process starts in the choroid as occult CNV.

When FA is used, RAP manifests as an occult form of CNV with PED and perilesional microhemorrhage and frequent intraretinal edema, associated to neurosensory retinal detachment (NSRD) and RCA in advanced stages.

eoangiography, truly improved the quality of the early frames of ICGA, making it possible to record a dynamic sequence of choroidal filling and improving visualization of this particular neovascular network corresponding to chorioretinal anastomosis(22).

Using SLO, other authors also described the use of high-speed ICGA for visualizing the feeding vessels in cases of OCNV, allowing their selective photocoagulation without the need to cover the entire membrane(31-33). However, the usefulness of ICGA in these cases is limited, since in many cases there is not only one but several feeding vessels; there may be failures in identification; and laser photocoagulation is often not enough to ensure definitive sealing of the afferent vessel(34).

The introduction of verteporfin photodynamic therapy (V-PDT) in the late 1990s allowed the treatment of subfoveal CNV. Different randomized clinical trials were performed, such as the TAP (Treatment of Age-Related Macular Degeneration with Photodynamic Therapy) and VIP (Verteporfin in Photodynamic Therapy) trials, which respectively, included patients with classic and occult CNV(35,36). New vessels were classified according to the FA findings, while ICGA was not performed because PDT became standard care treatment for classic subfoveal CNV.

Anti-VEGF therapy inaugurated a new era in the management of exudative AMD, because it was the first treatment to improve the mean visual acuity of eyes treated with intravitreous injections of anti-VEGF(37,38), which proved effective in both classic and occult subsets of CNV. In randomized clinical trials, CNV was classified according to the FA findings, and ICGA did not form part of the protocol. However, eyes with PEDs were not included in the initial randomized clinical trials, and it is known that such conditions show a poorer treatment response and predispose to RPE rupture(39).

New developments of OCT include better visualization of the choroid(40-43), and this noninvasive technology reduced the use of angiographic testing in routine clinical practice in patients with exudative AMD in which fixed-regimen monthly intravitreous reinjection are used. OCT brings another classification of the neovascular type secondary to AMD: type I is a lesion which grows under the RPE, type II is subretinal and type III corresponds to RAP.

ICGA remains necessary when the diagnosis proves uncertain despite FA and OCT; in certain cases of OCNV with PEDs (particularly in cases of suspected IPCV and RAP) or conditions such as central serous chorioretinopathy (CSC), which may require a different therapeutic approach from that used in neovascular AMD; and in the re-evaluation of non-responding patients.

Due to the economical and medical care burden implied by this type of treatment, an early and correct diagnosis is very important in order to prescribe adequate patient management.

 

Current Importance of ICGA in AMD

Nonexudative (Dry) AMD

In nonexudative AMD, ICGA has been used to improve drusen classification(44,45); to facilitate diagnosis of basal laminar drusen and the associated subretinal deposits(46); and to improve understanding of reticular pseudodrusen(47,48), due to the affinity of some lesions for this dye.
However, ICGA is not routinely used to study dry AMD(22), except in cases of drusenoid PED to confirm the absence of CNV (Figure 2).

Drusenoid PED could be defined as the progressive increase and confluence of small soft drusen. These appear as yellow-gray spots, sometimes with a characteristic hyperpigmented ring, due to their slow and progressive growth. OCT in application to these drusenoid forms allows us to identify elevations of the pigment epithelium, sometimes variable in number and sometimes as a single and larger PED, filled with hyper-reflecting material.

Drusenoid PEDs require no treatment, since they do not imply the presence of neovascular activity, and ICGA is important for establishing this differentiation (Figure 2). 

Figure 2. Drusenoid PEDs require no treatment, since they do not imply the presence of neovascular activity, and ICGA is important for establishing this differentiation. Drusenoid PEDs appear hyperfluorescent in the fluorescein study, while ICGA allows the exclusion of CNV if there is no hyperfluorescent component associated to the hypo- or isofluorescence of the PED.

 

Recently, ICGA has been reported to be useful for evaluating the atrophic areas present in Stargardt disease, in comparison with the atrophy of dry AMD(49). The hypocyanescence evidenced by ICGA in the atrophic areas (“dark atrophy”) has been shown to be more frequent in Stargardt disease than in atrophic AMD – suggesting a possible selective damage of the choriocapillaris in Stargardt disease.

 

Exudative (Wet) AMD

Using ICGA, CNV has been reclassified into three different morphologic types: focal spot or “hot spot”, plaques (well or poorly defined), and mixed (i.e., a combination of the previous two)(50,51). These alterations may be located at the margin of the lesion (marginal spot), above the lesion (overlying spot), or at a distance from the lesion (remote spot) (Figures 3-4).

Figure 3. Using ICGA, CNV has been reclassified as focal spots (or hot spots), plaques or mixed forms (a combination of the first two forms). Figure A shows a plaque with poorly defined margins, in the late phases of ICGA. Figure B shows two hot spots (the lower spot being more active).

 

 

Yannuzzi(23,60) reported that hypofluorescence is maintained throughout the duration of the test in non-vascularized PEDs. On analysing these changes in serous PED with CNV, this author established two well defined groups in 96% of the patients: (a) cases where ICGA shows a solitary and well defined hyperfluorescence spot which he refers to as focal CNV or a “hot spot”; and (b) a “plaque CNV”, which is a larger and poorly defined lesion with less intense fluorescence. In the remaining 4% of the cases the author observed no CNV with ICGA.

Figure 4. CNV may be located: (A) At the margin of the lesion (marginal spot); (B) Above the lesion (overlying spot); or (C) At a distance from the lesion (remote spot).

 

Plaques were found to be the most common type (61% of the cases) and had a poor visual prognosis, whereas the focal spots or “hot spots” (29%) had a better prognosis, and they were considered to be potentially treatable by ICG-guided laser photocoagulation(50,52).
Moreover, ICGA is useful to identify underlying causes of spontaneous submacular hemorrage, allowing for different approaches that may improve outcomes and safety(53).

 

Classic CNV

Classic CNV is better delimited by FA than by ICGA(54,55). ICGA does not offer relevant supplementary information unless associated OCNV is present. As a result, it is likewise not routinely used to study classic CNV.

 

Occult CNV (OCNV)

The term “occult” refers to a type of CNV that is difficult to visualize, analyse and localize on FA. This difficulty is due to their polymorphic features, but especially to the fact that they are ill-defined and ill-delimited(56). The fundamental advantage of ICGA is that it allows early detection and localization of OCNV and analysis of the filling and staining pattern of the CNV network of the lesions.

OCNV is very common, and most cases of CNV in AMD (60-85%) correspond to this type(57,58).

These cases may present a variable course:

  • A slow and almost asymptomatic course for many months.
  • A course characterized by exacerbation episodes.
  • A slowly progressing course with impairment of visual acuity, metamorphopsia and gradual growth of the lesion.

The natural outcome of the disease is generally poor, with severe central vision loss due to hemorrhages, sometimes associated with serous PEDs and/or tearing of the pigment epithelium (RPE Tear), and eventually, fibrous proliferation with formation of a fibrous scar.

OCNV can be grouped into two main types(56):

  1. OCNV without serous PED;
  2. OCNV associated with serous PED.

The classification proposed by the MPS (Macular Photocoagulation Study) investigators is similar and based on FA:

  • Type I: Fibro-vascular PED (Fibro-vascular pigment epithelium detachment with irregular and poorly demarcated hyperfluorescence);
  • Type II: Late leakage of an undetermined source (without PED and without classic CNV).

RAP or type 3 neovascularization in turn is added to the above(29,30).

The importance of ICGA in the study of OCNV is that it can afford an early diagnosis of the disease in cases that prove doubtful with FA and OCT particularly cases involving OCT based differential diagnosis with neurosensory detachment, which can also occur in CSC.
 

1. OCNV without serous PED

FA shows irregular and poorly delimited hyperfluorescence with late phase diffusion suggestive of OCNV. OCT in these cases is essential for evaluating the degree of activity due to the presence of neurosensory detachment and thickening of the pigment epithelium membrane – choriocapillary band, which is more suggestive of OCNV than of CSC in the absence of leakage points in FA. OCT is also fundamental for the follow-up of these patients.

2. OCNV with serous PED

Serous or serohematic PED appears in the ocular fundus as a yellow-orange elevation with margins that are very well defined by the solid adherence of the pigment epithelium to Bruch’s membrane at this level.
Serous PED may be avascular or vascular. For all eyes with PEDs, the purpose of the initial evaluation is to identify or exclude the presence of CNV, as this feature will change the treatment in the anti-VEGF era. ICGA is very useful in these cases for establishing the presence or absence of associated neovascularization, since such CNV may be masked by the hyperfluorescence of PED when FA is used (Figure 5).

Figure 5. ICGA is very useful in these cases for establishing the presence or absence of associated neovascularization, since CNV may be masked by the hyperfluorescence of PED when FA is used. When FA proves inconclusive, ICGA may confirm the absence of neovascularization.

 

When ICGA is used, PED appears as a hypofluorescent lesion in all phases of the angiogram. In some cases, as a result of the accumulation of lipids on Bruch’s membrane, the lesion may have an isofluorescent or even somewhat hyperfluorescent appearance, while in contrast CNV always appear hyperfluorescent. Gass(59) described a blood or pigment meniscus in a zone of the PED, forming a notch that affords a kidney-like shape and which represents the zone of firm adherence between the CNV and the pigment epithelium. This notch which gives the PED a kidney shape allows us to suspect where the CNV is located.

Yannuzzi(23,60) reported that hypofluorescence is maintained throughout the duration of the test in non-vascularized PEDs. On analysing these changes in serous PED with CNV, this author established two well defined groups in 96% of the patients: (a) cases where ICGA shows a solitary and well defined hyperfluorescence spot which he refers to as focal CNV or a “hot spot”; and (b) a “plaque CNV”, which is a larger and poorly defined lesion with less intense fluorescence. In the remaining 4% of the cases the author observed no CNV with ICGA.

ICGA is a very useful diagnostic technique and superior to FA in diagnosing and classifying CNV associated to serous PEDs. As can be seen in the figures, all types of neovascularization can be found associated to serous PED: classic, occult, RAP, and also quite often associated to IPCV (as shown in the Figure 6).

Figure 6. ICGA is very useful for diagnosing and classifying CNV associated to serous PED, due to its capacity to distinguish between the serous and the vascular component. All types of associated CNV may be found. Figure A shows classic CNV which is better visualized with FA in the early phases, since it is masked in the late phases by the hyperfluorescence of the entire lesion. The rest of the cases involve OCNV, where ICGA can show hot spots on the black image of the PED associated to RAP (yellow arrows) or polyps (red arrows).

 

OCTA used to enhance the detection of neovascularization could be possible in some cases where neovascularization is not clearly detected with OCT or dye-based angiographies. Nowadays, it could be interesting to use a multimodal approach in these cases(61).

 

Retinal Angiomatous Proliferation (RAP)

RAP has been described as a variant of exudative AMD characterized by the presumed retinal origin of CNV(62) (Figure 7). 

Figure 7. RAP has been described as a variant of exudative AMD characterized by the presumed retinal origin of CNV. (A) Image of retinal communication under Zeiss FF 450 plus IR (Carl Zeiss Meditec, Jena, Germany). (B) HRA (Heidelberg Retinal Angiography 2) image (HRA2, Heidelberg Engineering, Germany).

 

The pathogenesis and classification of RAP remain subject to debate(63).

Yannuzzi (62) was the first to describe this condition, establishing three different stages on the basis of the clinical and angiographic findings:

  • STAGE I. Proliferation of intraretinal capillaries; intraretinal neovascularization;
  • STAGE II. Subretinal new vessels without PED (IIA) or with PED (IIB);
  • STAGE III. Choroidal neovascularization (CNV);
  • Such CNV often consolidates as retinochoroidal anastomosis (RCA), which posteriorly has been classified as:
  • STAGE IV. Retinochoroidal anastomosis (RCA) (64,65).

RAP represents approximately 12-20% of all cases of exudative AMD (30,66,67), and currently the term type III neovascularization has been proposed for this condition, precisely in order to distinguish it from other types of CNV in exudative AMD (28,65,68).

This disorder is known to show a different natural course and response to treatment compared with other forms of AMD(69,70,71). The existence of retino-retinal anastomosis (RRA) and RCA(62,64) is characteristic of RAP type lesions and contributes to the presence of high-flow vascular alterations with double retinal and choroidal circulation. This and the presence of associated PED and cystoid macular oedema in the more advanced stages complicates the treatment of this disease(72) (Figure 8).

Figure 8. The existence of RRA is characteristic of RAP type lesions and contributes to the presence of high-flow vascular alterations with double retinal and choroidal circulation. This, and the presence of associated PED and cystoid macular edema, complicates the treatment of this disease.

 

In some cases, ICGA in the early phases allows us to identify RRA with a nutrient vessel and a draining vessel(73) – this representing a retinal neoangiogenic process that subsequently spreads to the subretinal space (Figure 8).

In some cases, ICGA in the early phases allows us to identify retino-retinal anastomosis with a nutrient vessel and a draining vessel (69) – this representing a retinal neoangiogenic process that subsequently spreads to the subretinal space (Fig.8).

In his description, Yannuzzi(62) defended this retinal onset of neovascularization, while Gass(40) suggested that the process starts in the choroid as OCNV.
When FA is used, RAP manifests as an occult form of CNV with PED and perilesional microhemorrhage and frequent intraretinal edema, associated to neurosensory detachment and RCA in advanced stages.

ICGA allows us to study the existence of single or multiple hot spots at neovascularization level. This is very useful in the earliest stages, especially to the effects of treatment, if V-PDT is to be associated. ICGA also allows us to observe spot persistence or disappearance after treatment, as well as the reappearance of spots in the case of patients who are re-evaluated due to a lack of response or resistance to therapy. In some early stage cases, ICGA can reveal the characteristic retino-retinal vascular communication in the early phases of the angiogram (Figure 9).

Figure 9. ICGA allows us to study the existence of single (A and C) or multiple (B) hot spots at neovascularization level (white arrows). In some early stage cases, ICGA can reveal the characteristic retino-retinal vascular communication in the early phases of the angiogram (red arrows).

 

It is important to distinguish RAP from other forms of AMD, not only in view of the poor vision prognosis if exhaustive patient follow-up and correct treatment are not carried out, but also due to the high risk of neovascularization in the other eye(70,74,75). In this context, ICGA helps us to establish a correct diagnosis (Figure 10).

Figure 10. It is important to distinguish RAP from other forms of AMD due to the high risk of neovascularization in the other eye when the disease is present in one eye. (A) RAP first manifesting in the left eye with spontaneous rupture of the RPE; we must keep a very close follow-up because of a possible involvement of the fellow eye. (B) After 6 months the fellow eye is affected, and ICGA helps us to establish a correct early diagnosis. We can see the communication in early phases and the small hot spot in late phases of ICGA.

 

The precise mechanism giving rise to this peculiar form of vascularization remains unclear, though some authors recently have suggested that there is ischemia at external retinal level, induced by diminished choroidal perfusion(76-78).

RAP requires more intensive management, often involving combined treatments, in order to control the activity of the disease(70, 79-82). Recently, favourable results have been published with the use of anti-VEGF drugs, though with a high recurrence rate after suspending the treatment(63,83-97).

V-PDT has been associated to anti-VEGF drugs and to intravitreous triamcinolone (IVT), with promising results(98-114). Combined treatment proves effective in maintaining or improving patient vision and in reducing exudation, without long-term adverse effects(110). Rouvas et al.(112) have reported stabilization in patients administered with ranibizumab alone and in patients administered with ranibizumab associated to V-PDT, with functional and anatomical improvements in the patients subjected to V-PDT with IVT.

Regarding the combined treatments, it must be taken into account that there is an increased risk of developing atrophy in patients with RAP subjected to V-PDT(112). Frequent geographic atrophy has been described in these patients after the treatment(77), and recently there have also been reports(78) of specific changes in the choroid and in drusen distribution, which may have long term diagnostic and therapeutic implications (Figure 11).

Figure 11. Regarding the combined treatments, it must be taken into account that there is an increased risk of developing atrophy in patients with RAP subjected to V-PDT. Frequent geographic atrophy has been described in these patients after the treatment (white arrow).

 

In coincidence with other authors(111), we perform ICGA-guided PDT in some patients in stages I and II, with the smallest possible spot covering the neovascularization, associated to anti-VEGF drugs. This technique would use smaller spots, thereby contributing to lessen the risk of long-term atrophy in these patients (Figure 12). 

Figure 12. We perform ICGA-guided PDT in some patients in stages I and II, with the smallest possible spot covering the neovascularization, associated to anti-VEGF drugs. This technique would use smaller spots, thereby contributing to lessen the risk of long-term atrophy in these patients. The figure shows the outcome, three years after the treatment.

 

It is also known that the prognosis is poorer in cases of large PEDs, where pigment epithelial rupture may be more frequent after treatment(115-117) (Figure 13), and that better functional results are obtained when these patients are diagnosed and treated in early stages of the process(118). The diagnosis should be, therefore, established as soon as possible, and ICGA is very important to this effect (Figure 14).

Figure 13. It is also known that the prognosis is poorer in cases of RAP (A, red arrow) with large PEDs, where pigment epithelial rupture may be more frequent after treatment. (A) PED is observed/marked with a thin wall (A, white arrow) where pigment epithelial rupture occurs after intravitreous injection of ranibizumab (B, white arrow).

 

Figure 14. The diagnosis therefore should be established as soon as possible, and ICGA is very important to this effect. (A) Stage I RAP is observed, easily identifiable with ICGA. (B) A patient with bilateral drusen where ICGA identifies an incipient communication (B1, white arrow) that produces alterations with intraretinal edema as evidenced by OCT (B1); ICGA proves negative in the other eye, however (B2).

 

OCT may be used to diagnose type 3 neovascularization in patients with neovascular AMD with relatively high concordance compared with ICGA-based diagnosis(119).

 

Idiopathic Polypoidal Choroidal Vasculopathy (IPCV) 

IPCV is characterized by the presence of a network of abnormal choroidal vessels with aneurysmal dilatations in the form of polyps visualized as red-orange nodules in the ocular fundus and located mainly at peripapillary level and in the macular zone(120-123). These choroidal vascular lesions can produce serous exudation and hemorrhage that have been associated to recurrent serohematic PEDs and neurosensory detachment(122,124-128) (Figure 15). 

Figure 15. (A and B) Peripapillary polyps with a large exudative component at superior temporal arcade level in the first patient. (C) Serohemorrhagic PED at macular level, with a small polyp, evidenced by ICGA (C, white arrow). (D) Circinate macular lesion centered by a network of polyps (D, white arrows).

 

The vascular lesions of IPCV can simulate an occult or minimally classic membrane with FA, while ICGA clearly reveals the polypoidal lesions and anomalous vascular network within the choroid. In concordance with this, a literature review published by Stanga et al. concluded that ICGA is highly recommended for the identification of choroidal lesions in patients with IPCV(51) (Figure 16).

Figure 16. Patients with OCNV in FA, where the polyps are seen by ICGA. The three patients responded poorly to anti-VEGF therapy and where therefore re-evaluated due to suspected IPCV. In case C, FA reveals a hyperfluorescent vascular PED (yellow arrow), while only the polyps at the margin of the PED appear hyperfluorescent with ICGA (white arrow).

 

ICGA is able to identify two elements in these anomalies: polypoidal structures projecting from the internal choroid towards the retina, and a branched choroidal vascular network (BVN) manifesting with early vascular hyperfluorescence(129-131). The underlying pathogenesis remains subject to debate (Figure 17). In this sense, some authors suggest that the polyps are a form of CNV(132-135), while others regard them as an alteration of the internal choroidal vessels(136-138).

Figure 17. ICGA is able to identify two elements in these anomalies: polypoidal structures projecting from the internal choroid towards the retina, and a BVN manifesting with early vascular hyperfluorescence. The figure shows this choroidal vascular network (short white arrows), with the formation of a polyp at marginal level (large white arrow).

 

The polyps are difficult to identify on FA, as they are located beneath the RPE. On ICGA, however, the polyps are clearly visible as single or multiple vascular aneurysms with a diameter of 100-500 μm which fill after a short delay and remain hyperfluorescent until the late phases. These lesions are often located in the juxtapapillary zone, but can also be found at macular level, in the vascular arches, and even at peripheral level. The choroidal vessels in the area of the polyps may appear irregular and dilated, forming a fine vascular network. The angiographic sequence would be as follows: filling of the fine choroidal vessels and polyps with early choroidal vascular hyperfluorescence when the dye flows through the network of anomalous vessels. This fluorescence is maintained over time and disappears as the dye washes out in the late phases, in the presence of polypoidal lesions without active leakage (Figure 18).

Figure 18. Patient diagnosed with IPCV subjected to laser treatment in the macular area. Scantly active polyps persist at inferior temporal arcade level, evidencing dye washout in the last phases of the angiogram. The angiographic sequence would be as follows: filling of the fine choroidal vessels and polyps with early choroidal vascular hyperfluorescence when the dye flows through the network of anomalous vessels. This fluorescence is maintained over time and disappears as the dye washes out in the late phases, in the presence of polypoidal lesions without active leakage.

 

Guidelines have recently been published for the clinical diagnosis and treatment of IPCV(139), in which ICGA is regarded as the gold standard for the diagnosis of this disorder. This panel of experts considers that such angiography should be performed when routine ophthalmoscopic examination indicates a serosanguineous maculopathy with one of  the following features:

  • Red-orange subretinal nodules visible with the ophthalmoscope.
  • Spontaneous massive subretinal hemorrhage.
  • Notched or hemorrhagic PED.
  • Occlusion of polyps with lowresponse rate to anti-VEGF therapy.

In conclusion, ICGA is the gold standard for diagnosis of IPCV, which is defined as the presence of single or multiple focal nodular areas of hyperfluorescence arising from the choroidal circulation within the first 6 minutes after injection of ICG, with or without an associated BVN(139).

These authors propose algorithms for both the diagnosis and the treatment of the disease. IPCV may be classified clinically as: quiescent (polyps in the absence of subretinal or intraretinal fluid or haemorrhage), exudative (exudation without haemorrhage, which includes variously sensory retinal thickening, neurosensory detachment, PED, and subretinal lipid exudation) or hemorrhagic (any subretinal or sub-RPR haemorrhage with or without other exudative characteristics). Treatment should be initiated for active and symptomatic IPCV, and can be considered for active asymptomatic presentations (Figure 19).

Figure 19. Treatment should be initiated for active and symptomatic IPCV, and can be considered for active asymptomatic presentations. The figure shows both eyes of the same patient. (A) The patient is symptomatic, with a visual acuity of 0.6 in the right eye. The polypoidal lesions can be observed with ICGA, and retinal thickening and intraretinal cysts can be seen by OCT. (B) The patient maintains a visual acuity of 1.0 in the left eye. ICGA shows a peripapillary polyp, likewise visible in the OCT acquisition, but there is no retinal edema, and the patient therefore remains asymptomatic.

 

ICGA is necessary not only for the diagnosis but also for continuous monitoring of the disease, which may present polyp relapse or the appearance of new polyps from the BVN, with phases of quiescence, exudation or hemorrhage (Figure 20). 

Figure 20. ICGA is necessary not only for the diagnosis but also for continuous monitoring of the disease, which may present polyp relapse or the appearance of new polyps from the BVN, with phases of quiescence, exudation or hemorrhage. (A) Late phases of FA showing the hyperfluorescent PEDs. (B) ICGA image. (B1 and B2) Only the PEDs are seen, and the BVN alteration is noted in early phases with ICGA. (B3 and B4) Formation of polyps from the BVN, visible with ICGA 6 months later.

 

Focal photocoagulation with ICGA-guided thermal laser still plays a role in the treatment of extrafoveal polyps, but it is not recommended in application to juxta- or subfoveal lesions(139,140). In these cases, PDT has been the treatment of choice, with maintenance or improvement of patient visual acuity in 80-95% of the cases(141-146) (Figure 21).

Figure 21. Focal photocoagulation with ICGA-guided thermal laser still plays a role in the treatment of extrafoveal polyps, but it is not recommended in application to juxta- or subfoveal lesions. (B) Resolution of the polyps with atrophy scarring secondary to thermal laser in extrafoveal polyps at inferior temporal arcade level and after PDT at macular level. OCT shows resolution of the retinal thickening, with disappearance of the hard exudate.

 

However, long-term follow-up studies have described frequent recurrences after PDT(142, 147). In some cases PDT is unable to fully occlude the polyps, though in other cases persistence of the BVN can give rise to new active polypoidal lesions. As a result, periodic controls with ICGA are needed during follow-up to determine whether there are persistent or recurrent lesions with polyps or BVN.

Although recurrence and retreatments do not seem to decrease best corrected visual acuity, long-term studies are still needed in order to evaluate the efficacy and safety of PDT treatment of IPVC(142). PDT can modified the natural progression of the polypoidal CNV as a result of ischemia and increased VEGF expression(148).

After the introduction of anti-VEGF therapy, different studies reported the efficacy of combined treatment with PDT and anti-VEGF agents for IPCV(149,150), as well as the way to minimize the complications associated with the use of PDT, and the need to use selective(151) or with reduced fluence PDT(152).

The Everest study(153) concluded that V-PDT combined with ranibizumab  0.5 mg or alone was superior to ranibizumab monotherapy. In our own experience with long-term patient follow-up(154), the association of ICGA-guided PDT for the occlusive therapy of polyps and intravitreous ranibizumab is effective in prolonging the recurrence-free interval, as well as in maintaining good visual acuity (Figure 22).

Figure 22. The association of ICGA-guided PDT for the occlusive therapy of polyps and intravitreous ranibizumab is effective in prolonging the recurrence-free interval, as well as in maintaining good visual acuity. Patient with a single juxtapapillary polyp associated to a large PED at macular level. Three PDT sessions are performed (A, B and C), without resolution of the condition. Indeed, worsening is noted, with the appearance of neurosensory detachment of the fovea (C). At the time where the polyp is not visualized by ICGA, and since the PED persists, three intravitreous injections of ranibizumab are administered, followed by resolution of the condition (E). The patient remained stable, with a visual acuity of 20/20 until month 36 of follow up.

 

Although the efficacy and safety of PDT and ranibizumab versus PDT and bevacizumab remain to be established, the experts have concluded(23) that PDT with standard fluence guided by ICGA with or without the combination of three intravitreous injections of ranibizumab 0.5 mg on a monthly basis is the most widely recommended treatment for this disease. ICGA-guided thermal laser photocoagulation can be considered for extrafoveal polyps.

The use of ICGA is very important for guiding laser treatment and PDT, both as regards the indication of such treatments and to minimize the spot size. This in turn contributes to limit as far as possible the risk of atrophy and other side effects derived from the multiple treatments usually required in the long-term follow-up of these patients.
A study with ICGA may in some cases help to unmask patients purportedly refractory to anti-VEGF therapy(155), because it allows a precise diagnosis and adequate treatment indication.

ICGA will retain a key role in the management of PCV because of its ability to better identify polyps although SS-OCTA is effective at detecting vascular network(156). However, SS-OCTA may be more sensitive than SD-OCT in detecting early recurrence.

On the other hand, the combination of En Face and cross-sectional OCTA images provides anatomical information about polypoidal structures that is comparable to ICGA. OCTA is safe and can provide supplementary information but due to its limitations it does not replace ICGA in a clinical setting nowadays(157-159).

 

>> References

References - Diagnostic usefulness of indocyanine green angiography (ICGA) in age-related macular degeneration (AMD)

References - Diagnostic usefulness of indocyanine green angiography (ICGA) in age-related macular degeneration (AMD)

English
  1. Congdon NG, Friedman DS, Lietman T. Important Causes of Visual Impairment in the World Today. JAMA. 2003;290(15):2057-2060.
     
  2. Eye Diseases Prevalence Research Group.  Prevalence of age-related macular degeneration in the United States.  Arch Ophthalmol. 2004;122:564-572.
     
  3. Bressler NM. Age-Related Macular Degeneration Is the Leading Cause of Blindness. JAMA. 2004;291(15):1900-1901.
     
  4. Alfredo Pece, Claudio Azzolini, Maurizio Battaglia Parodi, Ferdinando Bottoni, Paola Danzi, Simone Donati, Ugo Introini, Vincenzo Pucci, Francesco Semeraro, and Francesco Viola. Consensus on the diagnosis, treatment and follow-up of patients with age-related macular degeneration eligible for ranibizumab. Expert Review of Ophthalmology. 2012;7(3):219-225.
     
  5. Chappelow AV, Kaiser PK. Neovascular age-related macular degeneration: potential therapies. Drugs. 2008;68(8):1029–36.
     
  6. Stewart MW. The expanding role of vascular endothelial growth factor inhibitors in ophthalmology. Mayo Clin Proc. 2012;87(1):77–88.
     
  7. Waisbourd M, Loewenstein A, Goldstein M, Leibovitch I. Targeting vascular endothelial growth factor: a promising strategy for treating age-related macular degeneration. Drugs Aging. 2007; 24(8):643–62.
     
  8. Agarwal A, Rhoades WR, Hanout M, Soliman MK, Sarwar S, Sadiq MA, Sepah YJ, Do DV, Nguyen QD. Management of neovascular age related macular degeneration: current state-of-the art care optimizing visual outocomes and therapies in development. Clini Ophtahalmol. 2015 Jun;9:1001-15.
     
  9. Shoughy SS, Kozak I. Selective and complementary use of Optical Coherence Tomography and Fluorescein Angiography in retinal practice. Eye Vis (Lond). 2016;17;3:26.
     
  10. Hope-Ross M, Yannuzzi LA, Gragoudas ES, TFyer DR, Slakter JS, Sorenson JA, Krupsky S, Orlock DA, Puliafito CA. Adverse reactions due to indocyanine green. Ophthalmology. 1994;101:529-533.
     
  11. Cherrick GR, Stein SW, Leevy CM, Davidson CS. Indocyanine green: observations on its physical properties, plasma decay, and hepatic extraction. J Clin Invest. 1960 Apr;39:592-600.
     
  12. Baker KJ. Binding of sulfobromophthalein (BSP) sodium and indocyanine green (ICG) by plasma alpha-1 lipoproteins. Proc Soc Exp Biol Med. 1966 Aug-Sep;122(4):957-63.
     
  13. Brown N, Strong R. Infrared fundus angiography. Br J Ophthalmol. 1973 Oct;57(10):797-802.
     
  14. Fineman MS, Maguire JI, Fineman SW, Benson WE. Safety of indocyanine green angiography during pregnancy. Arch Ophthalmol. 2001; 119:353-355.
     
  15. Guyer DR and Yannuzzi LA. Occult choroidal neovascularization. In Flower RW. Indocyanine green angiography.  Yannuzzi LA, Flower RW, Slakter JS. Ed. Mosby. St. Louis, Missouri. 199:157-180.
     
  16. Guyer DR, Puliafito CA, Mones JM, Friedman E, Chang W, Verdooner SR.  Digital indocyanine-green angiography in chorioretinal disorders. Ophthalmology. 1992 Feb;99(2):287-91.
     
  17. Yannuzzi LA, Slakter JS, Sorenson JA, Guyer DR, Orlock DA. Digital indocyanine green videoangiography and choroidal neovascularization. Retina. 1992;12(3):191-223.
     
  18. Kogure K, David NJ, Yamanouchi U, Choromokos E. Infrared absorption angiography of the fundus circulation. Arch Ophthalmol. 1970 Feb;83(2):209-14.
     
  19. Flower RW. Injection technique for indocyanine green and sodium fluorescein dye angiography of the eye. Invest Ophthalmol. 1973; 12:881-895.
     
  20. Desmettre T, Devoisselle JM, Mordon S. Fluorescence properties and metabolic features of indocyanine green as related to angiography. Surv Ophthalmol. 2000; 45:15-27.
     
  21. Obana A, Miki T, Hayashi K, Takeda M, Kawamura A, Mutoh T, Harino S, Fukushima I, Komatsu H, Takaku Y. Survey of complications of indocyanine green angiography in Japan. Am J Ophthalmol. 1994;118:749-753.
     
  22. Cohen SY, Dubois L, Quentel G, Gaudric A. Is indocyanine green angiography still relevant? Retina. 2011 Feb;31(2):209-21.
     
  23. Yannuzzi LA, Slakter JS, Sorenson JA, Guyer DR, Orlock DA. Digital indocyanine green videoangiography and choroidal neovascularization. Retina. 1992;12:191–223.
     
  24. Hayashi K, Hasegawa Y, Tokoro T, De Laey JJ. Value of indocyanine green angiography in the diagnosis of occult choroidal neovascular membrane. Jpn J Clin Ophthalmol. 1988;42:827–829.
     
  25. Hayashi K, Hasegawa Y, Tokoro T, De Laey JJ. Clinical application of indocyanine green angiography to occult choroidal neovascularisation. Jpn J Clin Ophthalmol. 1989;43:57–65.
     
  26. Sorenson JA, Yannuzzi LA, Slakter JS, Guyer DR, Ho AC, Orlock DA. A pilot study of digital indocyanine green videoangiography for recurrent occult choroidal neovascularization in age-related macular degeneration. Arch Ophthalmol. 1994;112:473–479.
     
  27. Slakter JS, Yannuzzi LA, Sorenson JA, Guyer DR, Ho AC, Orlock DA. A pilot study of indocyanine green videoangiography-guided laser photocoagulation of occult choroidal neovascularization in age-related macular degeneration. Arch Ophthalmol. 1994;112:465–472.
     
  28. Yannuzzi LA, Negrao S, Iida T et al. Retinal angiomatous proliferation in age-related macular degeneration. Retina 2001;21:416–434.
     
  29. Kuhn D, Meunier I, Soubrane G, Coscas G. Imaging of chorioretinal anastomoses in vascularized retinal pigment epithelium detachments. Arch Ophthalmol. 1995;113:1392–1398.
     
  30. Freund KB, Ho IV, Barbazetto IA, Koizumi H, Laud K, Ferrara D, Matsumoto Y, Sorenson JA, Yannuzzi L. Type 3 neovascularization: the expanded spectrum of retinal angiomatous proliferation. Retina. 2008;28:201–211.
     
  31. Shiraga F, Ojima Y, Matsuo T, Takasu I, Matsuo N. Feeder vessel photocoagulation of subfoveal choroidal neovascularization secondary to age-related macular degeneration. Ophthalmology. 1998;105:662–669.
     
  32. Flower RW. Experimental studies of indocyanine green dye-enhanced photocoagulation of choroidal neovascularization feeder vessels. Am J Ophthalmol. 2000;129:501–512.
     
  33. Desatnik H, Treister G, Alhalel A, Krupsky S, Moisseiev J. ICGA-guided laser photocoagulation of feeder vessels of choroidal neovascular membranes in age-related macular degeneration. Indocyanine green angiography. Retina 2000; 20:143–150.
     
  34. Lafaut BA, Bartz-Schmidt KU, Vanden Broecke C, Aisenbrey S, De Laey JJ, Heimann K. Clinicopathological correlation in exudative age related macular degeneration: histological differentiation between classic and occult choroidal neovascularisation. Br J Ophthalmol. 2000;84:239–243.
     
  35. Bressler NM; Treatment of Age-Related Macular Degener- ation with Photodynamic Therapy (TAP) Study Group. Photodynamic therapy of subfoveal choroidal neovascularization in age-related macular degeneration with verteporfin: two-year results of 2 randomized clinical trials-TAP report 2. Arch Ophthalmol. 2001;119:198–207.
     
  36. Verteporfin In Photodynamic Therapy Study Group. Verteporfin therapy of subfoveal choroidal neovascularization in age- related macular degeneration: two-year results of a randomized clinical trial including lesions with occult with no classic choroidal neovascularization verteporfin in photodynamic therapy report 2. Am J Ophthalmol. 2001;131:541–560.
     
  37. Rosenfeld PJ, Brown DM, Heier JS, et al. Ranibizumab for neovascular age-related macular degeneration. N Engl J Med. 2006;355:1419–1431.
     
  38. Brown DM, Kaiser PK, Michels M, et al. Ranibizumab versus verteporfin for neovascular age related macular degeneration. N Engl J Med. 2006;355:1432–1444.
     
  39. Arias L. Treatment of retinal pigment epithelial detachment with antiangiogenic therapy. Clin Ophthalmol. 2010;4:369–374.
     
  40. Gass JD, Agarwal A, Lavina AM, Tawansy KA. Focal inner retinal hemorrhages in patients with drusen: an early sign ofoccult choroidal neovascularization and chorioretinal anastomosis. Retina. 2003;23:741–751.
     
  41. Matsumoto H, Sato T, Kishi S. Tomographic features of intraretinal neovascularization in retinal angiomatous proliferation. Retina 2010;30:425–430.
     
  42. Spaide RF. Enhanced depth imaging optical coherence tomography of retinal pigment epithelial detachment in age-related macular degeneration. Am J Ophthalmol. 2009;147:644–652.
     
  43. Keane PA, Sadda SR. Imaging chorioretinal vascular disease. Eye (Lond). 2010;24:422–427.
     
  44. Arnold JJ, Quaranta M, Soubrane G, Sarks SH, Coscas G. Indocyanine green angiography of drusen. Am J Ophthalmol. 1997;124:344–356.
     
  45. Chang AA, Guyer DR, Orlock DR, Yannuzzi LA. Age-dependent variations in the drusen fluorescence on indocyanine green angiography. Clin Experiment Ophthalmol. 2003;31:300–304.
     
  46. Quaranta M, Buglione M, Lo Schiavo Elia R, Coscas G, Soubrane G. Indocyanine green angiography of basal laminar drusen in the retinal pigment epithelium associated with vitelliform macular degeneration. J Fr Ophtalmol. 1998;21:185–190.
     
  47. Arnold JJ, Sarks SH, Killingsworth MC, Sarks JP. Reticular pseudodrusen. A risk factor in age-related maculopathy. Retina. 1995;15:183–191.
     
  48. Smith RT, Sohrab MA, Busuioc M, Barile G. Reticular macular disease. Am J Ophthalmol. 2009;148:733–743.
     
  49. Giani A, Pellegrini M, Carini E, Peroglio Deiro A, Bottoni F, Staurenghi G. The dark atrophy with indocyanine green angiography in Stargardt disease. Invest Ophthalmol Vis Sci. 2012 Jun 26;53(7):3999-4004.
     
  50. Guyer DR, Yannuzzi LA, Slakter JS, et al. Classification of choroidal neovascularizaton by digital indocyanine green videoangiography. Ophthalmology. 1996;103:2054–60.
     
  51. Stanga PE, Lim JI, Hamilton P. Indocyanine green angiography in chorioretinal diseases: indications and interpretation: an evidence-based update. Ophthalmology. 2003 Jan;110(1):15-21.
     
  52. Guyer DR, Yannuzzi LA, Ladas I, et al. Indocyanine green guided laser photocoagulation of focal spots at the edge of plaques of choroidal neovascularization. Arch Ophthalmol. 1996;114:693–7.
     
  53. Kim H, Lee SC, Kim SM, Lee JH, Koh HJ, Kim SS, Byeon SH, Kim M, Lee Cs. Identification of underlying causes of spontaneous submacular haemorrhage by indocyanine green angiography. Ophthalmologica. 2015;233(3-4):146-54.
     
  54. Hoang QV, Gallego-Pinazo R, Yannuzzi LA. Long-term follow-up of acute zonal occult outer retinopathy. Retina. 2013 Jul-Aug;33(7):1325-7.
     
  55. Sorenson JA, Yannuzzi LA, Slakter JS, Guyer DR, Ho AC, Orlock DA.  A pilot study of digital indocyanine green videoangiography for recurrent occult choroidal neovascularization in age-related macular degeneration. Arch Ophthalmol. 1994 Apr;112(4):473-9.
     
  56. Coscas G. Chapter III: Age-Related Macular Degeneration. In: Atlas of indocyanine green angiography fluorescein angiography, ICG angiography and OCT correlations. Coscas G.  pp 118. Ed. Elsevier  September 2005. ISBN: 2-84299-729-8.
     
  57. Freund KB, Yannuzzi LA, Sorenson JA. Age-related macular degeneration and choroidal neovascularization. Am J Ophthalmol. 1993 Jun;115(6):786-91.
     
  58. Haddad WM, Coscas G, Soubrane G. Eligibility for treatment and angiographic features at the early stage of exudative age related macular degeneration. Br J Ophthalmol. 2002 Jun;86(6):663-9.
     
  59. Gass JD. Serous retinal pigment epithelial detachment with a notch: a sign of occult choroidal neovascularization. 1984. Retina. 2003 Dec;23(6 Suppl):205-20.
     
  60. Yannuzzi LA, Hope-Ross M, Slakter JS, Guyer DR, Sorenson JA, Ho AC, Sperber DE, Freund KB, Orlock DA. Analysis of vascularized pigment epithelial detachments using indocyanine green videoangiography. Retina. 1994;14(2):99-113.
     
  61. Tan A, Simhaee D, Balaratnasingam C, Dansingani K, Yannuzzi L. A perspective on the nature and frequency of pigment epithelial detachment. Am J Ophthalmol. 2016 Dec;172:13-27.
     
  62. Yannuzzi LA, Negrao S, Iida T, et al. Retinal angiomatous proliferation in age-related macular degeneration. Retina. 2001; 21: 416-434.
     
  63. Parodi MB, Iacono P, Menchini F, Sheth S, Polini G, Pittino R, Bandello F. Intravitreal bevacizumab versus ranibizumab for the treatment of retinal angiomatous proliferation. Acta Ophthalmol. 2013 May;91(3):267-73.
     
  64. Yannuzzi LA. Degeneration: age-related macular degeneration. In: The Retinal Atlas. Philadelphia, PA: Saunders; 2010.
     
  65. Yannuzzi LA, Freund KB, Takahashi BS. Review of retinal angiomatous proliferation or type 3 neovascularization [editorial]. Retina. 2008;28:375–384.
     
  66. Massacesi AL, Sacchi L, Bergamini F, Bottoni F. The prevalence of retinal angiomatous proliferation in age-related macular degeneration with occult choroidal neovascularization. Graefes Arch Clin Exp Ophthalmol. 2008;246:89–92.
     
  67. Kuhn D, Meunier I, Soubrane G, Coscas G. Imaging of chorioretinal anastomoses in vascularised retinal pigment epithelial detachments. Arch Opthalmol. 1995,113:1392–1398.
     
  68. Querques G, Querques L, Forte R, Massamba N, Blanco R, Souied EH. Precursors of type 3 neovascularization: A Multimodal Imaging Analysis. Retina. 2013 Jun;33(6):1241-8.
     
  69. Hartnett ME, Weiter JJ, Staurenghi G, et al. Deep retinal vascular anomalous complexes in advanced age-related macular degeneration. Ophthalmology. 1996;103:2042–2053.
     
  70. Bottoni F, Massacesi A, Cigada M, et al. Treatment of retinal angiomatous proliferation in age-related macular degeneration: a series of 104 cases of retinal angiomatous proliferation. Arch Ophthalmol. 2005;123: 1644–1650.
     
  71. BresslerNM. Retinal anastomosis to choroidal neovascularization: a bum rap for a difficult disease. Arch Ophthalmol. 2005;123:1741–1743.
     
  72. Ghazi NG, Knape RM, Kirk TQ, Tiedeman JS, Conway BP. Intravitreal bevacizumab (avastin) treatment of retinal angiomatous proliferation. Retina. 2008 May;28(5):689-95.
     
  73. Axel-Siegel R, Bourla D, Priel E, Yassur Y, Weinberger D. Angiographic and flow patterns of retinal choroidal anastomoses in age-related macular degeneration with occult choroidal neovascularization. Ophthalmology. 2002 Sep;109(9):1726-36.
     
  74. Gross NE, Aizman A, Brucker A, Klancnik JM Jr, Yannuzzi LA. Nature and risk of neovascularization in the fellow eye of patients with unilateral retinal angiomatous proliferation. Retina. 2005;25(6):713–718.
     
  75. Campa C, Harding SP, Pearce IA, Beare NA, Briggs MC, Heimann H. Incidence of neovascularization in the fellow eye of patients with unilateral retinal angiomatous proliferation. Eye (Lond). 2010 Oct;24(10):1585-9.
     
  76. Koizumi H, Iida T, Saito M, Nagayama D, Maruko I. Choroidal circulatory disturbances associated with retinal angiomatous proliferation on indocyanine green angiography. Graefes Arch Clin Exp Ophthalmol. 2008;246(4):515–520.
     
  77. McBain VA, Kumari R, Townend J, Lois N. Geographic atrophy in retinal angiomatous proliferation. Retina. 2011; 31(6):1043–1052.
     
  78. Kim JH, Kim JR, Kang SW, Kim SJ, Ha HS. Thinner Choroid and Greater Drusen Extent in Retinal Angiomatous Proliferation than in Typical Exudative Age-Related Macular Degeneration. J Ophthalmol. 2013;155:743–749.
     
  79. Mendis R, Leslie T, McBain VA, Lois N. Combined therapy for retinal angiomatous proliferation with intravitreal triamcinolone and argon laser photocoagulation. Br J Ophthalmol. 2008;92:1154–1156.
     
  80. Boscia F, Furino C, Sborgia L, Reibaldi M, Sborgia C. Photodynamic therapy for retinal angiomatous proliferation and pigment epithelium detachment. Am J Ophthalmol. 2004; 138:1077–1079.
     
  81. Boscia F, Parodi MB, Furino C, Reibaldi M, Sborgia G. Photodynamic therapy with verteporfin for retinal angiomatous proliferation. Graefes Arch Clin Exp Ophthalmol. 2006;244: 1224–1232.
     
  82. Kuroiwa S, Arai J, Gaun S, Lida T, Yoshimura N. Rapidly progressive scar formation after transpupillary thermotherapy in retinal angiomatous proliferation. Am J Ophthalmol. 2003;23:417–420.
     
  83. Hemeida TS, Keane PA, Dustin L, Sadda SR, Fauzi AA. Long- term visual and anatomic outcomes following anti-VEGF monotherapy for retinal angiomatous proliferation. Br J Ophthalmol. 2009;6:701–705.
     
  84. Gharbiya M, Allievi F, Recupero V, Martini D, Mazzeo L, Gabrieli CB. Intravitreal bevacizumab as primary treatment for retinal angiomatous proliferation: twelve-month results. Retina. 2009 Jun;29(6):740-9.
     
  85. Rouvas AA, Papakostas TD, Ladas ID. Ranibizumab for retinal angiomatous proliferation. Graefes Arch Clin Exp Ophthalmol. 2009 Dec;247(12):1719-20.
     
  86. Engelbert M, Zweifel SA, Freund KB. "Treat and extend" dosing of intravitreal antivascular endothelial growth factor therapy for type 3 neovascularization/retinal angiomatous proliferation. Retina. 2009 Nov-Dec;29(10):1424-31.
     
  87. Montero JA, Fernandez MI, Gomez-Ulla F, Ruiz-Moreno JM. Efficacy of intravitreal bevacizumab to treat retinal angiomatous proliferation stage II and III. Eur J Ophthalmol. 2009;19:448–451.
     
  88. Konstantinidis L, Mameletzi E, Mantel I, Pournovas JA, Zografos L, Ambresin A. Intravitreal ranibizumab (Lucentis) in the treatment of retinal angiomatous proliferation (RAP). Graefes Arch Clin Exp Ophthalmol. 2009;9:1165–1171.
     
  89. Meyerle CB, Freund KB, Iturralde D, et al. Intravitreal bevacizumab (Avastin) for retinal angiomatous proliferation. Retina. 2007;27:451–457.
     
  90. Hemeida TS, Keane PA, Dustin L, Sadda SR, Fawzi AA. Long-term visual and anatomical outcomes following anti-VEGF monotherapy for retinal angiomatous proliferation. Br J Ophthalmol. 2010 Jun;94(6):701-5.
     
  91. Gupta B, Jyothi S, Sivaprasad S. Current treatment options for retinal angiomatous proliferans (RAP). Br J Ophthalmol. 2010 Jun;94(6):672-7.
     
  92. George S, Cooke C, Chakravarthy U. Exudative AMD subtypes and eligibility for treatment with ranibizumab. Eye (Lond). 2010 Jul;24(7):1247-51.
     
  93. Atmani K, Voigt M, Le Tien V, et al. Ranibizumab for retinal angiomatous proliferation in age-related macular degeneration. Eye. 2010;24:1193–1198.
     
  94. Querques G, Rousseau A, Forte R, et al. Longitudinal anatomical response of retinal-choroidal anastomosis to antivascular endothelial growth factor therapy. Retina. 2012;32:458–467.
     
  95. Hufendiek K, Hufendiek K, Panagakis G, Helbig H, Gamulescu MA. Visual and morphological outcomes of bevacizumab (Avastin®) versus ranibizumab (Lucentis®) treatment for retinal angiomatous proliferation. Int Ophthalmol. 2012 Jun;32(3):259-68.
     
  96. Kramann CA, Schöpfer K, Lorenz K, Zwiener I, Stoffelns BM, Pfeiffer N. Intravitreal ranibizumab treatment of retinal angiomatous proliferation. Acta Ophthalmol. 2012 Aug;90(5):487-91.
     
  97. Hsu TK, Liu JH, Lei J, Chao HM. Retinal angiomatous proliferation responds safely to a double dose (1.0 mg) of ranibizumab. Clin Exp Optom. 2013 Jan;96(1):112-6.
     
  98. Krebs I, Krepler K, Stolba U, Goll A, Binder S. Retinal angiomatous proliferation: combined therapy of intravitreal triamcinolone acetonide and PDT versus PDT alone. Graefes Arch Clin Exp Ophthalmol. 2008 Feb;246(2):237-43.
     
  99. Bakri SJ, Ekdawi NS. Intravitreal triamcinolone and bevacizumab combination therapy for refractory choroidal neovascularization with retinal angiomatous proliferation.Eye (Lond). 2008 Jul;22(7):978-80.
     
  100. Saito M, Shiragami C, Shiraga F, Nagayama D, Iida T. Combined intravitreal bevacizumab and photodynamic therapy for retinal angiomatous proliferation. Am J Ophthalmol. 2008 Dec;146(6):935-41.e1.
     
  101. Montero JA, Ruiz-Moreno JM, Sanabria MR, Fernandez-Munoz M. Efficacy of intravitreal and periocular triamcinolone associated with photodynamic therapy for treatment of retinal angiomatous proliferation. Br J Ophthalmol. 2009 Feb;93(2):166-70.
     
  102. Rouvas AA, Papakostas TD, Vavvas D, Vergados I, Moschos MM, Kotsolis A, Ladas ID. Intravitreal ranibizumab, intravitreal ranibizumab with PDT, and intravitreal triamcinolone with PDT for the treatment of retinal angiomatous proliferation: a prospective study. Retina. 2009 Apr;29(4):536-44.
     
  103. Shima C, Gomi F, Sawa M, Sakaguchi H, Tsujikawa M, Tano Y. One-year results of combined photodynamic therapy and intravitreal bevacizumab injection for retinal pigment epithelial detachment secondary to age-related macular degeneration. Graefes Arch Clin Exp Ophthalmol. 2009 Jul;247(7):899-906.
     
  104. Lo Giudice G, Gismondi M, De Belvis V, Cian R, Tavolato M, Galan A. Single-session photodynamic therapy combined with intravitreal bevacizumab for retinal angiomatous proliferation. Retina. 2009 Jul-Aug;29(7):949-55.
     
  105. Saito M, Shiragami C, Shiraga F, Kano M, Iida T. Comparison of intravitreal triamcinolone acetonide with photodynamic therapy and intravitreal bevacizumab with photodynamic therapy for retinal angiomatous proliferation. Am J Ophthalmol. 2010 Mar;149(3):472-81.
     
  106. Sahu AK, Narayanan R. Intravitreal ranibizumab, intravitreal ranibizumab with photodynamic therapy (PDT), and intravitreal triamcinolone with PDT for the treatment of retinal angiomatous proliferation. Retina. 2010 Jun;30(6):981.
     
  107. Viola F, Mapelli C, Villani E, Tresca Carducci F, Vezzola D, Ratiglia R. Sequential combined treatment with intravitreal bevacizumab and photodynamic therapy for retinal angiomatous proliferation. Eye (Lond). 2010 Aug;24(8):1344-51.
     
  108. Lee MY, Kim KS, Lee WK. Combination therapy of ranibizumab and photodynamic therapy for retinal angiomatous proliferation with serous pigment epithelial detachment in Korean patients: twelve-month results. Retina. 2011 Jan;31(1):65-73.
     
  109. Nakano S, Honda S, Oh H, Kita M, Negi A. Effect of photodynamic therapy (PDT), posterior subtenon injection of triamcinolone acetonide with PDT, and intravitreal injection of ranibizumab with PDT for retinal angiomatous proliferation. Clin Ophthalmol. 2012;6:277-82.
     
  110. Saito M, Iida T, Kano M. Combined intravitreal ranibizumab and photodynamic therapy for retinal angiomatous proliferation. Am J Ophthalmol. 2012 Mar;153(3):504-514.
     
  111. Lee MY, Kim KS, Lee WK. Combined intravitreal ranibizumab and photodynamic therapy for retinal angiomatous proliferation. Am J Ophthalmol. 2012 May;153(5):1004-5; author reply 1005.
     
  112. Rouvas AA, Chatziralli IP, Theodossiadis PG, Moschos MM, Kotsolis AI, Ladas ID. Long-term results of intravitreal ranibizumab, intravitreal ranibizumab with photodynamic therapy, and intravitreal triamcinolone with photodynamic therapy for the treatment of retinal angiomatous proliferation. Retina. 2012 Jun;32(6):1181-9.
     
  113. Shirakata Y, Shiragami C, Yamashita A, Nitta E, Fujiwara A, Shiraga F. One-year results of bevacizumab intravitreal and posterior sub-Tenon injection of triamcinolone acetonide with reduced laser fluence photodynamic therapy for retinal angiomatous proliferation. Jpn J Ophthalmol. 2012 Nov;56(6):599-607.
     
  114. Saito M, Iida T, Kano M. Two-year results of combined intravitreal anti-VEGF agents and photodynamic therapy for retinal angiomatous proliferation. Jpn J Ophthalmol. 2013 Mar;57(2):211-20.
     
  115. Forooghian F, Cukras C, Chew EY. Retinal angiomatous proliferation complicated by pigment epithelial tear following intravitreal bevacizumab treatment.Can J Ophthalmol. 2008 Apr;43(2):246-8.
     
  116. Gutfleisch M, Heimes B, Schumacher M, Dietzel M, Lommatzsch A, Bird A, Pauleikhoff D. Long-term visual outcome of pigment epithelial tears in association with anti-VEGF therapy of pigment epithelial detachment in AMD. Eye (Lond). 2011 Sep;25(9):1181-6.
     
  117. Introini U, Torres Gimeno A, Scotti F, Setaccioli M, Giatsidis S, Bandello F. Vascularized retinal pigment epithelial detachment in age-related macular degeneration: treatment and RPE tear incidence. Graefes Arch Clin Exp Ophthalmol. 2012 Sep;250(9):1283-92.
     
  118. Reche-Frutos J, Calvo-Gonzalez C, Pérez-Trigo S, Fernandez-Perez C, Donate-Lopez J, Garcia-Feijoo J. Ranibizumab in retinal angiomatous proliferation (RAP): influence of RAP stage on visual outcome. Eur J Ophthalmol. 2011 Nov-Dec;21(6):783-8.
     
  119. Kim JH, Chang YS, Kim JW, Lee TG, Kim H. Diagnosis of type 3 neovascularization based on optical coherence tomography images. Retina 2016;36: 1506-1515.
     
  120. Yannuzzi LA. Idiopathic polypoidal choroidal vasculopathy. Macula Society Meeting, 1982, Miami, FL.
     
  121. Kleiner RC, Brucker AJ, Johnston RL. The posterior uveal bleeding syndrome. Retina. 1990;10:9–17.
     
  122. Yannuzzi LA, Sorenson J, Spaide RF, Lipson B. Idiopathic polypoidal choroidal vasculopathy (IPCV). Retina. 1990;10: 1–8.
     
  123. Nakashizuka H, Mitsumata M, Okisaka S, et al. Clinicopathologic findings in polypoidal choroidal vasculopathy. Invest Ophthalmol Vis Sci. 2008;49:4729–4737.
     
  124. Stern RM, Zakov ZN, Zegarra H, Gutman FA. Multiple recurrent serosanguineous retinal pigment epithelial detachments in black women. Am J Ophthalmol. 1985;100:560–569.
     
  125. Yannuzzi LA, Ciardella A, Spaide RF, et al. The expanding clinical spectrum of idiopathic polypoidal choroidal vasculopathy. Arch Ophthalmol. 1997;115:478–485.
     
  126. Chang YC, Wu WC. Polypoidal choroidal vasculopathy in Taiwanese patients. Ophthalmic Surg Lasers Imaging. 2009; 40:576–58.
     
  127. Ciardella AP, Donsoff IM, Huang SJ, et al. Polypoidal choroidal vasculopathy. Surv Ophthalmol. 2004;49:25–37.
     
  128. Lafaut BA, Leys AM, Snyers B, et al. Polypoidal choroidal vasculopathy in Caucasians. Graefes Arch Clin Exp Ophthalmol. 2000;238:752–759.
     
  129. Yannuzzi LA; Ciardella A, Spaide RF, Rabb M, Orlock DA. The expanding clinical spectrum of idiopatic polypoidal choroidal vasculopathy. Arch Ophthalmol. 1997; 115: 478-85.
     
  130. Spaide RF, Yannuzzi LA, Slakter JS; Sorenson J, Orlach DA. Indocyanine green videoangiography of idiopatic polypoidal choroidal vasculopathy. Retina. 1995; 15: 100-110.
     
  131. Yuzawa M, Mori R, Kawamura A. The origins of polypoidal choroidal vasculopathy.Br J Ophthalmol. 2005; 89: 602-607.
     
  132. MacCumber MW, Dastgheib K, Bressler NM, Chan CC, Harris M, Fine S et al. Clinicopathologic correlation of the multiple recurrent serosanguineous retinal pigment epithelial detachments syndrome. Retina. 1994; 14: 143-152.
     
  133. Lafaut BA, Aisenbrey S, Van den Broecke C, Bartz-Schmidt KU, Heimann K. Polypoidal choroidal vasculopathy pattern in age-related macular degeneration. A clinicopathologic correlation. Retina. 2000; 20: 650-54.
     
  134. Terasaki H, Miyake Y, Suzuki T, Nakamura M, Nagasaka T. Polypoidal choroidal vasculopathy  treated with macular  translocation: clinical pathological correlation. Br J Ophthalmol. 2002; 86:321-27.
     
  135. Rosa RH Jr, Davis JL, Eifrig CW, Clinicopathologic correlation of idiopatic polypoidal choroidal vasculopathy. Arch Ophthalmol. 2002; 120: 502-508.
     
  136. Kuriwa S, Tateiwa H, Hisatomi T, Ishibashi T, Yoshimura N. Pathologic features of surgically excised polypoidal choroidal vasculopathy membranes. Clin Exp Ophthalmol.  2004;32:292-302.
     
  137. Okubo A, Sameshima M, Uemura A, Kanda S, Ohba N. Clinicopathological correlation of polypoidal choroidal vasculopathy revealed by ultrastructural study. Br J Ophthalmol 2002;86:1093-98.
     
  138. Nakajima M, Yuzawa M, Shimada H, Mori R. Correlation between indocyanine green angiographic findings and histopathology of  polypoidal choroidal vasculopathy. Jpn J Ophthalmol. 2004; 48: 249-55.
     
  139. Koh AH, Chen LJ, Chen SJ, Chen Y, Giridhar A, Iida T, Kim H, Yuk Yau Lai T, Lee WK, Li X, Han Lim T, Ruamviboonsuk P, Sharma T, Tang S, Yuzawa M; on behalf of the Expert PCV Panel. Polypoidal Choroidal Vasculopathy: Evidence-based Guidelines for Clinical Diagnosis and Treatment. Retina. 2013 Apr;33(4):686-716.
     
  140. Lee MW, Yeo I, Wong D, Ang CL. Argon laser photocoagulation for the treatment of polypoidal choroidal vasculopathy. Eye (Lond). 2009;23:145–148.
     
  141. Spaide RF, Donsoff I, Lam DL, et al. Treatment of polypoidal choroidal vasculopathy with photodynamic therapy. Retina. 2002; 22(5):529–535.
     
  142. E. Akaza, M. Yuzawa, Y. Matsumoto, S. Kashiwakura, K. Fujita, and R. Mori, “Role of photodynamic therapy in polypoidal choroidal vasculopathy,” Jpn J Ophthalmol. 2007;51(4):270–277.
     
  143. Gomi F, Ohji M, Sayanagi K, et al. One-year outcomes of photodynamic therapy in age-related macular degeneration and polypoidal choroidal vasculopathy in Japanese patients. Ophthalmology. 2008;115(1):141–146.
     
  144. Akaza E, Mori R, Yuzawa M. Long-term results of photodynamic therapy of polypoidal choroidal vasculopathy. Retina. 2008;28(5):717–722.
     
  145. Otani A, Sasahara M, Yodoi Y, et al. Indocyanine green angiography: guided photodynamic therapy for polypoidal choroidal vasculopathy. Am J Ophthalmol. 2007;144(1):7–e1.
     
  146. Silva RM, Figueira L, Cachulo ML, Duarte L, Faria de Abreu JR, Cunha-Vaz JG. Polypoidal choroidal vasculopathy and photodynamic therapy with verteporfin. Graefes Arch Clin Exp Ophthalmol. 2005 Oct;243(10):973-9.
     
  147. Kurashige Y, Otani A, Sasahara M, et al. Two-year results of photodynamic therapy for polypoidal choroidal vasculopathy. Am J Ophthalmol. 2008;146(4):513–22.
     
  148. Akaza E, Yuzawa M, Mori R. Three-year follow-up results of photodynamic therapy for polypoidal choroidal vasculopathy. Jpn J Ophthalmol. 2011 Jan;55(1):39-44.
     
  149. Kim SJ, Yu HG. Efficacy of combined photodynamic therapy and intravitreal bevacizumab injection versus photo- dynamic therapy alone in polypoidal choroidal vasculopathy. Retina. 2011; Oct;31(9):1827-34.
     
  150. Lai TY, Chan WM, Liu DT, Luk FO, Lam DS. Intravitreal bevacizumab (Avastin) with or without photodynamic therapy for the treatment of polypoidal choroidal vasculopathy.Br J Ophthalmol. 2008 May;92(5):661-6.
     
  151. Eandi CM, Ober MD, Freund KB, Slakter JS, Yannuzzi LA. Selective photodynamic therapy for neovascular age-related macular degeneration with polypoidal choroidal neovascularization. Retina. 2007 Sep;27(7):825-31.
     
  152. Ricci F, Calabrese A, Regine F, Missiroli F, Ciardella AP. Combined reduced fluence photodynamic therapy and intravitreal ranibizumab for polypoidal choroidal vasculopathy. Retina. 2012 Jul;32(7):1280-1288.
     
  153. Koh A, Lee WK, Chen LJ, Chen SJ, Hashad Y, Kim H, Lai TY, Pilz S, Ruamviboonsuk P, Tokaji E, Weisberger A, Lim TH. EVEREST study: efficacy and safety of verteporfin photodynamic therapy in combination with ranibizumab or alone versus ranibizumab monotherapy in patients with symptomatic macular polypoidal choroidal vasculopathy. Retina. 2012 Sep;32(8):1453-64.
     
  154. Fernández M, Gil M, Gomez-Ulla F, Charlón P. Verteporfin photodynamic therapy combined with intravitreal ranibizumab for polypoidal choroidal vasculopathy controversy concerning long-term followup. Case Rep Med. 2012;2012:897097.
     
  155. Stangos AN, Gandhi JS, Nair-Sahni J, Heimann H, Pournaras CJ, Harding SP Polypoidal choroidal vasculopathy masquerading as neovascular age-related macular degeneration refractory to ranibizumab. Am J Ophthalmol. 2010 Nov;150(5):666-73.
     
  156. Cheung CM, Yanagi Y, Mohla A, Lee SY, Mathur R, Chan CM, Yeo I, Wong TY. Characterization and differentiation of polypoidal choroidal vasculopathy using swept source optical coherence tomography angiography. Retina. 2016 Nov 8. 0:1–11 [Epub ahead of print].
     
  157. Novais EA, Ferrara D, Waheed NK. Optical coherence tomography in polypoidal choroidal vasculopathy disease. Clin Exp Ophthalmol. 2015 Dec;43(9):779-81.
     
  158. Inoue M, Balaratnasingam C, Freund KB. Optical coherence tomography angiography of polypoidal choroidal vasculopathyand polypoidal choroidal neovascularization. Retina. 2015 Nov;35(11):2265-74.
     
  159. Kokame GT, Hirai K, Yanagihara R. Polypoidal Choroidal Vasculopathy: Imaging by Indocyanine Green Angiography and En Face Optical Coherence Tomography. JAMA Ophthalmol. 2015 Nov;133(11):e151886.

Fundus autofluorescence in age-related macular degeneration

Fundus autofluorescence in age-related macular degeneration

English

Authors:

José Mª Ruiz-Moreno, MD, PhD.

President of the Spanish Society of Retina and Vitreous (SERV).
Professor of Ophthalmology, Castilla La Mancha University (UCLM).
Head of the Ophthalmology Department, Puerta de Hierro University Hospital. Madrid.
Medical Director Vissum Corporation. Spain.

 

Javier A Montero, MD, PhD

Head of the Ophthalmology Department, Rio Hortega Hospital, University of Valladolid. Valladolid.
Retina Unit, Oftalvist. Madrid. 
Associate Professor of Ophthalmology, European University, Madrid. Spain.

 

Jorge Ruiz-Medrano, MD.

Fellow of surgical and medical retina. Jules Gonin Eye Hospital. Fondation Asile des Aveugles. Lausanne. Switzerland.

 

Acknowledgements:

The authors have neither economical nor commercial interest in the devices and procedures described.

The authors wish to thank TOPCON Spain SA for its help providing the Spaide AF filter Set and adapting it to the Topcon TRC-50DX camera.

All the images presented in this chapter were obtained using this system and Spectralis OCT.

 

Updated/reviewed by the authors, July 2017.

Definitions

Fluorescence is the capability of absorbing light at a specific wavelength and releasing it at a longer, less energetic wavelength. This phenomenon raises an especial interest when the released radiation is found within the spectrum of visible light permitting its visualization, recording and measurement.

Autofluorescence is the spontaneous fluorescence that some substances present naturally.

Fluorophore is the part of a molecule that makes it fluoresce.

The human eye contains autofluorescent substances in the retina, especially within the retinal pigment epithelium (RPE). The main autofluorescent component of the RPE is lipofuscin (LF), containing at least ten different fluorophores presenting discrete emission spectra within the green, golden-yellow, yellow-green, and orange-red emitting range(1).

When LF granules are stimulated with light within the blue range, a characteristic yellow fluorescence is emitted(2).

 

Basic considerations on fundus autofluorescence

The RPE plays an important role in the physiopathology of age-related macular degeneration (AMD)(3, 4). The study of RPE can help to achieve a better understanding of AMD and to find new ways of early diagnosis as well as new prognostic and progression markers for this condition. LF originates from the constant phagocytosis of the shed outer segment disks of the photoreceptors and is accumulated in the cytoplasm of RPE cells(5,6). This accumulation is considered to be a hallmark of RPE aging (7). Some experimental studies have addressed the molecular mechanisms of the interaction of excessive LF with the normal cellular functions of RPE(8,9). According to these studies, A2-E (N-retinylidene-N-retinylethanol-amine) has been identified as the main fluorophore of LF. A2-E may play a toxic effect including phototoxic and detergent actions, as well as an inhibitory effect on lysosomal function(10,11). It has been suggested that the photo-oxidation by-products related to LF may trigger the complement cascade, thus contributing to the pathological chronic inflammation of the macular area(12). New studies have been designed to improve our understanding of potential underlying molecular mechanisms. The autofluorescence of LF, its distribution in post-mitotic human RPE cells and its accumulation with age have been extensively studied in post-mortem eyes with fluorescence microscopy(14,15). In the past few years new research on ocular autofluorescence in vivo has also been performed.

Ultraviolet light is frequently used to visualize LF by fluorescence microscopy ex vivo, since light absorption limits the transmission of ultraviolet light within the retina in human living eyes. However, due to the wide range of excitation of LF (from 300 nm to 600 nm), visible light can be used to visualize its fluorescence in vivo. The emitted spectrum ranges from 480 to 800 nm and peaks within the range of 600 to 640 nm(15).

Fundus autofluorescence (FAF) imaging of the human living eye is a relatively new imaging method that provides a topographic map of the distribution of LF in the RPE. The detection of FAF is limited by its low intensity (approximately two orders of magnitude lower than the peak background fluorescence of an ordinary fluorescein angiography), and by the FAF characteristics of the anatomic structures of the eye, including those of the optical media, especially of the lens(16).

 

Imaging methods

1. Fundus spectrophotometry

Fundus spectrophotometry was developed by Delori et al. and was designed to determine the spectrum of excitation and fluorescence emission from small areas of the retina (2º diameter)(17,18). The authors were able to determine the amount of FAF and compare it with in vitro fluorescence microscopy. They found out that the spectrum of in vivo excitation was slightly broader, and peaked at a longer wavelength than those of A2E and native LF, concluding that considering the spatial distribution, spectral characteristics and age relationship, LF is the main source of fluorescence in the FAF in vivo(18).

Currently there are two available systems to examine the FAF of the human eye in vivo in the clinical practice: confocal scanning laser ophthalmoscope and fundus camera.

 

Scanning laser ophthalmoscope

Confocal scanning laser ophthalmoscope (cSLO) was originally developed by Webb et al. using a low-energy laser source to scan the retina in two directions termed as x and y(19). The confocal nature of the optics ensures that the reflectance and fluorescence correspond to the same focal plane. cSLO overcomes the limitations of the low-intensity signal of FAF and the lens interferences. The defocused light is almost completely suppressed, thus reducing the FAF from the optical media anterior to the retina, such as the lens or the cornea. In order to reduce the background noise and to increase the contrast of the image, a series of FAF images are usually recorded(20,21). Following the aligning of the images in order to correct the movement of the eye during the acquisition, the final image is calculated (usually from 4-32 frames) and the values of the pixels are normalised. 

The FAF image can be obtained with low excitation energies within the limits of maximum retinal irradiance established by the American National Standard Institute and other international standards(22). cSLO enables the acquisition of FAF images from wide areas of the retina (55º with one frame and even larger areas using the composite mode)(20,22). Although limited by the optical properties of the human eye, SLO succeeds in imaging the posterior pole with a high contrast(23).

Currently, there are three different cSLO systems for FAF imaging: the Heidelberg Retinal Angiograph (HRA) (based on the HRA classic, HRA 2 and the Spectralis HRA) (Heidelberg Engineering, Dossenheim, Germany); the Rodenstock cSLO (RcSLO; Rodenstock, Weco, Düsseldorf, Germany); and the Zeiss prototype SM 30 4024 (ZcSLO; Zeiss, Oberkochen, Germany). HRA is the only currently commercially available system with the cSLO system to capture FAF images. HRA uses an excitation wavelength of 488 nm from an Argon laser or a solid-state laser. A barrier filter with a short-wavelenght cut-off at 500 nm is inserted just opposite the detector, blocking the laser light and letting the autofluorescent light through.

There are currently 3 kinds of cSLO-based FAF techniques:

  1. Short wavelength FAF (486 nm blue laser): The most extended cSLO since 2009, which basically excites the RPE LF;
     
  2. Long wavelength FAF (787 nm infrared laser): Less used, it excites the melanin of the RPE mainly;
     
  3. 518 nm green laser FAF has also been employed lately. This wavelength is absorbed by the macular pigment to a lesser extent, showing a clearer image of the foveal RPE. 

The device uses both lasers and it is able to simultaneously capture both images and use them to measure the distribution of the macular pigment. Blue FAF images acquired with a cSLO operating at an excitation wavelength of 486 nm present a dark area in the fovea caused by absorption of blue light by the macular pigment. However, green FAF images acquired at a wavelength of 518 nm show much less of a shadowing effect as green light is less absorbed by macula pigment (Figure 1). Both blue and green FAF images can be used to quantify the macular pigment distribution(24).

In patients with geographic atrophy (GA) that present with foveal sparing, the border of the atrophic area is easily recognizable on green FAF images. The quantification of areas of GA becomes more reliable when using green FAF images. The patterns of hyperfluorescence seem similar in blue and green FAF images. Meanwhile, it has been made possible to acquire real time images, a technique known as real-time averaging.

fundus autofluopresence_image1.png

Figure 1. Differences between blue and green FAF.

 

The Rodenstock cSLO and the Zeiss prototype SM 30 4024 have also been used to acquire clinical FAF images. Both systems use an excitation wavelength of 488 nm (the same as the HRA), and barrier filters at 515 nm and 521 nm, respectively(20,21,25). Bellmann et al. have noticed marked contrast and brightness differences as well as in the grey range (an important marker of the image quality between the different systems of cSLO). These limitations must be taken into consideration when comparing images from different cSLO systems(25).

The default software of the HRA system normalizes the pixel distribution of the final image in order to improve the distribution of the FAF intensity. Even though this final step facilitates the evaluation of the localized topographic differences, it allows a relative estimation of the intensities of the FAF. Thus, it should not be used for quantitative calculation and absolute comparison between different FAF images. The normalization of the average images can be easily turned off, and brightness and contrast can be manually adjusted to permit an adequate visualization of the distribution of FAF in areas with a very high or very low signal in order to improve the visualization of small details.

 

Fundus camera

Fundus cameras are widely used in clinical routine for imaging the retina as fundus photographs, reflectance photographs and fluorescein angiography. Fundus cameras use a single flash to capture images from large retinal areas. When confocal optics are not available, the FAF signal from all the ocular structures with fluorescent properties reaches the camera, and scattered light, anterior and posterior to the plane of interest can influence the detected signal(26-28). The lens contributes significantly to the FAF signal when similar wavelengths are used in the blue-range, as for the cSLO (wavelength = 488 nm), particularly in older patients with lens yellowing and nuclear opacities. Flashlight intensities and detector gain have to be set at relatively high levels in order to obtain reasonable FAF images. However, the signal-to-noise ratio decreases simultaneously which may result in reduced image quality.  To reduce interference from lens fluorophores that mainly emit in the range between 510 to 670 nm, Spaide modified the excitation filter (peak 580 nm, bandwidth 500-610 nm) and the barrier filter (peak 695 nm, bandwidth 675-715 nm). 

A further modification was introduced in 2007 using a slightly different filter set (excitation bandwidth 535-580 nm, emission bandwidth 615-715 nm)(29), thus improving signal-to-noise ratio and image quality. Furthermore, as this setup of the fundus camera uses different excitation and emission filters compared with the cSLO, it may even visualize other retinal fluorophores. However, a systematic comparison of different pathologies with clinico-pathological correlations between cSLO and fundus camera, particularly in patients with AMD, has not yet been performed.

In this chapter we will present a series of images of patients suffering from different forms of AMD using both fundus camera and Spectralis cSLO. 

Originally, a fundus camera that enabled imaging with a field of 13º was used. Afterwards, Spaide has obtained images of the spatial distribution of FAF intensities over larger retinal areas up to 50º with his new modified fundus camera(26-29). In the near future it will be possible to improve FAF imaging with the aid of new filters and some other innovations, and increased experience. Furthermore, it is already possible to visualise different fluorophores from the retina with the configuration of the fundus camera using excitation and emission filters for the cSLO.

 

Autofluorescence imaging in the human eye in vivo

FAF images show the spatial distribution of the intensity of FAF of each pixel in grey values (arbitrary values from 0 to 225); low intensities are commonly known as low pixel values (dark) and high intensities as high pixel values (light).

 

Normal fundus

FAF imaging shows a consistent pattern of FAF distribution in normal eyes(21). Such common findings have been reported in children as young as four years old(30). The macular FAF signal is reduced at the fovea because it is limited by the presence of lutein and zeaxanthin in the neurosensory retina. The signal is higher in the parafoveal area and tends to increase as we move away from it, peaking at the most peripheral retinal areas. It has been suggested that this FAF pattern is caused by the melanin deposition and density of LF granules at the different areas of the retina(18,31). The optic nerve head typically appears dark mainly due to the absence of RPE. The retinal vessels are associated with a markedly reduced FAF signal because of the blocked fluorescence (Figure 2).

fundus autofluopresence_image2.png

Figure 2. Colour fundus and FAF from a normal subject. (A) Right eye colour fundus and (B) FAF photographs. (C) Left eye colour fundus and (D) FAF photographs.

 

The common ratios of gGeographic atrophy (GA)rey intensity between the fovea and the perifoveal area have been established(32,33). Considering these findings, any deviation from the normal pattern in a specific location can be easily identified; hence the qualitative description of the local changes in the FAF is widely used. The changes in signal intensity are qualitatively described as decreased, normal, or increased as compared to the background signal of the same eye.

 

FAF imaging in AMD

When examined with FAF, the fundus of patients with AMD may show a range of signal changes(20,34-38). Assuming that RPE has an important role in the pathophysiology of AMD and that the major fluorophores in the retina are located within RPE cells, FAF imaging can show changes in the concentration and distribution of RPE LF and hence establish the condition of RPE in patients with AMD. Therefore, atrophic RPE typically appears as dark patches in FAF and can be clearly delineated, even better than in normal fundus photograph (Figure 3)(21,39).

fundus autofluopresence_image3.png

Figure 3. Atrophic area of RPE. These areas typically appear as dark patches in FAF images and can be clearly delineated. (A) Colour fundus and (B) FAF photographs.

 

All this information can be obtained from a quick and minimally invasive exploration with FAF. The decreased FAF intensity may also be associated with hyperpigmented areas due to the melanin absorption of light(36,40). However, it should be considered that other fluorophores than LF can be found in RPE and become more prominent in AMD patients, and hyperpigmented areas may also cause an increase in the signal, which is supposed to result from the accumulation of melanolipofuscin.Other changes in FAF which are not related to RPE defects may appear in AMD. Fresh haemorrhages typically appear dark due to blocked fluorescence (Figure 4). However, these haemorrhagic areas eventually synthesize substances and fluorophores, which are observed in the fundus as yellowish areas and in FAF images as increased signals (Figure 5)(41).

fundus autofluopresence_image4.png

Figure 4. Fresh haemorrhage in the left eye from a patient with choroidal neovascularization (CNV) secondary to AMD. Fresh haemorrhages typically appear dark due to blocked fluorescence. (A) Colour fundus and (B) FAF photographs.

 

fundus autofluopresence_image5.png

Figure 5. Haemorrhage in the left eye of a patient with CNV secondary to AMD. Fluorophores eventually appear in haemorrhagic areas which are observed as yellowish areas in the fundus and as an increased FAF signal. (A) Colour fundus photograph of the fresh haemorrhage. (B) Colour fundus photograph from the same eye one month later. (C) FAF of the haemorrhage.
 

Pigment epithelial and neurosensory detachment and areas with extracellular fluid accumulation associated with exudative lesions can be observed in FAF as increased or decreased signal intensity. Fluid accumulation under pigment epithelium detachment (PED), extracellular deposition of material under the RPE (drusen), and fluid originated from CNV can occur with increased, normal or decreased FAF intensity. This phenomenon is a consequence of the presence of unknown autofluorescent molecules other than LF, in the same spectral range than LF. FAF imaging alone may not distinguish between melanolipofuscin from RPE cells migrated into the neurosensory retina and LF within the normal RPE layer. It is always necessary to compare the FAF findings with those from other techniques such as fundus photograph, reflectance image, fluorescein angiography or optical coherence tomography (OCT)(36, 40).

 

FAF in early AMD

Early AMD is characterised by the appearance of localized RPE hypo or hyperpigmentation and drusen. Drusen are formed by the accumulation of extracellular deposits in the inner aspects of Bruch’s membrane(3). Depending on their size and morphology, they can be classified as hard or soft drusen. The molecular composition of drusen is quite complex and has not been completely elucidated (Figure 6).

fundus autofluopresence_image6.png

Figure 6. Small drusen from a patient with early AMD in the left eye. (A) Colour fundus and (B) FAF photographs.

 

FAF changes in early AMD have already been reported by several authors(9,21,26,28,34-37,42); all of them concluding that the changes in ophthalmoscopy and fluorescein angiography are not necessarily related to FAF, suggesting that FAF may provide new information regarding the stages and activity of the disease.

Differentiation between RPE LF and sub-RPE deposits with FAF images in vivo can be a hard work. An analysis of the variability of FAF in patients with early AMD was reported in an international workshop on FAF phenotype in early AMD. Among their conclusions, a new classification system with eight different FAF patterns was given(40).

Normal pattern characterized by a homogeneous background autofluorescence with a gradual fluorescence decrease in the inner macula towards the foveola (blocked fluorescence caused by yellow macular pigments). AF may be normal even in the presence of soft or hard drusen. 

Minimal change pattern characterized by a limited and irregular decrease or increase of background FAF, not associated to any obvious or important topographic pattern.

Focal increased pattern is defined by the presence of at least one well defined spot (<200 micron diameter) of markedly increased FAF much brighter than the surrounding background fluorescence. These areas may or may not correspond to large, soft drusen and to areas of hyperpigmentation.

Patchy pattern, characterized by the presence of at least one large area with well-defined borders (>200 micron diameter) of markedly increased FAF. Again, these areas may or may not correspond to large, soft drusen and areas of hyperpigmentation.

Linear pattern defined by the presence of at least one linear area of markedly increased FAF with well-defined borders. Linear structures of increased FAF usually correspond to hyperpigmented lines. This pattern shows multiple branching lines of increased FAF forming a lacelike pattern. The borders may be hard to define, and FAF may gradually decrease from the centre of the linear area towards the surrounding background. This pattern may correspond to hyperpigmentation or to non-visible abnormalities.

Reticular pattern is defined by the presence of multiple small, well-defined areas (<200 micron diameter) of decreased FAF. This pattern has been found to occur not only in the macular area, and may be associated with multiple small soft drusen, hard drusen or areas with pigment changes or non-visible abnormalities.

Speckled pattern is characterized by the simultaneous presence of different types of abnormalities in a large area. The changes reach beyond the macular area and may cover the entire posterior fundus. These abnormalities include multiple small areas of irregular increased or decreased FAF corresponding to hyper and hypopigmented areas and multiple subconfluent and confluent drusen.

The speckled pattern has been reported to be the most frequent (26%) followed by the patchy pattern (23%). The most infrequent patterns are the normal pattern (2%) and the lacelike pattern (2%). The study confirmed that visible drusen on fundus photography are not always correlated with noticeable FAF changes and that areas of increased FAF may or may not correspond to areas of hyperpigmentation or soft or hard drusen. Several authors have also mentioned the different FAF patterns in eyes with drusen.

Delori et al. described a pattern consisting of decreased FAF in the centre of the drusen surrounded in most of the cases by a ring of increased FAF(27). They also observed that the decreased drusen signal was not as intense as in the areas with RPE atrophy. The authors hypothesized that it might be caused by a displacement of the cytoplasm and LF granules in RPE cells instead of an actual RPE atrophy(36).

Von Rückmann et al. further reported that crystalline drusen are characterised by a decrease in FAF, signalling the onset of atrophy(21). Lois et al. confirmed that areas of confluent drusen are usually associated with focal, mildly increased FAF and that only large subfoveal soft drusen (drusenoid RPE detachments) topographically correspond with focal changes of FAF (Figure 7)(33).

fundus autofluopresence_image7.png

Figure 7. Drusen in a patient with early bilateral AMD. (A) Right eye colour fundus and (B) FAF photographs. (C) Left eye colour fundus and (D) FAF photographs.

 

Smith et al. reported their results after using image analysis software to study drusen and pigmented areas on fundus photographs from AMD patients(42). The authors initially used image analysis algorithms, including automated background levelling and thresholding. Areas of focally increased FAF intensities were compared to the normal background signal. By overlapping fundus photographs and FAF, the topographic correlation of drusen and pigmented areas with focally increased FAF signals was established.
Smith and co-workers reported that eyes with isolated drusen or pigment abnormalities were better correlated with FAF abnormalities than eyes with GA(42).

Regarding areas with changes in RPE, hypopigmented areas are usually associated with a corresponding decreased FAF signal, suggesting an absence or degeneration of RPE cells, with reduced content of LF granules (Figures 8 and 9). 

fundus autofluopresence_image8.png

Figure 8. RPE hypopigmentation in the macular area secondary to AMD. Hypopigmented areas are usually associated with correspondingly decreased FAF signals, suggestive of RPE cells loss or degeneration with reduced content of LF granules. (A) Colour fundus and (B) FAF photographs.

 

fundus autofluopresence_image9.png

Figure 9. RPE hypopigmentation in the macular area secondary to AMD. (A) Colour fundus and (B) FAF photographs.

 

However, hyperpigmented areas frequently show a higher FAF signal, which may be caused by a higher concentration of autofluorescent melanolipofuscin (Figure 10)(36).

fundus autofluopresence_image10.png

Figure 10. RPE hyperpigmentation in the macular area secondary to AMD. Areas with hyperpigmentation frequently show a higher FAF signal which may be caused by a higher amount of autofluorescent melanolipofuscin. (A) Colour fundus and (B) FAF photographs.

 

Reticular pseudodrusen are seen as small triangular deposits above the RPE in OCT imaging. Infrared pictures show them as small white spots with a surrounding darker halo and FAF typically show an orange skin pattern (Figure 11). 

fundus autofluopresence_image11.png

Figure 11. Reticular pseudodrusen. FAF and SD-OCT of a patient with bilateral reticular pseudodrusen.

 

FAF in advanced AMD

Advanced AMD is characterized by geographic atrophy (GA), choroidal neovascularization (CNV), pigment epithelial detachment (PED), RPE tears and disciform scars.

Geographic atrophy (GA)

GA is thought to be the natural end stage of the atrophic AMD process when CNV does not appear. GA occurs in areas where the RPE is dead and the outer neurosensory retina and choriocapillaris disappeared(43,44). Due to the loss of RPE and LF, the atrophic area appears dark in FAF imaging(36). High contrast between the atrophic and the non-atrophic retina defines the area of GA more precisely than colour fundus photographs, permitting a clearer and more specific study of GA, as well as its natural development and evolution (Figures 12 and 13)(39,45).

fundus autofluopresence_image12.png

Figure 12. GA secondary to AMD. The loss of RPE and LF causes a characteristic absence of FAF signal (a dark area in FAF images). (A) Right eye colour fundus and (B) FAF photographs. (C) Left eye colour fundus and (D) FAF photographs.

 

fundus autofluopresence_image13.png

Figure 13. GA secondary to AMD. The high contrast between atrophic and non-atrophic retina enables the delineation of the atrophic area more precisely than can be performed from conventional fundus photographs. (A) Colour fundus and (B) FAF photographs.

 

The GA patches usually become larger and coalesce as AMD progresses(46,47). An excessive accumulation of LF, and therefore an increased FAF in the junction are highly suggestive of the appearance or progression of pre-existing GA (Figure 14).

fundus autofluopresence_image14.png

Figure 14. GA secondary to AMD with increased autofluorescence in the junction. New areas of GA and the extension of pre-existing areas are characterised by an excessive accumulation of LF, and therefore an increased FAF signal. (A) Colour fundus and (B) FAF photographs.

 

Preliminary observations suggest that different phenotypes may appear associated with junction FAF changes(48)

A new classification for junction FAF patterns has been proposed in GA patients(Figure 15)(49).

fundus autofluopresence_image15.png

Figure 15. GA secondary to AMD with a banded pattern of increased autofluorescence in the junction. (A) Colour fundus and (B) FAF photographs.

 

Proposed classification for junction FAF patterns in GA(49):

  • Focal increased FAF is defined by single or multiple spots of focal markedly increased FAF localized at the border of the atrophic patch;
  • Band pattern of increased FAF is characterized by a continuous stippled band of increased FAF surrounding the entire atrophic area;
  • Patchy increased FAF are large patches of increased FAF outside the GA area. FAF tends to be less intense than that in the focal pattern described above;
  • Diffuse increased FAF is the most frequent pattern of increased FAF in eyes with GA.

 

FAF changes are not limited to the border of the atrophic area and may show inter individual differences that have been further classified into four subtypes:

  • Reticular pattern, characterised by several lines of increased FAF usually following a radial pattern;
  • Branching pattern shows a diffusely increased FAF with a fine branching pattern of increased FAF;
  • Fine granular pattern is defined by a large area of increased FAF with a granular appearance surrounding the GA area and a clear border between the granular increased FAF and the surrounding normal background FAF;
  • Fine granular with peripheral punctate spots pattern is characterised by diffuse FAF changes surrounding the atrophic area with elongated small lesions and increased FAF.
     

Refined phenotypes help to identify the prognosis and seem to be a prerequisite to determine specific genetic factors in a complex, multifactorial disease such as AMD. An analysis of the follow-up of junction FAF patterns in GA and the rate of progression of atrophic lesions revealed that variation in GA growth rates are dependent on the specific phenotype of FAF at baseline(50). Atrophy enlargement was slowest in eyes with normal FAF pattern (median, 0.38 mm2/year), followed by focal FAF pattern (median, 0.81 mm2/year), diffuse FAF pattern (median, 1.77 mm2/year) and banded FAF pattern (median, 1.81 mm2/year). The rate of progression of GA in eyes with patchy FAF pattern were not included in this analysis because of their low frequency, insufficient for statistical analysis. The rate of progression of band and diffuse FAF patterns were significantly higher compared to eyes without FAF abnormalities and focal FAF pattern.

Another interesting finding of this study was the identification of eyes with extremely rapid progression of the atrophy, showing distinct FAF features of atrophy that had not been previously reported.

The authors introduced the term diffuse trickling for a pattern associated with a significantly faster enlargement of atrophy. 

Areas with increased FAF and consequently higher concentrations of RPE LF precede the development of new areas of GA or the enlargement of the pre-existing atrophic areas(50).

The phenotypic features of FAF abnormalities may play a stronger influence on the progression of atrophy than any other previously reported risk factors such as smoking, arterial hypertension or diabetes. 

The different rates of enlargement of atrophy may be related to heterogeneity at a cellular and molecular level in the disease. The high degree of symmetry in GA suggests that genetic determinants may be involved, rather than nonspecific aging changes.

GA is a relatively slow progressing disease. Currently, FAF patterns are mostly accepted as prognostic predictors for GA progression. 

Risk factors for the progression of GA include the perimeter of damaged RPE surrounding the GA lesion, the number and size of foci and the FAF pattern. 

It has been shown that hyperautofluorescence of the junction zone surrounding regions of GA is related to a higher progression rate. Holz et al.(50) showed that progression rates in eyes with the banded and the diffuse FAF pattern (especially those with diffuse trickling pattern) were significantly higher compared to eyes with focal patterns or with no FAF abnormalities, and these changes were later confirmed by other authors(51).

Other authors evaluated initial GA lesions and grouped them into three categories: predominantly hyperautofluorescent (hyperAF), both hyper- and hypoautofluorescent (mixed AF), or predominantly hypoautofluorescent (hypoAF). In their series, the FAF characteristics were significantly dependent on the type of atrophic area: initial GA lesions were most commonly mixed AF, while drusen-associated atrophy was most commonly hypoAF(52). Progression of GA areas seems to be symmetrical between both eyes of the same patient and in some series it has been observed that lesion growth is faster in larger lesions, suggesting that progression rate is related to basal area(53). However these findings are controversial. Some groups support a linear model of growth over a quadratic model(54). A square root transformation of baseline GA area appears to ameliorate the effect of baseline area on GA progression. A classification based on the area and the irregularity of the lesion perimeter shows that there is an association with growth rates by providing a measure of the relative border of RPE at risk, since the extent of the junction zone of damaged RPE increases with non-circularity for a GA area(55,56). Measuring the lesion growth in mm2 may not accurately reflect the relative dynamics of progression of the lesion, and a concept such as proportion of growth might be more intuitive. Growth of atrophy is not linear within time, no matter what measurement is used (mm2, border, proportion or square root).

Caire et al. tried to identify the genetic risk factors that contribute to the presence and progression of established GA in a Spanish population. According to their results, genetic risk factors associated with the presence of GA (genetic polymorphisms within CFH, ARMS2 and FHR1-3) are not identical to those related to its progression (polymorphisms of CFB and CFH)(57)

Some samples of patients with GA can be seen in Figures 16, 17 and 18 (FAF defect progression during one year of follow-up).

fundus autofluopresence_image16.png

Figure 16. Patient with GA: FAF defect progression during one year of follow-up (example 1).

 

fundus autofluopresence_image17.png

Figure 17. Patient with GA: FAF defect progression during one year of follow-up (example 2).

 

fundus autofluopresence_image18.png

Figure 18. Patient with GA: FAF defect progression during one year of follow-up (example 3).

 

fundus autofluopresence_image19.png

Figure 19. FAF patterns in GA

 

fundus autofluopresence_image20.png

Figure 20. Atrophic AMD progression. FAF of a 78-year-old woman suffering from an atrophic form of AMD. Progression of the disease can be clearly appreciated over a 6 year follow-up (A to F).

 

fundus autofluopresence_image21.png

Figure 21. GA case 1: 88-year old woman showing a bilateral atrophic AMD. Although the atrophy is extensive, the foveal centre is still preserved and her visual acuity remains at 0.4 and 0.7 respectively.

 

fundus autofluopresence_image22.png

Figure 22. GA case 2: 71-year-old male suffering from dry AMD. FAF shows us the hyperfluorescent drusen and the limits of the GA which allows us to properly follow these patients.

 

fundus autofluopresence_image23.png

Figure 23. Study of GA with FAF and SD-OCT. Study of dry AMD combining FAF and SD-OCT. The disappearance of layers of the external retina is clearly visible.

 

Choroidal neovascularization (CNV)

CNV is considered to cause almost 90% of the cases of severe visual loss related to AMD(58). CNV is usually studied by fluorescein angiography and OCT to assess the extent, location and nature of the lesion(58). Fluorescein angiography shows changes in retinal vascularization, but does not reveal how deeply RPE is affected.FAF imaging shows RPE damage, with the advantage that is a non-invasive test, less time consuming than angiography. Several studies have reported that CNV may show irregular FAF alternating areas of increased, normal and decreased fluorescence intensity (Figure 24)(20,35-38). These studies have also reported that areas with previously high levels of FAF may show decreased FAF six months later(37). Data comparing FAF findings in occult and classic CNV are limited. These changes may be secondary to photoreceptor loss, RPE atrophy, replacement of normal phenotypes of RPE cells with scar and increased melanin deposition. These findings may have therapeutic implications and clarify long-term visual prognosis. For example, a person with an active CNV on fluorescein angiography and normal FAF may show a much better outcome than another with an abnormal basal FAF. Spital et al. reported that classic CNV usually shows more focal areas of decreased FAF than occult CNV(35). These findings have been confirmed by McBain et al.(61) who guessed that low FAF at the site of the CNV are related to blocked fluorescence induced by the presence of CNV in the subretinal space, rather than to severe damage to the RPE. Another study did not find significant differences in FAF patterns in early classic and occult CNV secondary to AMD (Figures 25 and 26)(60). A continuous preserved autofluorescence pattern was observed in the central macula in most of the cases. These findings suggest that neovascular complexes, regardless if classic or occult, would be external to the RPE in most cases. Additional studies with a higher number of patients and longer monitoring are required to verify with these changes in patients with CNV (Figure 27).

fundus autofluopresence_image24.png

Figure 24. Occult CNV. These lesions show irregular FAF intensities with alternating areas of increased, normal and decreased signal intensity. (A) Colour fundus and (B) FAF photographs. (C) and (D) Early frames fluorescein angiography. (E) Late frames fluorescein angiography.

 

fundus autofluopresence_image25.png

Figure 25. Retinal angiomatous proliferation (RAP). (A) Colour fundus and (B) FAF photographs.

 

fundus autofluopresence_image26_0.png

Figure 26. Classic CNV. (A) Colour fundus and (B) FAF photographs.

 

fundus autofluopresence_image27.png

Figure 27. CNV with chorioretinal anastomoses. (A) Colour fundus and (B) FAF photographs. (C) Early frame fluorescein angiography. (D) Late frame fluorescein angiography.

 

Pigment epithelial detachment (PED)

PED can show different FAF patterns(11,16,32,62). FAF can provide complementary information to that of fundus colour photograph and fluorescein angiography. In most of the cases, a moderately and diffusely increased FAF can be found, surrounded by a clearly defined ring with decreased fluorescence(35,36,62). Occasionally, intermediate or even decreased FAF can be found that may not correspond to the atrophic RPE or to fibrovascular scars. These changes in the FAF could correspond to different stages in the PED evolution(16). These findings in FAF should be compared with those in fluorescein angiography. We should bear in mind that areas with increased FAF do not always correspond to increased or decreased LF. Besides, the presence of other fluorophores, extracellular fluid or degraded photoreceptor remnants should be considered (Figure 28)(16).

fundus autofluopresence_image28.png

Figure 28. Retinal PED. (A) Colour fundus and (B) FAF photographs.

 

RPE tears

FAF imaging is a useful tool to diagnose RPE tears(35,63). RPE tears usually occur in association with PED in patients with neovascular AMD, either spontaneously or following therapy(58). FAF imaging reveals absence of autofluorescence in the area denuded from RPE. These areas are clearly identifiable by their very low signal, whereas a heterogeneous FAF signal is seen in the area where the RPE is rolled. Therefore, the exact location of the tear can be delineated in most cases.

 

Disciform scars

The appearance of disciform scars in FAF imaging depends on their duration and evolution(35,37). Disciform scars may show different variations and alterations of FAF signal. A decreased signal is typically observed in scarred and fibrotic areas. It has been reported that approximately 50% of the disciform scars may be surrounded by a rim of increased FAF(35,37). These areas of increased autofluorescence correspond to irregularly pigmented areas and may have been caused by a multi-layered RPE, a well-illustrated finding in histopathology (Figures 29 and 30)(36).

fundus autofluopresence_image29.png

Figure 29. CNV with fibrosis. FAF outlines the marked atrophic lesions in the RPE surrounding the CNV/fibrosis. These changes are inconspicuous in colour photographs. (A) Colour fundus and (B) FAF photographs.

 

fundus autofluopresence_image30.png

Figure 30. Fibrous scar secondary to CNV after treatment with anti-VEGF. The damaged RPE appears hyperpigmented in fundus photograph, whereas FAF imaging shows an increased signal. (A) Colour fundus and (B) FAF photographs.

 

Areas of GA and type II CNV may appear simultaneously (Figure 31). 

fundus autofluopresence_image31.png

Figure 31. GA and type II CNV. FAF images of a 67-year-old woman who presents a bilateral type II AMD CNV, along with an evident RPE atrophy.

 

FAF study also reveals the RPE changes after anti-VEGF injections to treat a type I CNV (Figure 32).

fundus autofluopresence_image32.png

Figure 32. CNV - RPE changes. FAF of a 82-year-old man after a series of 8 anti-VEGF injections to treat a type I AMD CNV of his left eye. The right eye shows mottled RPE alterations.

 

In cases of a type I CNV, the detached retina is seen as a hyperfluorescent area while blood spots produce the typical mask effect (Figure 33).

fundus autofluopresence_image33.png

Figure 33. CNV - detached retina. FAF image of a 75-year-old woman at the moment of the diagnosis of a type I CNV. The detached retina is seen as a hyperfluorescent area while blood spots produce the typical mask effect.

 

>> References

References - Fundus autofluorescence in age-related macular degeneration

References - Fundus autofluorescence in age-related macular degeneration

English
  1. Eldred GE, Katz ML. Fluorophores of the human retinal pigment epithelium: separation and spectral characterization. Exp Eye Res. 1988;47(1):71-86.
     
  2. Lamb LE, Simon JD. A2E: a component of ocular lipofuscin. Photochem Photobiol. 2004;79(2):127-36.
     
  3. Bird A. Age-related macular disease. Br J Ophthalmol. 1996;80(1):2-3.
     
  4. Boulton M, Yhaw-Barker P. The role of the retinal pigment epithelium: topographical variation and ageing changes. Eye. 2001;15(Pt3):384-89.
     
  5. Feeney-Burns L, Berman ER, Rothman H. Lipofuscin of human retinal pigment epithelium. Am J Ophthalmol. 1980;90(6):783-91.
     
  6. Wing GL, Blanchard GC, Weiter JJ. The topography and age relationship of lipofuscin concentration in the retinal pigment epithelium. Invest Ophthalmol Vis Sci. 1978;17(7):601-7.
     
  7. Terman A, Brunk UT. Oxidative stress, accumulation of biological ‘garbage’, and aging. Antioxid Redox Signal. 2006;8(1-2):197-204.
     
  8. Brunk UT, Wihlmark U, Wrigstad A, Roberg K, Nilsson SE. Accumulation of lipofuscin within retinal pigment epithelial cells results in enhanced sensitivity to photo-oxidation. Gerontology. 1995;41(Suppl2):201-12.
     
  9. Solbach U, Keilhauer C, Knabben H, Wolf S. Imaging of retinal autofluorescence in patients with age-related macular degeneration. Retina. 1997;17(5):385-89.
     
  10. Eldred GE, Katz ML. Fluorophores of the human retinal pigment epithelium: separation and spectral characterization. Exp Eye Res. 1988;47(1):71-86.
     
  11. Liu J, Itagaki Y, Ben-Shabat S, Nakanishi K, Sparrow JR. The biosynthesis of A2E, a fluorophore of aging retina, involves the formation of the precursor, A2-PE, in the photoreceptor outer segment membrane. J Biol Chem. 2000;275(38):29354-60.
     
  12. Zhou J, Jang YP, Kim SR, Sparrow JR. Complement activation by photooxidation products of A2E, a lipofuscin constituent of the retinal pigment epithelium. Proc Natl Acad Sci USA. 2006;103(44):16182-7.
     
  13. Sparrow JR, Boulton M. RPE lipofuscin and its role in retinal pathobiology. Exp Eye Res. 2005;80(5):595-606.
     
  14. Dorey CK, Wu G, Ebenstein D, Garsd A, Weiter JJ. Cell loss in the aging retina. Relationship to lipofuscin accumulation and macular degeneration. Invest Ophthalmol Vis Sci. 1989;30(8):1691-9.
     
  15. Marmorstein AD, Marmorstein LY, Sakaguchi H, Hollyfield JG. Spectral profiling of autofluorescence associated with lipofuscin, Bruch‘s Membrane, and sub-RPE deposits in normal and AMD eyes. Invest Ophthalmol Vis Sci. 2002;43(7):2435-41.
     
  16. Schmitz-Valckenberg S, Fleckenstein M, Scholl HP, Holz FG. Fundus autofluorescence and progression of age-related macular degeneration. Surv Ophthalmol. 2009;54(1):96-117.
     
  17. Delori FC. Spectrophotometer for noninvasive measurement of intrinsic fluorescence and reflectance of the ocular fundus. Appl Opt. 1994;33(31):7439-52.
     
  18. Delori FC, Dorey CK, Staurenghi G, Arend O, Goger DG, Weiter JJ. In vivo fluorescence of the ocular fundus exhibits retinal pigment epithelium lipofuscin characteristics. Invest Ophthalmol Vis Sci. 1995;36(3):718-29.
     
  19. Webb RH, Hughes GW, Delori FC. Confocal scanning laser ophthalmoscope. Appl Opt. 1987;26(8):1492-99.
     
  20. Bellmann C, Holz FG, Schapp O, Volcker HE, Otto TP. Topographie der Fundusautofluoreszenz mit einem neuen konfokalen Scanning-Laser-Ophthalmoskop. Ophthalmologe. 1997;94(6):385-91.
     
  21. Von Ruckmann A., Fitzke FW, Bird AC. Distribution of fundus autofluorescence with a scanning laser ophthalmoscope. Br J Ophthalmol. 1995;79(5):407-12.
     
  22. American Nationale Standards Institute (ANSI). American National Standard for the Safe Use of Lasers. ANSI Z136.1-2007. American Nationale Standards Institute (ANSI). Washington, USA. 2007.
     
  23. Solbach U, Keilhauer C, Knabben H, Wolf S. Imaging of retinal autofluorescence in patients with age-related macular degeneration. Retina. 1997;17(5):385-89.
     
  24. Trieschmann M, Heimes B, Hense HW, Pauleikhoff D. Macular pigment optical density measurement in autofluorescence imaging: comparison of one- and two-wavelength methods. Graefes Arch Clin Exp Ophthalmol. 2006;244(12):1565-74.
     
  25. Bellmann C, Rubin GS, Kabanarou SA, Bird AC, Fitzke FW. Fundus autofluorescence imaging compared with different confocal scanning laser ophthalmoscopes. Br J Ophthalmol. 2003;87(11):1381-86.
     
  26. Spaide RF. Fundus autofluorescence and age-related macular degeneration. Ophthalmology. 2003;110(2):392-9.
     
  27. Delori FC, Goger DG, Dorey CK. Age-related accumulation and spatial distribution of lipofuscin in RPE of normal subjects. Invest Ophthalmol Vis Sci. 2001;42(8):1855-66.
     
  28. Delori FC, Fleckner MR, Goger DG, Weiter JJ, Dorey CK. Autofluorescence distribution associated with drusen in age-related macular degeneration. Invest Ophthalmol Vis Sci. 2000;41(2):496-504.
     
  29. Spaide RF. Autofluorescence imaging with the fundus camera. In: Holz FG, Schmitz-Valckenberg S, Spaide RF, Bird AC, eds. Atlas of autofluorescence imaging. Berlin, Germany. Springer. 2007;5:49-54.
     
  30. Wabbels B, Demmler A, Paunescu K, Wegscheider E, Preising MN, Lorenz B. Fundus autofluorescence in children and teenagers with hereditary retinal diseases. Graefes Arch Clin Exp Ophthalmol. 2006;244(1):36-45.
     
  31. Weiter JJ, Delori FC, Wing GL, Fitch KA. Retinal pigment epithelial lipofuscin and melanin and choroidal melanin in human eyes. Invest Ophthalmol Vis Sci. 1986;27(2):145-52.
     
  32. Lois N, Halfyard AS, Bunce C, Bird AC, Fitzke FW. Reproducibility of fundus autofluorescence measurements obtained using a confocal scanning laser ophthalmoscope. Br J Ophthalmol. 1999;83(3):276-9.
     
  33. Lois N, Halfyard AS, Bird AC, Fitzke FW. Quantitative evaluation of fundus autofluorescence imaged “in vivo” in eyes with retinal disease. Br J Ophthalmol. 2000;84 (7):741-5.
     
  34. Lois N, Owens SL, Coco R, Hopkins J, Fitzke FW, Bird AC. Fundus autofluorescence in patients with age-related macular degeneration and high risk of visual loss. Am J Ophthalmol. 2002;133(3):341-9.
     
  35. Spital G, Radermacher M, Muller C, Brumm G, Lommatzsch A, Pauleikhoff D. Autofluoreszenz-Charakteristika von Lipofuszinbestandteilen bei unterschiedlichen Formen der späten altersabhängigen Makuladegeneration. Klin Monatsbl Augenheilkd. 1998;213(1):23-31.
     
  36. Von Ruckmann A., Fitzke FW, Bird AC. Fundus autofluorescence in age-related macular disease imaged with a laser scanning ophthalmoscope. Invest Ophthalmol Vis Sci. 1997;38(2):478-86.
     
  37. Von Ruckmann A., Schmidt KG, Fitzke FW, Bird AC, Jacobi KW. Dynamik der Einlagerung und des Abtransportes von Lipofuszin im retinalen Pigmentepithel bei altersbedingter Makuladegeneration. Klin Monatsbl Augenheilkd. 1998;213(1):32-7.
     
  38. Von Ruckmann A., Fitzke FW, Bird AC. Distribution of pigment epithelium autofluorescence in retinal disease state recorded in vivo and its change over time. Graefes Arch Clin Exp Ophthalmol. 1999;237(1):1-9.
     
  39. Schmitz-Valckenberg S, Jorzik J, Unnebrink K, Holz FG. Analysis of digital scanning laser ophthalmoscopy fundus autofluorescence images of geographic atrophy in advanced age-related macular degeneration. Graefes Arch Clin Exp Ophthalmol. 2002;240(2):73-8.
     
  40. Bindewald A, Bird AC, Dandekar SS, Dolar-Szczasny J, Dreyhaupt J, Fitzke FW, Einbock W, Holz FG, Jorzik JJ, Keilhauer C, Lois N, Mlynski J, Pauleikhoff D, Staurenghi G, Wolf S. Classification of fundus autofluorescence patterns in early age-related macular disease. Invest Ophthalmol Vis Sci. 2005;46(9):3309-14.
     
  41. Sawa M, Ober MD, Spaide RF. Autofluorescence and retinal pigment epithelial atrophy after subretinal hemorrhage. Retina. 2006;26(1):119-20.
     
  42. Smith RT, Chan JK, Busuoic M, Sivagnanavel V, Bird AC, Chong NV. Autofluorescence characteristics of early, atrophic, and high-risk fellow eyes in age-related macular degeneration. Invest Ophthalmol Vis Sci. 2006;47(12):5495-504.
     
  43. Sarks JP, Sarks SH, Killingsworth MC. Evolution of geographic atrophy of the retinal pigment epithelium. Eye. 1988;2(Pt5):552-77.
     
  44. Sarks SH. Ageing and degeneration in the macular region: a clinico-pathological study. Br J Ophthalmol. 1976;60(5):324-41.
     
  45. Deckert A, Schmitz-Valckenberg S, Jorzik J, Bindewald A, Holz FG, Mansmann U. Automated analysis of digital fundus autofluorescence images of geographic atrophy in advanced age-related macular degeneration using confocal scanning laser ophthalmoscopy (cSLO). BMC Ophthalmol. 2005;5(1):8.
     
  46. Dreyhaupt J, Mansmann U, Pritsch M, Dolar-Szczasny J, Bindewald A, Holz FG. Modelling the natural history of geographic atrophy in patients with age-related macular degeneration. Ophthalmic Epidemiol. 2005;12(6):353-62.
     
  47. Sunness JS, Gonzalez-Baron J, Applegate CA, Bressler NM, Tian Y, Hawkins B, Barron Y, Bergman A. Enlargement of atrophy and visual acuity loss in the geographic atrophy form of age-related macular degeneration. Ophthalmology. 1999;106(9):1768-79.
     
  48. Schmitz-Valckenberg S, Bindewald-Wittich A, Dolar-Szczasny J, Dreyhaupt J, Wolf S, Scholl HP, Holz FG. Correlation between the area of increased autofluorescence surrounding geographic atrophy and disease progression in patients with AMD. Invest Ophthalmol Vis Sci. 2006;47(6):2648-54.
     
  49. Bindewald A, Schmitz-Valckenberg S, Jorzik JJ, Dolar-Szczasny J, Sieber H, Keilhauer C, Weinberger AW, Dithmar S, Pauleikhoff D, Mansmann U, Wolf S, Holz FG. Classification of abnormal fundus autofluorescence patterns in the junctional zone of geographic atrophy in patients with age related macular degeneration. Br J Ophthalmol. 2005;89(7):874-8.
     
  50. Holz FG, Bindewald-Wittich A, Fleckenstein M, Dreyhaupt J, Scholl HP, Schmitz-Valckenberg S. Progression of geographic atrophy and impact of fundus autofluorescence patterns in age-related macular degeneration. Am J Ophthalmol. 2007;143(3):463-72.
     
  51. Jeong YJ, Hong IH, Chung JK, Kim KL, Kim HK, Park SP. Predictors for the progression of geographic atrophy in patients with age-related macular degeneration: fundus autofluorescence study with modified fundus camera. Eye. 2014;28:209-18. 
     
  52. Mauschitz MM, Fonseca S, Chang P et al. Topography of geographic atrophy in age-related macular degeneration. Invest Ophthalmol Vis Sci. 2012;53:4932-9.
     
  53. Batıoğlu F, Gedik Oğuz Y, Demirel S, Ozmert E. Geographic atrophy progression in eyes with age-related macular degeneration: role of fundus autofluorescence patterns, fellow eye and baseline atrophy area. Ophthalmic Res. 2014;52:53-9.
     
  54. Biarnés M, Arias L, Alonso J, Garcia M, Hijano M, Rodríguez A, Serrano A, Badal J, Muhtaseb H, Verdaguer P, Monés J. Increased Fundus Autofluorescence and Progression of Geographic Atrophy Secondary to Age-Related Macular Degeneration: The GAIN Study. Am J Ophthalmol. 2015;160:345-53 e5.
     
  55. Dreyhaupt J, Mansmann U, Pritsch M, Dolar-Szczasny J, Bindewald A, Holz FG. Modelling the natural history of geographic atrophy in patients with age-related macular degeneration. Ophthalmic Epidemiol. 2005;12:353-62.
     
  56. Domalpally A, Danis RP, White J, Narkar A, Clemons T, Ferris F, Chew E. Circularity index as a risk factor for progression of geographic atrophy. Ophthalmology. 2013;120:2666-71.
     
  57. Caire J, Recalde S, Velazquez-Villoria A, Garcia-Garcia L, Reiter N, Anter J, Fernandez-Robredo P, Alfredo García-Layana; Spanish Multicenter Group on AMD. Growth of geographic atrophy on fundus autofluorescence and polymorphisms of CFH, CFB, C3, FHR1-3, and ARMS2 in age-related macular degeneration. JAMA Ophthalmol. 2014;132:528-34.
     
  58. Holz FG, Pauleikhoff D, Spaide RF, Bird AC. Age-related macular degeneration. Berlin, Germany. Springer. 2004; 1-234 p.
     
  59. Dandekar SS, Jenkins SA, Peto T, Scholl HP, Sehmi KS, Fitzke FW, Bird AC, Webster AR. Autofluorescence imaging of choroidal neovascularization due to age-related macular degeneration. Arch Ophthalmol. 2005;123(11):1507-13.
     
  60. Vaclavik V, Vujosevic S, Dandekar SS, Bunce C, Peto T, Bird AC. Autofluorescence imaging in age-related macular degeneration complicated by choroidal neovascularization: a prospective study. Ophthalmology. 2008;115(2):342-6.
     
  61. McBain VA, Townend J, Lois N. Fundus autofluorescence in exudative age-related macular degeneration. Br J Ophthalmol. 2007;91(4):491-6.
     
  62. Schmitz-Valckenberg S, Holz FG, Bird AC, Spaide RF. Fundus autofluorescence imaging: review and perspectives. Retina. 2008;28(3):385-409.
     
  63. Karadimas P, Paleokastritis GP, Bouzas EA. Fundus autofluorescence imaging findings in retinal pigment epithelial tear. Eur J Ophthalmol. 2006;16(5):767-69.

Geographic Atrophy

Geographic Atrophy

English

Authors:

Fernanda Vaz, MD

Ophthalmology Department, Centro Hospitalar de Lisboa Ocidental, Lisbon, Portugal

Maria Picoto, MD

Ophthalmology Department, Hospital Beatriz Ângelo, Lisbon, Portugal

 

Updated/reviewed by the authors, July 2017.

Introduction

Geographic atrophy (GA) is considered the late stage of the dry form of age-related macular degeneration (AMD)(1).

GA is less common than neovascular AMD and it is responsible for 10-20% of cases of legal blindness in this condition(2,3,4), affecting more than 5 million people wordwilde(5).

Currently there is no approved or effective treatment to prevent either the onset or progression of GA, however, in recent years, significant progress has been made in understanding the pathogenesis of GA, which has led to a number of new potential therapies currently undergoing clinical trial evaluation(6,7,8).

 

Definition

Usually defined as any sharply delineated round or oval area of hypopigmentation, or apparent absence of the retinal pigment epithelium (RPE), in which choroidal vessels are more visible than in surrounding areas, that must be at least 175 μm in diameter (Figure 1)(1).

Figure 1. Fundus photography showing a sharply delineated area of hypopigmentation or apparent absence of RPE in both eyes of the same patient.

 

However larger dimensions have also been proposed for GA minimum limit. The Complications of Age-Related Macular Degeneration Prevention Trial (CAPT) specified a diameter ≥ 250 μm for definition of GA and in the AREDS2 study the minimum diameter of the area was increased to 433 μm. Currently, there is no international consensus on the minimum diameter for the diagnosis of GA(8).

In addition, changes have been proposed in the grading of AMD severity level regarding the presence of GA. In the AREDS grading system, the presence of GA qualifies patients as having either intermediate or advanced AMD, depending on whether or not there is central involvement(9).

In the classification system recently developed by the Beckman Initiative for Macular Research, the presence of any GA regardless of location qualifies as evidence of advanced AMD(10).So far there is no approved or effective treatment to prevent the onset and progression of GA

One eye with GA and choroidal new vessels (CNV) is considered as having exudative form(11).

 

Epidemiology and risk factors

The global prevalence of GA is 0.66% in all ages but it occurs in 0.34% between 65-74 years old, 1.3% between 75-84 and 4.4% over 85 years old.

The prevalence increases to 22% after 90 years of age(2).

It corresponds to half of exudative form prevalence(4,12-14).

A recent meta-analysis based on data in the United States reported that the incidences of GA and neovascular AMD were 1.9 and 1.8 per 1,000 American whites aged ≥ 50 years. Prevalence rates for the two forms of advanced AMD were comparable across all age groups(15).

For subjects with mild or intermediate AMD the 15-year cumulative incidence of neovascular AMD is approximately 2.0% and for progression to pure GA it is approximately 1.3%(16).

The most consistent risk factors are age, familiar history and tobacco smoking, like in other forms of AMD((2,4,13).

In AREDS, GA was also associated with: high BMI (body mass index); use of calcium channel blockers and ß-blockers; not using anti-acids; not using hormone replacement (woman); light iris color and less education level(14-17).

The prevalence of GA is lower in blacks than in whites (2.1% vs 4.8%) like in exudative forms(18,19).

In Rotterdam and Beaver Dam Eye Study, serum HDL cholesterol was directly associated with GA, however this association was not found in Blue Mountains Eye Studyl(20)

In this study, diabetes and the ratio total/HDL cholesterol were linked to increased risk of GA.

As complement system seems to play an essential role in this disease, the complement factor H (CFH) gene located at chromosome 1q32, and others as CFB, CFI, C2 and C3, have been implicated in the development of both forms of AMD(21-23)

Some studies have linked specifically 5p region and 4q 32 region with GAl(24,25).

Only recently, more detailed genome-wide association studies have investigated whether the 2 subtypes of advanced AMD segregate separately in families and associate with different disease variantsl(26).

The variants in the 10q26 locus confer increased risk for both advanced AMD subtypes, but imparts greater risk for CNV than for GA.

Other loci were detected with suggestive associations that differ for advanced AMD subtypes and deserve follow-up in additional studiesl(6).

 

Pathology

Accordingly to AREDS the most common sequence of events leading to GA is the progression of a large drusen to hyperpigmentation, followed by regression of the drusen, hypopigmentation and ultimately RPE cell death, with development of an atrophic area of retina and underlying choriocapillaris, sometimes preceded by the appearance of refractile deposits.

This evolution can be longer than 6 years(27-29).

The average time from entering the AREDS Study to initial appearance of GA has been shown to be approximately 5 to 6 years in the presence of confluent or large (> 125 μm) drusen and hyperpigmentation, as compared with 2.5 years in the presence of hypopigmentation(6).

The presence of large confluent drusen is a significant risk factor for developing GA(30).

Drusen were found in 100% of patients at the site of later GA development; drusen > 125 μm, confluent drusen and hyperpigmentation were present in > 90% of eyes, whereas drusen > 250 μm and hypopigmentation were present in > 80% eyes(31).

Less frequently, GA can follow a drusenoid RPE detachment, regression of a CNV membrane or a RPE rupture(11,32,33).

Higher prevalence of reticular pseudodrusen (RPD) has been reported in patients with GA compared with early or intermediate AMD. Also, the presence of RPD appear to represent a meaningful risk factor associated with progression to GA(34,36).

In some eyes, atrophy was related to a micro reticular pigment pattern distributed around the perimeter of the fovea(33).

Most histopathologic studies suggest that RPE cells are the primary target in GA and its death results in choriocapillaris atrophy(37,38).

Autofluorescent pigments as lipofuscin that accumulate in RPE cells, contribute to a decline of the cell function and degeneration conducting to GA(39).

Although RPE cells are the most involved, the outer nuclear layer is also severely affected with dysfunction and death of photoreceptors. Probably rods are the first affected photoreceptors(40,41).

The mechanisms of RPE death are best studied at junctional zone.

Here, lipofuscin may occupy 30% of RPE cell and may interfere with its metabolism, conducting to death.

These mechanisms include oxidative stress and inflammation(42-45).

Macrophages are often seen in areas of GA, apparently phagocytosing pigment and debris resulting of normal cells deletion(46).

 

Diagnosis

GA can be distinguished from other forms of dry AMD based on stereo biomicroscopy and color fundus photography(6).

The definition of GA is based on clinicohistopathologic investigations that have shown that clinically visible areas of atrophy are characterized by cell death in the RPE, outer neurosensory retina and choriocapillaris.

With modern in vivo imaging technology, these findings can be confirmed(6).

Spectral-domain optical coherence tomography (SD-OCT) and fundus autofluorescence (FAF) allow for noninvasive and rapid quantitative morphological assessment of GA in the clinical setting(6).

 

Fundus

Fundoscopy in GA typically shows a well-circumscribed oval or round area of pigment epithelium atrophy with sharply demarcated borders and increased visibility of choroidal vessels, usually sparing the fovea until late stages  (Figure 2).

Figure 2. Fundus photography showing a well-circumscribed round area of pigment epithelium atrophy, sparing the fovea in right and left eyes.

 

All precursor lesions of this final appearance can also be present: large drusen (>125 microns), focal pigmentation changes and refractile deposit(27,28).

 

Angiography

On fluorescein angiography, GA appears as a sharply delineated window defect due to atrophy of overlying layers of RPE (Figure 3).

Figure 3. Fluorescein angiography showing a sharply delineated window defect.

 

A prolonged choroid filling phase has been described as a clinical marker for changes in Bruch’s membrane and as a risk factor for development of GA(29).

Despite these aspects in GA, fluorescein angiography may be indicated only in atypical cases, in order to allow the correct diagnosis(47).

 

Optical coherence tomography (OCT)

OCT scan shows thinning of hyperreflective external band, corresponding to attenuation of RPE/Bruch’s complex, and deeper hyperreflectivity because of loss of outer layers including photoreceptors (Figure 4)(48,49).

Figure 4. Spectralis OCT: Thinning of hyperreflective external band (because of attenuation of RPE/Bruch’s complex) and deeper hyperreflectivity.

 

In high resolution OCT the atrophic area shows hyperreflective clumps at different levels, segmented plaques of the outer band and elevations with variable reflectivity(41).. In the perilesional area there are elevations of the outer retinal layers, as well as thickening of outer hyperreflective band. At the junction area the outer band shows different degrees of loss(50).

Recently, changes in OCT that precede the development of GA have been identified. These changes include the presence of hyperreflective foci in the retina overlying drusen, a subsidence of the inner nuclear layer (INL) and outer plexiform layer (OPL) with the development of hyporeflective, wedge-shaped bands, and increased signal transmission below the level of the RPE.

These anatomic changes might be used to identify patients early in the course of GA development who might still be at a reversible stage and therefore amenable to intervention(51-53).

OCT imaging may also contribute to a better understanding of the underlying pathologic mechanisms in AMD and GA, may suggest new biomarkers related to disease progression, and might potentially indicate new therapeutic targets in AMD(8).

 

Fundus autofluorescence (FAF)

FAF is the current standard imaging technology for the morphological assessment of GA(8).

Fundus spectrophotometric studies in vivo by Delori and co-workers have shown that FAF represents an accumulation of lipofuscin in the lysossomes of RPE cells, mainly derived from photoreceptors outer segments degradation.

The compound is found as micrometer-sized spherical particles and is characterized by yellow autofluorescence when exposed to blue light(54-56).

It has been shown with confocal scanning laser ophthalmoscopy (cSLO) that FAF response is very low or extinguished in areas of atrophy.

The lack of RPE cells or its low number and therefore of lipofuscin, (the dominant fluorophore) explain this reduction(57).

Increased FAF precedes development of GA(58,59).

FAF is increased in junctional zone around areas of atrophy, and intensity seems to correlate with extension of the atrophic area and also with reduction of retinal sensitivity detected by fundus perimetry (Figure 5)(60,61).

Figure 5. FAF showing increased fluorescence in the junctional zone around areas of atrophy.

 

The FAM-Study Group developed the classification of abnormal FAF according to distinct patterns in the junctional zone of the GA area and identified the following five primary phenotypes based on the presence of increased hyper-autofluorescence: None, Focal, Banded, Patchy and Diffuse(62).

Holz et al. showed that the phenotypic features of FAF abnormalities had a much stronger impact on atrophy progression than any other risk factor that has been addressed in previous studies on progression of GA attributable to AMD, and introduced the ‘diffuse trickling' pattern that is associated with an extremely rapid progression of atrophy(63).

These hyperfluorescent areas in the junctional zone are believed to represent regions of cells that are stressed and more likely to become atrophic(8).

 

Psychophysical Tests of Visual Function

GA develops gradually with the formation of scotomas that spare the fovea initially, expanding into the central visual field late in the course of disease(8).

The most common assessment of visual function, best corrected visual acuity (BCVA), usually fails to reveal the functional deficits experienced by patients with foveal-sparing GA(8).

Functional assessments beyond BCVA including multifocal electroretinography (mfERG), microperimetry, low-luminance visual acuity (LLVA), reading speed and contrast sensitivity, should be used to more fully assess visual impairment(8).

 

Clinical evolution

Eyes with GA may also develop CNV. In one study, 7% of eyes with GA developed CNV in 2 years.

The strongest risk factor for developing CNV in one eye with GA is the presence of CNV in the fellow eye(64).

The 4-year rate of developing CNV is 11% if the other eye has pure GA but increases to 34% if there is CNV.

As other forms of late AMD, GA tends to be bilateral (over 50% of cases) and there is high symmetry between eyes for total atrophic area, presence of peripapillary atrophy and enlargement rate((64-66).

On the other hand there is a high interindividual variability(67).

Longitudinal studies have shown that progression rates vary widely among patients(10).

The mean overall enlargement rate of atrophic area is 2.6 mm2/year.

Eyes with larger areas of atrophy at baseline tend to have larger enlargement rates(65).

The Beaver Dam Eye Study showed that eyes with multifocal disease had larger increase in area of GA and progressed to foveal involvement more frequently than eyes with single foci of disease over 5 years (1.2 vs 2.24 mm2)(68).

GA often first develops surrounding the fovea, sparing the central area. Because of that, the correlation between visual acuity and area of atrophy can be complex(69-71).

The loss of three lines (ETDRS scale) was observed by Suness et al. in 31% of studied eyes by 2 years and 53% by 4 years. More than half of patients with GA encroaching upon the fovea will suffer this severity of vision loss in the same time frame when compared with extrafoveal GA(71).

Even when GA is not center involving, patients can suffer significant deficits in visual function, such as compromised reading ability and impaired vision in dim lighting(72).

The occurrence of GA with concomitant CNV is associated with an even higher risk of severe vision than GA alone(73-75).

Suness et al. showed that 86% of eyes with GA that developed CNV lost 3 or more lines of visual acuity over 2 years compared with only 27% in eyes with GA that did not develop CNV(64).

The risk of visual acuity loss was higher in eyes with better visual acuity at baseline and with lightest iris color(71).

Usually vision loss is bilateral because of lesion symmetry, and evolution since first signs to legal blindness is quite variable(65,66).

Because of rods paucity, reduction of foveal cone function and also because of switching from central to eccentric fixation, other visual functions as the contrast sensitivity, reading rate and dark adaptation are decreased in GA(69,72,72).

The low luminance visual dysfunction and the reduction of the maximum reading rate seem to be significant risk factors for subsequent visual acuity loss(69).

 

Management

Prevention

In AREDS, dietary supplements as zinc, anti-oxidants, vitamins C and E and beta-carotene have been shown to reduce the risk of progression in participants in categories 3 and 4 to advanced AMD (25% in 5 years), however, in the group with GA away from the center (category 3), this reduction was not statistically significant.

Despite of that the AREDS Report nº 8 concluded that those with noncentral GA also should consider taking a supplement of antioxidants plus zinc(9,17).

Macular xantophylls and polyunsaturated fatty acids seem to be associated with a lower risk of advanced AMD(76,77).

Because of that antioxidant effect of macular pigments, lutein, zeaxanthin and omega-3 fatty acids have been tested in the AREDS 2 study(76,78).

The researchers also tried substituting lutein and zeaxanthin for beta-carotene, which prior studies had associated with an increased risk of lung cancer in smokers.

The AREDS 2 study found that while omega-3 fatty acids had no effect on the formulation, lutein and zeaxanthin together appeared to be a safe and effective alternative to beta-carotene.

Subgroup analysis indicated that those with the lowest intake of lutein and zeaxanthin, supplemental lutein and zeaxanthin were protective(79).

Low dietary glycemic index also seems to reduce the risk of evolution to advanced AMD(79).

Other behavioral factors such as stop smoking, as it represents a key modifiable risk factor and control of BMI may play an important role on prevention(80-82).

 

New treatments

So far there is no approved or effective treatment to prevent the onset and progression of GA(6,30).

Although the main target for GA remains unknown, several trials are investigating strategies regarding these pathways and therapeutic targets:

  1. Visual cycle toxic products (visual cycle modulators);
  2. Inflammation, complement, and ECM (mTOR inhibitors, complement inhibitors, MMP inhibitors);
  3. Lipoprotein accumulation (LDL lowering drugs);
  4. Beta-amyloid accumulation (anti-amyloid beta);
  5. Oxidative stress (antioxidants and neuroprotectant);
  6. Choriocapillaris atrophy (choroidal perfusion enhancers);
  7. RPE and photoreceptor loss (Stem cell therapy and neurotrophins);(6)

 

Visual cycle Inhibitors

Accumulation of lipofuscin and melanolipofuscin granules have been observed at the sites of RPE atrophy in GA and associated with GA pathogenesis(7).

ALK-001 is a modified vitamin A molecule, which slows the formation of lipofuscin and RPE apoptosis.

This compound was initially designed to treat Stargardt’s disease. It may also benefit GA by reducing toxic A2E, all-trans retinal, and lipofuscin accumulation.

A phase 1 study was designed to assess the safety and pharmacokinetics of oral ALK-001 capsules in 40 healthy volunteers with no results posted yet(6,7).

Fenretinide and Emixustat (ACU-4429) were also studied but there was no evidence of benefit(6,7,83-85).

 

Anti-inflammatory Agents

Chronic inflammation is believed to play an important role in AMD pathogenesis.

Copaxone (or Glatiramer Acetate) functions to induce suppressor T cells and down-regulates inflammatory cytokines.

A phase I trial was designed to test its safety as well as its efficacy in the prevention of GA progression or conversion of dry AMD to neovascular AMD.

The participant recruitment was suspended with no results posted yet(7).

Lampalizumab is an intravitreal (IVT) administered recombinant monoclonal antibody fragment directed against complement factor D in the alternative complement pathway.

In phase 2 trial (MAHALO study) monthly Lampalizumab showed 20.4% reduction rate in GA area at 18 months in patients with this form of dry AMD compared with monthly sham treatment(82).

Lampalizumab is currently being evaluated in a large, multicenter, phase 3 clinical trial for GA.

Primary outcomes will measure change in GA area after 48 weeks and BCVA up to 2 years after beginning the study(7,30,83).

LFG 316 is an antibody against the C5 portion of the complement pathway administered IVT(6).

The phase 2 trial designed to test the safety and efficacy of lower and higher doses of IVT LFG316 using 18 successive monthly injections in patients with GA was completed, but the results have not yet been published as reported on ClinicalTrials.gov.

APL2 is a derivative of compstatin, a small peptide inhibitor of complement factor C3, that is currently in phase 2 trial (FILLY) to assess the safety, tolerability and evidence of activity of multiple IVT injections of APL-2 in subjects with GA as reported on ClinicalTrials.gov.

Zimura (ARC-1905) is a chemically synthesized aptamer that inhibits complement factor C5.

Phase I trial for Other inflammation suppressor is fluocinolone acetonide, a glucocorticoid used as an intravitreal implant.

Although currently there is no drug proven effective, in the next decade some of the research lines will probably be able to find a more effective treatment for the atrophic form of AMD.

Dry AMD evaluated the safety and tolerability of IVT Zimura injection in patients with GA. The study was completed with no results posted, but Ophthotech plans to initiate Phase II/III clinical trial(7).

Flucinolone acetonide, Eculizumab and Sirolimus (Rapamicina) were also studied, but phase 2 trials failed to demonstrate efficacy(6,7,30).

 

Lipid Modulators

Lipids are also found in drusen and several studies demonstrate that lipids accumulate at the site of subsequent formation of AMD depositsl(86).

Barbosa et al. showed that statin use had a statistically significant effect in reducing AMD incidence in participants aged 68 years and olderl(87).

However the association of statin intake and AMD remains controversial as demonstrated by a 2012 Cochrane database review and to date no trials in GA with this target have been identifiedl(88).

 

Amyloid

The rationale for this new treatment strategy for AMD is related to the presence of amyloid β in drusen.

GSK933776 is a humanized monoclonal antibody intended to modulate amyloid-β levels in patients with GA secondary to AMD administered via intravenous infusion(6,7).

A phase 2 study is completed, however the study results have not been posted (ClinicalTrials.gov).

RN6G is a humanized monoclonal antibody intended to prevent accumulation of amyloid β-40 and β-42 and preserve the photoreceptors and the RPE delivered by intravenous injections(6,7).

A phase 2 study has terminated, but has not been successful, as reported on ClinicalTrials.gov.

MRZ-99030 is a dipeptide containing d-tryptophan and 2-amino-2-methylpropionic acid in clinical development for the topical treatment (eye drops) of glaucoma and AMD(7,89).

 

Neuroprotection

Neuroprotectants have been shown to protect photoreceptors in in vitro models of retinal degeneration and animal models of glaucoma and retinitis pigmentosa.

The mechanism by which neurotrophic factors, such as ciliary neurotrophic factor (CNTF) and basic fibroblast growth factor, protect photoreceptors is not fully understood, and it should be noted that the relevance of these animal models in AMD is not clear.

The RPE and retina both produce fibroblast growth factor and the retina produces CNTF(6).

Brimonidine stimulates the production of neurotrophic factors and can protect photoreceptor cells in animal models of retinal degeneration; it has been formulated as an intravitreal implant which delivers the drug to the retina over a 3-month period.

A Safety and Efficacy Study of Brimonidine Intravitreal Implant in Geographic Atrophy Secondary to Age-related Macular Degeneration (BEACON) (phase 2) is ongoing, with progression of the area of GA at month 12 being the primary endpoint (clinicalTrials.gov)(6).

According to the preliminary results launched at the AAO (American Academy og Ophthalmology) 2016 annual meeting, Brimonidine Intravitreal Implant (22-gauge polymer similar to the one used in ozurdex®, one injection twice a year) caused a significant slowing of disease progression with no adverse effects.

Allergan has completed enrollment of a follow-up study aiming to develop a more potent delivery method(90).

UF-021 (isopropyl unoprostone) is a prostaglandin analog that was investigated in Japan in a phase 2 clinical study in patients with mid- to late-stage retinitis pigmentosa(6).

A phase 2 trial for dry AMD in United States is completed but the study results have not been published yet (ClinicalTrials.gov).

A sustained-release platform with encapsulated human RPE cells engineered to release CNTF has been developed and was tested in a phase 2 study.

However, no benefit in the progression of lesion expansion was found(6,91).

Tandospirone (AL-8309B) a selective serotonin 1A agonist that protects the retina from light damage was investigated, but the phase 3 trial (GATE study) was discontinued in 2012 owing to lack of efficacy(6).

 

Oxidative stress protection
The AREDS trial has reported that daily high doses of antioxidants β-carotene, vitamins C and E and zinc reduced progression to advanced AMD by 25%.

However, the 25% benefit was reported only for the progression to neovascular AMD; there was no benefit in the progression to central GA. Furthermore, the AREDS antioxidant mineral formulation was reported as having no significant effect on the progression of GA in a more recent study(92).

The AREDS2 trial assessed the addition of lutein, zeaxanthin, and/or long-chain ω-3 fatty acids to the original AREDS formulation on progression to advanced AMD.

The study completed in early 2013 found that omega-3 fatty acids had no overall additional effect on the formulation, in a median follow-up period of 5 years(7,93).

Analyses of the AREDS participants over a 12-year period found that participants with the highest omega-3 fatty acids intake were 30% less likely than their peers to develop central GA and neovascular AMD(93).

Several studies are ongoing investigating the effects of ω-3 fatty acids, the ratio of omega-3 to omega-6 fatty acids, and lutein and zeaxanthin (AREDS2 study) in patients with AMD(6,94,95).

OT-551, a small molecule with antioxidant and antinflammatory properties was investigated, however phase 2 study results showed no significant effect on lesion enlargement, retinal sensitivity or total drusen area(96).

 

Choroidal Perfusion Enhancers

Choroidal circulation provides the nutrition and removes the waste from the retina/RPE.

As a consequence of a reduced choroidal blood flow, metabolic wastes are accumulated in photoreceptor cells, Bruch’s membrane and RPE cells.

Those events can lead to development of GA.

Therefore, improving choroidal blood flow could facilitate the removal of metabolic wastes from RPE, Bruch’ membrane and photoreceptor cells to halt AMD disease progression(7)..

MC-1101 is a topical agent with anti-inflammatory and antioxidative properties shown to increase choroidal blood flow; its intended action is to prevent rupture of Bruch’s membrane.

Phase Ib clinical trial showed that topical instillation of 1% MC-1101 produced no significant cardiovascular or ocular toxicity.

A phase II/III is ongoing, 60 patients will receive topical 1% ophthalmic solution, and be assessed for visual function over 24 months(6,7).

Trimetazidine was also studied but failed to prevent progression of GA(6).

 

Stem cells

Replenishing the lost or degenerating RPE cells in GA before the photoreceptors are irrevocably damaged with stem cell-derived RPE cells represents the forefront in the practice of regenerative medicine(7).

RPE can be differentiated from human embryonic stem cells (hESCs) or human induced pluripotent stem cells (iPSCs).

Both display RPE-like morphology and express typical RPE markers and have the ability to phagocytose photoreceptor segments.

When they were transplanted subretinally into rat or mouse model of RPE insufficiency, the grafted cells were retained and retinal function improved(7).

Two prospective phase 1/2 studies were done to assess safety and tolerability of subretinal transplantation of hESC-derived RPE in nine patients with Stargardt's macular dystrophy and nine with atrophic AMD.

Transplanted patients were followed up for a median of 22 months.

There was no evidence of adverse proliferation, rejection, or serious ocular or systemic safety issues related to the transplanted tissue.

BCVA, monitored as part of the safety protocol, improved in ten eyes, improved or remained the same in seven eyes, and decreased by more than ten letters in one eye, whereas the untreated fellow eyes did not show similar improvements in visual acuity.

Vision-related quality-of-life measures increased for general and peripheral vision, and near and distance activities in both diseases.

The results of this study provide the first evidence of the medium-term to long-term safety, graft survival, and possible biological activity of pluripotent stem cell progeny in individuals with any disease.

The results suggest that hESC-derived cells could provide a potentially safe new source of cells for the treatment of various unmet medical disorders requiring tissue repair or replacement(97).

There are currently 2 other clinical trials (phase 1/2) using hESC-RPE cells (MA09-hRPE).

Both are designed to evaluate safety and tolerability of subretinal injection or transplantation of MA09-RPE cells in patients with dry AMD.

Secondary outcomes will measure the mean change of BCVA, autofluorescence photography and reading speed(6,7).

In another study, clonogenic human central nervous system stem cells (HuCNS-SC) will be evaluated for treatment of dry AMD.

Phase I/II study will investigate the safety and preliminary efficacy of unilateral subretinal transplantation of HuCNS-SC cells in subjects with GA secondary to AMD(6,7).

 

Conclusion

GA is a devastating blinding disease without any approved or effective treatment currently available.

Numerous clinical trials are ongoing with the purpose of finding a viable solution to prevent or treat the disease.

The AREDS trials show that AREDS formulation reduces the risk of AMD progression by 25%, while transplanted hESC-derived RPE cells show long-term safety, graft survival and possible biological activity shown by improved visual acuity in GA patients(6,7,30).

Future studies should focus on understanding the pathogenesis of the disease, which remains unclear.

Moreover, the development of advanced imaging system will provide state-of-art tools for analyzing GA pathophysiology and testing new therapeutics(7).

 

>> References

References - Geographic Atrophy

References - Geographic Atrophy

English
  1. Bird AC, Bressler NM, Bressler SB, et al. An International Classification and Grading System for Age-Related Maculopathy and Age-Related Macular Degeneration. The International ARM Epidemiology Study Group. Surv Ophthalmol. 1995;39:367-374.
     
  2. Buch H, Vinding T, Nielsen NV, et al. 14-year incidence progression and visual morbidity of age-related maculopaty: The Copenhagen City Eye Study. Ophthalmology. 2005;112:787-798.
     
  3. Ferris FL, Fine SL, Hyman L. Age-related macular degeneration and blindness due to neovascular maculopathy. Arch Ophthalmol. 1984;102:1640-1642.
     
  4. Klein R, Klein BE, Linton KL. Prevalence of age-related maculopathy. The Beaver Dam Eye Study. Ophthalmology. 1992;99:933-943.
     
  5. Rudnicka AR, Jarrar Z, Wormald R, et al. Age and gender variations in age-related macular degeneration prevalence in populations of European ancestry: a meta-analysis. Ophthalmology. 2012;119:571–580.
     
  6. Holz FG, Strauss EC, Schmitz-Valckenberg S, et al. Geographic Atrophy Clinical Features and Potential Therapeutic Approaches. Ophthalmology. 2014;121:1079-1091.
     
  7. Hanus J, Zhao F, Wang S. Current Therapeutic Development for Atrophic Age-related Macular Degeneration. Br J Ophthalmol. 2016;100(6):744.
     
  8. Srinivas S, Chakravarthy U, et al. Clinical endpoints for the study of geographic atrophy secondary to age-related macular degeneration. Retina. 2016;36:1806–1822.
     
  9. Age-Related Eye Disease Study Research Group. Risk factors for the incidence of Advanced Age-Related Macular Degeneration in Age-Related Eye disease Study: AREDS report nº 8. Arch Ophthalmol. 2008; 119:1417-1436.
     
  10. Ferris FL III, Wilkinson CP, Bird A, et al. Clinical classification of age-related macular degeneration. Ophthalmology. 2013;120:844–851.
     
  11. Gass JDM. Drusen and disciform macular detachment and degeneration. Arch Ophthalmol. 1973; 90:206-217.
     
  12. Vingerling JR, Dielmans I, Hofman A, et al. The prevalence of age-related maculopathy in the Rotterdam Study. Ophthalmology. 1995;102:205-210.
     
  13. Smith W, Assink J, Klein R, et al. Risk factors for age-related macular degeneration: Pooled findings from three continents. Ophthalmology. 2001;108:697-704.
     
  14. Fraser-Bell S, Wu J, Klein R, et al. Smoking, Alcohol Intake, Estrogen Use, and Age-related Macular Degeneration in Latinos: The Los Angeles Latino Eye Study. Am J Ophthalmol. 2006;141:79-87.
     
  15. Rudnicka AR, Kapetanakis VV, Jarrar Z, et al. Incidence of late-stage age-related macular degeneration in American whites: systematic review and meta-analysis. Am J Ophthalmol. 2015;160:85–93 e83.
     
  16. Klein R, Klein BE, Knudtson MD, et al. Fifteen-year cumulative incidence of age-related macular degeneration: the Beaver Dam Eye Study. Ophthalmology. 2007;114:253-262.
     
  17. Age-Related Eye Disease Study Research Group. Risk factors associated with age-related macular degeneration: A case-control study in the Age-Related Eye Disease Study: AREDS report nº3. Ophthalmolgy.
     
  18. 2000;107:2224-2232.
     
  19. Friedman DS, Katz J, Bressler NM, et al. Racial differences in the prevalence of age-related macular degeneration: the Baltimore Eye Survey. Ophthalmology. 1999;106:1049-1055.
     
  20. Klein R, Klein BE, Knudtson M, et al. Prevalence of Age-related Macular Degeneration in 4 Racial/Ethnic Groups in the Multi-ethnic Study of Atherosclerosis. Ophthalmology. 2006;113:373-380.
     
  21. Klein R, Klein BE, Tomany SC. The Association of Cardiovascular Disease with the Long-term Incidence of Age-Related Maculopathy: The Beaver Dam Eye Study. Ophthalmology. 2002;110:1273-1280.
     
  22. Postel EA, Agarwal A, Caldwell J, et al. Complement Factor H Increases Risk for Atrophic Age-Related Macular Degeneration. Ophthalmology. 2006;113:1504-1507.
     
  23. Dominiek DG, Cornelia MD, Oostra BA, et al. Complement Component C3 and Risk of Age-Related Macular Degeneration. Ophthalmology. 2009;115:474-480.
     
  24. Cameron DJ, Yang Z, et al. HTRA1 variant confer similar risk to Geographic Atrophy and Neovascular Age-Related Macular Degeneration. Cell Cycle. 2007;6(9):1122-1125.
     
  25. MajewskiJ, Schultz DW, Weleber RG, et al. Age-Related Macular Degeneration: a genome scan in extended families. Am J Hum Genet. 2003;73:540-550.
     
  26. Abecasis GR, Yashar BM, Zhao Y, et al. Age-related macular degeneration: a high resolution genome scan for susceptibility loci in a population enriched for late-stage disease. Am J Hum Genet. 2004;74:482-494.
     
  27. Seddon JM, Ajani UA, Mitchell BD. Familial aggregation of age-related maculopathy. Am J Ophthalmol. 1997;123:199–206.
     
  28. Age-Related Eye Disease Study Research Group. The Age-Related Eye Disease Study Severity Scale for Age-Related Macular Degeneration: Age-Related Eye Disease Study (AREDS) report nº 17. Arch Ophthalmol. 2005;123:1484-1498.
     
  29. Klein ML, Ferris FLIII, Armstrong J, et al. (AREDS Research Group). Retinal Precursors and the Development of Geographic Atrophy in Age-Related Macular Degeneration. Ophthalmology. 2008;115:1026-1031.
     
  30. Pauleikhoff D, Spital G, Bird AC. A Fluorescein and Indocyanine Green Angiographic Study of Choriocapillaris in Age-related Macular Disease. Arch Ophthalmol. 1999;117:1353-1358.
     
  31. Kashani AH. Stem Cell Therapy in Nonneovascular Age-Related Macular Degeneration. Invest Ophthalmol Vis Sci. 2016;57(5):ORSFm1-9.
     
  32. Klein ML, Ferris FL 3rd, Armstrong J. Retinal precursors and the development of geographic atrophy in age-related macular degeneration. Ophthalmology. 2008;115(6):1026-31.
     
  33. Age-Related Eye Disease Study Research Group. Natural History of Drusenoid Pigment Epithelial Detachment in Age-Related Macular Degeneration: AREDS report nº 28. Ophthalmology. 2010;117:489-499.
     
  34. Casswell AG Kohen D, Bird AC. Retinal pigment epithelial detachments in the elderly: classification and outcome. Br J Ophthalmol. 1985;69:397-403.
     
  35. Finger RP, Wu Z, Luu CD, et al. Reticular pseudodrusen: a risk factor for geographic atrophy in fellow eyes of individuals with unilateral choroidal neovascularization. Ophthalmology. 2014;121:1252–1256.
     
  36. Marsiglia M, Boddu S, Bearelly S, et al. Association between geographic atrophy progression and reticular pseudodrusen in eyes with dry age-related macular degeneration. Invest Ophthalmology Vis Sci.
    2013;54:7362–7369.
     
  37. Kovach JL, Schwartz SG, Agarwal A, et al. The relationship between reticular pseudodrusen and severity of AMD. Ophthalmology. 2016;123:921–9236.
     
  38. Lutty G, Grunwald J, Majji AB,et al. Changes in choriocapillaris and retinal pigment epithelium in age-related macular degeneration. Mol Vis. 1999;5:35.
     
  39. McLeod DS, Taomoto M, Otsuji T, et al. Quantifying Changes in RPE and Choroidal Vasculature in Eyes with Age-Related Macular Degeneration. Invest Ophthalmol Vis Sci. 2002;43:1986-1993.
     
  40. Sparrow JR, Boulton M. Lipofuscin and its role in retinal pathobiology. Exp Eye Res. 2005;80:595-606.
     
  41. Curcio CA, Medeiros NE, Milican CL. Photoreceptor Loss in Age-Related Macular Degeneration. Invest Ophthalmol Vis Sci. 1996;37:1236-1249.
     
  42. Kim SY, Sadda S, Humayun MS, et al. Morphometric Analysis of the Macula in Eyes With Geographic Atrophy due to Age-Related Macular Degeneration. Retina. 2002;22:464-470.
     
  43. Holz FG, Schutt F, Kopitz J, et al. Inhibition of Lysosomal Degradative Functions in RPE Cells by a Retinoid Component of Lipofuscin. Invest Ophthalmol Vis Sci. 1999;40:737-743.
     
  44. Beatty S, Koh H, Phil M, et al. The Role of Oxidative Stress in the Pathogenesis of Age-Related Macular Degeneration. Surv Ophthalmol. 2000;45:115-134.
     
  45. Rózanowska M, Korytowski W, Rózanowski B, et al. Photoreactivity of Aged Human RPE Melanosomes: A Comparison with Lipofuscin. Invest Ophthalmol Vis Sci. 2002;43:2088-2096.
     
  46. Anderson DH, Mullins RF, Hageman GS, et al. A Role for Local Inflammation in the Formation of Drusen in the Aging Eye. Am J Ophthalmol. 2002;134:411-431.
     
  47. Coleman H R, Ferris III FL, Chew EY, et al. Age related macular degeneration. Lancet. 2008;372 (9652):1835-1845.
     
  48. Bischoff P, Speiser P. The role of angiography in age-related macular degeneration. Klin Monatsbl Augenheilkd. 1997;210:296-8.
     
  49. Shuman JS, Puliafito CA, Fujimoto JG, et al. Optical Coherence Tomography of Ocular Diseases. Second edition. Slack, 2004.
     
  50. Bearelly S, Chau FY, Koreishi A, et al. Spectral Domain Coherence Tomography Imaging of Geographic Atrophy Margins. Ophthalmology. 2009;116:1762-1769.
     
  51. Fleckenstein M, Issa PC, Holz FG, et al. High-Resolution Spectral Domain-OCT Imaging in Geographic Atrophy Associated with Age-Related Macular Degeneration. Invest Ophthalmol Vis Sci. 2008;49(9):4137-4144.
     
  52. Karl G Csaky. What Should We Be Measuring as a Geographic Atrophy Endpoint? AAO Subspeciality Day Retina Sillabus.2015; p154-156.
     
  53. Wu Z, Luu CD, Ayton LN, et al. Optical coherence tomography- defined changes preceding the development of drusen-associated atrophy in age-related macular degeneration. Ophthalmology. 2014;121:2415-2422.
     
  54. Ouyang Y, Heussen FM, Hariri A, et al. Optical coherence tomography-based observation of the natural history of drusenoid lesion in eyes with dry age-related macular degeneration. Ophthalmology. 2013;120:2656-2665.
     
  55. Delori FC, Dorey CK, Staurenghi G, et al. In Vivo Fluorescence of the Ocular Fundus Exhibits Retinal Pigment Epithelium Lipofuscin Characteristics. Invest Ophthalmol Vis Sci. 1995;36:718-729.
     
  56. Sparrow JR, Fishkin N, Zhou J, et al. A2E, a Byoproduct of the Visual Cycle. Vision Res. 2003;43:2983-2990.
     
  57. Frangieh GT, Green WR, Fine SL. A histopathological Study of Best’s Macular Dystrophy. Arch Ophthalmol. 1982;100:1115-1121.
  58. Sparrow JR, Boulton M. Lipofuscin and its role in retinal pathobiology. Exp Eye Res. 2005;80:595-606.
     
  59. Von Ruckmann A, Fitzke FW, Bird AC. Fundus Autofluorescence in Age-Related Macular Disease Imaged with Laser scanning Ophthalmoscope. Invest Ophthalmol Vis Sci. 1997;38:478-86.
     
  60. Holz FG, Bellman C, Staudt S, et al. Fundus Autofluorescence and Development of Geographic Atrophy in Age-Related Macular Degeneration. Invest Ophthalmol Vis Sci. 2001;42:1051-1056.
     
  61. Schmitz-Valkenberg S, Bültmann S, Dreyhaupy J, et al. Fundus autofluorescence and fundus perimetry in the junctional zone of geographic atrophy in patients with age-related macular degeneration. Invest Ophthalmol Vis Sci. 2004;45:4470-4476.
     
  62. Schmitz-Valckenberg S, Bindewald-Wittich A, Dolar-Szczasny J, et al. Fundus Autofluorescence and Fundus Perimetry in the Junctional Zone of Geographic Atrophy in Patients with Age-Related Macular Degeneration. Invest Ophthalmol Vis Sci. 2006;47:2648-2654.
     
  63. Bindewald A, Schmitz-Valckenberg S, Jorzik JJ et al. Classification of abnormal fundus autofluorescence patterns in the junctional zone of geographic atrophy in patients with age related macular degeneration. Br J Ophthalmol. 2005;89(7):874-8.
     
  64. Holz FG, Bindewald-Wittich A, Fleckenstein M et al. Progression of geographic atrophy and impact of fundus autofluorescence patterns in age-related macular degeneration. Am J Ophthalmol. 2007;143(3):463-72.
     
  65. Suness JS, Gonzalez-Baron J, Bressler NM, et al. The Development of Choroidal Neovascularization in Eyes with Geographic Atrophy Form of Age-Related Macular Degeneration. Ophthalmology. 1999;106:910-919.
     
  66. Suness JS, Margalit E, Bressler NM, et al. The Long-term Natural History of Geographic Atrophy from Age-Related Macular Degeneration: Enlargement of Atrophy and Implications for Interventional Clinical Trials. Ophthalmology. 2007;114:271-277.
     
  67. Schatz H, McDonald HR. Atrophic Macular Degeneration. Rate of Spread of Geographic Atrophy and Visual Loss. Ophthalmology. 1989;96:1541-1551.
     
  68. Bellmann C, Jorzik J, Spital G, et al. Symmetry of bilateral lesions in geographic atrophy in patients with age-related macular degeneration. Ophthalmology. 2002;120:579-584.
     
  69. Klein R, Meuer SM, Knudtson MD, Klein BE. The epidemiology of progression of pure geographic atrophy: the Beaver Dam Eye Study. Am J Ophthalmol. 2008;146:692–699.
     
  70. Suness JS, Rubin GS, Broman A, et al. Low luminance Visual Dysfunction as a Predictor of Subsequent Visual Loss From Geographic Atrophy in Age-Related Macular Degeneration. Ophthalmology. 2008;115:1480-1488.
     
  71. Maguire P, Vine AK. Geographic Atrophy of Retinal Pigment Epithelium. Am J Ophthalmol. 1986;102:621-625.
     
  72. Suness JS, Gonzalez-Baron J, Applegate CA, et al. Enlargement of Atrophy and Visual Acuity Loss in the Geographic Atrophy Form of Age-Related Macular Degeneration. Ophthalmology. 1999;106:1768-1779.
     
  73. Suness JS, Applegate CA, Bressler NM, et al. Visual Function Abnormalities and Prognosis in Eyes with Age-Related Geographic Atrophy of the Macula and Good Visual Acuity. Ophthalmology. 1997;104(10):1677-1691.
     
  74. Davis MD, Gangnon RE, Lee LY, et al. The Age-Related Eye Disease Study severity scale for age-related macular degeneration: AREDS Report No. 17. Arch Ophthalmol. 2005;123:1484–1498.
     
  75. Schatz H, McDonald HR. Atrophic macular degeneration. Rate of spread of geographic atrophy and visual loss. Ophthalmology. 1989;96:1541–1551.
     
  76. Macular Photocoagulation Study Group. Risk factors for choroidal neovascularization in the second eye of patients with juxtafoveal or subfoveal choroidal neovascularization secondary to age-related macular degeneration. Arch Ophthalmol. 1997;115:741–747.
     
  77. Coleman H, Chew E. Nutritional Supplementation in Age-Related Macular Degeneration. Curr Opin Ophthalmol. 2007;18:220-223.
     
  78. Age-Related Eye Disease Study Research Group. Risk Factors for the Incidence of Advanced Age-Related Macular Degeneration in Age-Related Eye Disease Study: AREDS report nº 23. Arch Ophthalmol. 2008;126:1274-1279.
     
  79. Chiu CJ, Milton RC, Klein R et al. Dietary Compound Score and Risk of Age-Related Macular Degeneration in the Age-Related Eye Disease Study. Ophthalmology. 2008;116: 939-946.
     
  80. Age-Related Eye Disease Study 2 Research Group. Lutein + zeaxanthin and omega-3 fatty acids for age-related macular degeneration: the Age-Related Eye Disease Study 2 (AREDS2) randomized clinical trial. JAMA. 2013;309:2005-15.
     
  81. Khan JC, Thurlby DA, Shahid H, et al. Smoking and age related macular degeneration: the number of pack years of cigarette smoking is a major determinant of risk for both geographic atrophy and choroidal neovascularisation. Br J Ophthalmol. 2006;90:75-80.
     
  82. Age-Related Eye Disease Study Research Group. Risk factors for the incidence of Advanced Age-Related Macular Degeneration in Age-Related Eye Disease Study: AREDS report nº 19. Ophthalmology. 2005;112:533-539.
     
  83. Age-Related Macular Degeneration PPP - Updated 2015. AAO Retina/Vitreous PPP Panel, Hoskins Center for Quality Eye Care.
     
  84. Jack LS, Sadiq MA, Do DV,Nguyen QD. Emixustat and Lampalizumab: Potential Therapeutic Options for Geographic Atrophy. Dev Ophthalmol. 2016;55:302-9.
     
  85. Mata NL, Lichter JB, Vogel R, et al. Investigation of oral fenretinide for treatment of geographic atrophy in age-related macular degeneration. Retina. 2013;33:498–507.
     
  86. http://www.businesswire.com/news/home/20160525006550/en/Acucela-Announces-Top-Line-Results-Phase-2b3-Clinical
     
  87. Elisa Buschini, Antonio M Fea, Carlo A Lavia, et al. Recent developments in the management of dry age-related macular degeneration. Clin Ophthalmol. 2015;9:563–574.
     
  88. Barbosa DT, Mendes TS, Cíntron-Colon HR, et al. Age-related macular degeneration and protective effect of HMG Co-A reductase inhibitors (statins): results from the National Health and Nutrition Examination Survey 2005-2008. Eye (Lond). 2014;28(4):472-80.
     
  89. Gehlbach P, Li T, Hatef E. Statins for age-related macular degeneration. Cochrane Database Syst Rev. 2012;(3):CD006927.
     
  90. Parsons CG, Ruitenberg M, Freitag CE, et al.  MRZ-99030 - A novel modulator of Aβ aggregation: I - Mechanism of action (MoA) underlying the potential neuroprotective treatment of Alzheimer's disease, glaucoma and age-related macular degeneration (AMD). Neuropharmacology. 2015;92:158-69.
     
  91. https://www.aao.org/interview/brimonidine-geographic-atrophy
     
  92. Zhang K, Hopkins JJ, Heier JS, et al. Ciliary neurotrophic factor delivered by encapsulated cell intraocular implants for treatment of geographic atrophy in age-related macular degeneration. Proc Natl Acad Sci U S A. 2011;108:6241–5.
     
  93. AREDS Research Group. Change in area of geographic atrophy in the Age-Related Eye Disease Study: AREDS report number 26. Arch Ophthalmol. 2009;127:1168–74.
     
  94. Sangiovanni JP, Agron E, Meleth AD, et al. {omega}-3 Long-chain polyunsaturated fatty acid intake and 12-y incidence of neovascular age-related macular degeneration and central geographic atrophy: AREDS report 30, a prospective cohort study from the Age-Related Eye Disease Study. Am J Clin Nutr. 2009;90(6):1601-7.
     
  95. Konstantin Petrukin. New Therapeutic Targets in Atrophic Age-Related Macular Degeneration. Expert Opin Ther Targets. 2007;11:625-639.
     
  96. Tanito M, Li F, Elliot MH, et al. Protective effect of TEMPOL derivates against Light-Induced Retinal Damage in Rats. Invest Ophthalmol Vis Sci. 2007;48:1900-1905.
     
  97. Wong WT, Kam W, Cunningham D, et al. Treatment of geographic atrophy by the topical administration of OT-551: results of a phase II clinical trial. Invest Ophthalmol Vis Sci. 2010;51(12):6131–6139.
     
  98. Schwartz SD, Regillo CD, Lam BL, et al. Human embryonic stem cell-derived retinal pigment epithelium in patients with age-related macular degeneration and Stargardt's macular dystrophy: follow-up of two open-label phase 1/2 studies. Lancet. 2015;385(9967):509–516.

Fundus autofluorescence patterns and optical coherence tomography in geographic atrophy secondary to AMD

Fundus autofluorescence patterns and optical coherence tomography in geographic atrophy secondary to AMD

English

Authors:

Jordi Monés, MD
Institut de la Màcula i de la Retina, Centro Médico Teknon, Barcelona. Spain

 

Marc Biarnés, OD, MPH
Institut de la Màcula i de la Retina, Centro Médico Teknon, Barcelona. Spain

 

Introduction

 

Geographic atrophy (GA) and choroidal neovascularization (CNV) represent the advanced forms of age-related macular degeneration (AMD).

GA is defined as a well circumscribed area of atrophy of the retinal pigment epithelium (RPE) where the large choroidal vessels can be seen by ophthalmoscopy and show thinning or absence of the RPE, closure of the choriocapillaris and degeneration of the overlying photoreceptors(1,2).

Visual loss in GA is due to areas of atrophy of the RPE larger than 175 µm and subsequent loss of tissue in the outer retina (photoreceptors) and choriocapillaris; these areas tend to coalesce progressively and may not affect the fovea until late in the course of the disease (the so-called “foveal sparing”), when visual acuity (VA) finally ensues.

Due to the large paracentral areas of atrophy but preservation of the fovea visual function is often very poor in spite of an apparently good VA.

Patients with a VA of 20/20 may be functionally blind (Figure 1).

Figure 1 - Foveal sparing may allow a 20/20 visual acuity in spite of very severe visual function impairment

 

Relatives of the patients, and even ophthalmologist, have often confused VA for visual function, thus, frequently these patients have felt poorly understood(1).

GA is responsible for one-third of the cases of end stage disease(3) and accounts for 20% of cases of severe visual loss due to the disorder(4).

At the age of 85 years or older incidence of GA is four times the one of CNV. GA is not a benign disease; the atrophy of the external layers may progress at a speed of 1.5-2.6 mm² per year
(Figure 2)(5).

Figure 2 - GA atrophy progression in one year of the right eye of the same patient (fundus autofluorescence images).

 

It remains a significant challenge because of its potential to cause blindness, its relentless progression and the lack of current effective treatment.

With the progressive increase of the longevity in developed countries GA secondary to AMD represents a true epidemic. Recent research has pointed towards lipofuscin, a fluorophore that accumulates in the RPE, as a triggering agent in the development of atrophy.

Lipofuscin derives mainly from phagocytosed photoreceptor outer segments and accumulates in RPE lysosomes, where one of its many compounds, A2E (N-retinylidene-N-retinylethanol-amine), increases the pH by inhibition of ATPase proton pump function, difficulting its phagocitation and inducing cellular apoptosis(6,7).

Using a fundus spectrophotometer, Delori et al.(8,9) were able to visualize lipofuscin due to its autofluorescent properties (when stimulated with blue light in the range of 488 nm, lipofuscin emits a yellow fluorescence).

Current developments allow the clinical in vivo visualization of the distribution of lipofuscin by means of confocal scanning laser ophthalmoscope (cSLO) or specific filters in the fundus cameras. by macular pigment, lutein and zeaxanthin.

A relatively high degree of inter-individual variability and technical difficulties limit the use of the absolute quantification of pixel gray values for longitudinal or transversal studies, and therefore the interpretation of the images is based on qualitative observations, ie decreased (dark), normal or increased (white) FAF, in a similar way to that of conventional fluorescein angiography.

 

Fundus autofluorescence

Fundus autofluorescence (FAF) is a novel, non-invasive method for imaging the fluorescence properties of lipofuscin (and possibly other molecules with a range of absorption and emission spectra close to that of this fluorophore) at the level of the RPE.

Using a commercially available cSLO, the distribution of FAF in the normal eye can be seen in Figure 3. It is characterized by a uniform grayish signal in the fundus and a marked dark appearance in the optic nerve (absence of RPE) and retinal vessels (absorption of fluorescence by hemoglobin and other blood contents).

Figure 3 - Normal fundus autofluorescence: uniform grayish signal in the fundus and a marked dark appearance in the optic nerve (absence of RPE) and retinal vessels (absorption of fluorescence by hemoglobin and other blood contents)

 

The macular area shows progressive diminished signal intensity towards the fovea because of absorption phenomena of short wavelengths A decreased signal is commonly due to RPE atrophy (absence of lipofuscin), an increase in RPE melanin content and absorption from extracellular material anterior to the RPE (intraretinal fluid, fibrosis and media opacities, being cataract a common cause of decreased FAF intensity and poor image quality, specially with cSLO).

On the other hand, an increased signal may be due to lipofuscin accumulation in the RPE (which is the main fluorophore in FAF imaging), presence of other fluorophores not in the RPE (drusen -including those in the optic nerve head-, older hemorrhages), lack of absorbing material and artifacts(10).

FAF imaging in patients with GA is characterized by a decreased signal with sharp borders corresponding to the area of atrophy on conventional retinography (Figure 4).

Figure 4 - Fundus autofluorecence of patients with GA secondary to dry AMD.
Atrophic areas show hipoautofluorescence. Borders of the lesion show hyperautofluorescence due to the accumulation of lipofuscin

 

However, many patients show increased FAF at the borders of atrophy (Figure 5), which has been histopathologically confirmed as areas of increased lipofuscin-filled RPE cells between atrophic and normal retina.

Figure 5 - Surrounding areas of accumulation of lipofuscin at the junction area show increased autofluorescence.
In pseudocolor 3D areas of higher accumulation of lipofuscin are seen as yellow spikes

 

This finding has not been identified by any other imaging modality.

Holz et al. showed in a longitudinal study that atrophy developed selectively in the junction areas of increased FAF(11) but not elsewhere, a finding that could not be confirmed on a small sample by Hwang et al.(12).

Based on this information, the FAM (Fundus autofluorescence in age-related macular degeneration) study initially described 8 patterns of FAF in the junction zone of GA(13), which were later modified to incorporate a nineth pattern (Figure 6)(14).

”Reproduced with permission from Rightslink (Elsevier)”.

Figure 6 - FAM-Study classification of fundus autofluorescence patterns of geographic atrophy in age-related macular degeneration(14).

 

According to FAF in the junction zone of atrophy, eyes are classified as none (when there is no increased FAF at the borders of the GA), localized (focal, banded, patchy) and diffuse (fine granular, branching, trickling, reticular and fine granular with punctuated spots)(14).

The relevance of these patterns relays in the fact that they may represent different phenotypic manifestations of the disease.

It has been shown in natural history studies that rates of growth differ between subtypes of FAF and that a strong correlation exists between FAF pattern and progression of atrophy in GA(14,15).

In the FAM study(14) 195 eyes of 129 patients with GA were followed a median of 1.80 years and classified according to FAF pattern at baseline.

Those without abnormal FAF at the borders of the lesion experienced the slowest progression over time (0.38 mm2 /year, n = 17) compared to those with the focal (0.81 mm2 / year, n = 14) and diffuse (1.77 mm2 / year, n = 112) subtype (p<0.0001).

FAF was more strongly associated with GA growth than other classic risk factors, such as size of baseline atrophy, smoking, age or family history.

Nowadays current devices of autofluorescence equipped with semiautomated software allow quantification of total area of GA and measurement of its progression in time (Figure 7).

Figure 7 - Semiautomated software allows quantification of total area of GA and measurement of its progression in time (Spectralis Heidelberg Retinal Angiograph/OCT; Heidelberg Engineering, Heidelberg, Germany)

 

Furthermore, studies using fine-matrix mapping(17) and SLO microperimetry(18) have found impaired rod photoreceptor function and photopic sensitivity respectively in areas of increased FAF in the junction zone, which underscores abnormalities associated with increased fundus autofluorescence.

Taken together, these results suggest that presence of increased FAF at the borders of GA is associated with a greater rate of progression of atrophy and that different patterns of FAF may reflect differences at the cellular and molecular level that may explain the different evolution of the disease process.

This information is relevant for understanding its physiopathology, natural history and to evaluate future therapeutic strategies.

 

Optical coherence tomography

Recent development of high resolution, high speed spectral domain optical coherence tomography (SD OCT) improves the visualization of the RPE, outer and inner segment of the photoreceptors and external limiting membrane over previous time-domain based technology.

Given the aforementioned reasons, simultaneous imaging techniques that combine SD-OCT and FAF are highly desirable for the evaluation of the atrophic and junction areas of patients with GA (Figure 8).

Figure 8 - High resolution optical coherence tomography correlated with the autofluorescence image.

 

Currently there are instruments that fulfill these criteria and allow the study of the correlation between areas of increased or decreased autofluorescence with the morphologic changes detected in the external retina by SD OCT.

SD OCT in atrophic areas of patients with GA show retinal thinning due to atrophy of the RPE and disappearance of the external retina, that includes inner and outer segment of the photoreceptors and, very frequently, the external limiting membrane(19); in the severe forms, the outer nuclear layer may be no longer identifiable and therefore the outer plexiform layer may be in direct contact with Bruch´s membrane.

The thinned retina permits the deeper penetration of light and a corresponding increased signal from the choroid within the atrophic area (Figure 9).

SD OCT is also useful to identify absence of exudative signs (intraretinal or subretinal pockets of fluid, RPE detachments, maintenance of the continuity of Bruch´s membrane)(20).

Several abnormalities have been found with SD OCT in the junction zone(19,21,22), such as disruption of external retinal layers with different shape of band endings, disappearance of the external limiting membrane and/or of the retina encompassing inner segments of the photoreceptor layer to Bruch´s membrane at the same or at a different transverse planes, small elevations of RPE thought to represent subRPE deposits or increased distance between inner and outer photoreceptors segment and RPE, presumably due to debris between these layers.

Smooth margins with no structural changes from normal to abnormal retina exist in the junction zone when there is no FAF abnormality, which underscores the significance of abnormal FAF(22) (Figure 9).

Figure 9 - SD OCT in normal retina (up left ). In the lower images, SD OCT of the junction area between preserved retina and the geographic atrophy. Atrophic areas show retinal thinning due to atrophy of the RPE and disappearance of the external retina that includes inner and outer segment of the photoreceptors and, very frequently, the external limiting membrane(16); The outer nuclear layer may be no longer identifiable and therefore the outer plexiform layer may be in direct contact with Bruch´s membrane. The thinned retina permits the deeper penetration of light and a corresponding increased signal from the choroid within the atrophic area.

 

In summary, precise quantification of the GA and its progression by fundus autofluorescence imaging and the detailed morphologic study of the external retina by the high resolution SD OCT allow to show the relative slow progressing abnormalities of the outer retinal layers in dry AMD.

This may prove essential for prognostic and interventional strategies, in order to detect potential benefit by slowing down the degenerative process or perhaps detect signs of RPE and/or photoreceptors rescue.

 

>> References

References - Fundus autofluorescence patterns and optical coherence tomography in geographic atrophy secondary to AMD

References - Fundus autofluorescence patterns and optical coherence tomography in geographic atrophy secondary to AMD

English
  1. Klein R, Davis MD, Magli YL, Segal P, Klein BE, Hubbard L. The Wisconsin age-related maculopathy grading system. Ophthalmology 1991; 98 (7): 1128-1134.
     
  2. Sarks JP, Sarks SH, Killingsworth MC Evolution of geographic atrophy of the retinal pigment epithelium. Eye 1988; 2 ( Pt 5): 552-577.
     
  3. Augood CA, Vingerling JR, de Jong PT, Chakravarthy U, Seland J, Soubrane G, Tomazzoli L, Topouzis F, Bentham G, Rahu M, Vioque J, Young IS, Fletcher AE. Prevalence of age-related maculopathy in older Europeans: the European Eye Study (EUREYE). Arch Ophthalmol 2006; 124 (4): 529-535.
     
  4. Ferris FL 3rd, Fine SL, Hyman L. Age-related macular degeneration and blindness due to neovascular maculopathy. Arch Ophthalmol 1984; 102 (11): 1640-1642.
     
  5. Sunness JS, Margalit E, Srikumaran D, Applegate CA, Tian Y, Perry D, Hawkins BS, Bressler NM. The long-term natural history of geographic atrophy from age-related macular degeneration: enlargement of atrophy and implications for interventional clinical trials. Ophthalmology 2007; 114 (2): 271-277.
     
  6. Monés J, Gómez-Ulla F. Degeneración macular asociada a la edad. Prous Science. Barcelone, Espagne. 2005; 467 p.
     
  7. Schmitz-Valckenberg S, Fleckenstein M, Scholl HP, Holz FG. Fundus autofluorescence and progression of age-related macular degeneration. Surv Ophthalmol 2009; 54 (1): 96-117.
     
  8. Delori FC. Spectrophotometer for non- invasive measurement of intrinsic fluorescence and reflectance of the ocular fundus. Appl Optics 1994; 33 (31): 7429-7452.
     
  9. Delori FC, Dorey CK, Staurenghi G, Arend O, Goger DG, Weiter JJ. In vivo fluorescence of the ocular fundus exhibits retinal pigment epithelium lipofuscin characteristics. Invest Ophthalmol Vis Sci 1995; 36 (3): 718-729.
     
  10. Holz FG, Schmitz-Valckenberg S, Spaide RF, Bird AC. Atlas of Fundus Autofluorescence Imaging. Springer. Berlin, Allemagne. 2007; 342 p.
     
  11. Holz FG, Bellman C, Staudt S, Schütt F, Völcker HE. Fundus autofluorescence and development of geographic atrophy in age-related macular degeneration. Invest Ophthalmol Vis Sci 2001; 42 (5): 1051-1056.
     
  12. Hwang JC, Chan JW, Chang S, Smith RT. Predictive value of fundus autofluorescence for development of geographic atrophy in age-related macular degeneration. Invest Ophthalmol Vis Sci2006; 47 (6): 2655-2661.
     
  13. Bindewald A, Schmitz-Valckenberg S, Jorzik JJ, Dolar-Szczasny J, Sieber H, Keilhauer C, Weinberger AW, Dithmar S, Pauleikhoff D, Mansmann U, Wolf S, Holz FG. Classification of abnormal fundus autofluorescence patterns in the junctional zone of geographic atrophy in patients with age related macular degeneration. Br J Ophthalmol 2005; 89 (7): 874-878.
     
  14. Holz FG, Bindewald-Wittich A, Fleckenstein M, Dreyhaupt J, Scholl HP, Schmitz-Valckenberg S; FAM-Study Group. Progression of geographic atrophy and impact of fundus autofluorescence patterns in age-related macular degeneration. Am J Ophthalmol 2007; 143 (3): 463-472.
     
  15. Schmitz-Valckenberg S, Bindewald-Wittich A, Dolar-Szczasny J, Dreyhaupt J, Wolf S, Scholl HP, Holz FG. Correlation between the area of increased autofluorescence surrounding geographic atrophy and disease progression in patients with AMD. Invest Ophthalmol Vis Sci 2006; 47 (6): 2648-2654.
     
  16. Scholl HP, Bellmann C, Dandekar SS, Bird AC, Fitzke FW. Photopic and scotopic fine matrix mapping of retinal areas of increased fundus autofluorescence in patients with age-related maculopathy. Invest Ophthalmol Vis Sci 2004; 45 (2): 574-583.
     
  17. Schmitz-Valckenberg S, Bültmann S, Dreyhaupt J, Bindewald A, Holz FG, Rohrschneider K. Fundus autofluorescence and fundus perimetry in the junctional zone of geographic atrophy in patients with age-related macular degeneration. Invest Ophthalmol Vis Sci 2004; 45 (12): 4470-4476.
     
  18. Wolf-Schnurrbusch UE, Enzmann V, Brinkmann CK, Wolf S. Morphologic changes in patients with geographic atrophy assessed with a novel spectral OCT-SLO combination. Invest Ophthalmol Vis Sci 2008; 49 (7): 3095-3099.
     
  19. Coscas G, Coscas F, Vismara S, Zourdani A, Li Calzi CI. Optical coherence tomography in age-related macular degeneration. 1st edition. Springer. Berlin, Allemagne. 2009. 414 p.
     
  20. Fleckenstein M, Charbel Issa P, Helb HM, Schmitz-Valckenberg S, Finger RP, Scholl HP, Loeffler KU, Holz FG. High-resolution spectral domain-OCT imaging in geographic atrophy associated with age-related macular degeneration. Invest Ophthalmol Vis Sci 2008; 49 (9): 4137-4144.
     
  21. Brar M, Kozak I, Cheng L, Bartsch DU, Yuson R, Nigam N, et al. Correlation between spectral-domain optical coherence tomography and fundus autofluorescence at the margins of geographic atrophy. Am J Ophthalmol 2009; 148 (3): 439-444.

Neovascular Phenotypes: Retinal Angiomatous Proliferation (RAP) or Type 3 Neovascularization

Neovascular Phenotypes: Retinal Angiomatous Proliferation (RAP) or Type 3 Neovascularization

English

Authors:

Rufino Silva MD, PhD

João Pedro Marques, MD, MSc

Department of Ophthalmology, Centro Hospitalar e Universitário de Coimbra (CHUC), Coimbra, Portugal
Association for Innovation and Biomedical Research on Light and Image (AIBILI), Coimbra, Portugal
Faculty of Medicine, University of Coimbra (FMUC), Coimbra, Portugal

 

Updated/reviewed by the authors, July 2017.

Introduction

Over 100 years ago, Oeller described for the first time the presence of anastomoses between the retinal and choroidal circulations in eyes with disciform scars(1).

Later, these were recognized in association with laser photocoagulation(2), radiotherapy(3), chorioretinal inflammatory diseases(4) and parafoveal telangiectasias(5).

The interest in this condition even led to anatomopathological studies in disciform scars from late age-related macular degeneration (AMD)(6).

In 1992, Hartnett et al.(7) described nine cases of retinal neovascularization, to which they referred as “deep retinal vascular anomalous complex”.

In 2000, Slakter et al.(8) described chorioretinal anastomosis in eyes with pigment epithelial detachment (PED) and indocyanine green angiography (ICG-A) hot spots but it was Yannuzzi et al.(9) who coined the term retinal angiomatous proliferation (RAP) when in 2001 the authors described chorioretinal anastomoses as a neovascular proliferation originating deeply in the retina.

In 2001, Yannuzzi et al. described chorioretinal anastomosis as neovascular proliferation with origin in the retina, and proposed the designation of RAP – retinal angiomatous proliferation(9).

Still, several authors(10-13) maintained the designation of chorioretinal anastomosis, proposing a choroidal origin for this clinical entity.

In 2008, Yannuzzi et al.(14) escribed 5 cases of RAP with the neovascular complex originating in the choroid instead of the retina, and proposed that RAP should be called type 3 neovascularization instead. The two terms have been used interchangeably ever since(15).

However, RAP is still the most common designation.

 

Classification

Based on the presumed origin and evolution of the neovascular process, a three-stage classification was proposed by Yannuzzi et al.(9), to characterize the clinical manifestations and progressive changes observed in RAP. However, it is often clinically difficult, if not impossible, to determine exactly when progression occurs from one stage to another.

 

Stage I – Intraretinal neovascularization.

The presence of small intraretinal hemorrhages is a hallmark of RAP and a very useful sign for its early clinical diagnosis. A small elevation of the inner/intermediate retina caused by angiomatous tissue may be observed under the slit lamp; this elevation may extend tangentially, assuming a telangiectatic appearance. Dilated retinal vessels may perfuse and drain the intraretinal neovascularization and form retino-retinal anastomoses (RRA)(9).

The fluorescein angiography (FA) of the intraretinal neovascular complex shows a focal hyperfluorescent area in front of the retinal pigment epithelium (RPE), mimicking classic choroidal or, more frequently, occult neovascularization. ICG-A may reveal a hot spot with staining and leakage.

 

Stage II – Subretinal Neovascularization

ovascular complex shows a focal hyperfluorescent area in front of the retinal pigment epithelium (RPE), mimicking classic choroidal or, more frequently, occult neovascularization. ICG-A may reveal a hot spot with staining and leakage.

 

Stage II – Subretinal Neovascularization

Involvement of the subretinal space with localized neurosensory retina detachment, edema and retinal hemorrhages at the edges may already be observed in colour fundus photography. Usually, a clear RRA can be seen, with a perfusing retina arteriole and draining venule communicating within the core of the subretinal neovascularization (Figure 1).

Stage III – Choroidal NeovascularizationNeovascular_Phenotypes_image1.png

Figure 1. RAP lesion – Colour fundus photography of a RRA in a patient with RAP.

 

In stage II, an associated serous PED can be observed in 94% of the eyes(9).

FA may reveal a well-limited hyperfluorescence area enclosed in the diffuse leakage area associated with the PED, identical to that occurring on a minimally classic choroidal neovascular membrane. ICG-A clearly establishes stage II as the presence of a hot spot with leakage, as well as a hyperfluorescent area associated with serous PED.

Leakage may not be revealed by FA, probably because fibrin from retinal exudates cannot be impregnated with fluorescein, opposing to what occurs with ICG-A(16).

In fact, FA is rarely useful in the differential diagnosis between classic, occult or minimally classic membranes and early RAP lesions (stages I and II)(9).

On the other hand, ICG-A reveals a slow-growing extra, juxta or subfoveal hot spot, even in asymptomatic patients with RAP stages I and II.

 

Stage III – Choroidal Neovascularization

Stage III is established when the clinical and angiographic examinations can clearly demonstrate the presence of choroidal neovascularization (CNV), sometimes with the appearance of a vascularized PED.

According to Gass(11), RAP would progress in 5 stages instead of 3, easily identifiable by FA and ICG-A:

I) Pre-clinical stage - Atrophy of the outer retina, with retinal capillaries moving closer to a choroidal neovascular complex located below the RPE - type 1 CNV – and no clinical signs of chorioretinal anastomosis. ICG-A would be necessary to identify type 1 neovascularization.

II) Early clinical signs - Involvement of the subretinal space with localized neurosensory detachment of the retina, oedema and retinal haemorrhages at the edges may already be observed in fundus colour photography.

Anastomosis between dilated capillaries of the deep retina and the choroidal neovascular complex is associated with small intraretinal hemorrhages, which would constitute the first clinical sign of chorioretinal anastomosis.

This stage may occur weeks or months before stage III, where subretinal CNV is already observed.

III) Proliferation of CNV over the RPE - subretinal neovascularization – type 2 CNV.

IV) Appearance of serous PED - caused by activation of newly formed subepithelial vessels.

V) Mixed neovascularization - piggyback-type neovascularization, with two levels – type 1 and type 2 with cicatricial disciform lesion, making chorioretinal anastomosis visible.

Stage III in the Gass classification(11) corresponds to stage I in the Yannuzzi classification(9), with stages IV and V in the Gass classification(11) corresponding to stages II and III, respectively. Currently, the Yannuzzi classification(9) is the most widely used.

Recently, an optical coherence tomography (OCT) based classification was developed by Su et al.(17) According to the authors, a precursor stage consisting of punctate hyperreflective foci in the outer retina is the first OCT finding:

Stage 1 consists of a larger intraretinal hyperreflective lesion associated with cystoid macular edema but without outer retinal disruption.

Stage 2 is notable for outer retinal disruption that occurs with RPE disruption in most of the cases. Cystoid macular edema is generally present.

Stage 3 is defined by an intraretinal hyperreflective lesion that extends through the RPE to vascularize a drusenoid/serous PED. Cystoid macular edema is generally present.

 

Clinical Findings

The presence of small hemorrhages, soft drusen, and pigmentary changes is well documented in eyes with RAP and in unaffected fellow eyes, both in caucasian and in asian populations(9,18-21).

Marques et al.(19) conducted a quantitative evaluation of the fundoscopic features of fellow eyes of RAP and compared it with a cohort of fellow eyes of typical exudative AMD.

The authors reported that the total number and area of drusen are significantly smaller in fellow eyes of RAP than in fellow eyes of typical AMD.

Aside from soft drusen, a high frequency of reticular pseudodrusen in eyes with RAP has also been described(18).

The presence of telangiectasias, microaneurysms or punctiform retinal hemorrhages and RRA associated with intraretinal fluid is highly suggestive of RAP (Figure 2), as is the sudden disappearance of a retinal vessel that appears to have moved deeper(8-14,19,22).

Hard exudates usually accumulate around the central retinal neovascular complex, bordering the PED in stage III RAP lesions.

The advent of OCT, fundus autofluorescence (FAF) and infrared (IR) imaging allowed for a better characterization of the structural changes associated with the disease(23). Figures 2, 3, 4, 5 and 6 show clinical examples of patients with RAP who underwent multimodal retinal imaging.

Figure 2. Multimodal retinal imaging of a RAP lesion. (A) Colour fundus photography displaying intraretinal hemorrhages, hard exudates and neurosensory retinal detachment. (B) FA shows a retinal hiperfluorescent spot apparently at the end of one retinal vessel, an intraretinal hemorrhage, neurosensory retinal detachment and PED. (C) Early phase ICG-A with an intraretinal hot spot (angiomatous proliferation) over diffuse choroidal hyperfluorescence. (D) Late phase ICG-A reveals a subfoveal hiperfluorescent plaque (subretinal neovascularization).

 

Figure 3. ICG-A of a RAP lesion. A RAP lesion with RRA, an apparently intraretinal angiomatous mass and a serous PED, imaged with ICG-A.

 

Figure 4. Multimodal retinal imaging of a RAP lesion. Colour fundus photography (A) with intraretinal hemorrhages, hard exudates and neurosensory detachment. Red-free imaging (B) with two juxtafoveal and extrafoveal small hemorrhages. FA (C) shows two juxtafoveal hyperfluorescent spots, neurosensory detachment and PED. Late ICG-A (D) reveals two juxtafoveal hyperfluorescent hot spots. FA (E) and ICG-A (F) after treatment with laser photocoagulation showing resolution of the exudation.

 

Figure 5. Bilateral RAP lesion: Colour fundus photography (A and B) and red-free images (C and D) clearly show small retinal hemorrhages and lipid exudation. (E and F) Late ICG-A of both eyes with a hot spot and subfoveal hypofluorescence (serous PED). Stratus OCT (G and H) reveals serous PED, neurosensory retinal detachment and intraretinal fluid in both eyes.

 

Figure 6. RAP lesion imaged with spectral-domain (SD)-OCT. SD-OCT of a RAP lesion with an area of intraretinal neovascularization (green arrow) in the deep retina with a RPE detachment (white arrow). Intact and ruptured portions of the Bruch’s membrane and subretinal fluid (yellow arrow) can also be seen.

 

The recent introduction of OCT angiography (OCTA) in clinical practice allowed for an unmet morphological characterization of RAP lesions.

Qualitative and quantitative analyses of type 3 neovascular complexes can be performed using OCTA(24,25).

Querques et al.(26) evaluated the features of RAP with this new imaging modality and found lesions emerging from the deep capillary plexus, forming a clear, tuft-shaped, high-flow network in the outer retinal segment in all eyes, abutting in the sub-RPE space.

The authors also reported a small, clew-like lesion present in the choriocapillaris and that, in some cases, this clew-like lesion seemed to be connected to the choroid through a small-caliber vessel.

Compared with conventional imaging, OCTA may improve the detection and delineation of vascular changes occurring in RAP(25).

Further longitudinal studies with OCTA may bring further insights on the pathophysiology of RAP and establish OCTA imaging as an additional biomarker to guide the treatment and monitoring of these lesions.

 

Differential diagnosis

Differential diagnosis is mandatory for:

  • Other forms of CNV, namely occult CNV
  • Polypoidal choroidal vasculopathy (PCV)
  • Macular telangiectasia (MacTel)

Other forms of CNV with ICG hot spots (occult CNV) and PCV are two important diagnoses that can be easily confounded with RAP.

Small intraretinal hemorrhages, sometimes punctiform, in patients with soft drusen, are very typical in RAP, as are telangiectasias and RRA.

On the contrary, retinal hemorrhages in PCV are normally larger, with round reddish-orange nodules usually being observed in the fundus(27).

OCT is a useful tool in the differential diagnosis between RAP, PCV and occult CNV(28,29).

In RAP, intraretinal hyperreflectivity may be observed, corresponding to angiomatous proliferation associated with intraretinal fluid and/or RPE detachment(30,31).

In PCV, polyps appear in OCT as abrupt protrusions from the RPE/Bruch’s membrane band, often associated with neurosensory detachment(27,28).

The double layer sign, RPE irregularity and a notched PED are hallmarks of PCV on OCT(32).

MacTel is a condition involving dilation of retinal capillaries located near the fovea, in one or both eyes.

RPE hyperplasia may also occur, with refractive punctiform deposits and macular leakage being observed in FA.

Migration of one or more venules to the deep retina can also be found(5).

Also, anastomoses between retinal vessels and the choroidal circulation have been described, as well as choroidal new vessels.

The most significant differences are the fact that MacTel are not associated with serous PED, the RPE is healthier and CNV occurs less frequently(5,22).

 

Natural history

Natural history may be highly variable and in many aspects similar to that of other neovascular AMD lesions.

RAP has been coined a bilaterally aggressive disease with predictable symmetry(9,33,34).

In fact, the vasogenic potential associated with RAP has been highlighted by several group(18,33,35), reporting annual and cumulative rates of neovascularization in fellow eyes far exceeding the typical forms of exudative AMD.

Yannuzzi et al.(9) reported an average of 15 months until the development of neovascularization in their cohort of fellow eyes of RAP.

In a cohort of 52 fellow eyes of RAP, Gross et al.(34) reported a cumulative incidence of neovascularization of 40% at 12 months, 56% at 24 months and 100% at 36 months.

Campa et al.(33) have also evaluated the cumulative incidence of neovascularization in a population of 37 fellow eyes of RAP but found a significantly inferior conversion rate (36.4% at 36 months).

Recently, a small cohort (20 fellow eyes of RAP) in a Japanese population(18) identified an incidence of neovascularization of 50% at ~49 months of follow-up (min 24; max 108 months).

RAP was historically associated with a poor natural history and high relapse rates. Kuhn et al.(10) studied 22 eyes and observed structural evolution towards classic membranes, signs of RPE rupture and fibrous scars in 36.4%, 4.6% and 31.8% of cases, respectively. The authors described a decrease in visual acuity (VA) in 77% of cases. Similarly, Hartnett et al.(36) reported that 9/11 patients they evaluated developed legal blindness.

A study by Viola et al.(37), where the authors investigated the natural history and visual outcome in 14 eyes with untreated RAP for an average of 20 months, found that, by the time of the final examination, VA had decreased to ≤20/200 in 11 eyes (69%), and 5/14 patients (36%) were legally blind. However, while several treatment approaches proved unsuccessful in the past(13,38), the blockade of vascular endothelium growth factor (VEGF) has considerably changed the paradigm of this condition, with recent studies showing similar functional results to typical exudative AMD(39,40).

An important feature of the natural history of RAP lesions is geographic atrophy (GA).

It was previously thought that GA in RAP was secondary to the treatments used(41,42).

However, in a recent study using FAF, McBain et al.(43) found high rates of GA in RAP, independently of treatment modalities (phtotodynamic therapy or anti-VEGFs) and other authors have reported similar results(39,44).

This finding may be a consequence of the reduced choroidal perfusion and reduced choroidal thickness in RAP, which can be the reason behind the high vasogenic potential reported with this disease(21,43,45).

 

Epidemiology

RAP is now considered a well-established neovascular phenotype of AMD, with distinct clinical and epidemiologic features, most likely due to genetic differences still not completely understood(46) .

Environmental risk factors such as age and arterial hypertension have been associated with the development of both RAP and typical exudative AMD(47) .

Even though the role of arterial hypertension is still matter of debate, it is now widely accepted that RAP patients are significantly older than typical exudative AMD patients – average age of 79 vs. 76 years(6,8,9,14,47).

The average and median ages of a series of 108 patients studied by Yannuzzi et al.(9) were 80 and 81 years, respectively.

Women are more frequently diagnosed with RAP, comprising 64.7% to 71% of the affected individuals(9-10,47).

In caucasian populations, RAP is estimated to represent 5-28% of wet AMD cases(9,15,34,40,48,49).

Using the ImageNet 1024 videoangiography system, our team observed a prevalence of 9.4% in a consecutive series of 563 patients(13).

Studies in asians suggest a considerably lower prevalence(35,50) and there are no known reports of RAP in blacks(14).

Prevalence appears to be greater in hyperpigmented eyes(8,9,51).

Rapid progress in identifying genetic risk factors for AMD susceptibility has been made over the last few years, including the identification of two major loci at chromosome 1q32 and 10q26(46,52).

Even though several studies have evaluated genetic risk factors for exudative AMD, very few have specifically addressed the genetic alterations found in RAP(46,53,54).

Wegscheider et al.(53) and Caramoy et al.(47) suggest that CFH gene variants are less frequent in patients with RAP than in other forms of exudative AMD, while ARMS2 variants, namely the single nucleotide polymorphism (SNP) A69S (rs10499024), appears to have a stronger association with this neovascular phenotype(46,54,55).

These findings stress the importance of detailed phenotyping in AMD in order to identify genetic biomarkers for the distinct AMD subtypes.

 

Anatomopathology

In 2000, Lafaut et al.(16) studied six “deep retinal vascular anomalous complex” lesion specimens from retinal translocation surgery in eyes evaluated with FA and ICG-A.

The deep vascular anomalous complex was located in front of the RPE, thus not representing CNV. Nodular fibrovascular complexes surrounded by a ring of diffuse drusen – basal linear deposits under electron microscopy –, pigment epithelium and an amorphous fibrous material containing photoreceptor outer segment debris were observed.

Even though the lesion was not covered by RPE, the latter was preserved around the lesion. The amorphous material did not cover the retinal surface of the membrane, which adhered directly to the outer nuclear layer in 50% of cases.

Anastomoses between fibrovascular nodules and the choroidal circulation were only observed in cases showing isolated disciform scars.

Avascular fibrocellular membranes were observed on the inside of Bruch’s membrane and, in three cases, on the choroidal side of diffuse drusen.

Fibrin was present in affected retinas; choroidal anastomoses were not observed. Still, the presence of retino-choroidal anastomoses could not be entirely excluded, since the specimens may not have included the entire lesion.

In a histopathological study performed in nine neovascular lesions classified as RAP, Shimada et al.(56)

found only intraretinal neovascularization in stage II cases.

In stage III cases, these authors found choroidal and intraretinal neovascularization; they concluded that their findings were in agreement with the classification proposed by Yannuzzi et al.(29).

Furthermore, the authors identified the expression of VEGF (in intraretinal neovascularization and in the RPE), CD68-positive macrophages (in the neovascularization area) and expression of hypoxia-inducing factors (HIF) alpha 1 and alpha 2 in neovascular endothelial cells.

They hypothesized that intraretinal neovascularization would appear before the occurrence of CNV and be associated with ischemia and increased expression of VEGF and inflammatory factors.

Monson et al.(57)described the histopathological characteristics of RAP in an 87-year-old woman.

The images obtained by fundus examination and FA were histopathologically consistent with a neovascular intraretinal angiomatous complex without subretinal pigment epithelial neovascularization.

Gass et al.(11) described chorioretinal anastomoses and atrophy of the outer nuclear layer in a pre-clinical case, with outer retinal capillaries moving close to a CNV focus, having proposed a choroidal origin for RAP, instead of the retinal origin initially proposed by Yannuzzi et al.(9).

The origin of the neovascularization process (choroidal or retinal) in RAP remains a controversial issue.

Regardless, the anatomopathological studies helped elucidate that RAP lesions are associated with ischemia, age-related macular alterations, increased VEGF production and macrophage expression, similar to other forms of AMD-related CNV.

 

Pathogenesis

Kuhn et al.(12)studied the evolution of RAP in the second eye of two patients.

The authors assumed the existence of two asymmetrical membranes, a smaller retinal membrane – the angiomatous anomaly – over a larger membrane – the choroidal membrane, secondary to failure of a diseased RPE, unable to modulate and inhibit neovascular factors.

Even though the exact mechanism and origin of RAP remains elusive, the important role of VEGF cannot be overemphasized. The presence of serous PED may increase hypoxia by moving the retina even further from the choroid(36).

RPE decompensation is probably behind the deposition of amorphous materials in the subretinal space (under the basal lamina) and consequent thickening of Bruch’s membrane, all contributing to a decrease in inner retinal oxygenation.

Choriocapillaris atrophy and hypoxia may lead to intense neovascular activity, with reactive synthesis of growth factors, such as VEGF(58).

Overexpression of VEGF is sufficient to produce intraretinal neovascularization and subretinal neovascular membranes in animal and human models(58,59).

The relatively good response of these lesions to ranibizumab, bevacizumab and aflibercept confirms the central role of VEGF in the pathogenesis of RAP lesions.

As described above, Yannuzzi et al.(9) initially hypothesized that the initial neovascular process in RAP occurs in the deep retina and consists of 3 stages. Gass(11) contested a retinal origin for anastomosis, based on the following facts:

  1. no communication between retinal vessels and the choroid could be found in surgical specimens;
  2. the location of vascular anomalous complexes on the outer nuclear layer of atrophic retinas is more compatible with a choroidal origin hypothesis;
  3. although the macular retina is not excised in surgery, the retinal neovascular complex disappears and no macular hole is left, which contradicts its retinal origin.

It is clear that no definitive sequential histopathological evidence exists to support an intraretinal versus a choroidal origin for RAP lesions.

In 2008, Yannuzzi et al.(14) re-evaluated the early stages of RAP lesions in five eyes, using FA, ICG-A, time-domain-OCT and frequency-domain -OCT, having concluded that the initial lesion may have its origin not only on the deep retinal capillaries but also in the choroid.

They described RAP as “type 3 neovascularization”, a type of neovascularization with preference for the retina, displaying the following manifestations:

  1. Focal neovascular proliferation from the deep retinal layer (originally RAP);
  2. Intraretinal neovascular extension from underlying occult type I CNV (originally occult chorioretinal anastomosis);
  3. De novo breaks in Bruch’s membrane with neovascular infiltration into the retina.

The term type 3 neovascularization helps to resolve the various conflicting theories and descriptions behind the origin of RAP lesions: neovascularization in RAP may originate not only from the deep retinal capillaries but also from the choroid.

However, the main issue is not the intraretinal or choroidal origin of neovascularization but its unique characteristic of presenting two neovascular foci: one located in the deep retina and the other at a choroidal level.

 

Treatment

Before the anti-VEGF era, several treatment methods that included direct laser photocoagulation of the vascular lesion, laser photocoagulation of the feeder retinal arteriole, surgical ablation, scatter grid-like laser photocoagulation, photodynamic therapy (PDT), transpupillary thermotherapy, intravitreal triamcinolone acetonide (IVTA) and combined regimens were used, yielding at most VA stabilization, marginally better VA or short-term VA improvements(12,13,60,61).

 

Thermal laser Photocoagulation

Kuhn et al.(10) photocoagulated hot spots in 28 eyes, reporting “occlusion” in 25% of cases, including one eye with recurrence.

No occlusion was observed in the remaining 75% of cases, despite multiple treatment sessions.

Treatment caused tearing of the RPE in 45% of cases and VA decreased in 86% of eyes.

The low success rate might have been related to inadequate laser penetration – given the serous PED height or absorption of radiation by the sub-RPE fluid –, incorrect location or the presence of multiple afferent vessels.

Slakter et al.(8) confirmed this poor prognosis, particularly in serous PED cases. Occlusion was not possible in 86% of eyes and progressive evolution to a disciform scar was noticed.

Clinical experience showed that some extrafoveal stage I lesions (according to the Yannuzzi classification(9)) are amenable to laser photocoagulation(62).

In a series of 108 eyes with RAP, Bottoni et al.(60) achieved full obliteration of chorioretinal anastomosis in 57.1% of these lesions.

However, the risk of complications needs to be evaluated and careful follow-up is mandatory, given the high rates of persistence and recurrence.

 

Surgical ablation

Borrillo et al.(63)performed vitrectomy with detachment of the posterior hyaloid face, surgical section of afferent arterioles and draining veins and membrane excision in 4 eyes with RAP and PED – stage II – with resolution of intraretinal edema, collapse of the PED after 7-10 days and an increase in the average VA (20/200 pre-operative vs. 20/70).

Shimada et al.(56) excised the neovascular complex in 9 eyes from 8 patients – stages II and III.

VA remained stable in the post-operative period; however, significant destruction of the RPE and the choriocapillaris occurred in patients with serous PED.

The authors concluded that surgery may stabilize VA in stage III but is not indicated in stage II.

Surgical section of the afferent vessel associated with intravitreal injection of triamcinolone was performed in one eye, with resolution of exudation and VA of 20/320 at 6 months(64).

Nakata et al.(65) found surgical ablation (even combined with PDT) not useful for the treatment of RAP lesions, given the high frequency of reperfusion from retinal inflow vessels associated with this procedure.

Shiragami et al.(66) observed recurrence of RAP lesions in all 7 cases treated, 2 to 13 months after surgical ablation.

Surgical ablation of RAP is no longer recommended for RAP lesions, considering the better outcomes achieved with other treatment modalities, such as antiangiogenic drugs.

 

Photodynamic therapy (PDT) alone or in combination with intravitreal triamcinolone acetonide (IVTA)

Studies of PDT in monkeys with neovascular complexes and chorioretinal anastomoses, showed occlusion of the neovascular complex but not of the anastomoses(67).

The authors hypothesized that the anastomoses would be responsible for repermeabilization of the neovascular complex.

Kusserow et al.(68) treated 6 eyes with predominantly classic membranes and chorioretinal anastomoses without success.

No improvements in VA were observed for any of the treated eyes and membranes continued to grow.

Our group reported stabilization or improvement in VA after PDT in 73.3% of eyes (<3 lines loss) at 12 months(13), thus representing a better outcome than the natural evolution(36).

However, a significant VA decline was observed in the second year, mainly due to recurrence(12).

Boscia et al.(69) treated 13 eyes with PDT, having concluded that PDT would only be useful in cases where serous PED represents less than 50% of the lesion.

In 2006, these authors referred that early treatment of eyes with smaller lesions using PDT with verteporfin potentially led to a beneficial effect on vision, whereas it might worsen the natural progression of larger lesions, with most eyes undergoing enlargement, disciform transformation or RPE tear(70).

Reported short-term results of non-randomized studies on RAP lesions treated with PDT and IVTA(41,71,72) revealed apparently better VA outcomes and/or a reduced number of treatment sessions compared to PDT alone.

However, recurrences were also frequent(42,61).

Krebs et al.(73) found no significant differences between the PDT monotherapy group and the combined PDT and IVTA group regarding evolution of distance VA, retinal thickness and lesion size, having concluded that new therapeutic strategies might be required in RAP lesions, probably including therapy with antiangiogenic agents.

 

Antiangiogenic agents

Better visual outcomes can be achieved by treating RAP lesions with intravitreal ranibizumab(39,40,44,74,75), bevacizumab(76,77) and most recently aflibercept(78).

Data from clinical trials suggests that the response of RAP lesions to CNV treatments may be similar to that of other variants of neovascular AMD(79).

Improved VA and short-term edema reduction or elimination, were also observed in combined treatments using PDT and ranibizumab(44,80) or bevacizumab(81.82).

However, no superiority of combined treatment with PDT and ranibizumab or bevacizumab has been demonstrated.

The Comparison of Age-Related Macular Degeneration Treatments Trials (CATT) study followed up a large cohort of patients with treatment-naïve exudative AMD eyes, who received randomly assigned ranibizumab or bevacizumab for 2 years. The cohort included eyes with classic and occult CNV as well as RAP.

The authors recently published a comparison between eyes with and without RAP regarding the 2-year visual and morphologic outcomes(39).

This study showed a rapid improvement in VA in eyes with RAP within the first 3 months of intravitreal anti-VEGF therapy that continued to improve and then stabilized during the first year.

However, in the second year, VA began to decline modestly such that there was no statistically significant difference between eyes with or without RAP in overall VA gain at the end of the second year.

The authors noticed that fewer injections were needed to treat RAP than the other types of CNV but the rate of GA was higher in the RAP group(39).

A recently published study(44) on the long-term results (≥36 months of follow up) of PRN intravitreal ranibizumab for the treatment of RAP lesions in clinical practice (n=79), showed that ∼40% of the eyes had stable or improved VA at the end of follow-up, and about one fifth of the individuals preserved reading vision.

Nevertheless, about 60% of the patients presented with some degree of GA at the last visit.

The presence of subretinal fluid at baseline was found to positively correlate with a favorable visual outcome, thus appearing to be an important prognostic factor for functional improvement.

 

>> References

References - Neovascular Phenotypes: Retinal Angiomatous Proliferation (RAP) or Type 3 Neovascularization

References - Neovascular Phenotypes: Retinal Angiomatous Proliferation (RAP) or Type 3 Neovascularization

English
  1. Oeller JN. [Atlas of rare ophthalmoscopic conditions and supplementary plates to the atlas of ophthalmoscopy]: Wiesbaden, Germany; 1904.
     
  2. Slusher MM, Tyler ME. Choroidoretinal vascular anastomoses. Am J Ophthalmol 1980;90:217-22.
     
  3. Boozalis GT, Schachat AP, Green WR. Subretinal neovascularization from the retina in radiation retinopathy. Retina 1987;7:156-61.
     
  4. Kennedy JE, Wise GN. Retinochoroidal vascular anastomosis in uveitis. Am J Ophthalmol 1971;71:1221-5.
     
  5. Gass JD, Oyakawa RT. Idiopathic juxtafoveolar retinal telangiectasis. Arch Ophthalmol 1982;100:769-80.
     
  6. Green WR, Gass JD. Senile disciform degeneration of the macula. Retinal arterialization of the fibrous plaque demonstrated clinically and histopathologically. Arch Ophthalmol 1971;86:487-94.
     
  7. Hartnett ME, Weiter JJ, Garsd A, Jalkh AE. Classification of retinal pigment epithelial detachments associated with drusen. Graefes Arch Clin Exp Ophthalmol 1992;230:11-9.
     
  8. Slakter JS, Yannuzzi LA, Schneider U, et al. Retinal choroidal anastomoses and occult choroidal neovascularization in age-related macular degeneration. Ophthalmology 2000;107:742-53; discussion 53-4.
     
  9. Yannuzzi LA, Negrao S, Iida T, et al. Retinal angiomatous proliferation in age-related macular degeneration. Retina 2001;21:416-34.
     
  10. Kuhn D, Meunier I, Soubrane G, Coscas G. Imaging of chorioretinal anastomoses in vascularized retinal pigment epithelium detachments. Arch Ophthalmol 1995;113:1392-8.
     
  11. Gass JD, Agarwal A, Lavina AM, Tawansy KA. Focal inner retinal hemorrhages in patients with drusen: an early sign of occult choroidal neovascularization and chorioretinal anastomosis. Retina 2003;23:741-51.
     
  12. Silva RM, Cachulo ML, Figueira J, de Abreu JR, Cunha-Vaz JG. Chorioretinal anastomosis and photodynamic therapy: a two-year follow-up study. Graefes Arch Clin Exp Ophthalmol 2007;245:1131-9.
     
  13. Silva RM, Faria de Abreu JR, Travassos A, Cunha-Vaz JG. Stabilization of visual acuity with photodynamic therapy in eyes with chorioretinal anastomoses. Graefes Arch Clin Exp Ophthalmol 2004;242:368-76.
     
  14. Yannuzzi LA, Freund KB, Takahashi BS. Review of retinal angiomatous proliferation or type 3 neovascularization. Retina 2008;28:375-84.
     
  15. Freund KB, Ho IV, Barbazetto IA, et al. Type 3 neovascularization: the expanded spectrum of retinal angiomatous proliferation. Retina 2008;28:201-11.
     
  16. Lafaut BA, Aisenbrey S, Vanden Broecke C, Bartz-Schmidt KU. Clinicopathological correlation of deep retinal vascular anomalous complex in age related macular degeneration. Br J Ophthalmol 2000;84:1269-74.
     
  17. Su D, Lin S, Phasukkijwatana N, et al. An Updated Staging System of Type 3 Neovascularization Using Spectral Domain Optical Coherence Tomography. Retina 2016;36 Suppl 1:S40-S49.
     
  18. Sawa M, Ueno C, Gomi F, Nishida K. Incidence and characteristics of neovascularization in fellow eyes of Japanese patients with unilateral retinal angiomatous proliferation. Retina 2014;34:761-7.
     
  19. Marques JP, Lains I, Costa MA, et al. Retinal Angiomatous Proliferation: A Quantitative Analysis of the Fundoscopic Features of the Fellow Eye. Retina 2015;35:1985-91.
     
  20. Kim JH, Lee TG, Kim JW, Kim CG, Cho SW, Han JI. Small retinal haemorrhages accompanied by macular soft drusen: prevalence, and funduscopic and angiographic characteristics. Br J Ophthalmol 2014;98:1066-72.
     
  21. Kim JH, Kim JR, Kang SW, Kim SJ, Ha HS. Thinner choroid and greater drusen extent in retinal angiomatous proliferation than in typical exudative age-related macular degeneration. Am J Ophthalmol 2013;155:743-9, 9 e1-2.
     
  22. Kim JH, Lee TG, Kim JW, Kim CG, Cho SW, Han JI. Small retinal haemorrhages accompanied by macular soft drusen: prevalence, and funduscopic and angiographic characteristics. Br J Ophthalmol 2014;98:1066-72.
     
  23. Rouvas AA, Papakostas TD, Ntouraki A, Douvali M, Vergados I, Ladas ID. Angiographic and OCT features of retinal angiomatous proliferation. Eye (Lond) 2010;24:1633-42; quiz 43.
     
  24. Kuehlewein L, Dansingani KK, de Carlo TE, et al. Optical Coherence Tomography Angiography of Type 3 Neovascularization Secondary to Age-Related Macular Degeneration. Retina 2015;35:2229-35.
     
  25. Tan AC, Dansingani KK, Yannuzzi LA, Sarraf D, Freund KB. Type 3 Neovascularization Imaged with Cross-Sectional and En Face Optical Coherence Tomography Angiography. Retina 2017;37:234-246.
     
  26. Querques G, Miere A, Souied EH. Optical Coherence Tomography Angiography Features of Type 3 Neovascularization in Age-Related Macular Degeneration. Dev Ophthalmol 2016;56:57-61.
     
  27. Koh AH, Expert PCVP, Chen LJ, et al. Polypoidal choroidal vasculopathy: evidence-based guidelines for clinical diagnosis and treatment. Retina 2013;33:686-716.
     
  28. Liu R, Li J, Li Z, et al. Distinguishing Polypoidal Choroidal Vasculopathy from Typical Neovascular Age-Related Macular Degeneration Based on Spectral Domain Optical Coherence Tomography. Retina 2016;36:778-86.
     
  29. Mrejen S, Sarraf D, Mukkamala SK, Freund KB. Multimodal imaging of pigment epithelial detachment: a guide to evaluation. Retina 2013;33:1735-62.
     
  30. Brancato R, Introini U, Pierro L, et al. Optical coherence tomography (OCT) angiomatous prolifieration (RAP) in retinal. Eur J Ophthalmol 2002;12:467-72.
     
  31. Politoa A, Napolitano MC, Bandello F, Chiodini RG. The role of optical coherence tomography (OCT) in the diagnosis and management of retinal angiomatous proliferation (RAP) in patients with age-related macular degeneration. Ann Acad Med Singapore 2006;35:420-4.
     
  32. De Salvo G, Vaz-Pereira S, Keane PA, Tufail A, Liew G. Sensitivity and specificity of spectral-domain optical coherence tomography in detecting idiopathic polypoidal choroidal vasculopathy. Am J Ophthalmol 2014;158:1228-38 e1.
     
  33. Campa C, Harding SP, Pearce IA, Beare NA, Briggs MC, Heimann H. Incidence of neovascularization in the fellow eye of patients with unilateral retinal angiomatous proliferation. Eye (Lond) 2010;24:1585-9.
     
  34. Gross NE, Aizman A, Brucker A, Klancnik JM, Jr., Yannuzzi LA. Nature and risk of neovascularization in the fellow eye of patients with unilateral retinal angiomatous proliferation. Retina 2005;25:713-8.
     
  35. Fujimura S, Ueta T, Takahashi H, Obata R, Smith RT, Yanagi Y. Characteristics of fundus autofluorescence and drusen in the fellow eyes of Japanese patients with exudative age-related macular degeneration. Graefes Arch Clin Exp Ophthalmol 2013;251:1-9.
     
  36. Hartnett ME, Weiter JJ, Staurenghi G, Elsner AE. Deep retinal vascular anomalous complexes in advanced age-related macular degeneration. Ophthalmology 1996;103:2042-53.
     
  37. Viola F, Massacesi A, Orzalesi N, Ratiglia R, Staurenghi G. Retinal angiomatous proliferation: natural history and progression of visual loss. Retina 2009;29:732-9.
     
  38. Silva R, Cachulo ML, Fonseca P, et al. Age-related macular degeneration and risk factors for the development of choroidal neovascularisation in the fellow eye: a 3-year follow-up study. Ophthalmologica 2011;226:110-8.
     
  39. Daniel E, Shaffer J, Ying GS, et al. Outcomes in Eyes with Retinal Angiomatous Proliferation in the Comparison of Age-Related Macular Degeneration Treatments Trials (CATT). Ophthalmology 2016;123:609-16.
     
  40. Gharbiya M, Parisi F, Cruciani F, Bozzoni-Pantaleoni F, Pranno F, Abdolrahimzadeh S. Intravitreal Anti-Vascular Endothelial Growth Factor for retinal angiomatous proliferation in treatment-naive eyes: Long-term Functional and Anatomical Results Using A Modified PrONTO-Style Regimen. Retina 2014;34:298-305.
     
  41. Sutter FK, Kurz-Levin MM, Fleischhauer J, Bosch MM, Barthelmes D, Helbig H. Macular atrophy after combined intravitreal triamcinolone acetonide (IVTA) and photodynamic therapy (PDT) for retinal angiomatous proliferation (RAP). Klin Monbl Augenheilkd 2006;223:376-8.
     
  42. Montero JA, Ruiz-Moreno JM, Sanabria MR, Fernandez-Munoz M. Efficacy of intravitreal and periocular triamcinolone associated with photodynamic therapy for treatment of retinal angiomatous proliferation. Br J Ophthalmol 2009;93:166-70.
     
  43. McBain VA, Kumari R, Townend J, Lois N. Geographic atrophy in retinal angiomatous proliferation. Retina 2011;31:1043-52.
     
  44. Marques MF, Marques JP, Gil JQ, et al. Long-Term Management of RAP Lesions in Clinical Practice: Treatment Efficacy and Predictors of Functional Improvement. Ophthalmic Res 2016;55:119-25.
     
  45. Yamazaki T, Koizumi H, Yamagishi T, Kinoshita S. Subfoveal Choroidal Thickness in Retinal Angiomatous Proliferation. Retina 2014;34:1316-22.
     
  46. Hayashi H, Yamashiro K, Gotoh N, et al. CFH and ARMS2 variations in age-related macular degeneration, polypoidal choroidal vasculopathy, and retinal angiomatous proliferation. Invest Ophthalmol Vis Sci 2010;51:5914-9.
     
  47. Caramoy A, Ristau T, Lechanteur YT, et al. Environmental and genetic risk factors for retinal angiomatous proliferation. Acta Ophthalmol 2014;92:745-8.
     
  48. Bressler NM. Photodynamic therapy of subfoveal choroidal neovascularization in age-related macular degeneration with verteporfin: two-year results of 2 randomized clinical trials-tap report 2. Arch Ophthalmol 2001;119:198-207.
     
  49. Massacesi AL, Sacchi L, Bergamini F, Bottoni F. The prevalence of retinal angiomatous proliferation in age-related macular degeneration with occult choroidal neovascularization. Graefes Arch Clin Exp Ophthalmol 2008;246:89-92.
     
  50. Maruko I, Iida T, Saito M, Nagayama D, Saito K. Clinical characteristics of exudative age-related macular degeneration in Japanese patients. Am J Ophthalmol 2007;144:15-22.
     
  51. Spaide RF. Fundus autofluorescence and age-related macular degeneration. Ophthalmology 2003;110:392-9.
     
  52. Cantsilieris S, White SJ, Richardson AJ, Guymer RH, Baird PN. Comprehensive analysis of Copy Number Variation of genes at chromosome 1 and 10 loci associated with late age related macular degeneration. PLoS One 2012;7:e35255.
     
  53. Wegscheider BJ, Weger M, Renner W, et al. Association of complement factor H Y402H gene polymorphism with different subtypes of exudative age-related macular degeneration. Ophthalmology 2007;114:738-42.
     
  54. Ohkuma Y, Hayashi T, Sakai T, et al. Retinal angiomatous proliferation associated with risk alleles of ARMS2/HTRA1 gene polymorphisms in Japanese patients. Clin Ophthalmol 2014;8:143-8.
     
  55. Tanaka K, Nakayama T, Yuzawa M, et al. Analysis of candidate genes for age-related macular degeneration subtypes in the Japanese population. Mol Vis 2011;17:2751-8.
     
  56. Shimada H, Kawamura A, Mori R, Yuzawa M. Clinicopathological findings of retinal angiomatous proliferation. Graefes Arch Clin Exp Ophthalmol 2007;245:295-300.
     
  57. Monson DM, Smith JR, Klein ML, Wilson DJ. Clinicopathologic correlation of retinal angiomatous proliferation. Arch Ophthalmol 2008;126:1664-8.
     
  58. Aiello LP, Avery RL, Arrigg PG, et al. Vascular endothelial growth factor in ocular fluid of patients with diabetic retinopathy and other retinal disorders. N Engl J Med 1994;331:1480-7.
     
  59. Tolentino MJ, Miller JW, Gragoudas ES, et al. Intravitreous injections of vascular endothelial growth factor produce retinal ischemia and microangiopathy in an adult primate. Ophthalmology 1996;103:1820-8.
     
  60. Bottoni F, Massacesi A, Cigada M, Viola F, Musicco I, Staurenghi G. Treatment of retinal angiomatous proliferation in age-related macular degeneration: a series of 104 cases of retinal angiomatous proliferation. Arch Ophthalmol 2005;123:1644-50.
     
  61. Reche-Frutos J, Calvo-Gonzalez C, Donate-Lopez J, et al. Retinal angiomatous proliferation reactivation 6 months after high-dose intravitreal acetonide triamcinolone and photodynamic therapy. Eur J Ophthalmol 2007;17:979-82.
  62. Johnson TM, Glaser BM. Focal laser ablation of retinal angiomatous proliferation. Retina 2006;26:765-72.
     
  63. Borrillo JL, Sivalingam A, Martidis A, Federman JL. Surgical ablation of retinal angiomatous proliferation. Arch Ophthalmol 2003;121:558-61.
     
  64. Boscia F, Furino C, Prascina F, Delle Noci N, Sborgia L, Sborgia C. Combined surgical ablation and intravitreal triamcinolone acetonide for retinal angiomatous proliferation. Eur J Ophthalmol 2005;15:513-6.
     
  65. Nakata M, Yuzawa M, Kawamura A, Shimada H. Combining surgical ablation of retinal inflow and outflow vessels with photodynamic therapy for retinal angiomatous proliferation. Am J Ophthalmol 2006;141:968-70.
     
  66. Shiragami C, Iida T, Nagayama D, Baba T, Shiraga F. Recurrence after surgical ablation for retinal angiomatous proliferation. Retina 2007;27:198-203.
     
  67. Criswell MH, Ciulla TA, Lowseth LA, Small W, Danis RP, Carson DL. Anastomotic vessels remain viable after photodynamic therapy in primate models of choroidal neovascularization. Invest Ophthalmol Vis Sci 2005;46:2168-74.
     
  68. Kusserow C, Michels S, Schmidt-Erfurth U. [Chorioretinal anastomosis as unfavourable prognostic factor during photodynamic therapy]. Ophthalmologe 2003;100:197-202.
     
  69. Boscia F, Furino C, Sborgia L, Reibaldi M, Sborgia C. Photodynamic therapy for retinal angiomatous proliferations and pigment epithelium detachment. Am J Ophthalmol 2004;138:1077-9.
     
  70. Boscia F, Parodi MB, Furino C, Reibaldi M, Sborgia C. Photodynamic therapy with verteporfin for retinal angiomatous proliferation. Graefes Arch Clin Exp Ophthalmol 2006;244:1224-32.
     
  71. Mantel I, Ambresin A, Zografos L. Retinal angiomatous proliferation treated with a combination of intravitreal triamcinolone acetonide and photodynamic therapy with verteporfin. Eur J Ophthalmol 2006;16:705-10.
     
  72. van de Moere A, Kak R, Sandhu SS, Talks SJ. Anatomical and visual outcome of retinal angiomatous proliferation treated with photodynamic therapy and intravitreal triamcinolone. Am J Ophthalmol 2007;143:701-4.
     
  73. Krebs I, Krepler K, Stolba U, Goll A, Binder S. Retinal angiomatous proliferation: combined therapy of intravitreal triamcinolone acetonide and PDT versus PDT alone. Graefes Arch Clin Exp Ophthalmol 2008;246:237-43.
     
  74. Lai TY, Chan WM, Liu DT, Lam DS. Ranibizumab for retinal angiomatous proliferation in neovascular age-related macular degeneration. Graefes Arch Clin Exp Ophthalmol 2007;245:1877-80.
     
  75. Konstantinidis L, Mameletzi E, Mantel I, Pournaras JA, Zografos L, Ambresin A. Intravitreal ranibizumab (Lucentis) in the treatment of retinal angiomatous proliferation (RAP). Graefes Arch Clin Exp Ophthalmol 2009;247:1165-71.
     
  76. Montero JA, Fernandez MI, Gomez-Ulla F, Ruiz-Moreno JM. Efficacy of intravitreal bevacizumab to treat retinal angiomatous proliferation stage II and III. Eur J Ophthalmol 2009;19:448-51.
     
  77. Ghazi NG, Knape RM, Kirk TQ, Tiedeman JS, Conway BP. Intravitreal bevacizumab (avastin) treatment of retinal angiomatous proliferation. Retina 2008;28:689-95.
     
  78. Matsumoto H, Sato T, Morimoto M, et al. Treat-and-Extend Regimen with Aflibercept for Retinal Angiomatous Proliferation. Retina 2016;36:2282-2289.
     
  79. Scott AW, Bressler SB. Retinal angiomatous proliferation or retinal anastomosis to the lesion. Eye (Lond) 2010;24:491-6.
     
  80. Rouvas AA, Papakostas TD, Vavvas D, et al. Intravitreal ranibizumab, intravitreal ranibizumab with PDT, and intravitreal triamcinolone with PDT for the treatment of retinal angiomatous proliferation: a prospective study. Retina 2009;29:536-44.
     
  81. Saito M, Shiragami C, Shiraga F, Nagayama D, Iida T. Combined intravitreal bevacizumab and photodynamic therapy for retinal angiomatous proliferation. Am J Ophthalmol 2008;146:935-41 e1.
     
  82. Shima C, Gomi F, Sawa M, Sakaguchi H, Tsujikawa M, Tano Y. One-year results of combined photodynamic therapy and intravitreal bevacizumab injection for retinal pigment epithelial detachment secondary to age-related macular degeneration. Graefes Arch Clin Exp Ophthalmol 2009;247:899-906.

Neovascular Phenotypes: Polypoidal Choroidal Vasculopathy

Neovascular Phenotypes: Polypoidal Choroidal Vasculopathy

English

Authors:

Rufino Silva, MD, PhD1,2,3; Adrian Koh, MD, FRCOphth, FRCS (Edin), MMed (Ophthalmology)2

1- Faculty of Medicine. University of Coimbra., Institute for Biomedical Imaging and Life Sciences. (FMUC-IBILI). Portugal

2- Department of Ophthalmology. Centro Hospitalar e Universitário de Coimbra(CHUC). Portugal.

3- Association for Innovation and Biomedical Research on Light and Image (AIBILI). Coimbra. Portugal

4- Founding Partner & Senior Consultant, Eye & Retina Surgeons, Singapore

 

Updated/reviewed by the authors, July 2017.

Introduction

Polypoidal choroidal vasculopathy (PCV) was described for the first time in 1982(1).

Different names were proposed like posterior uveal bleeding syndrome(2) or multiple recurrent retinal pigment epithelium (RPE) detachments in black women(3).

It has a characteristic imaging expression on indocyanine green angiography (ICG), peculiar characteristics in optical coherent tomography (OCT) and apparently different responses to treatments when compared to occult or classic choroidal neovascularization.

Diagnosis is based on ICG and confirmed with characteristic fundus and OCT changes.

The primary abnormality was initially thought to involve the choriocapillaris, with the characteristic lesion consisting of a branching vascular network of vessels terminating, in one or more aneurismal dilatations, known as polyps.

Clinically a reddish orange, nodular or spheroid, polyp-like structure may be observed.

The natural course of the disease often follows a remitting-relapsing course, and clinically, it is associated with chronic, multiple, recurrent serosanguineous detachments of the RPE and neurosensory retina. In the majority of cases, there is significant long-term visual morbidity caused by fibrosis, RPE atrophy and secondary choroidal neovascularisation.

Cheung et al. reported that 50% of patients had reduction of vision to worse than 20/200 after mean follow up of 3 years(4).

A more recent knowledge has expanded the spectrum of PCV allowing us a clearer characterization of PCV. However, the classification of PCV as distinct subtype of exudative age-related macular degeneration (AMD) easily differentiated from other diseases and other subtypes of choroidal neovascularization associated with AMD is still controversial.

Polypoidal choroidal vasculopathy has also been described in different pathological conditions including central serous chorioretinopathy(5), circumscribed choroidal hemangioma(6); melanocytoma of the optic nerve(7); pathological myopia and staphyloma(8); or choroidal osteoma(9). PCV seems to behave more like a neovasculopathy occurring in a variety of different diagnoses than a distinct abnormality of the inner choroidal vasculature(10)

Most recently, Freund and others have described PCV as part of the pachychoroid spectrum, characterised by the following clinical and OCT features: thick choroid on enhanced depth imaging (EDI-OCT), dilated choroidal vessels in Haller layer, attenuation of the choriocapillaris and inner choroid (Sattler’s layer) and loss of fundus tessellation.

While it is plausible that some patients are predisposed to development of PCV due to underlying pachychoroid, it is unclear if the mechanism responsible for inner choroidal thinning and ischemia is largely mechanical (due to compression from the enlarged choroidal vessels) or primarily driven by hypoxic and inflammatory changes in the choriocapillaris.

 

Pathogenesis

The pathogenesis of PCV is not completely understood. It is widely accepted to be originated at the inner choroidal level.

Polyps may develop from a choroidal vascular network or from a plaque of occult new vessels(11,14).

Yuzawa et al.(13) described the filling of PCV lesion simultaneously with the surrounding choroidal arteries suggesting that PCV lesions grow from inner choroidal vessels.

Few clinicopathological studies have been reported.

MacCumber et al.(14) examined an enucleated high myopic eye with rubeosis and vitreous haemorrhage from a diabetic patient without diabetic retinopathy, with high blood pressure and history of multiple, bilateral, recurrent neurosensory and pigment epithelium detachments (PED).

Bruch’s membrane was crossed by choroidal vessels and an extensive fibrovascular proliferation was disclosed within Bruch’s membrane and the inner retinal space. They did not observe choroidal saccular dilatations except that some choroidal veins were quite large.

Inflammation was expressed by the presence of B and T lymphocytes at the level of choroid and fibrovascular tissue and the expression of intercellular adhesion cytokines like ICAM-1 was shown.

Lafaut et al.(15) reported the histopathologic features of surgically removed submacular tissue from an elderly patient with a pattern of PCV on ICG.

A thick fibrovascular membrane located on the choroidal side of the RPE was described. The RPE layer was discontinuous whereas on its choroidal side an almost intact layer of diffuse drusen was observed.

A group of dilated thin-walled vessels were found and located directly under diffuse drusen within a sub-RPE, intra-Bruch’s fibrovascular membrane.

Dilatations appeared to be of venular rather than arteriolar origin and some lesions were associated with lymphocytic infiltration.

The presence of choroidal infiltration by inflammatory cells was also referred by Rosa et al.(16)

Okubo et al.(17) described unusually dilated venules adjacent to an arteriole with marked sclerotic changes and newly formed capillaries within the wall of the degenerate arteriole and near the dilated venule.

Therasaki et al.(18) described clusters of dilated thin-walled vessels surrounded by macrophages and fibrinous material in neovascular membranes obtained from submacular surgery for PCV.

Hyalinization of choroidal vessels and massive exudation of fibrin and blood plasma were observed in all the five specimens of PCV lesions studied by Nakashizuka et al.(19).

They also found some blood vessels located above the RPE in two of the five eyes. Immunohistochemically, CD68-positive cells were described by them around the hyalinized vessels.

There were no alpha-SMA-positive cells in the vessels of PCV. CD34 staining showed endothelial discontinuity.

Vascular endothelial cells within the PCV specimens were negative for VEGF. HIF-1alpha positive inflammatory cells were located in the stroma of specimens(19).

Hyalinization of choroidal vessels, like arteriosclerosis, seems to be characteristic of PCV(19).

All these previous histopathological studies identifying a large spectrum features (like dilated choroidal vessels, intra-Bruch’s neovascularization, inflammatory cells, drusen material, thick membranes, single saccular dilatations or clusters of dilated thin walled vessels) may partially being expressing the influence of disease stage(20).

For many authors(11,15,21-25) and following the results of clinicopathological studies, PCV may represent a subtype of exudative AMD.

Yannuzzi et al.(24,25) found a prevalence of 7.8% of PCV in a population with signs of exudative AMD and Laffaut et al.(15) described the presence of late ICG hyperfluorescent plaques in 58% of 45 cases with PCV, proposing that PCV should be considered a subtype of exudative AMD.

Many other authors(23,24,25) describe PCV cases with subretinal neovascularization.

Ahuja et al.(26) described a prevalence of PCV in 85% of a consecutive series with 16 eyes diagnosed as exudative AMD and showing a PED greater than 2 mm of diameter, haemorrhage or retinal neurosensory exudation.

According to Yannuzzi(24,25)PCV and exudative AMD may differ in mean age of onset (PCV affected patients are younger), presence of exudative peripapillary lesions (more frequent in PCV), prevalence of soft drusen (greater in AMD patients) and ethnicity (PCV more prevalent in non-white population).

PCV also has less tendency to develop fibrous proliferation and a higher incidence of PED and neurosensory detachments.

PCV and AMD share common genetic factors, which suggests that PCV and wet AMD are similar in some pathophysiologic aspects.

A common genetic background may exist between typical exudative AMD and PCV patients. Complement pathway plays a substantial role in the pathogenesis of PCV, like in AMD.

A consistent association of the ARMS2/HTRA1 locus with both neovascular AMD and PCV suggests that the two diseases at least in part share molecular mechanisms(27).

The nonsynonymous variant I62V in the complement factor H gene is strongly associated with PCV and may be a plausible candidate for a causal polymorphism leading to the development of PCV, given its potential for functional consequences on the CFH protein(28).

The LOC387715/HTRA1 variants are associated with PCV and wet AMD in the Japanese population. The associations are stronger in AMD than in PCV(29).

LOC387715 rs10490924 correlates with vitreous hemorrhage and also with the lesion size. Rs10757278 on chromosome 9p21 was shown to be significantly associated with the risk of PCV in Chinese Han patients(30).

Among the patients with AMD and PCV, those with a homozygous HTRA1 rs11200638 risk allele seem to have larger CNV lesions(31).

Another susceptibility gene in PCV, the elastin gene (ELN), has shown to be more associated with PCV than with AMD(32).

With meta-analyses, variants in four genes were found to be significantly associated with PCV: LOC387715 rs10490924 (n=9, allelic odds ratio [OR]=2.27, p<0.00001), HTRA1 rs11200638 (n=4, OR=2.72, p<0.00001), CFH rs1061170 (n=4, OR=1.72, p<0.00001), CFH rs800292 (n=5, OR=2.10, p<0.00001), and C2 rs547154 (n=3, OR=0.56, p=0.01). LOC387715 rs10490924 was the only variant showing a significant difference between PCV and wet AMD (n=5, OR=0.66, p<0.00001)33.

PCV has been referred to be associated with other ocular disorders like macroaneurysms or inflammatory diseases(11,34).

However, this association is still inconclusive and deserves further investigation.

A relation between the retinal vascular changes in hypertensive retinopathy, like vascular remodelling, aneurismal dilatations and focal vascular constrictions, and choroidal alterations in PCV was proposed by Ross et al.(34).

 

Epidemiology

PCV is usually diagnosed in patients between 50 and 65 years old but the age of diagnoses may range between 20 and more than 80 years.

Most patients with PCV are likely to have AMD signs.

Prevalence of PCV in patients with AMD signs ranged between 4.8% and 23% in different series and different countries(3,11-14,21,35).

It is considered to be more prevalent in Asian population(35) and African American than in Caucasian(3,11,12,21) as it seems to preferentially affect pigmented individuals.

Preference for female gender is referred in Caucasian(11,12,21) whilst in Asian population male are more affected(13,35).

Bilateral involvement is common and may be as high as 86%(21).

 

Natural evolution

The disease has a remitting-relapsing course and is associated with chronic, multiple recurrent serosanguineous detachment of the neurosensory retina and RPE, with long-term preservation of vision(11,24,25).

Visual acuity (VA) loss is associated with central macular involvement and may range from mild to severe VA loss or blindness.

Treatment of central macular lesions with photodynamic therapy (PDT) and more recently with antiangiogenic drugs has precluded a better knowledge of natural history of PCV with macular involvement.

Approximately half the patients with PCV lesions in the posterior pole may have a favourable course without treatment(36).

In the remaining half the disorder may persist for a long time with occasional repeated bleeding and leakage, resulting in severe macular damage and VA loss.

Eyes with a cluster of grape-like polypoidal dilations of the vessels may have a higher risk for severe visual loss(37,38).

Choroidal vascular lesions may be located in the peripapillary area, in central macula or in midperiphery.

The analysis of VA outcome must consider location of polyps and/or abnormal vascular network. Most of the PCV natural history series describe lesions in the posterior pole, differentiating macular from extramacular and/or peripapillary polyps and macular involvement ranged from 25% to 94% (12,35,38,39).

Kwok et al.(38) followed the natural history of nine eyes with macular involvement after a follow-up ranging from 5 to 60 months and found VA improvement of two lines in only one eye (11.1%), VA change of one line in one eye, and VA decrease of two lines in seven eyes (77.7%).

Uhyama et al.(36) followed 14 eyes with PCV (13 with macular involvement) for a mean period of 39.9 months and described VA improvement of two lines in five eyes (35.7%) and VA decrease of two lines in four eyes (28.5%).

A favorable course was demonstrated in 50% of the cases with the remaining half of the cases showing recurrent leakage and hemorrhages and progressive VA loss.

Lesions may grow by enlargement with proliferation and hypertrophy of the vascular component but, apparently, not by confluence.

Polyps may bleed, grow, regress or leak and a choroidal neovascularization may appear

A massive spontaneous choroidal hemorrhage is rare but may constitute a severe complication associated with blindness(37).

Progression to RPE atrophy is common and may be related with resolution of PED, chronic or recurrent leakage with PED or neurosensory detachment, autoinfarction, regression or flattening of the lesion.

Chronic atrophy and foveal cystic degeneration is associated with severe VA loss(2,11,15,23,24,25,39).

 

Diagnosis

The current gold standard of diagnosis of PCV is based on ICG imaging (Figure 1) and may be complemented with OCT, fluorescein angiography and fundus findings (Figures 2, 3 and 4).

Clinical examination may show one or more reddish-orange, spheroid, subretinal mass located at the macular or juxtapapillary area (Figure 2).

This mass may correspond to the anteriorly projection of multiple polyps and is very suggestive of PCV.

Also very suggestive of PCV is the serous or serous-hemorrhagic PED and/or neurosensory detachment associated with extensive subretinal haemorrhages and circinate hard exudates (Figures 1,2 and 4).

Fluorescein angiography alone is not useful for PCV diagnosis.

Neurosensory detachment and serous or sero-hemorrhagic PED may suggest the diagnosis but polypoidal lesions are only visualized if the overlying pigment epithelium is atrophic.

Intermediate or late leakage on fluorescein angiography is very often diagnosed as occult choroidal neovascularization with late leakage from undetermined source or may be confused with chronic central serous chorioretinopathy.

The characteristic PCV lesion in ICG is sub-RPE network of vessels ending, in the great majority of cases, in aneurysmatic dilatations, which elevate the overlying RPE, giving rise to the characteristic sharp inverted V-shaped or thumb-like protrusions seen on the OCT.

Other important signs on OCT suggestive of underlying PCV are the irregular undulation of the RPE separated from the Bruch’s membrane (often referred to as the ‘double layer sign’) and adjacent PED.

The lesion may be juxtapapillary, macular or may be rarely located in the midperiphery.

Juxtapapillary lesions often show, in early ICG images, a radial arching pattern, and the vascular channels may be interconnected by smaller spanning branches more numerous at the edge of the lesions(39).

Interestingly, juxtapapillary lesions have been reported to be more common in patients of Caucasian and Afro-Caribbean descent compared to macular distribution of polyps being more commonly seen in East Asian patients.

When the polypoidal lesion is located in the macular area, the vascular network often arise in the macula and follows an oval distribution(39).

The area surrounding the vascular network is hypofluorescent during early phases of ICG and in late phase ICG often shows a reversal of the pattern: the area surrounding the polypoidal lesion becomes hyperfluorescent and the centre shows hypofluorescence.

In very late phases ICG shows disappearance of the fluorescence (washout) in non-leaking lesions (Figure 4-d)(11,22,24,25).

OCT and particularly 3-D OCT is very useful for diagnosis confirmation.

Documentation of polyps (number, location, size) and associated features (neurosensory detachment, serous or haemorrhagic PED, haemorrhage) may be assessed by OCT. Particular findings like the undulation of RPE (corresponding to the branch vascular network), double layer sign (the inner layer is RPE and the outer layer is the inner boundary of the Bruch/choroid complex), Bruch membrane thinning, the identification of polyps under a detached RPE and thicker choroid compared with typical neovascular AMD can be assessed as well.

Polypoidal lesion and branching vascular network identified in ICG may be visualized on spectral domain OCT in near 95% of the cases as areas of moderate reflectivity between the clearly delineated abnormal section of RPE and Bruch’s membrane(40).

Polypoidal lesions are visualized as anterior protrusions of a highly reflective RPE line.

With “en face OCT”(41)branching vascular networks were detected as elevations of the RPE.

Serous PED may be seen as round protrusions of the RPE and are often accompanied by adjacent smaller round protrusions of the RPE, consistent with polypoidal lesions.

These protrusions of the RPE are often fused and typically appeared as a ‘snowman’.

Subsequent longitudinal examination reveals the polypoidal lesions to be sharp protrusions of the RPE with moderate inner reflectivity.

Consistent with the location of the branching vascular network, a highly reflective line may be seen often just beneath the slightly elevated reflective line of RPE(40,41).

Optical coherence tomography angiography is able to detect the BVN in all cases and is located in the space between the RPE and Bruch's membrane. Polypoidal lesions can be detected in up to 85% of the cases, showing high flow signal. They are located just below the top of the pigment epithelial detachment42

Figure 1 - PCV. Red-free (A) image with circinate lipidic exudation. Intermediate phase ICG shows an abnormal choroidal vascular network and multiple polyps in the centre of the circinate exudation (B).

 

Figure 2 - Fundus colour photography with lipidic exudation and a reddish-orange, spheroid, subretinal mass located at the macular area associated with macular edema. Late ICG (top, right) reveals the presence of two polypoidal lesions in the papilomacular bundle. On OCT (bottom) the polyp is well delineated and a sub-foveal neurosensory detachment is observed.

 

Figure 3 - Same eye of Figure 2 after one PDT session. The two polyps show fluorescein staining (A; fluorescein angiography late phase). A complete polyps resolution is observed on late ICG (B). OCT (C): An intermediate reflectivity is now registered inside the polyp limits (no fluid) associated with a complete resolution of the neurosensory detachment.

 

Figure 4 - Fundus colour photography (A) reveals the presence of circinate lipidic exudation surrounding a reddish-orange lesion temporal to the fovea. Late phase fluorescein angiography (B) shows a diffuse leakage with juxtafoveal involvement and two focal areas of staining and leakage. Two polypoidal foci separated by an abnormal choroidal vascular network are observed on ICG intermediate phase (C). Late phase ICG (D) shows an hiperfluorescent plaque, a more temporal hot spot (active polyp) and a more nasal hypofluorescent spot with hyperflluorescent borders (apparently inactive polyp).

 

Differential diagnosis

Differential diagnosis of PCV with central serous chorioretinopathy, AMD with choroidal neovasculazrization, inflammatory conditions and tumors is not always easy and needs ICG for a clear differentiation.

Retinal vascular lesions such as retinal artery macroaneurysms and retinal microaneurysms may be mistaken for polypoidal lesions based on ICG findings, but can be easily differentiated with good stereopairs made available during the ICGA.

Polypoidal choroidal vasculopathy is a primary cause of macular serous retinal detachment without hemorrhage in patients over 50 years of age in Asian population(43).

Since clinical and fluorescein angiographic findings are often indistinguishable among central serous chorioretinopathy, PCV, and occult choroidal neovascularization, ICG might help to establish a more definitive diagnosis(22,44).

Central serous chorioretinopathy shows staining or late leakage but not an abnormal choroidal vascular network neither polyps.

Both these conditions are part of the pachychoroid spectrum of diseases characterised by loss of fundus tessellation, increased choroidal thickness, dilated large choroidal vessels in Hallers layer of the choroid, and attenuation of the choriocapillaris and Sattler’s layer.

The differential diagnosis becomes more challenging when lipid exudation and small PED are associated. ICG may be helpful differentiating PED from polypoidal lesions.

Small PED from central serous chorioretinopathy become hypofluorescent in late phases ICG and hyperfluorescent in late phases fluorescein angiography.

In contrast, polypoidal lesions are usually hyperfluorescent in late phases ICG because of its vascular nature(22).

PCV represents a subtype of type 1 CNV in AMD(11,15,23,24,25).However, some features distinguish PCV from other subtypes of CNV: eyes with PCV are characterized by a higher incidence of neurosensory detachments, greater neurosensory detachment height, and less intraretinal edema than eyes with occult or predominantly classic CNV(45).

Non polypoidal lesions in exudative AMD patients tend to produce small calibre vessels that are associated with grayish membranes not easily observed clinically, in contrast with the reddish-orange lesions clinically observed in PCV and corresponding to vascular saccular polypoid lesions(24,45,47).

Stromal choroidal fibrosis is common in predominantly classic and occult lesions but is quite rare in PCV.

PED associated with CNV in AMD has a poor prognosis whilst PED in PCV lesions virtually never forms fibrotic scars(22,24,39).

The natural evolution of CNV in AMD eyes to fibrosis and disciform scar is not observed in PCV eyes.

Tumoral lesions like choroidal circumscribed hemangioma, renal cell carcinoma or metastasis from carcinoid syndrome may also be confused with PCV(39). Again ICG is essential for differentiation.

Choroidal hemangiomas show, in general, a rapid filling of dye in very early phases and a washout in late phases.

ICG characteristic lesions of PCV are not observed in choroidal or metastatic tumors and ultrasound is also effectiveFigure 4. for characterization of the tumoral mass.

Inflammatory lesions, like posterior scleritis, multifocal choroiditis, panuveitis, acute posterior multifocal placoid pigment epitheliopathy, Harada disease, sympathetic uveitis and birdshot chorioretinopathy may also be confused with PCV.

PCV is not associated with anterior or posterior uveitis neither with leakage nor staining of the optic disc in fluorescein angiography(4,22,24,39).

Lipid deposition, often observed in PCV is not commonly seen in inflammatory conditions.

Scleral or choroidal thickening and liquid in the subtenon space have never been described in PCV eyes(24,39).

 

Treatment

Treatment of PCV lesions is only recommended when central vision is being threatened by persistent and progressive exudative changes.

Otherwise, a conservative approach is recommended.

Different treatment modalities like laser photocoagulation, transpupillary thermotherapy, PDT and surgery or intravitreal antiangiogenic drugs have been reported.

To date, only one randomized, controlled clinical trial (RCT) has been published to prove the efficacy and safety of ranibizumab vs PDT or PDT plus ranibizumab (the EVEREST trial).

There are currently two ongoing major RCT which will provide more robust evidence supporting the most optimal therapy for the condition.

Direct laser photocoagulation of leaking polyps has proven short-term safety and efficacy for extrafoveal lesions(46-49).

Other authors, however, described poor results(50) and persistence or even increasing exudation in up to 44% of the cases(15), or VA decrease of 2 or more lines in near half of the eyes and legal blindness in up to two third of the eyes(35).

Yuzawa et al.(49) reported good efficacy of laser photocoagulation in near 90% of the eyes if all the polyps and abnormal vascular network were treated.

If the treatment involved only the polyps, more than half of the eyes suffered VA decrease related with exudation, recurrences or foveal scars. Direct laser photocoagulation of feeder vessels, identified in ICG, was also reported as showing VA improvement of 2 or more lines in 8 out of 15 eyes(51).

Considering the possibility of using other treatment modalities, laser photocoagulation should be reserved for well defined extrafoveal active polyps.

Transpupillary thermotherapy has shown to be useful in PCV(52).

A large number of reports on PDT in subfoveal PCV has been published((23,26,53-57).

Results at one year and even at two years are apparently superior to those of PDT in predominantly classic or occult CNV.

A longer follow-up shows a trend to a progressive but not significant VA decline at two(57) and 3 years follow-up(58)

The rate of eyes gaining a significant amount of vision dropped from 26% in the first year to 15% in the third year, and the rate of eyes with significant VA loss (3 or more ETDRS lines) increased from 17% to 26%.

A high rate of recurrence (44%) occurred during the 3-years follow-up but it was not associated with significant VA decline(58).

This high rate of recurrences may be associated with a poor response of the vascular network to PDT(59).

VA decline at 3 years may partially be explained by photoreceptors death due to chronic or recurrent neurosensory detachment, massive hemorrhages and progressive RPE atrophy.

The number of treatments decreases markedly after the first year from an average of 2.0 to 0.4 and 0.5 in the second and third years respectively(58).

Complications like subretinal haemorrhages, haemorrhagic PED or even massive hemorrhages have been associated to PDT in PCV eyes(60).

However, all of these complications may also occur without any treatment.

Surgical removal of polypoidal lesions and associated hemorrhages has been reported(15-19,61) 

with and without macular translocation.

Considering the potential alternative treatments and the high rate of serious complications, macular translocation is no longer being considered for PCV(62).

The relatively large vascular lesions of PCV patients needs to be considered if a vitrectomy is planned in cases without associated massive hemorrhage.

Intravitreal antiangiogenic drugs, like bevacizumab(63) and ranibizumab(64) have been used for treating PCV eyes.

Ranibizumab short-term results(64) seem to be promising in terms of maintenance or improvement of VA and resolution of subretinal fluid and PED.

Polyps were reported to disappear in 69% of the cases at 3 months when using ranibizumab(64) but not with bevacizumab(63).

Intravitreal bevacizumab appeared also to result in stabilisation of vision and reduction of exudative retinal detachment in PCV patients in short-term evaluation.

However, it had limited effectiveness in causing regression of the polypoidal lesions(63).

Some exudative AMD eyes refractory to ranibizumab or bevacizumab may be, in fact, PCV cases.

Combined treatments associating PDT and intravitreal ranibizumab or bevacizumab have been reported to show good efficacy in these refractory cases of AMD(65).

The EVEREST study is the first multicenter, double-masked,ICG-guided randomized controlled trial with an angiographic treatment outcome designed to assess the effect of Visudyne® (PDT) alone or in combination with Lucentis® (ranibizumab) compared with Lucentis® alone in patients with symptomatic macular PCV.

A total of 61 PCV patients of Asian ethnicity from 5 countries (Hong Kong, Taiwan, Korea, Thailand, and Singapore) participated in the study.

The six months EVEREST study results(66) suggests that in a majority of patients, Visudyne® therapy, with or without Lucentis®, may lead to complete regression of the polyps that can cause vision loss in patients with PCV.

A complete polyp regression (primary endpoint) was achieved in 77.8% of patients who received the Visudyne® – Lucentis® combination, in 71.4% of Visudyne® monotherapy patients and in 28.6% of patients in the Lucentis® monotherapy group (p=0.0018 for combination, p=0.0037 for Visudyne® vs Lucentis®).

Best corrected VA from baseline to month six improved in average in all groups with patients in the combination group achieving the highest gain (+10.9 letters from baseline). Lucentis® monotherapy patients gained +9.2 letters, and Visudyne® monotherapy patients +7.5 letters. Differences between the groups are not statistically significant.

All therapies were well tolerated and the safety findings were consistent with the established safety profiles of Visudyne® and Lucentis®.

Three multicentre randomized clincial trials are beeing runned with Ranibizumab (EVEREST II) and Aflibercept (PLANET AND ATLANTIC) comparing the efficacy of anti-VEGF alone or in combination with PDT.

The results will help to clarify if a combined treratment is necessary for polyps closure, resolution of exudation and visual acuity imporovement. 

The EVEREST II is a phase 4, 2-year randomized clinical trial comparing ranibizumab alone with a combined teraphy of ranibizumab plus verteporfin PDT in 321 Asian patients.(71,41)

The PLANET randomized clinical trial enrolled 331 patients in Asia and in two European countries. 

It is a phase 3-4, 1-year study comparing aflibercept alone with aflibercept plus verteporfin PDT in patients with PCV.(72)

The ATLANTIC study is a phase 4, 1-year randomized clinical trial, being runned in Portugal and Spain, comparing intravitreal treat and extend aflibercept monotherapy with aflibercept treat and extend regimen with adjunctive PDT in patients with PCV. (73)

 

Combined therapies

The adjunctive use of intravitreal triamcinolone acetonide with PDT did not appear to result in additional benefit in the PDT treatment of PCV.

A retrospective analysis of PCV patients who underwent PDT with or without intravitreal triamcinolone acetonide, with a follow-up of 2 or more years, was reported in 27 eyes and showed to have a non-sustained effect after 1 year.(67)

The combination therapy between PDT and an anti-VEGF showed encouraging results including improved vision, reduced incidence of subretinal hemorrhage and reduced recurrence of polyps, when compared to PDT monotherapy for PCV. Reduced-fluence PDT combined with intravitreous bevacizumab for PCV improved vision and reduced complications.(68) even at two years.(69)

Combined treatment with PDT plus anti-VEGF should be considered in eyes diagnosed with leakage from the branching vascular network and the polyps, in eyes with visible exudation associated with PED and  when ICGA results do not permit clearly to distinguish between PCV and CNV and when both conditions co-exist.(70)

The results of the three ongoing multicenter and randomized studies will certainly help to clarify the need to use or not combined therapies in the treatment of PCV.

 

Conclusion

PCV may be considered a subtype of exudative AMD.

ICG is always mandatory for the diagnosis of PCV, showing a unique lesion – an abnormal inner choroidal vascular network with polypoidal structures at the borders.

OCT and fundus findings may complement the diagnosis.

PCV needs to be differentiated from other forms of exudative AMD, central serous chorioretinopathy, inflammatory conditions and some choroidal tumors.Photodynamic therapy with Visudyne®, alone or in combination with antiangiogenic drugs, seems to be necessary for a complete resolution of the polypoidal lesions.

 

>> References

References - Neovascular Phenotypes: Polypoidal Choroidal Vasculopathy

References - Neovascular Phenotypes: Polypoidal Choroidal Vasculopathy

English
  1. Yannuzzi LA. Idiopathic polypoidal choroidal vasculopathy. The Macula Society Meeting. Miami, USA, February 5, 1982. 1982;1.
     
  2. Kleiner RC, Brucker AJ, Johnston RL. The posterior uveal bleeding syndrome. Retina 1990;10(1):9-17.
     
  3. Stern RM, Zakov ZN, Zegarra H, Gutman FA. Multiple recurrent serosanguineous retinal pigment epithelial detachments in black women. Am J Ophthalmol 1985;100(4):560-569.
     
  4. Cheung CM, Yang E, Lee WK, Lee GK, Mathur R, Cheng J, Wong D, Wong TY, Lai TY. The natural history of polypoidal choroidal vasculopathy: a multi-center series of untreated Asian patients. Graefes Arch Clin Exp Ophthalmol 2015;253(12):2075-85.
     
  5. Yannuzzi LA, Freund KB, Goldbaum M, Scassellati-Sforzolini B, Guyer DR, Spaide RF, Maberley D, Wong DW, Slakter JS, Sorenson JA, Fisher YL, Orlock DA. Polypoidal choroidal vasculopathy masquerading as central serous chorioretinopathy. Ophthalmology 2000;107:767–777.
     
  6. Li H, Wen F, Wu D. Polypoidal choroidal vasculopathy in a patient with circumscribed choroidal hemangioma. Retina 2004;24:629–631.
     
  7. Bartlett HM, Willoughby B, Mandava N. Idiopathic polypoidal choroidal vasculopathy in a patient with melanocytoma of the optic nerve. Retina 2001;21:396–399.
     
  8. Mauget-Faÿsse M, Cornut PL, Quaranta El-Maftouhi M, Leys A. Polypoidal choroidal vasculopathy in tilted disk syndrome and high myopia with staphyloma. Am J Ophthalmol 2006;142:970–975.
     
  9. Fine HF, Ferrara DC, Ho IV, Takahashi B, Yannuzzi LA. Bilateral choroidal osteomas with polypoidal choroidal vasculopathy. Retin Cases Brief Rep 2008;2:15–17.
     
  10. Samira K, Engelbert ME, Yutaka  I,  Freund KB. Polypoidal choroidal vasculopathy: Simultaneous Indocyanine Green Angiography and Eye-Tracked Spectral Domain Optical Coherence Tomography Findings. Retina 2012;32:1057-1068.
     
  11. Ciardella AP, Donsoff IM, Huang SJ, Costa DL, Yannuzzi LA. Polypoidal choroidal vasculopathy. Surv Ophthalmol 2004;49(1):25-37.
     
  12. Scassellati-Sforzolini B, Mariotti C, Bryan R, Yannuzzi LA, Giuliani M, Giovannini A. Polypoidal choroidal vasculopathy in Italy. Retina 2001;21(2):121-125.
     
  13. Yuzawa M, Mori R, Kawamura A. The origins of polypoidal choroidal vasculopathy. Br J Ophthalmol 2005;89(5):602-607.
     
  14. MacCumber MW, Dastgheib K, Bressler NM, Chan CC, Harris M, Fine S, Green WR. Clinicopathologic correlation of the multiple recurrent serosanguineous retinal pigment epithelial detachments syndrome. Retina 1994;14(2):143-152.
     
  15. Lafaut BA, Leys AM, Snyers B, Rasquin F, De Laey JJ. Polypoidal choroidal vasculopathy in Caucasians. Graefes Arch Clin Exp Ophthalmol 2000;238:752–9.
     
  16. Rosa RH Jr, Davis JL, Eifrig CW. Clinicopathologic reports, case reports, and small case series: clinicopathologic correlation of idiopathic polypoidal choroidal vasculopathy. Arch Ophthalmol 2002;120(4):502-508.
     
  17. Okubo A, Sameshima M, Uemura A, Kanda S, Ohba N. Clinicopathological correlation of polypoidal choroidal vasculopathy revealed by ultrastructural study. Br J Ophthalmol 2002;86 (10):1093-1098.
     
  18. Terasaki H, Miyake Y, Suzuki T, Nakamura M, Nagasaka T. Polypoidal choroidal vasculopathy treated with macular translocation: clinical pathological correlation. Br J Ophthalmol 2002;86(3):321-327.
     
  19. Nakashizuka H, Mitsumata M, Okisaka S, Shimada H, Kawamura A, Mori R, Yuzawa M. Clinicopathologic findings in polypoidal choroidal vasculopathy. Invest Ophthalmol Vis Sci 2008;49(11):4729-4737.
     
  20. Okubo A, Hirakawa M, Ito M, Sameshima M, Sakamoto T. Clinical features of early and late stage polypoidal choroidal vasculopathy characterized by lesion size and disease duration. Graefes Arch Clin Exp Ophthalmol 2008;246(4):491-499.
     
  21. Guyomarch J, Jean-Charles A, Acis D, Donnio A, Richer R, Merle H. Vasculopathie polypoidale choroidienne idiopathique: aspects cliniques et angiographiques. J Fr Ophtalmol 2008;31(6 Pt 1):579-584.
     
  22. Spaide RF, Yannuzzi LA, Slakter JS, Sorenson J, Orlach DA. Indocyanine green videoangiography of idiopathic polypoidal choroidal vasculopathy. Retina 1995;15(2):100-110.
     
  23. Silva RM, Figueira J, Cachulo ML, Duarte L, Faria JR., Cunha-Vaz JG. Polypoidal choroidal vasculopathy and photodynamic therapy with verteporfin. Graefes Arch Clin Exp Ophthalmol 2005;243(10):973-979.
     
  24. Yannuzzi LA, Wong DW, Sforzolini BS, Goldbaum M, Tang KC, Spaide RF, Freund KB, Slakter JS, Guyer DR, Sorenson JA, Fisher Y, Maberley D, Orlock DA. Polypoidal choroidal vasculopathy and neovascularized age-related macular degeneration. Arch Ophthalmol 1999;117(11):1503-1510.
     
  25. Yannuzzi LA, Ciardella A, Spaide RF, Rabb M, Freund KB, Orlock DA. The expanding clinical spectrum of idiopathic polypoidal choroidal vasculopathy. Arch Ophthalmol 1997;115(4):478-485.
     
  26. Ahuja RM, Stanga PE, Vingerling JR, Reck AC, Bird AC. Polypoidal choroidal vasculopathy in exudative and haemorrhagic pigment epithelial detachments. Br J Ophthalmol 2000;84(5):479-484.
     
  27. Cheng Y, Huang L, Li X, Zhou P, Zeng W, Zhang C. Genetic and functional dissection of ARMS2 in age-related macular degeneration and polypoidal choroidal vasculopathy. PLoS One 2013;8: e53665.
     
  28. Kondo N, Honda S, Kuno S, Negi A. Coding variant I62V in the complement factor H gene is strongly associated with polypoidal choroidal vasculopathy. Ophthalmology 2009;116(2):304-310.
     
  29. Gotoh N, Nakanishi H, Hayashi H, Yamada R, Otani A, Tsujikawa A, Yamashiro K, Tamura H, Saito M, Saito K, Iida T, Matsuda F, Yoshimura N. ARMS2 (LOC387715) variants in Japanese patients with exudative age-related macular degeneration and polypoidal choroidal vasculopathy. Am J Ophthalmol 2009;147(6):1037-1041.
     
  30. Zhang X, Wen F, Zuo C, Li M, Chen H, Wu K. Association of genetic variation on chromosome 9p21 with polypoidal choroidal vasculopathy and neovascular age-related macular degeneration. Invest Ophthalmol Vis Sci 2001;52:8063-8067.
     
  31. Gotoh N, Yamada R, Nakanishi H, Saito M, Iida T, Matsuda F, Yoshimura N. Correlation between CFH Y402H and HTRA1 rs11200638 genotype to typical exudative age-related macular degeneration and polypoidal choroidal vasculopathy phenotype in the Japanese population. Clin Experiment Ophthalmol 2008; 36(5):437-442.
     
  32. Kondo N, Honda S, Ishibashi K, Tsukahara Y, Negi A. Elastin gene polymorphisms in neovascular age-related macular degeneration and polypoidal choroidal vasculopathy. Invest Ophthalmol Vis Sci 2008;49(3):1101-1105.
  33. Chen H, Liu K, Chen LJ, Hou P, Chen W, Pang CP. Genetic associations in polypoidal choroidal vasculopathy: a systematic review and meta-analysis. Mol Vis 2012;18:816-829.
     
  34. Ross RD, Gitter KA, Cohen G, Schomaker KS. Idiopathic polypoidal choroidal vasculopathy associated with retinal arterial macroaneurysm and hypertensive retinopathy. Retina 1996;16(2):105-111.
     
  35. Wen F, Chen C, Wu D, Li H. Polypoidal choroidal vasculopathy in elderly Chinese patients. Graefes Arch Clin Exp Ophthalmol 2004;242(8):625-629.
     
  36. Uyama M, Wada M, Nagai Y, Matsubara T, Matsunaga H, Fukushima I, Takahashi K, Matsumura M. Polypoidal choroidal vasculopathy: natural history. Am J Ophthalmol 2002;133 (5):639-648.
     
  37. Yang SS, Fu AD, McDonald HR, Johnson RN, Ai E, Jumper JM. Massive spontaneous choroidal hemorrhage. Retina 2003;23(2):139-144.
     
  38. Kwok AK, Lai TY, Chan CW, Neoh EL, Lam DS. Polypoidal choroidal vasculopathy in Chinese patients. Br J Ophthalmol 2002;86(8):892-897.
     
  39. Klais CM, Ciardella A, Yannuzzi LA. Polypoidal choroidal vasculopathy. In: Virgil Alfaro D, Liggett PE, Mieler WF, Quiroz-Mercado H, Jager RD, Tano Y, eds. Age-Related Macular Degeneration: A Comprehensive Textbook. Philadelphia, USA. Lippincott-Raven. 2006;6:53-63.
     
  40. Ojima Y, Hangai M, Sakamoto A, Tsujikawa A, Otani A, Tamura H, Yoshimura N. Improved visualization of polypoidal choroidal vasculopathy lesions using spectral-domain optical coherence tomography. Retina 2009;29(1):52-59.
     
  41. Kameda T, Tsujikawa A, Otani A, Sasahara M, Gotoh N, Tamura H, Yoshimura N. Polypoidal choroidal vasculopathy examined with en face optical coherence tomography. Clin Experiment Ophthalmol 2007;35(7):596-601.
     
  42. Wang M, Zhou Y, Gao SS, Liu W, Huang Y, Huang D, Jia Y. Evaluating Polypoidal Choroidal Vasculopathy With Optical Coherence Tomography Angiography.Invest Ophthalmol Vis Sci. 2016 Jul 1;57(9):OCT526-32. doi: 10.1167/iovs.15-18955.
     
  43. Hikichi T, Ohtsuka H, Higuchi M, Matsushita T, Ariga H, Kosaka S, Matsushita R. Causes of macular serous retinal detachments in Japanese patients 40 years and older. Retina 2009;29(3):395-404.
     
  44. Yannuzzi LA, Sorenson J, Spaide RF, Lipson B. Idiopathic polypoidal choroidal vasculopathy (IPCV). Retina 1990;10(1):1-8.
     
  45. Ozawa S, Ishikawa K, Ito Y, Nishihara H, Yamakoshi T, Hatta Y, Terasaki H. Differences in macular morphology between polypoidal choroidal vasculopathy and exudative age-related macular degeneration detected by optical coherence tomography. Retina 2009;29(6):793-802.
     
  46. Moorthy RS, Lyon AT, Rabb MF, Spaide RF, Yannuzzi LA, Jampol LM. Idiopathic polypoidal choroidal vasculopathy of the macula. Ophthalmology 1998;105(8):1380-1385.
     
  47. Guyer DR, Yannuzzi LA, Ladas I, Slakter JS, Sorenson JA, Orlock D. Indocyanine green-guided laser photocoagulation of focal spots at the edge of plaques of choroidal neovascularization. Arch Ophthalmol 1996;114(6):693-697.
     
  48. Gomez-Ulla F, Gonzalez F, Torreiro MG. Diode laser photocoagulation in idiopathic polypoidal choroidal vasculopathy. Retina 1998;18(5):481-483.
     
  49. Yuzawa M, Mori R, Haruyama M. A study of laser photocoagulation for polypoidal choroidal vasculopathy. Jpn J Ophthalmol 2003;47(4):379-384.
     
  50. Yamanishi A, Kawamura A, Yuzawa M. Laser photocoagulation for idiopathic polypoidal choroidal vasculopathy. Jpn J Clin O 1998;52:1691-1694.
     
  51. Nishijima K, Takahashi M, Akita J, Katsuta H, Tanemura M, Aikawa H, Mandai M, Takagi H, Kiryu J, Honda Y. Laser photocoagulation of indocyanine green angiographically identified feeder vessels to idiopathic polypoidal choroidal vasculopathy. Am J Ophthalmol 2004;137(4):770-773.
     
  52. Mitamura Y, Kubota-Taniai M, Okada K, Kitahashi M, Baba T, Mizunoya S, Yamamoto S. Comparison of photodynamic therapy to transpupillary thermotherapy for polypoidal choroidal vasculopathy. Eye (Lond) 2009;23(1):67-72.
     
  53. Gomi F, Ohji M, Sayanagi K, Sawa M, Sakaguchi H, Oshima Y, Ikuno Y, Tano Y. One-year outcomes of photodynamic therapy in age-related macular degeneration and polypoidal choroidal vasculopathy in Japanese patients. Ophthalmology 2008;115(1):141-146.
     
  54. Mauget-Faysse M, Quaranta-El MM, De La ME, Leys A. Photodynamic therapy with verteporfin in the treatment of exudative idiopathic polypoidal choroidal vasculopathy. Eur J Ophthalmol 2006;16(5):695-704.
     
  55. Chan WM, Lam DS, Lai TY, Liu DT, Li KK, Yao Y, Wong TH. Photodynamic therapy with verteporfin for symptomatic polypoidal choroidal vasculopathy: one-year results of a prospective case series. Ophthalmology 2004;111(8):1576-1584.
     
  56. Sayanagi K, Gomi F, Sawa M, Ohji M, Tano Y. Long-term follow-up of polypoidal choroidal vasculopathy after photodynamic therapy with verteporfin. Graefes Arch Clin Exp Ophthalmol 2007;245(10):1569-1571.
     
  57. Tsuchiya D, Yamamoto T, Kawasaki R, Yamashita H. Two-year visual outcomes after photodynamic therapy in age-related macular degeneration patients with or without polypoidal choroidal vasculopathy lesions. Retina 2009;29(7):960-965.
     
  58. Silva Leal S, Silva Rufino, Figueira J, Cachulo ML, Pires I, Faria de Abreu JR, Cunha-Vaz JG. Photodynamic therapy with verteporfin in polypoidal choroidal vasculopathy: Results after 3 years of follow-up. Retina 2010;30(8):1197-205.
     
  59. Lee WK, Lee PY, Lee SK. Photodynamic therapy for polypoidal choroidal vasculopathy: vaso-occlusive effect on the branching vascular network and origin of recurrence. Jpn J Ophthalmol 2008;52(2):108-115.
     
  60. Sayanagi K, Gomi F, Sawa M, Ohji M, Tano Y. Long-term follow-up of polypoidal choroidal vasculopathy after photodynamic therapy with verteporfin. Graefes Arch Clin Exp Ophthalmol 2007;245(10):1569-1571.
     
  61. Shiraga F, Matsuo T, Yokoe S, Takasu I, Okanouchi T, Ohtsuki H, Grossniklaus HE. Surgical treatment of submacular hemorrhage associated with idiopathic polypoidal choroidal vasculopathy. Am J Ophthalmol 1999;128(2):147-154.
     
  62. Fujii GY, Pieramici DJ, Humayun MS, Schachat AP, Reynolds SM, Melia M, De Juan E Jr. Complications associated with limited macular translocation. Am J Ophthalmol 2000;130(6):751-762.
     
  63. Lai TY, Chan WM, Liu DT, Luk FO, Lam DS. Intravitreal bevacizumab (Avastin) with or without photodynamic therapy for the treatment of polypoidal choroidal vasculopathy. Br J Ophthalmol 2008;92(5):661-666.
     
  64. Reche-Frutos J, Calvo-Gonzalez C, Donate-Lopez J, Garcia-Feijoo J, Leila M, Garcia-Sanchez J. Short-term anatomic effect of ranibizumab for polypoidal choroidal vasculopathy. Eur J Ophthalmol 2008;18(4):645-648.
     
  65. Reche-Frutos J, Calvo-Gonzalez C, Donate-Lopez J, Garcia-Feijoo J, Leila M, Garcia-Sanchez J. Short-term anatomic effect of ranibizumab for polypoidal choroidal vasculopathy. Eur J Ophthalmol 2008;18(4):645-648.
     
  66. QLT INC: 6-month results from EVEREST study evaluating Visudyne(R) therapy in patients with polypoidal choroidal vasculopathy. Available at: http://www.4-traders.com/QLT-INC-10600/news/QLT-INC-6-month-results-from-EVEREST-study-evaluating-Visudyne-R-therapy-in-patients-with-polypoi-13292269/. Accessed May 19, 2017.
     
  67. Lai TY, Lam CP, Luk FO, Chan RP, Chan WM, Liu DT, Lam DS. Photodynamic therapy with or without intravitreal triamcinolone acetonide for symptomatic polypoidal choroidal vasculopathy. J Ocul Pharmacol Ther 2010;26:91-95.
     
  68. Ricci F, Calabrese A, Regine F, Missiroli F, Ciardella AP. Combined reduced fluence photodynamic therapy and intravitreal ranibizumab for polypoidal choroidal vasculopathy. Retina 2012;32:1280-1288.
     
  69. Yoshida Y, Kohno T, Yamamoto M, Yoneda T, Iwami H, Shiraki K.  Two-year results of reduced-fluence photodynamic therapy combined with intravitreal ranibizumab for typical agerelated macular degeneration and polypoidal choroidal vasculopathy. Jpn J Ophthamol 2013;57:283-293.
     
  70. Koh AH; Expert PCV Panel, Chen LJ, Chen SJ, Chen Y, Giridhar A, Iida T, Kim H, Yuk Yau Lai T, Lee WK, Li X, Han Lim T, Ruamviboonsuk P, Sharma T, Tang S, Yuzawa M. Polypoidal choroidal vasculopathy: Evidence-Based Guidelines for Clinical Diagnosis and Treatment. Retina 2013;33:686-716.
     
  71. Visual Outcome in Patients With Symptomatic Macular PCV Treated With Either Ranibizumab as Monotherapy or Combined With Verteporfin Photodynamic Therapy. (EVEREST II). https://clinicaltrials.gov/ct2/show/NCT01846273
  72. Aflibercept in Polypoidal Choroidal Vasculopathy (PLANET) (https://clinicaltrials.gov/ct2/show/NCT02120950)
  73. Marques JP, Farinha C, Costa MA, Ferrão A, Sandrina Nunes, Rufino Silva. Protocol for a randomised, double-masked, sham-controlled phase 4 study on the efficacy, safety and tolerability of intravitreal aflibercept monotherapy compared with aflibercept with adjunctive photodynamic therapy in polypoidal choroidal vasculopathy: the ATLANTIC study. BMJ Open 2017 August 28, 7 (8): e015785

Serous PED

Serous PED

English

Authors:

Ugo Introini, MD, Giuseppe Casalino, MD
Department of Ophthalmology and University Vita-Salute

Scientific Institute San Raffaele Milano, Italy

Chair: Prof Francesco Bandello

 

Updated/reviewed by the author, July 2017.

Serous PED in AMD

 

Retinal pigment epithelial detachment (PED) is part of age-related macular degeneration (AMD) clinical spectrum.

However, different types of PED have been reported in the literature and they have been either related or not with AMD.

Serous PED is defined as an area of sharply demarcated, dome-shaped serous elevation of the retinal pigment epithelium (RPE). The histopathology of serous PED is consistent with the detachment of the RPE basement membrane, along with the overlying RPE from the remaining Bruch’s membrane due to accumulation of fluid(1).

The presence of this lesion is a negative prognostic factor for AMD in terms of visual acuity outcome.

While no definite therapeutic indications have been set so far, early detection of serous PED is important for the prognosis and the management of patients with AMD.

In AMD, serous PED can be either associated or not with choroidal new vessels (choroidal neovascularization – CNV). However, the vascularised type is by far the most observed.

Several theories regarding the relationship between serous PED and CNV have been proposed.

To explain its pathogenesis, Gass theorized the growth of new vessels from the choroid (Type 1 neovascularization (NV)) inside the Bruch’s membrane thickness, that actively leak, increasing the hydrostatic pressure and causing RPE detachment among the less adherent layers(2).

This concept has been later supported by the evidence that the development of CNV comes with inflammatory mechanisms that add more damage to Bruch’s membrane, supporting RPE separation from the inner collagenous layer(3-5).

When the growth of new vessels starts from the inner retina, more recently described as Type 3 NV and also known as retinal angiomatous proliferation (RAP), it has been hypothesized that the serous PED formation, which is very frequently associated, can be related to RPE invasion by the neovascular complex6-8).

By contrast, other authors observed that the presence of PED can represent a pre-existing condition that can promote CNV growth through a further Bruch’s membrane damage, expression of the same ongoing disease(9,10).

Although the pathogenesis of the PED is not completely understood, from these studies the formation of NV seems to be a pivotal moment.

At fundus examination, serous PED appears as a round or oval, distinct dome-shaped area of regular detachment of the RPE and the overlying neurosensory retina, with yellow to orange color and smooth surface. Margins are typically sharply demarcated; and focal RPE atrophy and pigment figures are frequently observed(9,11).

However, the concomitant presence of NV can generate a variety of associated ophthalmoscopic aspects, such as hemorrhagic and exudative components, areas of irregular elevation of the RPE and serous detachment of the surrounding neuroretina.

Presentation of Type 1 NV located at the margin of the PED may vary,  usually resulting in a reniform or notched aspect, or a flat-sided RPE detachment(12).

Serous PED may be imaged by fluorescein angiography (FA), indocyanine green angiography (ICGA) and optical coherence tomography (OCT).

FA represents, however, the gold standard for diagnosis of serous PED.

Examined by FA, serous PED classically shows an early uniform hyperfluorescence of the entire lesion, slightly delayed compared to the background fluorescence, that progressively increases in brightness as the examination progresses (pooling).

Serous PED hyperfluorescence typically does not change in size or shape during the angiographic phases.

FA can also demonstrate the presence of NV, usually associated to serous PED as Type 1 NV, like areas of indistinct late subretinal staining, more evident when located at the margin of the RPE detachment or corresponding to the “notch”(11).

The presence of NV can be also deducted by the presence of an hemorrhagic component of the PED, the dark meniscus described by Gass(12).

However, a more precise localization of the neovascular component can be obtained with digital ICGA. Indocyanine green molecule has biophysical properties that, unlike fluorescein, make it useful to enhance vessels anatomy through RPE, blood and turbid exudation.

In detail, ICGA enables to better delineate the presence and the type of new vessels associated with a serous PED, and for this reason is considered a fundamental tool in the management of this disease(13-15).

On ICGA, serous PED appears as an hypofluorescent lesion, with sharply delineated margins, that remains constantly hypofluorescent during all the phases of the examination(16).

When the new vessels are not present, no signs of localized hyperfluorescent areas are detectable; the outline of the PED is sharply round and it is therefore considered a pure serous PED.

In AMD patients, Yannuzzi found an incidence of 4% of non-vascularized PED among serous PED(15).

When the neovascular component is present, it has been suggested the term vascularized PED(15),.which accounts to approximately 24% of newly diagnosed exudative AMD(17).

New vessels associated with serous PED are represented in different subtypes.

High-speed videoangiography with scanning laser ophthalmoscope appears as a precious tool that allows the ophthalmologist to identify the new vessel pattern and their angiographic behaviour(18).

Recognize the different types of NV, by distinguishing angiographic findings, is mandatory for the distinct natural course, visual prognosis and different response to the treatment of the three main kinds of new vessels associated to serous PED in AMD.

The most common type of new vessels associated with serous PED are those occurring from the choroid beneath the RPE monolayer(15-,17).

These new vessels have been recently classified as Type 1 NV and are by far the most common type of NV in AMD(19)(Figure 1).

Figure 1 - Vascularized PED with Type 1 NV. (A) FA, (B)ICGA and (C) OCT.

 

In the early phases, ICGA shows the NV feeder artery that arises from the choroidal circulation, and subsequently the draining venule.

At the same time the capillary network of the neovascular membrane can be detected.

Unlike fluorescein, indocyanine green leaks slightly and the NV hyperfluorescence is usually minimal, with the exception of some cases that show an intense leakage, considered as very active new vessels.

Frequently, in the late phases, a well-defined area of mild hyperfluorescence corresponding to the NV network can be appreciable.

The second type of new vessels complicating serous PED are the RAP(7,20-22), also referred as Type 3 NV(19).

These vascular lesions, as reported by various authors, may involve the outer retina and the RPE, through a progression that has been hypothesized to originate from the retinal circulation and/or choroid.

ICGA typically shows the presence of a “hot-spot”, due to the early hyperfluorescence of the intraretinal neovascular complex, that increases during the angiography, with an intense leakage in the late phases.

Its brightness is enhanced by the surrounding hypofluorescence of the underlying PED (Figure 2).

Figure 2 - Vascularized PED with Type 3 NV (RAP). FA (A) and  ICGA early (B) and late phases (C and D).

 

In the late stages of the disease, the choroidal neovascular complex is typically connected with one or more retinal vessels which appear tortuous and dilated(7,22,23).

The Type 3 NV may be single or multiple, its origin is typically extrafoveal, and an intraretinal hemorrhage in correspondence of the neovascular lesion is frequently observed(20).

The third type of new vessels associated with serous PED in AMD is consistent with polypoidal choroidal vasculopathy (PCV)(24).

PCV is a peculiar form of CNV, characterized by the presence of orange, aneurismal, polyp-like round dilatations at the border of a branching vascular network of choroidal origin.

Although PCV affects more frequently middle-age black and asian populations, its clinical spectrum is expanded to whites, where it has been found to be present in 8-13% of patients with concomitant AMD lesions.

In these cases, when the manifestations attributable to both PCV and AMD are present, some authors consider PCV as a subtype of CNV in AMD(24,25).

Hemorrhagic manifestation is common in patients with PCV.

Serous PED associated with PCV shows frequently a blood level in the lower portion of the detachment.
ICGA is the state-of-the-art examination to distinguish the typical features of the two vascular components.

The vascular network is characterized by the presence of one or more aneurismal lesions that show a bright fluorescence since the early phases, followed in the late phases by a clearing of the dye, called “wash-out”, typical of this disease (Figure 3).

Figure 3 - Vascularized PED with PCV: FA (A) and early (B), mid (C) and late (D) phases of ICGA.

 

Nevertheless, some polyp-like structures can actively leak showing late staining of their walls and surrounding exudation.

The polypoidal lesions are usually located at the margin of the serous PED(26).

Recognition of these lesions is critical because of their different clinical course, prognosis and treatment response as compared to the other neovascular AMD subtypes.

OCT provides images that allow an exact correlation with the angiographic findings.

In cross-sectional OCT scans, serous PED appears as an optically empty dome-shaped elevation of the external high reflective band – the RPE, that steeply detaches from Bruch’s membrane(26).

The overlying retina, usually adherent to the bullous PED, at the margins of the lesion, can be slightly detached from the underlying RPE.

More additional information can be provided by OCT in vascularized PED(28).

The tomographic sections, guided by FA and ICGA in the area corresponding to the CNV, show a smoother elevation of the RPE, continuous with the serous detachment, with a deeper backscattering, due to the presence of the fibrovascular tissue.

Hyporeflective areas of homogeneous optically empty spaces referable to fluid accumulation are frequently present in the intraretinal and subretinal spaces(29).

Intraretinal optically empty spaces are more pronounced when the serous PED is associated with a Type 3 NV, especially with cystic shape (Figure 4).

By positioning the scan line corresponding to the “hot-spot”, the neovascular abnormality is represented as a dense or hyperreflective pre-epithelial zone in the inner retinal layers, where the outer hyperreflectant layers are no longer detectable(30).

The RPE close to that lesion shows frequently effractions or interruptions in its hyperreflective layer(31).

The retinal topographic measurement sustains an increased retinal thickness.

In eyes with serous PED and PCV, the polypoidal lesions show a sharp protrusion of RPE, similar to the PED but steeply sloped.

The polyps cavity, usually optically empty, is contiguous to irregular RPE elevation, expression of the occult neovascular component of the lesion(32,33).

Subretinal and intraretinal fluid, observed as hypofluorescent optical empty areas, are related to the PCV activity (Figure 4-B).

Figure 4 - Type 3 NV (RAP) (left) and PCV (right): OCT and ICGA patterns.

 

The recent introduction of OCT angiography (OCT-A) has made it possible to image perfusion of the different retinal layers, without injection of the dye and by utilizing endoluminal flow as intrinsic contrast.

Figure 5 provides an example of how OCT-A image the choroidal network in a case of PCV.

serou_ped_image5.png

Figure 5 - Serous PED and recurrent PCV (FA, OCT, ICGA and OCT-A). On the top, OCT scan acquired simultaneously with FA; On the bottom, appearance of the polyps profile on OCT-A (left) and ICGA (right).

 

However, the static nature of this examination and the presence of possible artefacts are important limitations that should be acknowledged.

With regard to serous PED, shadowing due to loss of signal transmission in correspondence of the PED may make it difficult to detect the CNV complex on OCT-A (Figure 6).

serou_ped_image6.png

Figure 6 - Vascularized PED and serous PED with Type 1 NV (FA, ICGA and OCT-A) . On the top, FA (left) shows dye pooling due to presence of a serous PED; ICGA (top middle and right) shows a neovascular network on the edge of the PED (notch); On the bottom, OCT-A does not provide a definite imaging of the neovascular net.

 

Serous PED natural course depends on the presence or not of the neovascular component(34).

In pure serous PED there is generally a slow enlargement of the lesion, with a minimal progression of visual loss over a long period (months or years).

However, many can subsequently develop neovascularization, which make it worse(35).

Natural course in vascularized PED may vary, and it is related to the type of new vessels associated.

The most common acute complication of PED is the tearing of the RPE(36-39).

It usually occurs at the edge of the PED, at the intersection of the detached and attached RPE.

Clinically, RPE tear or rip appears as a well-defined area of bare choroid, contiguous to a darker hyperpigmented rugate area, that corresponds to the mound of the RPE that has torn away(40,41).

The ripped RPE usually rolls towards the CNV, and its propensity to tear can be predicted by the observation of pre-tear characteristics, such as an increase in the size and a modification in the shape, the presence of small holes at the PED margins, the presence of hemorrhages or subretinal fluid, but the most noteworthy aspects are the irregular filling of the PED visible at the FA, height of PED > 580 nm, duration > 4.5 months, hyperreflective radial lines on near reflectance imaging, smaller ratio of vascularized PED and anti-vascular endothelial growth factor (anti-VEGF) therapy(42-46).

RPE tears occurs either spontaneouslyor after a treatment, formally laser photocoagulation, photodynamic therapy and intravitreal injection of steroids or anti-VEGF agents(47- 58).

The exact pathogenesis of RPE tears is poorly understood.

Concerning PEDs’ natural course, it has been hypothesized that tangential shearing forces in the PED can cause the break of the RPE basement membrane at the edge of the detachment; however, it is more likely the result of several variables, where the presence of a CNV plays a major role.

Several causal relationships have been reported for RPE tears occurring after treatment, including the heat generated by photocoagulation, the abrupt increase of intra-PED fluid, a contraction of the associated CNV and the concomitant sudden resolution of the sub RPE fluid.

The combining presence of vitreomacular traction and the deformation of the globe due to the mechanical trauma by the needle have also been reported as causative agents(59).

After RPE rips, the majority of patients complain of a sudden severe visual decrease.

In a small percentage of eyes, where the tear spares the fovea, patients can experience a temporary preservation of good visual function(60).

However, in the long term, the progression of a subretinal scar leads to a severe visual decrease. In the prognosis of serous PED, it must be also considered the high risk of bilateral involvement(61).

Treatment of serous PED, associated or not with CNV, has always been a challenge and so far there are no recommended guidelines for their management.

Pure serous PEDs have been treated in the past with laser grid or scattered photocoagulation, nevertheless with disappointing results(61).

Treatment of serous PED, associated or not with CNV, has always been challenging and there are no recommended guidelines for their management so far. Pure serous PEDs have been treated in the past with laser grid or scattered photocoagulation, with disappointing results(62).

No other approaches have been attempted to treat these lesions.

When a neovascular net is present, treatment of serous PED has been focused on CNV management. However, given that vascularized PEDshave never been included in the major randomized controlled trials, we need to make the treatment decision based on small series published, which are often retrospective and do involve different therapeutic approaches.

Now, in the anti-VEGF therapy era, all the previous employed treatments appear unsatisfactory.

Laser photocoagulation has been widely employed and might still have a limited indication when an ICGA-well defined CNV lies remote to the detached RPE(63).

Verteporfin photodynamic therapy (PDT) alone has been proved to be harmful, increasing the risk of RPE tear, hemorrhages and sudden visual acuity decrease(39,48,49,64) (Figure 7).

serou_ped_image7.png

Figure 7 - Vascularized PED with CNV (Type 1 NV) before (left) and after (right) PDT: RPE tear (FA and OCT).

 

However, PDT combined with intravitreal triamcinolone acetonide injection (IVTA) has been demonstrated as potentially capable of visual acuity stabilization and recurrences reduction(65).

Nevertheless, the high rate complications (cataract and glaucoma) has reduced the use of intravitreal triamcinolone.

After the encouraging results obtained with the anti-VEGF intravitreal therapy in the treatment of occult CNV, the use of anti-VEGF treatment have been extended to vascularized PED with disappointing results(66-69).

Both acute complications and poor anatomical response to the treatment frequently invalidate our attempts to heal the lesion. RPE tears and subretinal hemorrhages have been reported to complicate intravitreal ranibizumab and bevacizumab treatments(51-57).

Moreover, the sub-RPE fluid hardly responds to the anti-VEGF therapy, possibly due to the hydroconductivity changes of Bruch’s membrane(69).

In a retrospective case series of 328 patients treated with bevacizumab, ranibizumab, pegaptanib and PDT+IVTA respectively, after a mean follow-up of 42.4 weeks, the authors reported a significant stabilization of visual acuity in each group, better in the bevacizumab and ranibizumab ones compared to the other two, and an overall RPE tears frequency of 12.5%.

However, they conclude that with these treatments, only a partial regression of the lesions can be obtained, and the risk of RPE tears is not avoided(69).

Another retrospective study(58). reviewed patients outcome of vascularized PED treated with PDT alone, PDT combined with IVTA or intravitreal anti-VEGF injections alone (bevacizumab or ranibizumab) and showed better functional results for the anti-VEGF treatment group.

Moreover, in this series, Type 1 NV with vascularized PED as compared to Type 3 NV, along with a better visual acuity at baseline, showed greater risk of acute RPE tear after treatment(58).

In a recent prospective study(19), treatment of PED associated with subfoveal Type 1 NV with intravitreal ranibizumab injection and with a 3-monthly loading phase and a pro re nata strategy led to only partial results over a 24-month follow-up.

Recently several studies have investigated the efficacy of intravitreal aflibercept therapy of PED in AMD showing good anatomical response with improvement or no significant change in visual acuity(70-73).

Moreover, intravitreal aflibercept has been shown to be a promising treatment in PED resistant to intravitreal ranibizumab treatment(72-74)..

Several studies have identified factors which might influence response of PED to anti-VEGF treatment(75-78).

Dirani et al.(75) showed that better visual improvement was associated with lower baseline visual acuity, presence of subretinal fluid and RAP. Moreover, in their series, PED reduction was associated with higher PED at baseline, predominantly serous PED and use of aflibercept.

Cho et al.(76), in a recent case series, found that lower PED height at baseline, PCV or RAP compared to typical neovascular AMD, serous PED compared to fibrovascular PED, and aflibercept compared to ranibizumab, have an higher chance of PED resolution during anti-VEGF treatment of PEDs.

However, in a recent post hoc analysis of a phase III randomized controlled trial(77), it has been shown that, at 24 months after initiation of anti-VEGF treatment, around half of the patients presenting PED at baseline showed complete resolution of PED regardless of PED status and height at baseline.

Recently Chen et al.(78) outlined the importance of differentiating eyes presenting RAP as they havebetter anatomical and functional outcomes with fewer injections as compared to PED with Type 1 NV.

In the future, new combination therapies and new therapeutic strategies, along with identification of new clinical biomarkers of response to treatment, will help to improve the prognosis of the patients affected by vascularized PED.
 

>> References

References - Serous PED

References - Serous PED

English
  1. Murphy RP, Yeo JH, Green WR, et al. Dehiscences of the pigment epithelium. Trans Am Ophthalmol Soc 1985;83:63-81.
     
  2. Gass JD. Pathogenesis of disciform detachment of the neuroepithelium. Am J Ophthalmol 1967;63(Suppl):S1–S139.
     
  3. Anderson DH, Mullins RF, Hageman GS, et al. A role for local inflammation in the formation of drusen in the aging eye. Am J Ophthalmol 2002;134:411-31.
     
  4. Hutchinson AK, Grossniklaus HE, Capone A. Giant-cell reaction in surgically excised subretinal neovascular membrane. Arch Ophthalmol 1993;111:734-5.
     
  5. Oh H, Takagi H, Takagi C, et al. The potential angiogenic role of macrophages in the formation of choroidal neovascular membranes. Invest Ophthalmol Vis Sci 1999;40:1891-8.
     
  6. Hartnett ME, Weiter JJ, Garsd A, et al. Classification of retinal pigment epithelial detachments associated with drusen. Graefes Arch Clin Exp Ophthalmol 1992;230:11-9.
     
  7. Kuhn D, Meunier I, Soubrane G, et al. Imaging of chorioretinal anastomoses in vascularized retinal pigment epithelium detachments. Arch Ophthalmol 1995;113:1392-8.
     
  8. Slakter JS, Yannuzzi LA, Schneider U, et al. Retinal choroidal anastomoses and occult choroidal neovascularization in age-related macular degeneration. Ophthalmology 2000;107:742-53, discussion 753-4.
     
  9. Green WR, McDonnell PJ, Yeo JH. Pathologic features of senile macular degeneration. Ophthalmology 1985;92:615-27.
     
  10. Casswell AG, Kohen D, Bird AC. Retinal pigment epithelial detachments in the elderly: classification and outcome. Br J Ophthalmol 1985;69:397-403.
     
  11. Yannuzzi LA. Retinal pigment epithelial detachment, in Yannuzzi LA (ed): Laser Photocoagulation of the Macula. Philadelphia, Lippincott, 1989, pp 49-63.
     
  12. Gass JD. Serous retinal pigment epithelial detachment with a notch. A sign of occult choroidal neovascularization. Retina 1984;4:205-20.
     
  13. Yannuzzi LA, Slakter JS, Sorenson JA, et al. Digital indocyanine green videoangiography and choroidal neovascularization. Retina 1992;12:191-223.
     
  14. Guyer DR, Yannuzzi LA, Slakter JS, et al. Digital indocyanine-green videoangiography of occult choroidal neovascularization. Ophthalmology 1994;101:1727-35, discussion 1735-7.
     
  15. Yannuzzi LA, Hope-Ross M, Slakter JS, et al. Analysis of vascularized pigment epithelial detachments using indocyanine green videoangiography. Retina 1994;14:99-113.
     
  16. Flower RW, Csaky KG, Murphy RP. Disparity between fundus camera and scanning laser ophthalmoscope indocyanine green imaging of retinal pigment epithelium detachments. Retina 1998;18:260-8.
     
  17. Cohen SY, Creuzat-Garcher C, Darmon J, et al. Types of choroidal neovascularisation in newly diagnosed exudative age-related macular degeneration. Br J Ophthalmol 2007;91: 1173-1176.
     
  18. Wolf S, Remky A, Elsner AE, et al. Indocyanine green video angiography in patients with age-related maculopathy-related retinal pigment epithelial detachments. Ger J Ophthalmol 1994;3:224-7.
     
  19. Parodi MB, Iacono P, Papayannis A, Kontadakis SD, Cascavilla M, Pierro L, Gagliardi M, Bandello F. Intravitreal ranibizumab for pigment epithelium detachment with subfoveal occult choroidal neovascularization: a prospective 24-month case series. Am J Ophthalmol 2013;155(1):103-108.
     
  20. Yannuzzi LA, Negrao S, Iida T, et al. Retinal angiomatous proliferation in age-related macular degeneration. Retina 2001; 21:416–434.
     
  21. Gass JDM, Agarwal A, Lavina AM, et al. Focal inner retinal hemorrhages in patients with drusen. An early sign of occult choroidal neovascularization and chorioretinal anastomosis. Retina 2003; 23:741–751.
     
  22. Freund KB, Ho IV, Barbazetto IA, et al. Type 3 neovascularization: the expanded spectrum of retinal angiomatous proliferation. Retina 2008;28(2):201-11.
     
  23. Axer-Siegel R, Bourla D, Priel E, Yassur Y, Weinberger D. Angiographic and flow patterns of retinal choroidal anastomoses in age-related macular degeneration with occult choroidal neovascularization. Ophthalmology 2002;109: 1726-1736.
     
  24. Yannuzzi LA, Wong DW, Sforzolini BS, et al. Polypoidal choroidal vasculopathy and neovascularized age-related macular degeneration. Arch Ophthalmol 1999;117:1503–1510.
     
  25. Yannuzzi LA, Ciardella A, Spaide RF. The expanding clinical spectrum of idiopathic polypoidal choroidal vasculopathy. Arch Ophthalmol 1997;115:478–85.
     
  26. Spaide RF, Yannuzzi LA, Slakter JS. Indocyanine green videoangiography of idiopathic polypoidal choroidal vasculopathy. Retina 1995;15:100–10.
     
  27. Sato T, Iida T, Hagimura N, et al. Correlation of optical coherence tomography with angiography in retinal pigment epithelial detachment associated with age-related macular degeneration. Retina 2004;24:910-4.
     
  28. Coscas F, Coscas G, Souied E, et al. Optical coherence tomography identification of occult choroidal neovascularization in age-related macular degeneration. Am J Ophthalmol 2007;144:592–599.
     
  29. Coscas G. Optical coherence tomography in age-related macular degeneration. (ed) Springer Medizin Verlag Heildelberg 2009: pp 201-203.
     
  30. Brancato R, Introini U, Pierro L, et al. Optical coherence tomography in Retinal angiomatous proliferation Eur J Ophthalmol 2002; 12:467-472.
     
  31. Coscas G. Optical coherence tomography in age-related macular degeneration. (ed) Springer Medizin Verlag Heildelberg 2009: pp 277-279.
     
  32. Iijima H, Imai M, Gohdo T, Tsukahara S. Optical coherence tomography of idiopathic polypoidal choroidal vasculopathy. Am J Ophthalmol 1999;127:301–305.
     
  33. Otsuji T, Takahashi K, Fukushima I, Uyama M. Optical coherence tomographic findings of idiopathic polypoidal choroidal vasculopathy. Ophthalmic Surg Lasers 2000;31:210-214.
     
  34. Klein ML, Obertynski H, Patz A, et al. Follow-up study of detachment of the retinal pigment epithelium. Br J Ophthalmol 1980;64:412-6.
     
  35. Pauleikhoff D, Loeffert D, Spital G, et al. Pigment epithelial detachment in the elderly. Clinical differentiation, natural course and pathogenetic implications. Graefes Arch Clin Exp Ophthalmol 2002;240:533–538.
     
  36. Bird AC, Marshall J. Retinal pigment epithelial detachments in the elderly. Trans Ophthalmol Soc UK 1986;105:674-82.
     
  37. Gass JD. Pathogenesis of tears of the retinal pigment epithelium. Br J Ophthalmol 1984;68:513-9.
     
  38. Lafaut BA, Aisenbrey S, Vanden Broecke C, et al. Clinicopathological correlation of retinal pigment epithelial tears in exudative age related macular degeneration: pretear, tear, and scarred tear. Br J Ophthalmol 2001;85:454-60.
     
  39. Zayit-Soudry S, Moroz I, Loewenstein A. Retinal pigment epithelial detachment. Surv Ophthalmol 2007; 52(3):227-243.
     
  40. Hoskin A, Bird AC, Sehmi K. Tears of detached retinal pigment epithelium. Br J Ophthalmol 1981;65:417-22.
     
  41. Giovannini A, Amato G, Mariotti C, et al. Optical coherence tomography in the assessment of retinal pigment epithelial tear. Retina 2000;20:37–40.
     
  42. Coscas G, Koenig F, Soubrane G. The pretear characteristics of pigment epithelial detachments. A study of 40 eyes. Arch Ophthalmol 1990;108:1687–169.
     
  43. Chang LK, Sarraf D. Tears of the retinal pigment epithelium: an old problem in a new era. Retina 2007;27(5):523–534.
     
  44. Nagiel A, Freund KB, Spaide RF, Munch IC, Larsen M, Sarraf D. Mechanism of retinal pigment epithelium tear formation following intravitreal anti-vascular endothelial growth factor therapy revealed by spectral-domain optical coherence tomography. Am J Ophthalmol 2013;156(5):981–988.e982.
     
  45. Doguizi S, Ozdek S. Pigment epithelial tears associated with anti-VEGF therapy: incidence, long-term visual outcome, and relationship with pigment epithelial detachment in age-related macular degeneration. Retina 2014;34(6):1156–1162.
     
  46. Clemens CR, Bastian N, Alten F, Milojcic C, Heiduschka P, Eter N. Prediction of retinal pigment epithelial tear in serous vascularized pigment epithelium detachment. Acta Ophthalmol 2014;92(1):e50–56.
     
  47. Gass JD. Retinal pigment epithelial rip during krypton red laser photocoagulation. Am J Ophthalmol 1984;98:700-6.
     
  48. Pece A, Introini U, Bottoni F, et al. Acute retinal pigment epithelial tear after photodynamic therapy. Retina 2001;21:661-5.
     
  49. Gelisken F, Inhoffen W, Partsch M, et al. Retinal pigment epithelial tear after photodynamic therapy for choroidal neovascularization. Am J Ophthalmol 2001;131:518-20.
     
  50. Michels S, Aue A, Simader C, et al. Retinal pigment epithelium tears following verteporfin therapy combined with intravitreal triamcinolone. Am J Ophthalmol 2006;141:396–398.
     
  51. Dhalla MS, Blinder KJ, Tewari A, et al. Retinal pigment epithelial tear following intravitreal pegaptanib sodium. Am J Ophthalmol 2006; 141(4): 752–754.
     
  52. Nicolo M, Ghiglione D, Calabria G. Retinal pigment epithelial tear following intravitreal injection of bevacizumab (Avastin). Eur J Ophthalmol 2006;17:770–773.
     
  53. Weinberger AW, Thiel M, Mohammadi B, et al. Retinal pigment epithelium tears after intravitreal bevacizumab in pigment epithelium detachment. Am J Ophthalmol 2007; 144(2): 294–296.
     
  54. Lee GKY, Lai TYY, Chan WM, Lam DSC. Retinal pigment epithelial tear following intravitreal ranibizumab injections for neovascular age-related macular degeneration. Graefes Arch Clin Exp Ophthalmol 2007; 245(8): 1225-7.
     
  55. Carvounis PE, Kopel AC, Benz MS. Retinal pigment epithelium tears following ranibizumab for exudative agerelated macular degeneration. Am J Ophthalmol 2007;143: 504–505.
     
  56. Bakri SJ, Kitzmann AS. Retinal pigment epithelial tear after intravitreal ranibizumab. Am J Ophthalmol 2007;143:505–507.
     
  57. Chang LK, Sarraf D. Tears of the retinal pigment epithelium. An old problem in a new era. Retina 2007; 27: 523-34.
     
  58. Introini U, Torres Gimeno A, Scotti F, Setaccioli M, Giatsidis S, Bandello F. Vascularized retinal pigment epithelial detachment in age-related macular degeneration: treatment and RPE tear incidence. Graefes Arch Clin Exp Ophthalmol 2012 Sep;250(9):1283-92.
     
  59. Meyer CH, Toth CA. Retinal pigment epithelial tear with vitreomacular traction: a novel pathogenic feature. Graefes Arch Clin Exp Ophthalmol 2001;239:325–333.
     
  60. Bressler NM, Finklestein D, Sunness JS, et al. Retinal pigment epithelial tears through the fovea with preservation of good visual acuity. Arch Ophthalmol 1990;108:1694-7.
     
  61. Chang B, Yannuzzi LA, Ladas ID, et al. Choroidal neovascularization in second eyes of patients with unilateral exudative age-related macular degeneration. Ophthalmology 1995;102:1380-6.
     
  62. Yannuzzi LA. [Retinal pigment epithelial detachment]. J Fr Ophtalmol 1989;12:761-74.
     
  63. Brancato R, Introini U, Bolognesi G, et al. ICGA-guided laser photocoagulation of occult choroidal neovascularization in age-related macular degeneration. Retina 2000;20:134-42.
     
  64. Goldstein M, Heilweil G, Barak A, Loewenstein A. Retinal pigment epithelial tear following photodynamic therapy for choroidal neovascularization secondary to AMD. Eye 2005; 19(12):1315–1324.
     
  65. Axer-Siegel R, Ehrlich R, Avisar I, et al. Combined photodynamic therapy and intravitreal triamcinolone acetonide injection for neovascular age-related macular degeneration with pigment epithelium detachment. Ophthalmic Surg Lasers Imaging 2006; 37(6):455–461.
     
  66. Lai TY, Chan WM, Liu DT, Lam DS. Ranibizumab for retinal angiomatous proliferation in neovascular age-related macular degeneration. Graefes Arch Clin Exp Ophthalmol 2007; 245(12):1877–1880.
     
  67. Meyerle CB, Freund KB, Iturralde D, et al. Intravitreal bevacizumab (Avastin) for retinal angiomatous proliferation. Retina 2007; 27(4):451–457.
     
  68. Kook D, Wolf A, Neubauer AS, et al. Retinal pigment epithelial tears after intravitreal injection of bevacizumab for AMD. Frequency and progress. Ophthalmologe 2008; 105(2):158–164.
     
  69. Lommatzsch A, Heimes B, Gutfleisch M, et al. Serous pigment Epithelial detachment in age-related macular degeneration: comparison of different treatments. Eye 2009; 23(12):2163-8.
     
  70. Waldstein SM, Simader C, Staurenghi G, et al. Morphology and Visual Acuity in Aflibercept and Ranibizumab Therapy for Neovascular Age-Related Macular Degeneration in the VIEW Trials. Ophthalmology 2016;123(7):1521-9.
     
  71. He L, Silva RA, Moshfeghi DM, et al. Aflibercept for the treatment of retinal pigment epithelial detachments. Retina 2016;36:492–498.
     
  72. de Massougnes S, Dirani A, Ambresin A, Decugis D, Marchionno L, Mantel I. Pigment epithelial detachment response to aflibercept in neovascular age-related macular degeneration refractory to ranibizumab: Time Course and Drug Effects. Retina 2016;36(5):881-8.
     
  73. Kumar N, Marsiglia M, Mrejen S, et al. Visual and anatomical outcomes of intravitreal aflibercept in eyes with persistent subfoveal fluid despite previous treatments with ranibizumab in patients with neovascular age-related macular degeneration. Retina 2013;33:1605–1612.
     
  74. Patel KH, Chow CC, Rathod R, et al. Rapid response of retinal pigment epithelial detachments to intravitreal aflibercept in neovascular age-related macular degeneration refractory to bevacizumab and ranibizumab. Eye (Lond) 2013;27:663–667.
     
  75. Dirani A, Ambresin A, Marchionno L, Decugis D, Mantel I. Factors Influencing the Treatment Response of Pigment Epithelium Detachment in Age-Related Macular Degeneration. Am J Ophthalmol 2015;160(4):732-8.
     
  76. Cho HJ, Kim KM, Kim HS, Lee DW, Kim CG, Kim JW. Response of Pigment Epithelial Detachment to Anti-Vascular Endothelial Growth Factor Treatment in Age-Related Macular Degeneration. Am J Ophthalmol 2016;166:112-9.
     
  77. Sarraf D, London NJ, Khurana RN, et al. Ranibizumab Treatment for Pigment Epithelial Detachment Secondary to Neovascular Age-Related Macular Degeneration: Post Hoc Analysis of the HARBOR Study. Ophthalmology 2016;123(10):2213-24.
    ​​​​​​​
  78. Chen X, Al-Sheikh M, Chan CK, et al. Type 1 versus type 3 neovascularization in pigment epithelial detachments associated with age-related macular degeneration after anti-vascular endothelial growth factor therapy: A Prospective Study. Retina 2016;36 Suppl 1:S50-S64.

Preventive AMD Treatments

Preventive AMD Treatments

English

Authors:

Maria João Veludo, MD
Lisbon Hospital Center - Lisbon, Portugal

 

Filomena Costa e Silva, MD
Prof. Dr. Fernando Fonseca Hospital, Amadora, Lisbon, Portugal

 

Susana Teixeira, MD
Prof. Dr. Fernando Fonseca Hospital, Amadora, Lisbon, Portugal

 

Updated/reviewed by the authors, July 2017.

Introduction

A better understanding of the pathophysiological processes occurring in “retinal aging” and age-related macular degeneration (AMD) has been achieved in recent years, leading to the emergence of new treatments and consequent long-term improvements in the patients quality of life.

AMD is one of the leading causes of severe, irreversible vision impairment in developed countries, in individuals over 50 years of age.

Approximately 1.75 million people over 40 in the United States suffer from neovascular AMD or geographic atrophy; 7.3 million patients display large drusen (≥125 microns) in one or both eyes(1).

In the United States, AMD causes approximately 46% of severe visual loss cases (visual acuity of 20/200 or worse) in patients over 40(2).

Although, an estimated 80% of AMD patients display the non-neovascular form of this disease, the neovascular form is responsible for almost 90% of cases of severe visual loss (visual acuity of 20/200 or worse) caused by AMD(3).

Data from three population-based studies – the Beaver Dam Eye Study, the Rotterdam Study and the Blue Mountains Eye Study – have led to an estimated prevalence of advanced AMD of 0-2% in patients aged 55–64, increasing to 13% in patients over 85(4).

Since there is no significant cure for AMD, prevention may be the first and logic approach to reduce vision loss, justifying an intensive search for some kinds of intervention able to prevent the onset of AMD or to delay its progression to more advanced and severe forms.

Age is the main risk factor for AMD; all population-based studies confirm that the prevalence of AMD increases with age in white individuals(5,6,7).

Belonging to the female gender may also constitute a risk factor in individuals aged over 75 years(4).

Several studies also demonstrated that effective control of modifiable risk factors, such as smoking, hypertension and body-mass index, could reduce the risk of developing AMD by half(8).

Since the early 90’s, when “large population studies” appeared, several hypotheses have been formulated around the idea that nutritional supplements, such as antioxidants, vitamins and/or minerals may be able to reduce the risk of AMD development.  

 

AREDS 1

Design implications and study categories

AREDS (Age-related Eye Disease Study) was a prospective, multicentric, randomised clinical trial conducted between 1992 and 2006, mainly sponsored by the National Eye Institute (NEI) of the National Institutes of Health (NIH).

This study was designed to evaluate the clinical aspects, natural course and risk factors associated with age-related cataract and AMD, as well as the effects of antioxidant vitamins and minerals on these two ocular conditions.

Eligible patients were aged 55-80 by occasion of enrolment and required to be free of any illness or condition that would make long-term follow-up or compliance with study medications unlikely or difficult.

Participants were placed in one of several AMD categories according to fundus photographs graded by a central reading centre, best corrected visual acuity and ophthalmic examination(9):

AREDS category 1 – (No AMD) – this was the AREDS control group, consisting of patients with no or a few small drusen (<63 microns in diameter).

AREDS category 2 – (Early AMD) – characterised by a combination of multiple small drusen, a few intermediate drusen (63 to 124 microns in diameter) or retinal pigment epithelium (RPE) abnormalities.

AREDS category 3 – (Intermediate AMD) – characterised by extensive intermediate drusen, at least one large drusen (>125 microns in diameter) or geographic atrophy not involving the centre of the fovea.

AREDS category 4 – (Advanced/Late AMD) – characterised by one or more of the following (in the absence of other causes), in one eye:

  • Geographic atrophy of the RPE and choriocapillaris, including the centre of the fovea;
  • Neovascular maculopathies, such as the following: choroidal neovascularization;
  • Serous and/or haemorrhagic detachment of the sensory retina or the RPE;
  • Hard exudates in the retina;
  • Subretinal and sub-RPE fibrovascular proliferation;
  • Disciform scar.

 

Risk factors and categories

AREDS Report no. 18 described a simplified clinical scale that defines risk categories for the development of advanced AMD(10).

The grading system described assigns one risk factor to each eye for the presence of one or more large drusen (125 microns) and one risk factor for the presence of any pigment abnormality.

If no large drusen are present, the presence of intermediate drusen in both eyes is counted as one risk factor.
Advanced AMD in one eye is counted as two risk factors; if this is observed together with large drusen and hypo/hyperpigmentary changes in the RPE, four risk factors are considered, which corresponds to the highest risk level for patients with AMD.

Risk factors are added for both eyes, leading to a 5-stage risk of developing advanced AMD in at least one eye increases as follows(10):

  • Stage 0 (zero factors) – 0.5% in five years;
  • Stage 1 (one factor) – 3% in five years;
  • Stage 2 (two factors) – 12% in five years;
  • Stage 3 (three factors) – 25% in five years;
  • Stage 4 (four factors) – 50% in five years.

 

Results

AREDS results show an overall beneficial effect for high doses of antioxidant vitamin (vitamins C, E and beta-carotene) and zinc supplements in reducing the progression of intermediate or advanced AMD to advanced AMD in the fellow eye, corresponding to 25%.

Therefore, a formulation has been proposed(11):

 

AREDS 1 formulation

  • Antioxidant vitamins – 500 mg of vitamin C
  • 400 IU of vitamin E
  • 15 mg of beta-carotene
  • 80 mg of zinc oxide and 2 mg of cupric oxide

This formulation has been shown to reduce the risk of developing advanced AMD and the associated visual loss by as much as 25%, over 5 years, in individuals with moderate to high risk of AMD (AREDS categories 3 and 4).

These findings were accompanied by a 19% reduction in the risk of moderate vision loss (loss of three or more lines on the visual acuity chart), at 5 years(11).

However, this formulation is not recommended for smokers, since beta-carotene has been shown to increase the risk of lung cancer(12,13).

 

AREDS 2

The Age-related Eye Disease Study 2 (AREDS2), initiated in 2006 enrolled 4000 patients with non-neovascular AMD consisting of large drusen in both eyes or advanced AMD in one eye and large drusen in the fellow eye (AREDS categories 3 and 4).

The aim of this study was to evaluate the effect of dietary supplements – xanthophylls (10 mg of lutein and 2 mg of zeaxanthin) and/or long-chain omega-3 polyunsaturated fatty acids (1 g of docosahexaenoic acid [DHA] and eicosapentaenoic acid [EPA]) – on the progression to advanced AMD.

An additional goal of the study was to assess whether forms of the AREDS nutritional supplement with reduced zinc and/or no beta-carotene works as well as the original supplement in reducing the risk of progression to advanced AMD.

 

AREDS 2 formulation

  • Antioxidant vitamins – 500 mg of vitamin C
  • 400 IU of vitamin E
  • 15 mg of beta-carotene
  • 80 mg of zinc oxide and
  • 2 mg of cupric oxide
  • Macular pigments – xanthophylls
  • 10 mg of lutein
  • 2 mg of zeaxanthin
  • 1 g of omega-3 fatty acids (DHA +EPA)

This study also investigated whether the current AREDS formulation might be modified by eliminating beta-carotene. As previously mentioned, beta-carotene, which is not present within the eye, constitutes a problem for smokers, due to the high incidence of lung cancer in this patient group.

A secondary randomisation in AREDS 2 evaluated the possibility of eliminating and/or lowering the amount of zinc in the AREDS formulation, since zinc levels in the current formulation are considered high and available evidence suggests that the body is only able to absorb 25 mg of zinc per day.

Results found that addition of lutein + zeaxanthin, DHA + EPA, or both to the AREDS formulation did not further reduce risk of progression to advanced AMD. However, because of potential increased incidence of lung cancer in former smokers, lutein + zeaxanthin could be an appropriate carotenoid substitute in the AREDS formulation(14).

 

Ocular Micronutrition

Carotenoids

Macular pigments: lutein and zeaxanthin

Over 600 carotenoids have been identified to the present date; however, only about 50 are found in normal human diets.

Vegetables are the only source of carotenoids. Macular pigments (MP) consist of two natural xanthophylls from the carotenoid family: lutein (L) and zeaxanthin (Z), as well as the products of their transformation in the body, notably meso-zeaxanthin(15).

These pigments are not synthesised by the human body and have been investigated for their ability to promote visual health.

Macular pigments are found in photoreceptor axons in the pigment epithelium and in the external segment.

The latter contains very high levels of polyunsaturated fatty acids, at a high risk of oxidation(16-18). Macular concentrations of L and Z decrease with age, which exacerbates the harmful effects of blue light on photoreceptors.

Their protective mechanism is unknown; however, two mechanisms have been proposed: macular pigments may act as an optical filter, due to their ability to absorb blue light, and they are strong antioxidants, neutralising free radicals generated by light(19-20).

Several observational studies have demonstrated a correlation between plasma L and Z levels, macular pigment density and a lower risk of AMD.

Increased intake of L and Z supplements resulted in increased plasma levels, which were positively and significantly associated with the optical density of macular pigment(21).

Furthermore, various well-conducted population-based longitudinal studies have suggested that high dietary antioxidant levels, specifically L and Z, may have protective and beneficial effects, delaying progression to advanced AMD.

The following are some examples of observational studies investigating the relationship between dietary and/or serum antioxidant levels and the risk of AMD.

EDCCS (Eye Disease Case-Control Study): this study, published in 1993, demonstrated a high significant inverse relationship (p= 0.001) between the prevalence of AMD and serum L and Z concentrations: the risk of neovascular lesions was 70% lower in subjects with the highest serum L and Z concentrations, compared to those with the lowest levels (odds ratio (OR): 0.4; 95% confidence interval (CI): 0.2-0.6; p= 0.0001)(22).

In 1994, the same group concluded that the risk of developing the most severe form of macular degeneration was 43% lower in individuals who consume large amounts of fruits and vegetables rich in L and Z (6 mg/day) (OR: 0.57; 95% CI: 0.35-0.92; p= 0.02)(23).

Another observational study conducted in 2002 showed that average L and Z levels were 32% lower in AMD eyes than in control eyes, in elderly subjects, provided the latter were not consuming high doses of L supplements. This study demonstrated that average levels of macular pigments in patients who had begun to regularly consume supplements containing high doses of L (> 4 mg/day) after the initial diagnosis of AMD were within the normal range and significantly higher than in AMD patients not consuming this supplementation(24).

In 2006, the POLA study (Pathologies Oculaires Liées a l’Age) found that high plasma L and total L and Z concentrations are associated with a reduced risk of age-related maculopathy (ARM)(25).

CAREDS (Carotenoids in Age-related Eye Disease Study): in this observational study, conducted in 2006, no statistically significant differences in intermediate AMD prevalence were found between subjects with high and low dietary intakes of L plus Z. However, analysis of a sub-group of women under 75 with stable L plus Z intakes revealed a reduced risk of this subtype of AMD in association with a high dietary intake of those antioxidants. A diet rich in L plus Z may protect against intermediate AMD(26).

In 2007, AREDS (Age-related Eye Disease Study report 22), a case-control study with 4519 participants, concluded that a high dietary intake of L and Z was independently associated with a decreased likelihood of neovascular AMD, geographic atrophy, and large and extensive drusen(27).

Up to the present date, seven important interventional studies have investigated the role of L supplementation in AMD patients(28-34).

Of the aforementioned studies, two assume major importance.

LAST (Lutein Antioxidant Supplementation Trial) was the first study to show that L supplementation improved visual function in AMD patients. Furthermore, it reinforced the notion that AMD is a nutrition-responsive disease. The results of this trial were confirmed in 2004.

In this trial, 90 AMD patients received either a daily supplement consisting of 10 mg of L, a supplement consisting of 10 mg of L and a mixed antioxidant formula (containing vitamin A, beta-carotene, vitamin C, vitamin E, vitamin B complex, copper, zinc, manganese, magnesium, selenium and other minerals), or placebo, for 12 months.

Results showed that patients receiving the L supplement displayed significant improvements in several objective visual function measures (contrast sensitivity or visual acuity) when compared to the placebo group. Slightly better results were observed in subjects consuming the combined supplement(33).

CARMA (Carotenoids in Age-related Maculopathy Study) was an important European intervention study, published in 2008(34).

This randomised, double-blind clinical trial of antioxidant supplementation versus placebo enrolled 433 patients from two centres in Ireland, with signs of early AMD of sufficient severity in at least one eye, or any level of AMD in one eye and late AMD (neovascular AMD or central geographic atrophy) in the fellow eye.

The aim of the CARMA Study was to investigate whether administration of 12 mg of L and 2 mg of Z, in combination with antioxidants (120 mg of vitamin C, 15mg of vitamin E, 20mg of zinc and 0.4 mg of copper), had a beneficial effect on visual function and/or was able to delay progression of early to late disease stages.

The primary outcome was improved or distance visual acuity was preserved at 12 months.
Although no beneficial effects were demonstrated in the primary outcome measure at the stated end point (12 months), secondary outcomes favoured the supplemented group(34).

In 2013, AREDS 2 presented an exploratory analysis of the 1114 participants who received lutein/zeaxanthin added to an AREDS formulation without beta carotene versus the 1117 participants who received only beta carotene in the AREDS formulation.

This revealed that those receiving lutein/zeaxanthin without beta carotene supplement had a 18% reduction in the risk for late AMD and 22% reduction for neovascular AMD, as well as 6% reduction in the risk for geographic AMD (which yielded a P value of 0.67).

A pre-specified comparison of those receiving lutein/zeaxanthin of those who did receiving lutein/zeaxanthin revealed 10% reduction in the risk for progression to late AMD.

A second pre-specified analysis showed that those in the lowest quintile of dietary intake of lutein/zeaxanthin who received lutein/zeaxanthin along with the original AREDS formulation had a 26% reduced risk for progression to late AMD relative to participants receiving the original AREDS formulation, without lutein/zeaxanthin(35).

 

Key points

The aforementioned findings are consistent with the hypothesis that low L and Z levels in the human macula may represent a pathogenic risk factor for AMD development.

Based on the analyses from AREDS2 we recommend the patient with intermediate AMD (bilateral large drusen) or late AMD in one eye be given the AREDS2 formulation.

They also suggest that supplementation may prevent damage from oxidation and harmful wavelength of light and contribute to maintaining eye health.

 

Antioxidants

Vitamin and mineral supplements

Antioxidants are recommended for AMD due to oxidative stress on photoreceptors in the retina and the fact that cumulative damage caused by blue light enhances free radical production.

It has been proposed that antioxidants may prevent cellular damage in the retina by reacting with free radicals(36).

The substances that possess antioxidant activity are vitamins C and E, beta-carotene and some minerals, such as zinc, copper, selenium and manganese.

Some studies indicate that diets rich in antioxidants may protect against the appearance of signs of early AMD; in common perception, a diet rich in antioxidants is capable of protecting against AMD.

Randomised control trials and observational studies have been conducted in well-nourished Western populations; however, the role of dietary antioxidants in the primary prevention of AMD remains unclear.

In the 90’s, several studies reported a protective effect against AMD development for high intakes of antioxidant vitamins and minerals.

In 1993, the EDCCS (Eye Disease Case-Control Study) performed a comparison between 421 patients with neovascular AMD and 615 control subjects. The results revealed that high plasma levels of antioxidants (vitamins A, C, E, selenium and carotenoids) are associated with a lower risk of developing neovascular AMD. Additional carotenoid intakes, particularly of those present in the retina, are associated with a lower risk of developing AMD(22).

In 1994, the authors of the Baltimore Longitudinal Study on Aging, a study involving 976 patients, reported a protective effect against AMD for high plasma concentrations of vitamin E. The authors also found an antioxidant combination of vitamin C, vitamin E and beta-carotene to be protective(37).

In 1998, the Beaver Dam Eye Study, in which a cohort of 1,700 patients was subject to a 5-year follow-up eye examination, showed that a high intake of carotenoids and vitamin E is associated with a lower risk of developing large drusen. High dietary zinc intakes would be associated with a lower number of RPE anomalies(38).

In 2001, AREDS report no. 8, a large multicentric, randomised clinical trial, revealed that the risk of progression to advanced AMD was reduced by 28% in patients with intermediate AMD treated with high doses of antioxidant supplements (vitamins C and E, zinc and β-carotene), when compared to the placebo group (OR: 0.72; 99% CI: 0.52-0.98). This study did not specifically investigate whether antioxidant supplements were effective in the primary prevention of early AMD in individuals without signs of this condition(11).

In 2004, AREDS report no. 13 evaluated mortality rates in patients with ocular disorders taking high doses of antioxidants or zinc. Results showed that mortality was lower in patients taking zinc (alone or with antioxidants) (12% reduction), when compared to those not taking this mineral (RR: 0.73; 95% CI: 0.61-0.89)(39).

In 2005, the Rotterdam Study, a population-based study involving 4170 participants, showed that an above-average intake of the 4 AREDS trial nutrients protected against AMD development or early AMD, as indicated by large drusen, and was associated with a 35% reduction in the risk of AMD (HR: 0.65; 95% CI: 0.46-0.92)(40).

In 2007, Chong and colleagues undertook a systematic review and meta-analysis of nine prospective cohort studies and three randomised clinical trials. The results from the first studies indicated that vitamin A, vitamin C, vitamin E, zinc, L, Z, alpha-carotene, beta-carotene, beta-cryptoxanthin and lycopene have little or no effect in the primary prevention of early AMD. The three randomised clinical trials failed to show that antioxidant supplements prevented early AMD(32).

In 2008, a systematic review and meta-analysis undertaken with the objective of examining available evidence as to whether antioxidant vitamin or mineral supplements are able to prevent AMD development or delay its progression was published online. No evidence was found that antioxidant (vitamin E or beta-carotene) supplements are able to prevent AMD (RR 1.03; 95% CI; 0.74-1.43). Some evidence was found that antioxidant (beta-carotene, vitamin C and E) and zinc supplements were able to delay progression to advanced AMD and prevent loss of visual acuity in individuals displaying signs of the disease (adjusted OR = 0.68; 95% CI: 0.53-0.57, and 0.77; 95% CI: 0.62-0.96, respectively)(41).

In AREDS 2, smokers were not randomized to beta-carotene, and there was a doubling of the risk of developing lung cancer in subjects randomized to beta-carotene, compared with those randomized to lutein/zeaxanthin. More than 90% of these subjects affected with lung cancer were former smokers. A large proportion of the persons affected with AMD were former or current smokers. The totality of evidence from these analysis support the substitution of lutein/zeaxanthin for beta-carotene(35).

 

Key points

According to current evidence, antioxidant vitamin supplements are unable to prevent AMD.

High-dose antioxidant supplementation may increase the risk of lung cancer in smokers (beta-carotene), heart failure in individuals with vascular disease or diabetes (vitamin E) and hospitalisation in patients with genitourinary conditions.

The totality of evidence from all analysis supports that lutein/zeaxanthin together appeared to be a safe alternative to beta-carotene.

 

Dietary fatty acids

Docosahexaenoic acid (DHA) and eicosapentaenoic acid (EPA)

The role of fatty acids in AMD was initially investigated because of the hypothesis that AMD and cardiovascular disease may share a similar pathogenesis and fat intake has been associated with atherosclerosis and cardiovascular disease.

Fatty acids may be divided into three types:

  • Saturated fat from dairy products and meat.
  • Monounsaturated fatty acids (MUFA) from olive oil.
  • Polyunsaturated fatty acids (PUFA), especially from fish and seafood.

Omega-3 fatty acids, also known as Long-Chain Polyunsaturated Fatty Acids (LCPUFAs), are essential to human health.

Omega-3 fatty acids include alpha-linolenic acid (a short-chain omega-3 fatty acid) and long-chain omega-3 fatty acids docosahexaenoic acid (DHA) and eicosapentaenoic acid (EPA).

Omega-3 fatty acids, especially DHA, have morphological, functional and protective roles in the retina:

  1. Morphological role – DHA is the main PUFA found within the outer segments of rods and has anti-apoptotic, anti-inflammatory and antiangiogenic functions.
  2. Functional role – DHA provides an adequate environment for conformational changes in rhodopsin.
  3. Protective role – DHA protects against aging of the retina and may reduce lipofuscin accumulation in the RPE and lipid deposits in Bruch’s membrane.

Several epidemiological studies have evaluated the relationship between total and specific fat intake and the risk of advanced AMD.

Results confirm that higher intakes of vegetable and animal fat are associated with a greater risk of advanced AMD. In 2008, a systematic review and meta-analysis was undertaken with the objective of evidencing the role of dietary omega-3 fatty acid and fish intakes in the primary prevention of AMD.

This review included three randomised, controlled, prospective cohort trials(42-44), three case-control studies(45-47) and three cross-sectional studies(45,48,49).

The results of these nine studies demonstrated that high dietary omega-3 fatty acid intakes were associated with a 38% reduction in the risk of late AMD (OR: 0.62; 95 % CI: 0.48-0.82).

Eating fish at least twice a week was associated with a reduced risk of both early AMD (OR: 0.76; 95% CI: 0.64-0.90) and late AMD (OR: 0.67; 95% CI: 0.53-0.85).

Several other relevant studies evidence this fact:

A prospective study conducted by Cho et al. in 2001 evidenced a positive association between total fat intake and incidence of AMD. A diet rich in fat increases the risk of advanced AMD. Nevertheless, eating fish 4 or more times a week (fish is a major source of DHA) decreases the relative risk of AMD by 35%(42).

A case-control study conducted by SanGiovanni et al. concluded that higher omega-3 and fish intakes are associated with a decreased risk of neovascular AMD(50).

The objective of AREDS report n. 20 was to evaluate the association between lipid intake and AMD severity at baseline.

The results of this study showed that total dietary intake of total long-chain omega-3 polyunsaturated fatty acids (LCPUFA) was inversely associated with neovascular AMD (OR: 0.61; 95 % CI: 0.41-0.90), the same occurring for docosahexaenoic acid, a retinal omega-3 LCPUFA (OR: 0.54; 95% CI: 0.36-0.80), when the highest and lowest intake quintiles were compared, after adjustment for total energy intake and covariates.

Higher fish intakes, both total and broiled/baked, were also inversely associated with neovascular AMD (OR: 0.61; 95% CI: 0.37-1.00, and OR: 0.65; 95% CI: 0.45-0.93, respectively).

Dietary intake of arachidonic acid was directly associated with the prevalence of neovascular AMD (OR: 1.54; 95% CI: 1.04-2.29).

No statistically significant relationships were found for other lipids or groups(47).

In AREDS report n. 23, reduced likelihood of progression from bilateral drusen to central geographic atrophy was observed in individuals reporting the highest EPA intakes (OR: 0.44; 95% CI: 0.23-0.87) and EPA + DHA intakes (OR: 0.45; 95% CI: 0.23-0.9). DHA levels were associated with central geographic atrophy in age, gender and calorie adjusted models (OR: 0.51; 95% CI: 0.36-1.00). However, this statistical relationship was not observed in multivariable models. This study suggested that dietary intake, of long-chain omega-3 polyunsaturated fatty acids, is associated with a decreased risk of progression from bilateral drusen to central geographic atrophy(51).

European study Nat-2, performed at the University of Créteil, a double-blind, randomised, parallel, comparative study, compared oral DHA supplementation with placebo in the prevention of exudative AMD in 298 patients with any type of drusen in the study eye and wet AMD in the fellow eye. Nat-2 supplementation consisted of 10 mg of L, 2 mg of Z, 1 mg of omega-3 (DHA plus EPA), 500 mg of vitamin C, 400 IU of vitamin E, 25 mg of zinc and 2 mg of copper. Patients took no other supplements and were followed for three years (2004-2008). The first study results included in NAT-2 report no. 1, revealed high HDL and low PUFA levels in exudative AMD patients. These findings confirmed the benefits of DHA supplementation in these AMD patients(52).

Two important prospective observational studies clearly reveal that fish consumption and omega-3 fatty acid intake decrease the risk of AMD: The Blue Mountains Eye Study and the Melbourne Collaborative Cohort Study.

Blue Mountains Eye Study: The objective of this longitudinal study was to investigate the association between baseline dietary fatty acids and 10-year incidence of AMD in an elderly Australian cohort. Nutrient intakes were estimated through a semi-quantitative food frequency questionnaire.

The risk of incidence of early AMD was lower in individuals consuming 1 to 2 servings of nuts per week (RR: 0.65; 95% CI: 0.47-0.91). These results were similar to those obtained for dietary consumption of long-chain omega 3 PUFAs, which also show a lower risk of incidence of early AMD in participants eating 1 serving of fish per week (RR: 0.69; 95% CI: 0.49-0.98).
Participants consuming below-average amounts of linoleic acid contributed the most to this association (RR: 0.57; 95 % CI: 0.36-0.89).

Nut consumption was associated with a lower risk of pigmentary abnormalities in non-smokers, individuals with below-average total to high-density lipoprotein serum cholesterol ratios, and individuals with above-average beta-carotene intakes(53).

Melbourne Collaborative Cohort Study: the aim of this study, carried out in 1990-1994, was to investigate the relationship between past dietary fat intake and the ticipants aged 58-69.

The corresponding results showed that a higher dietary intake of trans unsaturated fats was associated with an increased prevalence of late AMD. Comparing the highest and lowest trans fat intake quartiles, the OR for late AMD was 1.76 (95% CI: 0.92-3.37; p = 0.02), whereas a higher intake of omega-3 fatty acids was inversely associated with early AMD (OR for highest quartile versus lowest quartile: 0.85; 95% CI: 0.71-1.02; p = 0.03). The prevalence of late AMD was lower for olive oil intakes equal to or higher than 100 mL/week versus less than 1 mL/week (OR: 0.48; 95% CI: 0.22-1.04; P = 0.03). No significant associations were found between fish, total fat, butter and margarine intakes and AMD(54).

In 2009, the SanGiovanni AREDS Group investigated the relationship between dietary omega-3 LCPUFA intake and progression to advanced AMD in 1837 AREDS participants with a moderate risk for developing sight-threatening AMD (1211 participants in category 3a and 626 participants in category 4a).

It was observed that participants reporting the highest baseline omega-3 LCPUFA intakes were approximately 30% less likely to develop advanced AMD by the end of the 12-year follow-up period than those reporting the lowest omega-3 LC-PUFA intakes.

Results for central geographic atrophy and neovascular AMD were similar; the corresponding multivariate OR were 0.65 (95% CI: 0.45-0.92; p ≤ 0.02) and 0.68 (95% CI: 0.49-0.94; p ≤ 0.02)(55).

In 2013 based on the analysis from AREDS2, omega-3 LCPUFA appear to have no role and did not reduce the risk of advanced AMD, even when evaluating participants with the lowest level of dietary intake of omega-3 LCPUFA(14,35).

 

Key point

It is important to note that the observational data overwhelmingly suggest that eating fish has a favourable effect.

 

Response to AREDS Supplements According to Genetic Factors

The impact of nutritional supplements on rate of progression to advanced AMD for patients within specific genotype groups and the need to genotype all patients, taking the AREDS supplements, has been a subject of debate.

The first evidence of a differential treatment effect on progression according to genotype demonstrated that the odds ratio of progression for the combined antioxidant and zinc group versus the placebo group was lower for non-risk CFH (complement factor H) subjects compared with high-risk subjects. More recent publications evaluated similar relationships between treatment and genotype; however, these studies revealed conflicting results.

An initial publication by Awh et al.(56) reported the benefit of zinc in reducing progression to advanced AMD among 995 subjects with no risk alleles for CFH and one or two risk alleles for ARMS2 (age-related maculopathy sensitivity 2). A more recent publication from the same group suggested a differential impact on disease progression according to the number of risk alleles for these SNPs: the detriment posed by a CFH risk allele was exacerbated and the harmful effect of the ARMS2 risk allele was alleviated in subjects receiving supplementation with zinc, both alone or as a component of the AREDS combination supplement. A survival analysis approach using the eye as the Unit of analysis, based on a larger sample of the AREDS population, concluded that the effectiveness of antioxidant and the zinc supplementation appears to differ by genotype(57).

We need more results but genetic testing is important in research and additional studies are needed.

 

Key point

The combination of antioxidants and zinc, as found in both the AREDS and AREDS2 supplements, remains the only proven beneficial formulation regardless of genotype, with no apparent indication for treatment with antioxidants or zinc alone.

 

>> References

References - Preventive AMD Treatments

References - Preventive AMD Treatments

English
  1. Meyers SM, Greene T, Gutman FA. A twin study of age-related macular degeneration. Am J Ophthalmol 1995;120(6):757-766.
     
  2. Congdon N, O’Colmain B, Klaver CC, Klein R, Munoz B, Friedman DS, Kempen J, Taylor HR, Mitchell P. Causes and prevalence of visual impairment among adults in the United States. Arch Ophthalmol 2004;122(4):477-485.
     
  3. Ferris FL, III, Fine SL, Hyman L. Age-related macular degeneration and blindness due to neovascular maculopathy. Arch Ophthalmol 1984;102(11):1640-1642.
     
  4. Smith W, Assink J, Klein R, Mitchell P, Klaver CC, Klein BE, Hofman A, Jensen S, Wang JJ, de Jong PT. Risk factors for age-related macular degeneration: Pooled findings from three continents. Ophthalmology 2001;108(4):697-704.
     
  5. Bird AC, Bressler NM, Bressler SB, Chisholm IH, Coscas G, Davis MD, de Jong PT, Klaver CC, Klein BE, Klein R. An international classification and grading system for age-related maculopathy and age-related macular degeneration. The International ARM Epidemiological Study Group. Surv Ophthalmol 1995;39(5):367-374.
     
  6. Mitchell P, Smith W, Attebo K, Wang JJ. Prevalence of age-related maculopathy in Australia. The Blue Mountains Eye Study. Ophthalmology 1995;102(10):1450-1460.
     
  7. VanNewkirk MR, Nanjan MB, Wang JJ, Mitchell P, Taylor HR, McCarty CA. The prevalence of age-related maculopathy: the visual impairment project. Ophthalmology 2000;107(8):1593-1600.
     
  8. Tomany SC, Wang JJ, van Leeuwen R, Klein R, Mitchell P, Vingerling JR, Klein BE, Smith W, de Jong PT. Risk factors for incident age-related macular degeneration: pooled findings from 3 continents. Ophthalmology 2004;111(7):1280-1287.
     
  9. The Age-Related Eye Disease Study Research Group. The Age-Related Eye Disease Study (AREDS): design implications. AREDS Report No. 1. Control Clin Trials 1999;20(6):573-600.
     
  10. Ferris FL, Davis MD, Clemons TE, Lee LY, Chew EY, Lindblad AS, Milton RC, Bressler SB, Klein R, Age-Related Eye Disease Study (AREDS) Research Group. A Simplified Severity Scale for Age-Related Macular Degeneration: AREDS Report No. 18. Arch Ophthalmol 2005;123(11):1570-1574.
     
  11. Age-Related Eye Disease Study Research Group. A randomized, placebo-controlled, clinical trial of high-dose supplementation with vitamins C and E, beta carotene, and zinc for age-related macular degeneration and vision loss: AREDS Report No. 8. Arch Ophthalmol 2001;119(10):1417-1436. Erratum in: Arch Ophthalmol 2008;126(9):1251.
     
  12. The Alpha-Tocopherol, Beta Carotene Cancer Prevention Study Group. The effect of vitamin E and beta carotene on the incidence of lung cancer and other cancers in male smokers. N Engl J Med 1994;330(15):1029-1035. 
     
  13. Omenn GS, Goodman GE, Thornquist MD, Balmes J, Cullen MR, Glass A, Keogh JP, Meyskens FL, Valanis B, Williams JH, Barnhart S, Hammar S. Effects of a combination of beta carotene and vitamin A on lung cancer and cardiovascular disease. N Engl J Med 1996;334(18):1150-1155.
     
  14. Age-Related Eye Disease Study 2 Research Group. Lutein + Zeaxanthin and Omega -3 Fatty Acids for Aged-Related Macular degeneration. The Age -Related Eye Disease Study 2 (AREDS2) Randomized Clinical Trial. JAMA 2013;309(19):2005-2015.
     
  15. Bone RA, Landrum JT, Tarsis SL. Preliminary identification of the human macular pigment. Vision Res 1985;25(11):1531-1535.
     
  16. Rapp LM, Maple SS, Choi JH. Lutein and zeaxanthin concentrations in rod outer segment membranes from perifoveal and peripheral human retina. Invest Ophthalmol Vis Sci 2000;41(5):1200-1209.
     
  17. Snodderly DM, Handelman GJ, Adler AJ. Distribution of individual macular pigment carotenoids in central retina of macaque and squirrel monkeys. Invest Ophthalmol Vis Sci 1991;32(2):268-279.
     
  18. Wisniewska A, Subczynski WK. Distribution of macular xanthophylls between domains in a model of photoreceptor outer segment membranes. Free Radic Biol Med 2006;41(8):1257-1265.
     
  19. Beatty S, Boulton M, Henson D, Koh HH, Murray IJ. Macular pigment and age related macular degeneration. Br J Ophthalmol 1999;83(7):867-877.
     
  20. Landrum JT, Bone RA, Kilburn MD. The macular pigment: a possible role in protection from age-related macular degeneration. Adv Pharmacol 1997;38:537-556.
     
  21. Johnson EJ, Hammond BR, Yeum KJ, Qin J, Wang XD, Castaneda C, Snodderly DM, Russell RM. Relation among serum and tissue concentrations of lutein and zeaxanthin and macular pigment density. Am J Clin Nutr 2000;71(6):1555-1562.
     
  22. Eye Disease Case-Control Study Group. Antioxidant status and neovascular age-related macular degeneration. Arch Ophthalmol 1993;111(1):104-109. Erratum in: Arch Ophthalmol 1993;111(11):1499. Arch Ophthalmol 1993;111(9):1228. Arch Ophthalmol 1993;111(10):1366.
     
  23. Seddon JM, Ajani UA, Sperduto RD, Hiller R, Blair N, Burton TC, Farber MD, Gragoudas ES, Haller J, Miller DT. Dietary carotenoids, vitamins A, C, and E, and advanced age-related macular degeneration. Eye Disease Case-Control Study Group. JAMA 1994;272(18):1413-1420. Erratum in: JAMA 1995;273(8):622.
     
  24. Bernstein PS, Zhao DY, Wintch SW, Ermakov IV, McClane RW, Gellermann W. Resonance Raman measurement of macular carotenoids in normal subjects and in age-related macular degeneration patients. Ophthalmology 2002;109(10):1780-1787.
     
  25. Delcourt C, Carriere I, Delage M, Barberger-Gateau P, Schalch W. Plasma lutein and zeaxanthin and other carotenoids as modifiable risk factors for age-related maculopathy and cataract: the POLA Study. Invest Ophthalmol Vis Sci 2006;47(6):2329-2335.
     
  26. Moeller SM, Parekh N, Tinker L, Ritenbaugh C, Blodi B, Wallace RB, Mares JA. Associations between intermediate age-related macular degeneration and lutein and zeaxanthin in the Carotenoids in Age-related Eye Disease Study (CAREDS): ancillary study of the Women’s Health Initiative. Arch Ophthalmol 2006;124(8):1151-1162.
     
  27. SanGiovanni JP, Chew EY, Clemons TE, Ferris FL, III, Gensler G, Lindblad AS, Milton RC, Seddon JM, Sperduto RD. The relationship of dietary carotenoid and vitamin A, E, and C intake with age-related macular degeneration in a case-control study: AREDS Report No. 22. Arch Ophthalmol 2007;125(9):1225-1232.
     
  28. Bartlett HE, Eperjesi F. Effect of lutein and antioxidant dietary supplementation on contrast sensitivity in age-related macular disease: a randomized controlled trial. Eur J Clin Nutr 2007;61(9):1121-1127.
     
  29. Massacesi AL, Faletra R, Gerosa F, Staurenghi G, Orzalesi N. The effect of oral supplementation of macular carotenoids (lutein and zeaxanthin) on the prevention of age-related macular degeneration: A 18 months follow up study. ARVO. Fort Lauderdale, USA, April 29-May 4, 2001. Invest Ophthalmol Vis Sci 2001;42(4 Suppl.):S234.
     
  30. Olmedilla B, Granado F, Blanco I, Vaquero M, Cajigal C. Lutein in patients with cataracts and age-related macular degeneration: a long-term supplementation study. J Sci Food Agric 2001;81(9):904-909.
     
  31. Richer S. Part II: ARMD--pilot (case series) environmental intervention data. J Am Optom Assoc 1999;70(1):24-36.
     
  32. Chong EW, Wong TY, Kreis AJ, Simpson JA, Guymer RH. Dietary antioxidants and primary prevention of age related macular degeneration: systematic review and meta-analysis. BMJ 2007;335(7623):755.
     
  33. Richer S, Stiles W, Statkute L, Pulido J, Frankowski J, Rudy D, Pei K, Tsipursky M, Nyland J. Double-masked, placebo-controlled, randomized trial of lutein and antioxidant supplementation in the intervention of atrophic age-related macular degeneration: the Veterans LAST study (Lutein Antioxidant Supplementation Trial). Optometry 2004;75(4):216-230.
     
  34. Neelam K, Hogg RE, Stevenson MR, Johnston E, Anderson R, Beatty S, Chakravarthy U. Carotenoids and co-antioxidants in age-related maculopathy: design and methods. Ophthalmic Epidemiol 2008;15(6):389-401. 
     
  35. Age-Related Eye Disease Study 2 (AREDS2) Research Group, Chew EY, Clemons TE, Sangiovanni JP, et al. Secondary Analyses of the effects of lutein /zeaxanthin on Age-Related Macular Degeneration Progression: AREDS2 report No. 3. JAMA Ophthalmol 2014;132(2):142-149.
     
  36. Seddon JM, Hennekens CH. Vitamins, minerals, and macular degeneration. Promising but unproven hypotheses. Arch Ophthalmol 1994;112(2):176-179.
     
  37. West S, Vitale S, Hallfrisch J, Munoz B, Muller D, Bressler S, Bressler NM. Are antioxidants or supplements protective for age-related macular degeneration? Arch Ophthalmol 1994;112(2):222-227.
     
  38. VandenLangenberg GM, Mares-Perlman JA, Klein R, Klein BE, Brady WE, Palta M. Associations between antioxidant and zinc intake and the 5-year incidence of early age-related maculopathy in the Beaver Dam Eye Study. Am J Epidemiol 1998;148(2):204-214.
     
  39. Clemons TE, Kurinij N, Sperduto RD. Associations of mortality with ocular disorders and an intervention of high-dose antioxidants and zinc in the Age-Related Eye Disease Study: AREDS Report No. 13. Arch Ophthalmol 2004;122(5):716-726.
     
  40. Van Leeuwen R, Boekhoorn S, Vingerling JR, Witteman JC, Klaver CC, Hofman A, de Jong PT. Dietary intake of antioxidants and risk of age-related macular degeneration. JAMA 2005;294(24):3101-3107.
     
  41. Evans J. Antioxidant supplements to prevent or slow down the progression of AMD: a systematic review and meta-analysis. Eye (Lond) 2008;22(6):751-760.
     
  42. Cho E, Hung S, Willett WC, Spiegelman D, Rimm EB, Seddon JM, Colditz GA, Hankinson SE. Prospective study of dietary fat and the risk of age-related macular degeneration. Am J Clin Nutr 2001;73(2):209-218.
     
  43. Arnarsson A, Sverrisson T, Stefansson E, Sigurdsson H, Sasaki H, Sasaki K, Jonasson F. Risk factors for five-year incident age-related macular degeneration: the Reykjavik Eye Study. Am J Ophthalmol 2006;142(3):419-428.
     
  44. Chua B, Flood V, Rochtchina E, Wang JJ, Smith W, Mitchell P. Dietary fatty acids and the 5-year incidence of age-related maculopathy. Arch Ophthalmol 2006;124(7):981-986.
     
  45. Seddon JM, Rosner B, Sperduto RD, Yannuzzi L, Haller JA, Blair NP, Willett W. Dietary fat and risk for advanced age-related macular degeneration. Arch Ophthalmol 2001;119(8):1191-1199.
     
  46. Seddon JM, George S, Rosner B. Cigarette smoking, fish consumption, omega-3 fatty acid intake, and associations with age-related macular degeneration: the US Twin Study of Age-Related Macular Degeneration. Arch Ophthalmol 2006;124(7):995-1001.
     
  47. SanGiovanni JP, Chew EY, Clemons TE, Davis MD, Ferris FL, III, Gensler GR, Kurinij N, Lindblad AS, Milton RC, Seddon JM, Sperduto RD. The relationship of dietary lipid intake and age-related macular degeneration in a case-control study: AREDS Report No. 20. Arch Ophthalmol 2007;125(5):671-679.
     
  48. Mares-Perlman JA, Brady WE, Klein R, VandenLangenberg GM, Klein BE, Palta M. Dietary fat and age-related maculopathy. Arch Ophthalmol 1995;113(6):743-748.
     
  49. Delcourt C, Carriere I, Cristol JP, Lacroux A, Gerber M. Dietary fat and the risk of age-related maculopathy: the POLANUT study. Eur J Clin Nutr 2007;61(11):1341-1344.
     
  50. SanGiovanni JP, Chandra SR, Chew EY, Friberg TR, Klein ML, Kurinij N, Seddon JM. Dietary omega-3 long chain polyunsaturared fatty acids and risk for age-related macular degeneration. ARVO. Fort Lauderdale, USA, May 4-9, 2003. Invest Ophthalmol Vis Sci 2003;44(4 Suppl.):E-Abstract 2112.
     
  51. SanGiovanni JP, Chew EY, Agron E, Clemons TE, Ferris FL, III, Gensler G, Lindblad AS, Milton RC, Seddon JM, Klein R, Sperduto RD. The relationship of dietary omega-3 long-chain polyunsaturated fatty acid intake with incident age-related macular degeneration: AREDS Report No. 23. Arch Ophthalmol 2008;126(9):1274-1279.
     
  52. Souied E, Benlian P, Leveziel N, Zourdani A, Lablache-Combier M, Allaire C, Paccou B, Reynoird O, Carriere I, Coscas G, Delcourt C, Soubrane G. NAT-2. Report 1: High HDL and low PUFAs levels in exudative AMD patients. 112ème Congrès de la Société Française d’Ophtalmologie. Paris, France, 6-10 mai 2006. J Fr Ophtalmol 2006;29(HS 1):1S235.
     
  53. Tan JS, Wang JJ, Flood V, Mitchell P. Dietary fatty acids and the 10-year incidence of age-related macular degeneration: the Blue Mountains Eye Study. Arch Ophthalmol 2009;127(5):656-665.
     
  54. Chong EW, Robman LD, Simpson JA, Hodge AM, Aung KZ, Dolphin TK, English DR, Giles GG, Guymer RH. Fat consumption and its association with age-related macular degeneration. Arch Ophthalmol 2009;127(5):674-680.
     
  55. SanGiovanni JP, Agron E, Clemons TE, Chew EY. Omega-3 long-chain polyunsaturated fatty acid intake inversely associated with 12-year progression to advanced age-related macular degeneration. Arch Ophthalmol 2009;127(1):110-112.
     
  56. Awh CC, Lane AM, Hawken S, Zanke B, Kim IK. CFH and ARMS2 genetic polymorphisms predict response to antioxidants and zinc in patients with Age-Related Macular Degeneration. Ophthalmology 2013;120:2317-23. 
     
  57. Seddon JM, Silver RE, Rosner B. Response to AREDS supplements according to genetic factors. Br J Ophthalmol 2016;100(12):1731-1737.

Laser photocoagulation

Laser photocoagulation

English

Author:

Rufino Silva, MD, PhD
Coimbra University Hospital. Faculty of Medicine. Coimbra. Portugal 

 

Last revision: October 2011 by Rufino Silva

Introduction

Currently, three types of therapy are approved for the treatment of exudative age-related macular degeneration (AMD): laser photocoagulation, photodynamic therapy with Verteporfin and intravitreal injections of antiangiogenic agents (Ranibizumab and Pegaptanib).

Other treatments revealed to be ineffective or even more aggressive than natural disease progression. Examples of these later are radiotherapy, surgical removal of the subfoveal membranes, alpha 2 Interferon, transpupillary thermal therapy and anecortave acetate injections.

 

Laser photocoagulation in exudative AMD

The Macular Photocoagulation Study Group (MPS, 1982-1997)(1-16) has performed several randomized, double-blind, placebo-controlled studies in patients with exudative AMD.

These studies showed that laser photocoagulation might be effective in reducing loss of vision in cases of well defined exudative AMD lesions.

The importance of well-defined limits resulted from the absolute need to treat the entire lesion in order to maximally reduce recurrence and persistence rates, generally associated to greater loss of vision.

For extrafoveal and juxtafoveal lesions, results were evaluated for well-defined lesions before any differentiation based on angiographic patterns (occult or classic) was made.

In the subfoveal lesion study, results were evaluated for well-defined lesions with a classic component.

In addition to defining which cases to treat and the benefits expected from laser photocoagulation, the MPS defined the angiographic characteristics of neovascular lesions and guidelines for treating each type of membrane (extra, juxta or subfoveal), including preparation for treatment, treatment techniques, the wavelength to be selected, post-treatment care, special circumstances and expected complications.

Other studies of extra, juxta and subfoveal lesions with less impact on everyday clinical practice were performed by various authors using the same treatment technique studied by the MPS (for extra and juxtafoveal lesions) or macular grid photocoagulation (subfoveal lesions)(17,18,9).

 

Macular Photocoagulation Study Group (MPSG): extrafoveal lesions.

The first study was performed by the MPSG in patients with well-defined extrafoveal neovascular lesions (located 200 to 2500 µm from the foveal centre), with drusen, age ≥ 50 years and VA ≥ 20/100.

No differentiation was made between classic and occult membranes in this study.

Choroidal neovascularization was angiographically defined as the presence of leakage in the external retina.

Patients were enrolled in the study between 1979 and 1982 and treated with blue-green Argon laser.

The first results were published in the latter year (MPS, 1982)(1,2).

The MPS concluded that laser photocoagulation with blue-green or green Argon laser of sufficient intensity to produce nearly white lesions in the retina and cover the entire neovascular lesion, as well as adjoining blood, reduces the risk of additional and severe loss of vision, when compared to natural progression of the disease.

The benefits of laser were greater during the first year following treatment, having persisted after 5 years(5).

The probability of stabilizing or increasing VA doubled for treated eyes; a 58% reduction in the risk of severe loss of vision (6 lines in the ETDRS scale) was also observed.

After 5 years, 48% of treated eyes and 62% of non-treated eyes had lost ≥ 6 lines.

These results show reduced efficacy when evaluated in terms of the number needed to treat(20).

It was necessary to treat 7 patients for one patient to benefit from the treatment.

Average VA after 5 years was 20/125 in the treated group and 20/200 in the non-treated group (MPS, 1982, 1986).

After 5 years, 54% of treated eyes had shown recurrence with severe loss of vision most of them occurring in the first 2 years after treatment, they have been responsible for the majority of cases of severe loss of vision in the treated group.

Smokers had a greater risk of recurrence (recurrence was observed in 85% of patients smoking more than 10 cigarettes per day, compared to 51% of non-smokers).

Similar results were obtained in two other studies performed in the United Kingdom(17) and France(19).

 

Macular Photocoagulation Study Group (MPSG): juxtafoveal neovascular lesions

A second study was performed with juxtafoveal membranes and using a Krypton laser(11).

Inclusion criteria allowed treatment of well-defined choroidal neovascular lesions located 1 to 199 µ from the foveal centre, or 200 to 2500 µm from the foveal centre but showing blood or pigmentation less than 200 microns from the foveal centre (resulting in a barrier effect in fluorescence).

As opposed to the study with Argon laser in extrafoveal membranes, this trial did not require treatment of the entire area where bleeding occurred. Under no circumstances should treatment reach the foveal centre.

After 3 years, severe loss of vision (≥6 lines) had occurred in 49% of treated eyes and 58% of non-treated eyes(14).

The efficacy of this treatment in terms of the number needed to treat(20) was very low: 11.1.

This treatment reduced the risk of severe loss of vision by 10%.

However, this benefit was not observed in patients with hypertension or taking antihypertensive medication.

Nevertheless, the MPS maintained the indication to treat for these non-normotensive patients due to the absence of similar findings in other MPS studies.

After 5 years, the number of eyes with final VA ≥ 20/40 was double for treated eyes(14).

Persistence (incomplete treatment) and recurrence (neovascularization later than six weeks after treatment) were responsible for the majority of loss of vision in the treated group.

The MPS reclassified membranes as 100% classic, classic with an occult component and 100% occult.

Results were better in classic membranes: 54% of treated eyes and 72% of non-treated eyes lost 6 or more VA lines.

No statistically significant differences were observed between the treated and non-treated groups in cases of mixed membranes (only the classic component was treated) and 100% occult membranes.

Therefore, treatment of occult membranes and the classic component of mixed membranes was not effective in reducing loss of vision(14,16).

In conclusion, the MPS showed that laser photocoagulation of well-defined extrafoveal choroidal membranes and classic extra and juxtafoveal membranes secondary to AMD may prevent of delay loss of vision in patients fulfilling the inclusion criteria.

 

Macular Photocoagulation Study Group (MPSG): subfoveal neovascular lesions

In 1986, the MPS started two studies(12,13) to determine the efficacy of laser photocoagulation in subfoveal choroidal neovascularization.

In the first study, the effect of laser photocoagulation (Argon or Krypton) was evaluated in eyes with subfoveal exudative AMD not previously treated; in the second study, the efficacy of laser treatment in subfoveal recurrence in eyes with extra or juxtafoveal membranes was evaluated.

The results of this study and treatment recommendations generated a great deal of controversy.

In fact, treated eyes displayed a very marked loss of vision immediately after treatment.

After 4 years, 30% of treated eyes and 60% of non-treated eyes displayed VA ≤ 20/400, whereas 45% of non-treated eyes and 23% of treated eyes has suffered severe loss of vision.

The efficacy of this treatment in terms of the number needed to treat was 4.5(20).

A large percentage of ophthalmologists did not agree with the MPS recommendations for treating subfoveal lesions.

In fact, patients were losing 3 lines immediately after treatment.

The MPS re-evaluated treatment efficacy in terms of lesion size and difference from baseline VA, having established treatment groups and criteria according to these two variables(13).

Ophthalmologists could advise their patients and help them choose whether or not to undergo treatment according to lesion size and baseline VA.

With the emergence of photodynamic therapy with Verteporfin, laser photocoagulation for subfoveal lesion became obsolete.

It remains indicated only for extrafoveal lesions and the angiographic control should be performed 15 days after treatment.

 

Treatment

Preparation for treatment

Regarding preparation for treatment, the MPS recommended that patients should be informed that photocoagulation causes permanent paracentral scotoma in cases of juxta and extrafoveal choroidal membranes.

Patients should also be informed that they may continue to lose vision, even under the best treatment conditions, and that treatment does not cure AMD but it is only a means of reducing the risk of marked loss of visual acuity.

A fluorescein angiography (FA) should be performed 72 to 96 hours before photocoagulation in order to select treatable cases and to guide the ophthalmologist during treatment.

Patients should undergo treatment as quickly as possible, since neovascular lesions may grow 10 to 18µ per day(21).

Most neovascular lesions are extra or juxtafoveal at the onset, becoming subfoveal with rapid growth towards the fovea.

 

Treatment technique

The MPS recommends that treatment should be performed so that a white lesion in the retina is obtained.

The neovascular lesion should be surrounded by laser marks with a diameter of 200µ and duration of 0.2 to 0.5 seconds.

After surrounding the perimeter of the neovascular lesion, its central part is covered with 200µ burns; the remaining lesion is covered with 200 to 500µ burns, with duration of 0.5 to 1.0 seconds.

In cases of juxtafoveal lesions, the foveal centre should be preserved, although it should be ensured that the entire lesion is treated.

If bleeding extends to the area under the fovea, treatment should include the entire neovascularization area and stop at the limit of the fovea.

Since the emergence of new treatments, namely intravitreal antiangiogenic treatments, laser photocoagulation of juxtafoveal lesions has become controversial.

The MPS demonstrated that the wavelength selected does not affect laser results.

Laser treatment should avoid retinal blood vessels and the optic nerve (treatment should start 10-200 µm from the optic nerve), as well as preserve at least 1.5 hours of the papillomacular bundle (no peripapillary treatment).

Treatment of serous pigment epithelial detachment (PED) could be indicated when photocoagulation is used to treat subfoveal lesions including serous PED as a component(1,3,8,10,11).

 

Post-treatment follow-up

Follow-up of treated patients was also recommended and defined by the MPS.

In addition to self-evaluation, it is necessary to perform medical examinations and control FA 2 to 3 weeks, 4 to 6 weeks, 3 to 4 months and 6, 8, 9 and 12 months after treatment.

Recurrence is rare after 2 years.

The greater risks exist 6 weeks to 12 months after treatment. Detection through biomicroscopy without FA is sometimes difficult.

Angiography allows detection of approximately 12% of the cases that go unnoticed in medical examinations.

Recurrence and persistence rates are much greater in cases of choroidal neovascular or disciform lesions caused by AMD in the non-treated eye.

Other factors that appear to increase recurrence rates include smoking, hypertension and choroidal neovascularization with reduced pigmentation(1,2,8,11).

 

Treatment complications

Laser photocoagulation treatment may also lead to complications, including choroidal haemorrhage (rarer if spots ≥ 200 microns and time intervals ≥ 0.2 seconds are used), premacular fibrogliosis, accidental treatment of the fovea in extrafoveal or juxtafoveal lesions (minimised by retrobulbar anaesthesia, drawing of lesion limits and correct identification of the fovea), rupture of the pigment epithelium (more frequent in cases of PED) and atrophy of the RPE in the area adjoining the laser scar (immediately after treatment or years later)(2,3,7,8).

 

Treatment of occult membranes

The MPS also defined guidelines regarding occult membranes.

When extrafoveal and juxtafoveal neovascular lesions caused by AMD started to be studied no distinction was made between classic and occult membranes.

Subsequent analysis of all angiography results obtained during study of juxtafoveal lesions revealed that treatment was effective for classic neovascular lesions with no occult component.

In cases where an occult component (not treated) coexisted with a classic component no benefits were gained from treatment(15).

Photocoagulation may be reasonably considered in cases of well-defined, symptomatic, occult neovascular lesions with no classic component, in order to reduce the risk of membrane growth towards the fovea.

However, little knowledge exists of the natural progression of these occult membranes and it would not be wrong to delay treatment while examining patients at regular intervals (of months, albeit varying according to the type of membrane), in order to wait for the appearance of a classic membrane that would benefit from laser photocoagulation treatment.

Only 25% of occult choroidal membranes maintained baseline VA values after 3 years and approximately 50% suffer severe loss of vision within the same time period(13).

 

Other laser photocoagulation studies

In a retrospective study, Soubrane et al.(18) demonstrated the absence of benefits for the treatment of extrafoveal and juxtafoveal occult neovascular lesions.

Scatter or grid photocoagulation revealed to be ineffective in ill-defined neovascular lesions(18).

In an attempt to preserve the foveal centre, Coscas et al(19) described a form of ring treatment for subfoveal membranes.

This randomized, placebo-controlled study included eyes with VA 20/200–20/1000.

Treatment surrounded the 400 central µ of the central avascular area.

After one year, baseline VA has been maintained or increased in 41% of treated eyes and only 20% of non-treated eyes.

This technique was not well received. Several groups determined the efficacy of laser photocoagulation in eyes with AMD and PED.

In MPS studies, cases with PED were excluded. The Moorfields Macular Study Group(17) showed that grid laser photocoagulation of “pure” PED (with no clinical or angiographic signs of choroidal membrane) had worsened prognosis in terms of VA.

With the advent of indocyanine green (ICG) angiography, enhanced imaging of occult CNV allowed a characterization of at least 2 forms of occult CNV: a plaque of late staining and a focal area of active vessel proliferation or a so-called “hot spot”(22,23).

ICG-laser photocoagulation was used in several centers(23-28), in uncontrolled studies, to treat these hot spots with apparently relative success.

Polypoidal choroidal vasculopathy (PCV) and retinal angiomatous proliferation (RAP), two AMD sub-types, represented the great majority of these treated cases.

Laser photocoagulation in RAP lesions have shown very poor results with a high rate of persistence and recurrences(29-30).

Better results may be obtained in early lesions with extrafoveal hot spot resulting in stabilization of the pathology and visual acuity.

However, an accurate follow-up is mandatory after the treatment due to the high rate of recurrences.

Direct laser photocoagulation of polypoidal lesions has shown controversial results(31-34).

Treatment of leaking polyps has proven short-term safety and efficacy for extrafoveal lesions(31-32).

Yuzawa et al(33) reported good efficacy of laser photocoagulation in near 90% of the eyes if all the polyps and abnormal vascular network were treated.

If the treatment involved only the polyps more than half of the eyes suffered VA decrease related with exudation, recurrences, or foveal scars.

Considering the possibility of using other treatment modalities, laser photocoagulation should be reserved for well defined extrafoveal active polyps.

 

Conclusion

Laser photocoagulation remains currently indicated for the treatment of well-defined extrafoveal choroidal membranes.

For classic juxtafoveal membranes, laser photocoagulation could theoretically be considered as an option for cases in which the entire neovascular lesion can be treated without damaging the fovea.

However, considering the great incidence of persistence and recurrences, intravitreal antiangiogenic agents are the first treatment option.

Photodynamic therapy with verteporfin and antiangiogenic agents eliminated all indications for the treatment of subfoveal neovascular lesions, with laser photocoagulation.

>> References

References - Laser photocoagulation

References - Laser photocoagulation

English
  1. Macular Photocoagulation Study Group. Argon laser photocoagulation for senile macular degeneration. Results of a randomized clinical trial. Arch Ophthalmol 1982; 100: 912-918.
     
  2. Macular Photocoagulation Study Group. Argon laser photocoagulation for neovascular maculopathy: three-year results from randomized clinical trials. Arch Ophthalmol 1986; 104: 694-701.
     
  3. Macular Photocoagulation Study Group. Krypton laser photocoagulation for idiopathic neovascular lesions. Results of a randomized clinical trial. Arch Ophthalmol 1990; 108: 832-837.
     
  4. Macular Photocoagulation Study Group. Persistent and recurrent neovascularization after Krypton laser photocoagulation for neovascular lesions of age-related macular degeneration. Arch Ophtahlmol 1990b; 108: 825-833.
     
  5. Macular Photocoagulation Study Group. Argon laser photocoagulation for neovascular maculopathy: five-year results from randomized clinical trials. Arch Opthalmol 1991; 109: 1109-1114.
     
  6. Macular photocoagulation Study Group. Laser photocoagulation of subfoveal neovascular lesions in age-related macular degeneration: results of a randomized clinical trial. Arch Ophtahlmol 1991b; 109:1220-1231.
     
  7. Macular Photocoagulation Study Group. Laser photocoagulation of subfoveal recurrent neovascular lesions in age-related macular degeneration: Results of a randomized clinical trials. Arch Opthalmol 1991c; 109: 132-1241.
     
  8. Macular photocoagulation Study Group. Subfoveal neovascular lesions in age-related macular degeneration: guidelines for evaluation and treatment in the Macular Photocoagulation Study. Arch. Ophtahlmol 1991d; 109:1242-1257.
     
  9. Macular Photocoagulation Study Group. Five-year follow-up of fellow eyes of patients with age-related macular degeneration and unilateral extrafoveal choroidal neovascularization. Arch Ophthalmol 1993; 111:1189-1199.
     
  10. Macular Photocoagulation Study Group. Laser photocoagulation of subfoveal neovascular lesions of age-related macular degeneration. Arch Ophthalmol 1993b; 111:1200-1209.
     
  11. Macular Photocoagulation Study Group. Laser photocoagulation for juxtafoveal choroidal neovascularization. Arch Opthalmol 1994a; 112: 500-509.
     
  12. Macular Photocoagulation Study Group. Persistent and recurrent neovascularization after laser photocoagulation for subfoveal choroidal neovascularization of age-related macular degeneration. Arch Opthalmol 1994b; 112:489-499.
     
  13. Macular Photocoagulation Study Group. Visual outcome after laser photocoagulation for subfoveal choroidal neovascularization secondary to age-related macular de-generation: the influence of initial lesion size and initial visual acuity. Arch Ophthalmol 1994c; 112: 480-488.
     
  14. Macular Photocoagulation Study Group. Laser photocoagulation for juxtafoveal choroidal neovascularization: five-year results from randomized clinical trials. Arch Ophthalmol 1994d; 112: 500-509.
     
  15. Macular Photocoagulation Study Group. Occult choroidal neovascularization: influence of visual outcome in patients with age-related macular degeneration. Arch Ophthalmol 1996; 114: 400-412.
     
  16. Macular photocoagulation Study Group. Risk factors for choroidal neovascularization in the second eye of patients with juxtafoveal or subfoveal neovascularization secondary to age-related macular degeneration. Arch Ophthalmol.1997; 115:741-747.
     
  17. Morfields Macular Study Group. Treatment of senile disciform macular degeneration: a single blind randomized trial by argon laser photocogulation. Br J Ophthalmol 1982; 66:745-753.
     
  18. Soubrane G, Coscas G, Français C, Koenig F. Ocult subretinal new vessels in age-related macular degeneration: natural history and early laser treatment. Ophthalmology 1990; 97:649-657.
     
  19. Coscas G, Soubrane G, Ramahefasolo. Perifoveal laser treatment for subfoveal choroidal new vessels in age-related macular degeneration. Results of a randomized clinical trial. Arch Ophthalmol 1991; 109:1258-1265.
     
  20. Vartner P. Applying number-need-to treat (NNT) analysis to ophthalmic clinical trials. Optom Vis Sci 2005; 83: 919-930.
     
  21. Vander JF, Morgan CM, Schatz H, 1989. Growth rate of subretinal neovascularization in age-related macular degeneration. Ophthalmology 96:1422-1426.
     
  22. Guyer DR, Yannuzzi LA, Slakter JS, et al. Classification of choroidal neovascularization by digital indocyanine green videoangiography. Ophthalmology 1996;103:2054–60.
     
  23. Fernandes LHS, Freund KB, Yannuzzi LA, Spaide RF, Huang SJ, Slakter JS, Sorenson JA, 2002. The nature of focal areas of hyperfluorescence or “hot spots” imaged with indocyanine green angiography. Retina. 2002; 22:557-568.
     
  24. Slakter JS, Yannuzzi LA, Sorenson JA, et al. A pilot study of indocyanine green videoangiography-guided laser treatment of occult-choroidal neovascularization in age-related macular degeneration. Arch Ophthalmol 1994;112:465–72.
     
  25. Regillo SD, Benson WE, Maguire JI, Annesley WH. Indocyanine green angiography and occult-choroidal neovascularization. Ophthalmology 1994;101:280–8.
     
  26. Lim JI, Sternberg PJ, Capone AJ, et al. Selective use of indocyanine green angiography and occult-choroidal neovascularization. Am J Ophthalmol 1995;120:75– 87.
     
  27. Introini U, Brancato R, Pece A. ICG-guided laser photocoagulation of occult-CNV. Ophthalmologica 1998;215:295–300.
     
  28. Guyer DR, Yannuzzi LA, Ladas I, et al. Indocyanine-green guided laser photocoagulation of focal spots at the edge of plaques of choroidal neovascularization. Arch Ophthalmol 1996;114:693–
     
  29. Kuhn D, Meunier I, Soubrane G, Coscas G. Imaging of chorioretinal anastomoses in vascularized retinal pigment epithelium detachments. Arch. Ophthalmol. 1995; 113: 1392-1398.
     
  30. Johnson TM, Glaser BM. Focal laser ablation of retinal angiomatous proliferation. Retina 2006; 26 (7): 765-72.
     
  31. Guyer DR, Yannuzzi LA, Ladas I, Slakter JS, Sorenson JA, Orlock D. Indocyanine green-guided laser photocoagulation of focal spots at the edge of plaques of choroidal neovascularization. Arch Ophthalmol 1996; 114 (6): 693-697.
     
  32. Gomez-Ulla F, Gonzalez F, Torreiro MG. Diode laser photocoagulation in idiopathic polypoidal choroidal vasculopathy. Retina 1998; 18 (5): 481-483.
     
  33. Yuzawa M, Mori R, Haruyama M. A study of laser photocoagulation for polypoidal choroidal vasculopathy. Jpn J Ophthalmol 2003; 47 (4): 379-384.
     
  34. Yamanishi A, Kawamura A, Yuzawa M. Laser photocoagulation for idiopathic polypoidal choroidal vasculopathy. Jpn J Clin O 1998; 52: 1691-1694.

Photodynamic Therapy

Photodynamic Therapy

English

Authors:

Rita Flores, MD

Centro Hospitalar de Lisboa Central, Lisboa, Portugal ( CHLC )

Nova Medical School, Faculdade de Ciência Médicas de Lisboa, Portugal ( FCML )

 

Rufino Silva, MD, PhD

Centro Hospitalar e Universitário de Coimbra (CHUC). Coimbra, Portugal

Faculty of Medicine, University of Coimbra. Portugal (FMUC)

Association for Innovation and Biomedical Research on Light and Image (AIBILI). Coimbra. Portugal 

Updated/reviewed by the authors, July 2017.

Introduction

Photodynamic therapy was approved in 2000 as an alternative treatment for patients with AMD of the exudative form, having been the first effective pharmacological treatment for this form of the disease.

Until then, laser photocoagulation was only successful in treating a small percentage of neovascular lesions (juxtafoveal and extrafoveal), excluding subfoveal lesions, which are more frequent.

With the emergence of antiangiogenic therapies, photodynamic therapy has been used less frequently.

However, it remains useful in three situations: in patients with systemic or ocular contraindications regarding intravitreal administration of antiangiogenic drugs, as an adjuvant, in combination with other drugs, and in the treatment of polypoidal choroidal vasculopathy and central serous chorioretinopathy.
 

Mechanism of action

Experimental studies(1,2) suggest that photodynamic therapy (PTD) causes endothelial cell lesions, with formation of clots and selective vascular occlusion. Endothelial cell membrane lesions appear to be caused by free radicals released when verteporfin is activated by non-thermal laser light. These free radicals react with endothelial cell membranes and circulating blood cells, inducing platelet activation and local clot formation.

The mechanisms by which PTD induces tissue destruction are not exactly known. Three related mechanisms of action have been proposed: cellular, vascular and immune(3)

The cellular mechanism, which is the most relevant, corresponds to the cytotoxic effects of free radicals on mitochondria, the endoplasmic reticulum and lysosomes.

When exposed to these radicals, endothelial cell membranes rupture, exposing the basal membrane, which causes platelet adhesion and aggregation.

Activated platelets release mediators such as histamine, thromboxane and TNF-α.

These mediators trigger a sequence of events, namely vasoconstriction, thrombosis, increased vascular permeability, blood stasis and hypoxia.

The proposed immune mechanism is based on the high concentrations of cytokines observed in patients subject to PDT, such as interleukin 2 and TNF–α.

It is equally admitted that PDT may decrease immune response by reducing antigen-presenting cell activity.

Standard treatment consists of endovenous infusion of verteporfin at a dose of 6 mg/m2 body surface, for 10 minutes.

Fifteen minutes after starting the infusion, the patient is treated with a diode laser with wavelength of 689 nm and light intensity of 600 mw/cm2, at a radiation dose of 50 J/cm2, with an exposure time of 83 seconds and a spot diameter corresponding to the diameter of the largest lesion plus 1mm.

These parameters have been studied and appear to be ideal, allowing maximum vascular effect with minimum photoreceptor and pigment epithelial cell damage.

Verteporfin activation by the diode laser induces temporary closure of the choroidal neovascular complex, through the mechanisms already described, causing little damage to adjacent retinal structures.

This characteristic doubtlessly represented a therapeutic advantage, since it allowed treatment of lesions whose location or size prevented use of other available therapies, namely conventional laser photocoagulation.

However, photodynamic therapy does entail some damage, although induced retinal lesions are smaller than that occurring following thermal laser photocoagulation. Laser fluence reduction protocols have been proposed in the attempt to reduce the extent of this damage.

Therapy schemes with more intense treatment regimes, including treatment every 2 months in the first 6 months, were also tested.

The efficacy and safety of the latter regime were compared with those of the standard regime(4).

No statistically significant differences were found between the two regimes in terms of visual improvement, number of retreatments and safety.

The intensive treatment regime in the first 6 months appears to be more effective in preventing severe loss of visual acuity; however, the difference observed after 24 months is not statistically significant, with loss of visual acuity greater than 6 lines being observed in 25% of patients treated with the intensive regime and 38% of patients treated with the standard regime.
 

Main clinical trials

The efficacy of PDT was evaluated in several multicentric, randomized clinical trials in patients with AMD with choroidal neovascularization, of which the following should be highlighted:

  • Treatment of AMD with PDT (TAP studies)(5,6,7,8,9)
  • Verteporfin in PDT (VIP studies)(10,11)
  • Verteporfin in Minimally Classic Choroidal Neovascularization (VIM studies)(12)
  • Visudyne in Occult Classic Choroidal Neovascularization (VIO study)(13)
  • Meta-analysis of the TAP and VIP Studies(14)
  • TAP Extension(15)

Many studies were subsequently performed in order to study and compare several therapeutic modalities, of which the following should be highlighted:

  • Anti-Vascular endothelial growth factor (VEGF) Antibody for the Treatment of Predominantly Classic Choroidal Neovascularization (CNV) in AMD (Anchor Study(16,17,18))
  • Ranibizumab Combined with Verteporfin Photodynamic Therapy in Neovascular AMD (Focus(19))
  • Summit Clinical Trial Program, which includes 3 studies: the Mont Blanc, Denali and Everest Studies(20)
     

TAP Study

This study provided the main evidence of PDT efficacy. It included two multicentric, double-blind, randomized, placebo-controlled studies, in Europe and the United States of America (Table 1).

Four hundred and two patients with classic subfoveal choroidal neovascularization were treated with PDT, while 207 patients were treated with placebo. The primary endpoint was the percentage of eyes for which losses of less than 15 ETDRS letters from baseline were observed at 12 and 24 months. PDT was significantly more effective than the placebo, both at 12 months (61% versus 46%) and 24 months (53% versus 38%).

These results were more significant in predominantly classic membranes.

Study Number of patients
Pred. classic
Verteporfin
Pred. classic
Placebo
 

p

All membranes
Verteporfin
All membranes
Placebo
p
TAP: 12 months
N=609
67.3% 39.8% <0.001 61.2% 46.4% <0.001
TAP: 24 months
N=609
59.1% 31.3% <0.001 53% 37.7% <0.001
TAP: 36 months
N=476
58.1% -        
TAP: 48 months 57% -

 

     

Table 1 - TAP study: percentage of eyes with loss <3 lines in the ETDRS chart

 

VIP Study

In this study, the efficacy and safety of Photodynamic Therapy were evaluated in patients with occult lesions (Table 2).

Results after 12 months were somewhat disappointing; however, efficacy was demonstrated in the treated group at 24 months (46.2% versus 33.3%). Subgroup analysis led to the conclusion that greater benefits were achieved in patients with small lesions (less than 4 disc areas) and/or visual acuity worse than 20/50. In these patient subgroups, the differences between the PDT group and the placebo group had greater statistical significance (51% versus 25%).

  

  

  

  

  

  

 

  

  

  

  

  

  

 

  

  

  

  

  

  

 

  

  

  

  

  

  

 

  

  

  

  

  

  

Study
MTRI
verteporfin
MTRI
Placebo
 

p

VIP 12 months 49% 45% Ns
VIP 24 months 45% 32% 0.032
VIM 12 months
300 mw/cm2
86% 53% 0.002
VIM 12 Months
600 mw/cm2
72% 53%

0.08

VIM 24 months
300 mwcm2
74% 38% 0.003
VIM 24 months
600 mwcm2
47% 38% 0.45

Table 2 - VIP and VIM studies: percentage of eyes with loss <3 lines in the ETDRS chart.

 

VIM Study

The objective of this study was to determine the efficacy of photodynamic therapy in minimally classic membranes (where the classic component represents less than 50% of the neovascular lesion) sized below six disc areas (Table 2).

Additionally, the efficacy of reducing fluence to 50% (25J/cm2) relatively to standard parameters (50J/cm2) was also analysed.

In the standard laser light activation protocol, a wavelength of 689 nm and an intensity of 600 mw/cm2 are used for 83 seconds to achieve a fluence value of 50J/cm2.

In this study, no statistically significant efficacy was found at 12 and 24 months in the group of patients treated with the standard protocol. On the contrary, better results were observed for patients treated with the reduced fluence protocol, in terms of the primary endpoint (loss of visual acuity of less than 15 letters).

Based on these results, the study authors advise treatment of small minimally classic lesions with PDT, concluding that the reduced fluence protocol may be beneficial. The percentage of conversion of minimally classic lesions into predominantly classic lesions was also studied and treatment efficacy was demonstrated, irrespective of the fluence used.

The reduced fluence issue will also be referred in the Denali study.

Two other studies – VALIO (Verteporfin Therapy with Altered Light in Occult choroidal neovascularization) and VER (Verteporfin Early Retreatments) were also performed.

In the VALIO study, the efficacy of laser treatment at 15 and 30 minutes was evaluated and compared.

Since no statistically significant differences were observed between these two therapeutic modalities, it was decided to maintain the 15 minutes used in standard treatment.

The objective of the VER study was to determine whether it would be beneficial to reduce treatment intervals to 6 weeks in the first 6 months.

Since no increase in efficacy was found relatively to the standard regime (treatment every 3 months), it was advised that the usual treatment regime be maintained.

  

  

  

  

  

  

 

  

  

  

  

  

  

 

  

  

  

  

  

  

 

  

  

  

  

  

  

 

  

  

  

  

  

  

Study
MTRI
verteporfin
MTRI
Placebo
 

p

VIP 12 months 49% 45% Ns
VIP 24 months 45% 32% 0.032
VIM 12 months
300 mw/cm2
86% 53% 0.002
VIM 12 Months
600 mw/cm2
72% 53%

0.08

VIM 24 months
300 mwcm2
74% 38% 0.003
VIM 24 months
600 mwcm2
47% 38% 0.45

Table 2 - VIP and VIM studies: percentage of eyes with loss <3 lines in the ETDRS chart.

 

VIO Study

The VIO study was designed to determine PDT indications in occult lesions with no classic component. Although the complete results report has not been published, the primary endpoint had not been reached at 12 and 24 months; therefore, no significant benefits were demonstrated for the treatment of occult membranes with PDT.

These results led the EMEA to remove occult membranes from the list of photodynamic therapy indications (April 2007).

 

Meta-analysis of the TAP and VIP studies

The meta-analysis of the TAP and VIP studies was a retrospective analysis in which lesion size, composition and visual acuity at baseline were considered, as well as possible relations between these parameters and study results.

The objective of this meta-analysis was to explain the apparent discrepancies found between the TAP and VIP study results, considering the following:

  • in the TAP study, treatment was found to be beneficial in predominantly classic and occult lesions, whereas it was found not to be beneficial in minimally classic lesions;
  • in the VIP study, treatment of occult lesions was found to be more beneficial in small lesions (≤4 disc areas) and/or visual acuity <20/50.

This meta-analysis revealed that the most important factor in predicting final visual acuity in patients treated with PDT appears to be lesion size.

Therefore, treatment of small lesions (≤4 disc areas) will be beneficial for all types of lesions, including occult lesions with no classic component, provided lesions are recent.

Regarding classic membranes, treatment benefits extend to lesions > 4 DA and non-recent lesions.

 

TAP Extension

Some patients that completed the 2-year TAP were enrolled in a 3-year extension study, for a total duration of 5 years (60 months), under an open-label regime.

The main objective of this study was to obtain long-term visual acuity and 5-year safety data in patients with subfoveal choroidal neovascularization treated with photodynamic therapy.

Patients having completed month 24 of the TAP study were eligible to participate in the study extension, irrespectively of having been included in the treatment or the placebo group and of lesion characteristics at baseline. In the TAP study extension, visual outcomes remained stable between month 24 and month 60, even in patients with low retreatment rates.

No safety problems were found leading to contraindications being associated to retreatment with photodynamic therapy in the 5 years of study duration. No safety problems were found in bilateral treatment.

 

PDT Safety

The most complete and extensive PDT safety data were published in the meta-analysis of the TAP and VIP studies, where a comparison with placebo was performed. PDT is considered a safe treatment, with rare side effects, of little significance (Table 3).

Choroidal hypoperfusion associated to PDT has been documented in fluorescein and ICG angiography in the first days after treatment and, more rarely, in the following months. Controversy exists regarding the cumulative effect of treatment in permanent occlusion of the normal choriocapillaris and the association between this hypoperfusion and eventual functional consequences(21).

Ocular effects

Non-specific visual disorders

Transient loss of visual acuity (18% vs. 0%)

Severe loss of visual acuity (≥ 20 letters

up to 7 days after PDT) (0.7% vs. 0%)

Scotomatous alterations (6% vs. 3.4%)

Systemic effects

Injection site reactions (13% vs. 5.6%)

Lower back pain (2.4% vs. 0%)

Hypersensitivity reactions (3% vs. 0%)

Sleep pattern alterations (1.6% vs. 0%)

Table 3 - PDT adverse effects

 

Combined treatments

Combined approaches for treating exudative AMD have been investigated as a mean of improving treatment efficacy and reducing treatment frequency.

Many non-randomized studies reported successful treatment using combinations of PDT, corticosteroids and antiangiogenic agents(22,23,24).

The Focus trial(19) showed that combination therapy using PDT and Ranibizumab was superior to PDT alone in efficacy and also reduced the need for repeat PDT sessions.

A merely illustrative comparison of the Anchor(16) and Focus(19) trials showed more favourable results in terms of visual acuity gain in the Anchor patients, which included only treatment naïve patients, suggesting that adding PDT to Ranibizumab may not increase the visual acuity gain.

The SUMMIT program, which includes three randomized clinical trials - DENALI, EVEREST and MONT BLANC, was designed to compare a combination therapy with PDT and ranibizumab with ranibizumab monotherapy.

The DENALI study(26) is a two-year, randomized, double-blind multicentric study conducted at 45 centres in the United States and five centres in Canada.

Enrolled patients with subfoveal CNV of all angiographic subtypes were randomized to receive either ranibizumab monotherapy, a combination of ranibizumab and standard fluence (ST) PDT or a combination of ranibizumab and reduced-fluence (RF) PDT.

This study investigated the efficacy and safety of combined therapy involving PDT and antiangiogenic drugs, namely ranibizumab 0.5 mg, administered intravitreally.

Combining verteporfin PDT with ranibizumab 0,5 mg (with 3 ranibizumab loading doses followed by additional injections on a monthly as-needed basis) can improve visual acuity from baseline at month 12 by 5,3 letters for verteporfin ST PDT and 4,4 letters for verteporfin RF PDT combination therapy versus 8,1 letters for ranibizumab alone.

Althought the primary objective (to demonstrate non-inferiority of at least one of the verteporfin combination arms to ranibizumab monotherapy) was not met, combination therapy reduced the number of injections required: 5,1 verteporfin SF PDT and 5,7 verteporfin RF PDT combination therapy versus 10,5 for ranibizumab alone.

Reduced fluence did not provide a clinical benefit over standart fluence in verteporfin PDT combination arms.

MONT BLANC, a similar study conducted at 50 centres throughout Europe, enrolled subjects with subfoveal CNV of all angiographic subtypes, who were randomized to receive either ranibizumab monotherapy or ranibizumab in combination with standard fluence PDT.Preliminary visual acuity results at 12 months revealed the non-inferiority of the combined treatment (PDT+Ranibizumab), when compared with Ranibizumab alone; the number of treatments and safety evaluation were similar in both groups.

These results and those from Focus trial suggest that PDT with standard fluence may be useful in combination with Ranibizumab for treating predominantly classic, minimally classic or occult AMD lesions.

Certain angiographic lesion subtypes, such as retinal angiomatous proliferation (RAP) and polypoidal choroidal vasculopathy appear to respond differently to PDT treatment(24,25) when compared to predominantly classic, minimally classic or occult lesions.

It is unclear whether they are more likely to benefit from a combination therapy. Polypoidal choroidal vasculopathy (PCV) may be considered as a well-defined subtype of AMD with a distinct natural history characterized by multiple recurrences and specific response to treatment. PCV often follows a remission-relapsing course and usually has a good visual prognosis. However, up to half of patients may have persistent bleeding and leakage, leading to vision loss.

The EVEREST study (part of the SUMMIT programme) is being performed in Asia and is designed to compare and evaluate the efficacy and safety of verteporfin PDT alone or in combination with ranibizumab, with that of ranibizumab monotheraphy for symptomatic macular PCV.

EVEREST trial(27) demonstrated that verteporfin PDT combined with ranibizumab or alone was statistically superior to ranibizumab monotherapy in achieving complete polyp regression in PCV patients (primary end point). The proportion of patients with at least a complete regression of polyps at any time-point during the study was significantly larger with verteporfin PDT combined with ranibizumab (83,32%) or alone (85,7%) versus ranibizumab monotherapy (42,9%). At month 6, a decrease in mean polyp area from baseline was seen in all three treatment groups. The largest decrease was seen with verteporfin PDT combined with ranibizumab followed by verteporfin PDT monotherapy, and ranibizumab monotherapy.

Subsequent Roundtable meetings(28) with international experts in retinal diseases had been held annually since 2007 and had formulate practical guidelines on diagnosis and management of PVC.

When considering PCV, ICGA is strictly necessary to confirm or to exclude the diagnosis. Then treatment should be considered for active symptomatic PCV and can be considered for active asympomatic PCV. The ICGA –guided thermal laser photocoagulation may be considered for extrafoveal polyps.

For the inicial treatment of active juxtafoveal and subfoveal PCV, the recommendation is either combination of standard fluence verteporfin PDT and 3x antiVEGF intravitreal injections at monthy intervals or ICGA-guided standart fluence verteporfin PDT. The combination treatment should be considered when there is leakage from polyps and from associated branching vascular network, or when there is large amount of subretinal fluid or exudation associated with PED. Other conditions that suggest the choice of combination treatment are ICGA images ambiguous or combining features of PCV and CNV. On the other hand antiVEGF monotherapy may be considered for initial treatment if verteporfin PDT is contraindicated or is not possible.

Monthy monitoring includes visual acuity, slit-lamp biomicroscopy and OCT. Three months after inicial treatment FA, ICGA and OCT shoud be performed.

If there is incomplete regression of polyps at this time, retreatment with verteporfin PDT alone or with antiVEGF in association shoud be considered. If there is complete regression of polyps at 3 months detected by ICGA but there is leakage on FA, subsequent antiangiogenic treatment is recommended.

Some manuscripts, studies and retrospective reports demonstrated that total polyp regression or complete disappearance of PCV lesions occurred in 56–95% of ≥200 eyes treated with verteporfin PDT(29).These studies indicated that many verteporfin-treated patients had stable or improved  vision
(Table 4), with outcomes that compared favourably with the natural history of PCV.s

 

  

  

  

  

  

  

 

  

  

  

  

  

  

 

  

  

  

  

  

  

 

  

  

  

  

  

  

 

  

  

  

  

  

  

Authors
Spaide
2002

Chan
2004

Silva
2005
Hussain
2005
Mauget-Faÿsse 2006 Eandi
2007
Gomi
2008

Akaza
2008

Type R P R P P R P P
N 16 22 21 9 30 30 36 57
Age (average) 70.5 66.6 75.6 67.2 67 75 72 71
VA increase 12M 56.3% 45.5% 28.6% 0.0% - 50.0% 25.0% 12.0%
VA stabilization 12M 31.3% 50.0% 57.1% 100.0% - 30.0% 67.0% 77.0%
VA decrease 12M 12.5% 4.5% 14.3% 0.0% - 20.0% 8.0% 11.0%
VA increase 24M     0.0%         9.0%
VA stabilization 24M     100.0%         70.0%
VA decrease 24M     0.0%         22.0%
Comments
Average follow-up 12M
 
6 eyes at 24M
 
Mean VA improved from 0.50 to 0.38 logMAR
     

R – Retrospective; P – Prospective.

Table 4 - PDT and PCV. Results from different studies.

 

More recently aflibercept has shown good results in treating PCV. These results appear to be superior to those obtained with ranibizumab with a complete regression of polypoidal lesions ranging from 55.4%. to 69.2% at one year, and with a mean number of 7 injections(30, 31).

Three randomised clinical trials on naïve PCV patients are beeing runned – Everest II, Planet  and Atlantic.  The Everest II is a phase 4, 2-year RCT comparing Ranibizumab alone with a combined teraphy of Ranibizumab plus Verteporfin PDT in 321 Asian patients. Estimated primary completion date is April 2017(32). The Planet RCT enrolled 331 patients in ASIA and 2 European countries. It is a phase 3-4, 1-year study comparing Aflibercept alone with Aflibercept plus Verteporfin PDT in patients with PCV.  Estimated Primary completion date is August 2016(33) Atlantic Study, is a phase 4, 1-year RCT, being runned in Portugal and Spain, comparing intravitreal treat and extend aflibercept monotherapy with aflibercept treat and extend regimen with adjunctive PDT in patients with PCV.  Estimated primary completion date is November 2017(34)

 

>> References

References - Photodynamic Therapy

References - Photodynamic Therapy

English
  1. Miller JW, Walsh AW, Kramer M, Hasan T, Michaud N, Flotte TJ, Haimovici R, Gragoudas ES. Photodynamic therapy of experimental choroidal neovascularization using lipoprotein-delivered benzoporphyrin. Arch Ophthalmol 1995; 113 (6): 810-8.
     
  2. Kramer M, Miller JW, Michaud N, Moulton RS, Hasan T, Flotte TJ, Gragoudas ES. Liposomal benzoporphyrin derivative verteporfin photodynamic therapy. Selective treatment of choroidal neovascularization in monkeys. Ophthalmology 1996; 103 (3): 427-38.
     
  3. Henderson BW, Dougherty TJ. How does photodynamic therapy work? Photochem Photobiol 1992; 55 (1): 145-57.
  4. Schmidt-Erfurth U, Sacu S; Early Retreatment Study Group. Randomized multicenter trial of more intense and standard early verteporfin treatment of neovascular age-related macular degeneration. Ophthalmology 2008; 115 (1): 134-40.
     
  5. Photodynamic therapy of subfoveal choroidal neovascularization in age-related macular degeneration with verteporfin: one-year results of 2 randomized clinical trials--TAP report. Treatment of age-related macular degeneration with photodynamic therapy (TAP) Study Group. Arch Ophthalmol 1999; 117 (10): 1329-45.
     
  6. Bressler NM; Treatment of Age-Related Macular Degeneration with Photodynamic Therapy (TAP) Study Group. Photodynamic therapy of subfoveal choroidal neovascularization in age-related macular degeneration with verteporfin: two-year results of 2 randomized clinical trials-tap report 2. Arch Ophthalmol 2001; 119 (2): 198-207.
     
  7. Bressler NM, Arnold J, Benchaboune M, Blumenkranz MS, Fish GE, Gragoudas ES, Lewis H, Schmidt-Erfurth U, Slakter JS, Bressler SB, Manos K, Hao Y, Hayes L, Koester J, Reaves A, Strong HA; Treatment of Age-Related Macular Degeneration with Photodynamic Therapy (TAP) Study Group. Verteporfin therapy of subfoveal choroidal neovascularization in patients with age-related macular degeneration: additional information regarding baseline lesion composition’s impact on vision outcomes-TAP report No. 3. Arch Ophthalmol 2002; 120 (11): 1443-54.
     
  8. Blumenkranz MS, Bressler NM, Bressler SB, Donati G, Fish GE, Haynes LA, Lewis H, Miller JW, Monés JM, Potter MJ, Pournaras C, Reaves A, Rosenfeld PJ, Schachat AP, Schmidt-Erfurth U, Sickenburg M, Singerman LJ, Slakter JS, Strong A, Vannier S; Treatment of Age-Related Macular Degeneration with Photodynamic Therapy (TAP) Study Group. Verteporfin therapy for subfoveal choroidal neovascularization in age-related macular degeneration: three-year results of an open-label extension of 2 randomized clinical trials--TAP Report no. 5. Arch Ophthalmol 2002; 120 (10): 1307-14.
     
  9. Bressler NM, Bressler SB, Haynes LA, Hao Y, Kaiser PK, Miller JW, Naor J, Potter MJ, Pournaras CJ, Reaves A, Rosenfeld PJ, Schmidt-Erfurth U, Slakter JS, Strong A, Vannier S. Verteporfin therapy for subfoveal choroidal neovascularization in age-related macular degeneration: four-year results of an open-label extension of 2 randomized clinical trials: TAP Report No. 7. Arch Ophthalmol 2005; 123 (9): 1283-5.
     
  10. Blinder KJ, Bradley S, Bressler NM, Bressler SB, Donati G, Hao Y, Ma C, Menchini U, Miller J, Potter MJ, Pournaras C, Reaves A, Rosenfeld PJ, Strong HA, Stur M, Su XY, Virgili G; Treatment of Age-related Macular Degeneration with Photodynamic Therapy study group; Verteporfin in Photodynamic Therapy study group. Effect of lesion size, visual acuity, and lesion composition on visual acuity change with and without verteporfin therapy for choroidal neovascularization secondary to age-related macular degeneration: TAP and VIP report no. 1. Am J Ophthalmol 2003; 136 (3): 407-18.
     
  11. Verteporfin in Photodynamic Therapy (VIP) Study Group. Verteporfin therapy of subfoveal choroidal neovascularization in age-related macular degeneration: two year results of a randomized clinical trial including lesions with occult with no classic choroidal neovascularization. — Verteporfin in photodynamic therapy report 2. Am J Ophthalmol 2001; 131: 541-60.
     
  12. Azab M, Boyer DS, Bressler NM, Bressler SB, Cihelkova I, Hao Y, Immonen I, Lim JI, Menchini U, Naor J, Potter MJ, Reaves A, Rosenfeld PJ, Slakter JS, Soucek P, Strong HA, Wenkstern A, Su XY, Yang YC; Visudyne in Minimally Classic Choroidal Neovascularization Study Group. Verteporfin therapy of subfoveal minimally classic choroidal neovascularization in age-related macular degeneration: 2-year results of a randomized clinical trial. Arch Ophthalmol 2005; 123 (4): 448-57.Cruess AF, Zlateva G, Pleil AM, Wirostko B. Photodynamic therapy with verteporfin in age-related macular degeneration: a systematic review of efficacy, safety, treatment modifications and pharmacoeconomic properties. Acta Ophthalmol 2009; 87 (2): 118-32.
     
  13. Azab M, Benchaboune M, Blinder KJ, Bressler NM, Bressler SB, Gragoudas ES, Fish GE, Hao Y, Haynes L, Lim JI, Menchini U, Miller JW, Mones J, Potter MJ, Reaves A, Rosenfeld PJ, Strong A, Su XY, Slakter JS, Schmidt-Erfurth U, Sorenson JA; Treatment of Age-Related Macular Degeneration with Photodynamic Therapy (TAP) Study Group; Verteporfin in Photodynamic Therapy (VIP) Study Group. Verteporfin therapy of subfoveal choroidal neovascularization in age-related macular degeneration: meta-analysis of 2-year safety results in three randomized clinical trials: Treatment Of Age-Related Macular Degeneration With Photodynamic Therapy and Verteporfin In Photodynamic Therapy Study Report no. 4. Retina 2004; 24 (1): 1-12.
     
  14. Kaiser PK. Treatment of Age-Related Macular Degeneration with Photodynamic Therapy (TAP) Study Group. Verteporfin therapy of subfoveal choroidal neovascularization in age-related macular degeneration: 5-year results of two randomized clinical trials with an open-label extension: TAP report no. 8. Graefes Arch Clin Exp Ophthalmol 2006; 244 (9): 1132-42.
     
  15. Brown DM, Michels M, Kaiser PK, Heier JS, Sy JP, Ianchulev T; ANCHOR Study Group. Ranibizumab versus verteporfin photodynamic therapy for neovascular age-related macular degeneration: Two-year results of the ANCHOR study. Ophthalmology 2009; 116 (1): 57-65.e5.
     
  16. Brown DM, Kaiser PK, Michels M, Soubrane G, Heier JS, Kim RY, Sy JP, Schneider S; ANCHOR Study Group. Ranibizumab versus verteporfin for neovascular age-related macular degeneration. N Engl J Med 2006; 355 (14): 1432-44.
     
  17. Bressler NM, Chang TS, Fine JT, Dolan CM, Ward J; Anti-VEGF Antibody for the Treatment of Predominantly Classic Choroidal Neovascularization in Age-Related Macular Degeneration (ANCHOR) Research Group. Improved vision-related function after ranibizumab vs photodynamic therapy: a randomized clinical trial. Arch Ophthalmol 2009; 127 (1): 13-21.
     
  18. Antoszyk AN, Tuomi L, Chung CY, Singh A; FOCUS Study Group. Ranibizumab combined with verteporfin photodynamic therapy in neovascular age-related macular degeneration (FOCUS): year 2 results. Am J Ophthalmol 2008; 145 (5): 862-74.
     
  19. MONT BLANC Study Results. 17th Congress of the European Society of Ophthalmology, 2009, Amsterdam, the Netherlands.
     
  20. Schmidt-Erfurth U, Kiss C, Sacu S. The role of choroidal hypoperfusion associated with photodynamic therapy in neovascular age-related macular degeneration and the consequences for combination strategies. Prog Retin Eye Res 2009; 28 (2): 145-54.
     
  21. Ahmadieh H, Taei R, Soheilian M, Riazi-Esfahani M, Karkhaneh R, Lashay A, Azarmina M, Dehghan MH, Moradian S. Single-session photodynamic therapy combined with intravitreal bevacizumab and triamcinolone for neovascular age-related macular degeneration. BMC Ophthalmol 2007; 7:10.
     
  22. Augustin AJ, Puls S, Offermann I. Triple therapy for choroidal neovascularization due to age-related macular degeneration: verteporfin PDT, bevacizumab, and dexamethasone. Retina 2007; 27 (2): 133-40.
     
  23. Liggett PE, Colina J, Chaudhry NA, Tom D, Haffner G. Triple therapy of intravitreal triamcinolone, photodynamic therapy, and pegaptanib sodium for choroidal neovascularization. Am J Ophthalmol 2006; 142 (6): 1072-4.
     
  24. Silva RM, Cachulo ML, Figueira J, de Abreu JR, Cunha-Vaz JG. Chorioretinal anastomosis and photodynamic therapy: a two-year follow-up study. Graefes Arch Clin Exp Ophthalmol 2007; 245 (8): 1131-9.
     
  25. Silva RM, Faria de Abreu JR, Travassos A, Cunha-Vaz JG. Stabilization of visual acuity with photodynamic therapy in eyes with chorioretinal anastomoses. Graefes Arch Clin Exp Ophthalmol 2004; 242 (5): 368-76.
     
  26. Kaiser PK, Boyer DS, Cruess AF, Slakter JS, Pilz S, Weisberger A; DENALI Study Group. Verteporfin plus ranibizumab for choroidal neovascularization in age-related macular degeneration: twelve-month results of the DENALI study. Ophthalmology. 2012 May;119(5):1001-10.
     
  27. Koh A, Lee WK, Chen LJ, Chen SJ, Hashad Y, Kim H, Lai TY, Pilz S, Ruamviboonsuk P, Tokaji E, Weisberger A, Lim TH. EVEREST study: efficacy and safety of verteporfin photodynamic therapy in combination with ranibizumab or alone versus ranibizumab monotherapy in patients with symptomatic macular polypoidal choroidal vasculopathy. Retina. 2012 Sep;32(8):1453-64.
     
  28. Koh AH; Expert PCV Panel, Chen LJ, Chen SJ, Chen Y, Giridhar A, Iida T, Kim H, Yuk Yau Lai T, Lee WK, Li X, Han Lim T, Ruamviboonsuk P, Sharma T, Tang S, Yuzawa M. Polypoidal choroidal vasculopathy: evidence-based guidelines for clinical diagnosis and treatment. Retina. 2013 Apr;33(4):686-716.
     
  29. Leal S, Silva R, Figueira J, Cachulo ML, Pires I, de Abreu JR, Cunha-Vaz JG. Photodynamic therapy with verteporfin in polypoidal choroidal vasculopathy: results after 3 years of follow-up. Retina. 2010 Sep;30(8):1197-205.
     
  30. Oishi, A., Tsujikawa, A., Yamashiro, K., Ooto, S., Tamura, H., Nakanishi, H., Ueda-Arakawa, N., Miyake, M., Akagi-Kurashige, Y., Hata, M., Yoshikawa, M.,Kuroda, Y., Takahashi, A., Yoshimura, N., 2015. One-year result of aflibercept treatment on age-related macular degeneration and predictive factors for visual outcome. Am. J. Ophthalmol. 159, 853e860 e851.
     
  31. Yamamoto, A., Okada, A.A., Kano, M., Koizumi, H., Saito, M., Maruko, I., Sekiryu, T.,Iida, T., 2015. One-year results of intravitreal aflibercept for polypoidal choroidal vasculopathy. Ophthalmology 122, 1866e1872.32- Visual Outcome in Patients With Symptomatic Macular PCV
     
  32. Treated With Either Ranibizumab as Monotherapy or Combined With Verteporfin Photodynamic Therapy. (EVEREST II). . https://clinicaltrials.gov/ct2/show/NCT01846273
     
  33. Aflibercept in Polypoidal Choroidal Vasculopathy (PLANET) (https://clinicaltrials.gov/ct2/show/NCT02120950)
     
  34. Randomized, Double-masked, Sham-controlled Phase 4 Study, Efficacy, Safety, and Tolerability of Intravitreal Aflibercept Monotherapy Compared to Aflibercept With Adjunctive Photodynamic Therapy in Patients With Polypoidal Choroidal Vasculopathy (ATLANTIC). https://clinicaltrials.gov/ct2/show/NCT02495181.

Anti-VEGF in the treatment of AMD

Anti-VEGF in the treatment of AMD

English

Authors:

Paulo Rosa, MD
Gama Pinto Ophthalmology Institute, Lisbon, Portugal.

João P Figueira, MD
Coimbra University Hospital - Coimbra, Portugal

 

Background

Since the begining of the last century much attention has been focused on tumour vascularization.

Warren Lewis(1), in 1922, and Gordon Ide, in 1939(2), had already considered the hypothesis of synthesis of a vascular growth factor by tumour cells.

Synthesis of a vascular growth factor in the retina was proposed for the first time in 1948(3), in diabetic eyes by Isaac Michelson.

In the beginning of 1970, Folkman and his research group demonstrated that tumour growth is directly related to tumour vascularization, which, in turn, depends on the expression of certain growth factors(4).

 

Vascular endothelial growth – VEGF

Harold Dvorak’s group(5), who identified the Tumour Vascular Permeability Factor (VPF), which causes vascular hyperpermeability, in tumour cells from guinea pigs.

In 1989, 3 groups, including Ferrara et al, published articles highlighting a molecule with pro-mitotic properties in endothelial cells.

Ferrara et al. identified this protein in bovines, having named it Vascular Endothelial Growth Factor (VEGF), the term by which it has been known since then(6,7,8).

VEGF-A, a molecule involved in eye diseases such as Age-related Macular Degeneration (AMD) and diabetic retinopathy, is part of a family of genes that also includes VEGF-B, C and D, and the viral homologue VEGF-E, in addition to the Placental Growth Factor – PIGF.

VEGF-A, which has been extensively studied, is a dimeric 36-46 kd glycosylated protein with an N-terminal signal sequence and a heparin-binding domain(9,10).

Four different VEGF-A isoforms have been identified in humans with varying numbers of amino acids: VEGF121, VEGF 165, VEGF 189 and VEGF 206.

They arise from alternative splicing of mRNA.

The longer forms are matrix-bound and the shorter forms are freely diffusible.

VEGF 165 is the dominant isoform in ocular neovascularization processes(11).

 

VEGF receptors

Three VEGF receptors have been identified: VEGFR-1 (fms-like tyrosine kinase-1 or Flt-1), VEGFR-2 (kinase insert domain-containing receptor or KDR) and VEGFR-3 (fms-like tyrosine kinase-4 or Flt-4), which is a receptor for VEGF-C and VEGF-D.

VEGF-A binds both to the R1 and R2 receptors.

VEGFR-2 is considered the main VEGF mediator in endothelial cells.

Its activation induces NO (nitric oxide) production, cell membrane and cytoskeleton reorganisation and proliferation and migration of endothelial cells.

It is also involved in the activation of the phosphatidylinositol 3-kinase (PI3)Akt pathway, which is a crucial signal transduction pathway in the process leading to endothelial cell survival induced by VEGF-A(12).

 

Physiology of VEGF

VEGF-A is an important permeability inducer and is about 50,000 times more potent than histamine.

It is also a potent mitogen in endothelial cells and may have an important role in maturing of new blood vessels through pericytes(13).

VEGF-A is involved in physiological angiogenesis in adults, for example, in the female reproduction cycle(14).

In addition, VEGF-A mRNA is expressed in various healthy human adult tissues that do not show angiogenesis, such as the epithelium of the choroid plexus in the brain, the glomerular epithelium in the kidney, the gastrointestinal mucosa and hair follicles(15).

It has been suggested that VEGF-A maintains the integrity of endothelial cells via anti-apoptotic signalling(15).

VEGF-A has been recognised as an important neuroprotectant in the central nervous system.

VEGF-A exposure resulted in a dose-dependent reduction in retinal neuronal apoptosis(16).

Although mechanistic studies have suggested that VEGF-induced volumetric blood flow to the retina may be partially responsible for neuroprotection, ex vivo retinal cultures have revealed a direct neuroprotective effect for VEGF-A.

VEGF receptor-2 expression has been detected in several neuronal cell layers of the retina, and functional analyses have shown that VEGFR-2 is involved in retinal neuroprotection(16).

It has been shown that VEGF-A is secreted by Retinal Pigment Epithelial (RPE) cells, on their basal side, i.e. the side adjacent to the choriocapillaris, and the 3 VEGFRs are expressed in choriocapillaris endothelial cells, on the side facing retinal pigment epithelial cells.

It has long been known that loss of RPE cells in the human eye causes atrophy of the choriocapillaris.

These findings are consistent with a role of VEGF-A secreted by RPE cells as a permeability/survival factor for quiescent choriocapillaris endothelium(17).

Since VEGF is highly regulated by hypoxia, a feedback mechanism must exist in these epithelia to promote physiological formation of new blood vessels when tissue oxygenation is low.

Unbalances in this mechanism may cause serious diseases, such as Exudative Age-related Macular Degeneration (AMD)(15).

 

VEGF and pathology

The predominant role of VEGF-A in the development of pathological angiogenesis, such as that occurring in tumours and ischaemic and inflammatory processes was widely demonstrated in the last decade(18).

In hypoxic states, VEGF is secreted by RPE cells(19).

This factor induces endothelial cell proliferation and increases vascular permeability.

It has been shown in several models that VEGF-A is required and sufficient for development of new blood vessels in the retina and the iris.

s already mentioned, VEGF-A has been identified as a primordial factor in the neovascular response induced by retinal ischaemia.

Therefore, VEGF-A levels are increased in the vitreous and retina of patients with neovascularization secondary to proliferative diabetic retinopathy, venous occlusion or retinopathy of prematurity(20-23). In clinical practice, observed blood VEGF levels are increased in AMD patients(24).

Many studies have revealed VEGF overexpression in neovascular membranes during autopsy procedures or after surgical extraction(25,26).

Since 1996, immunohistochemistry studies of frozen sections of neovascular membranes have shown significant VEGF levels in highly vascularized regions, although lower immunoreactivity has been observed in fibrotic membrane regions(28,29).

Drusens and basal linear deposits have also been associated with high VEGF levels(30).

Therefore, vascular endothelial growth factor A (VEGF-A) regulates angiogenesis and vascular permeability in the eye, both in physiological and pathological processes.

This growth factor selectively influences endothelial cell growth, being particularly responsible for increased vascular permeability.

It also plays a role in the survival of many cells. Inhibition of neovascularization – the cause of exudative or neovascular AMD – was the basis of some disease-modifying therapies, since anti-VEGFs may delay or even halt disease progression.

The vascular endothelial growth factor is a secreted protein that induces angiogenesis and increases vascular permeability and inflammation, which appear to contribute to neovascular AMD progression. Naturally, VEGF is the target of investigational drugs for the treatment of AMD(31,32).

It is possible to inhibit every step of the angiogenesis cascade induced by VEGF: VEGF synthesis may be inhibited by inhibiting the synthesis of the corresponding mRNA or by inhibiting transcription(33).

The effect of VEGF may also be directly inhibited , by inhibiting protein action.

This is the mechanism used in anti-VEGF therapies(34,35). Angiogenesis may also be inhibited after VEGF binding, as occurs with anecortave acetate and squalamine lactate(36-38).

Treatment of AMD with anti-VEGFs is thus considered to be a turning point since its emergence has allowed a more direct approach to choroidal neovascularization and its selective inhibition.

Therefore, anti-VEGF treatments offer new hope to thousands of neovascular AMD patients, a disease that used to be understood as an untreatable condition associated with ageing before the emergence of anti-VEGF drugs.

These drugs are particularly effective in the early stages of the disease, when newly formed blood vessels are less mature: inhibition of their growth allows photoreceptors to remain viable, as well as reducing the risk of central fibrosis and delaying progressive loss of vision.

Three drugs in this class are currently used in the treatment of AMD: pegaptanib (Macugen®), ranibizumab (Lucentis®) and bevacizumab (Avastin®), of which only the first two have been approved for this therapeutic indication.

 

Ranibizumab (Lucentis®)

Ranibizumab is a Fab fragment of a recombinant humanized monoclonal antibody with high affinity for VEGF-A (the ranibizumab binding site has an affinity for binding VEGF-A 140-fold higher than that displayed by the bevacizumab binding site) specifically studied for the treatment of AMD(39,40,41).

Ranibizumab has a solid clinical development program for this therapeutic indication, involving over 7,000 patients.

Ranibizumab binds to an amino acid chain common to all VEGF-A isoforms, thus rendering them inactive, reducing retinal and choroidal angiogenesis and halting the increase in capillary permeability.

It has been shown in animal models that ranibizumab effectively penetrates the retina and the subretinal space after intravitreal injection.

Its systemic half-life is short (2-3 hours, following intravitreal administration) and systemic clearance is fast, which makes its administration safe.

The average vitreous elimination half-life is approximately10 days(42,43). Ranibizumab has been approved for all types of exudative/neovascular AMD lesions: classic, predominantly classic, minimally classic and occult lesions with no classic component, up to 12 disc areas (DA), where the neovascular component is ≥ 50% of the entire lesion.

The recommended dose is 0.5 mg.

Treatment includes a loading phase, consisting of 3 monthly injections, in the first 3 months, and a maintenance phase, where retreatment is decided according to disease progression, mostly evaluated in monthly visits through VA and OCT criteria, at least during the initial stage or recent neovascularization activity(44).

Phase III clinical trials MARINA and ANCHOR, which supported ranibizumab approval for the treatment of AMD, demonstrated that treatment with monthly intravitreal injections for a 12 months period was associated with a significant increase in visual acuity, compared to photodynamic therapy and placebo(45).

After 12 months, 25-40% of patients treated with ranibizumab showed gains of ≥ 15 letters (ETDRS), compared to 5-6% of the control group patients (p<0.001).

Similar results were confirmed after 2 years.

Both these studies have established ranibizumab as the first therapy not only capable of preventing loss of vision but also of improving vision in a substantial percentage of patients: 33% of the patients treated with ranibizumab in the MARINA study and 41% in the ANCHOR study showed visual gains of at least 15 letters(45, 46,47,48, 49).

Subsequent studies (PIER, SUSTAIN, EXCITE) were aimed at defining flexible and individual dose regimes for the maintenance stage of treatment with ranibizumab, allowing an effective approach to maintaining visual gains, practical in terms of hospital follow-up and with maximum systemic and ocular safety(50,51,52).

These visual gains translate into real benefits for patients.

This effect was evaluated through 3 VFQ-25 sub-scales (near vision, distance vision and vision-related dependency); in fact, patients treated with ranibizumab showed improvements in these 3 sub-scales (MARINA and ANCHOR endpoints).

Specifically regarding dependency, ranibizumab allowed patients to become more independent in their daily activities.

Overall average VFQ scores increased by 4.6 points in the Lucentis® 0.5 group, compared to a 4.4-point decrease observed in the placebo group(53).

The PIER study evaluated an alternative therapeutic regime consisting of monthly injections, in the first 3 months, followed by quarterly injections, corresponding to a total of 6 injections within a year.

After an initial gain of 4.8 letters at month 3, patients treated with ranibizumab had lost an average of 0.2 letters at month 12, whereas patients in the control group lost 16.3 letters.

These results indicate that individual treatment criteria should be adopted during the maintenance stage, allowing an effective approach to maintaining visual gains, as well as allowing follow-up in clinical practice, with maximum systemic and ocular safety.

Vision is expected to be maintained in 90-95% of patients; a minimum gain of 3 lines should be observed in 30-40% of patients treated with ranibizumab(50).

In the EXCITE study, the quarterly treatment regime used in the PIER study (0.3 mg and 0.5 mg) was directly compared with a monthly regime (0.3 mg).

An average increase in VA was observed in all treatment groups during the 12 months of study duration. At month 12, compared to month 3, VA gains had decreased slightly with the quarterly regime (by -2.2 and -3.1 letters with ranibizumab 0.3 mg and 0.5 mg, respectively), having slightly increased (by +0.9 letters) with monthly administration of 0.3 mg of ranibizumab(52).

PrONTO, a small prospective, unicentric, open-label, non-randomized study sponsored by the investigator, evaluated the efficacy of 3 consecutive monthly injections, followed by individual retreatment based on OCT results (at intervals ≥ 1 month).

Retreatment criteria were: loss of 5 letters in VA, presence of fluid in the macula detected by OCT; increase ≥ 100 µm in central retinal thickness (CRT); de novo classic choroidal neovascularization; de novo macular haemorrhage; or persistent macular fluid detected by OCT.

Despite similar VA outcomes to those observed in the MARINA and ANCHOR studies having been observed with a smaller number of intravitreal injections, comparisons are limited by substantial differences in study design.

Although being a small, open-label trial, this study suggests that individual retreatment based on OCT results allows visual gains to be maintained with a smaller number of injections(45,54) .

The SAILOR-cohort 1 study evaluated the efficacy and safety of 3 consecutive monthly injections followed by quarterly monitoring visits, injections according to VA criteria (loss of > 5 letters from the maximum previous VA score) and OCT, if available (increase > 100 µm in CRT from the lowest previous measurement).

Additional visits/injections would take place if required.

Average VA increased from baseline after the first 3 injections, having subsequently decreased to an average gain of 2.3 letters for both ranibizumab doses, a better outcome than that observed for the PIER study, albeit suboptimal compared to those observed in the ANCHOR and MARINA studies.

These results indicate that quarterly visits are not sufficient to monitor and evaluate disease progression(45,55).

The objective of the SUSTAIN study was to evaluate the efficacy of 3 consecutive monthly injections followed by monthly monitoring and treatment according to the following criteria: loss of > 5 letters from the maximum previous VA score, in the first 3 months; or increase > 100 µm in CRT from the lowest previous measurement, in the first 3 months.

It was observed at month 12 that the majority of visual gains achieved in the first 3 months had been maintained.

Although this study consisted only of an interim analysis of 69 patients, the corresponding results suggest that efficacy outcomes may be maintained by a flexible regime with a smaller number of intravitreal injections and monthly monitoring.

However, some VA loss occurred after month 3, whereas fixed monthly injections led to additional VA gains during the maintenance stage(51,56).

In summary, the best VA outcomes were achieved with the monthly regime.

The poorest, albeit variable, efficacy outcomes were observed in studies with < 5 intravitreal injections.

The PrONTO and SUSTAIN studies demonstrated that monthly monitoring is required to maintain efficacy benefits, when compared to the SAILOR-cohort 1 study, which included compulsory quarterly monitoring visits, although more frequent follow-up was performed in many patients.

Therefore, ranibizumab emerges as the first approved neovascular AMD therapy (FDA approval in June 2006) able to improve visual acuity, having thus been recommended as first line therapy by many Ophthalmological Societies (e.g., the Royal College of Ophthalmologists, the German Ophthalmologists Association, etc.) and NICE (National Institute for Health and Clinical Excellence)(57).

Extension study HORIZON was performed in order to evaluate efficacy and safety after the first 2 years.

This study was designed as a post-marketing surveillance to monitor the safety and tolerability of Lucentis®, with a follow-up period of up to 3 years.

HORIZON enrolled 853 patients who had already completed one of the 2-year randomized Lucentis® trials, ANCHOR, MARINA or FOCUS(58,59).

While participating in the ANCHOR, MARINA or FOCUS studies, patients received monthly injections (active treatment with Lucentis® or Visudyne®, or sham).

During the HORIZON study, patients attended fixed quarterly visits; however, visit frequency could be increased by the investigator if they deemed it necessary to see the patient more often.

Lucentis® 0.5 mg injections were given on an as-needed basis, when the investigator felt that the patient would benefit from Lucentis® treatment.

The interval between injections was at least 30 days.

After 2 years (preliminary results), 69% of the 600 initial Lucentis®-treated patients received their injections.

Visual Acuity was available for 384/600 patients.

Among these 384 patients, median Snellen VA had increased by 3 lines, from 20/100 to 20/50, during the initial 2-year trial, having subsequently decreased by 2 lines from the HORIZON baseline to 20/80, at year 2 of the HORIZON study.

Overall, the safety profile of Lucentis® was very good and consistent with previous pivotal clinical trials of Lucentis®.

In general, better VA and anatomical outcomes after the first 2 years delayed the need for subsequent retreatment.

Additionally, the need for early AMD treatment was somewhat confirmed.

ome loss of previously achieved VA gains occurred, eventually related to sub-treatment during the extension period.

Loss of visual acuity and the need for retreatment during the HORIZON study shows that the disease remains active after the first two years of monthly injections, evidencing the need for careful patient monitoring, as well as timely retreatment.

In clinical trials, the benefits of ranibizumab regarding visual acuity were independent of the type of CNV lesion.

Additionally, these benefits were associated with a low rate (< 0.1%) of severe adverse events (endophthalmitis, retinal detachment, traumatic cataract).

Less severe ocular adverse events occurred in less than 2% of patients, including intraocular inflammation and increase in intraocular pressure. In all clinical trials, Lucentis® revealed to be a well-tolerated drug, with no statistically significant differences observed in ocular adverse events between treatment arms.

The results of the SAILOR study suggest a possible increase in the risk of de novo cardio vascular adverse events (CVA) in patients treated with ranibizumab with previous history of CVA or its risk factors (e.g., cardiac arrhythmias), although the differences observed were not statistically significant.

Safety monitoring during the post-marketing period has confirmed the good ocular and systemic safety profile of ranibizumab, whose risk management plan has been strictly implemented.

Other clinical trials are in course for other therapeutic indications, namely Diabetic Macular Oedema, Central Retinal Vein Occlusion and other ocular pathologies involving choroidal neovascularization, whose preliminary results have revealed to be promising.

 

Pegaptanib ( Macugen®)

Pegaptanib sodium (Macugen®, OSI-Eyetech Pharmaceuticals, Pfizer), was the first anti-VEGF inhibitor available for the treatment of choroidal neovascularization(60).

This medicine is part of a new drug set called aptamers.

The aptamers are synthetic oligonucleotides which acquire a specific tridimensional shape and allow high specificity and affinity to a great extent of therapeutic agents.

These compounds are chemically synthetised with the use of nucleotide bases and the use of reverse transcription and PCR - polymerase chain reaction technology(61).

Pegaptanib sodium is a 28-base ribonucleic acid (RNA) oligonucleotide with two branched 20KDa polyethylene glycol (PEG) moieties attached in order to increase the half-life of the drug in the vitreous cavity.

The RNA sugar background is modified to prevent its degradation by endogenous endo and exo-nucleases(62).

Pegaptanib sodium specifically targets the VEGF165 isoform(63).

The pharmacokinetics of pegaptanib following intravitreous injection were profiled in a study of 147 subjects with exsudative AMD (Apte RS, 2007).

Either 1 or 3 mg of pegaptanib sodium per study eye was administered every 6 weeks for 54 weeks.

For the 1 mg dose, mean maximal plasma concentrations were 20 – 24 ng/ml, and pegaptanib was measurable (> 8 ng/ml) in the plasma for up to 1 week after injection.

The mean apparent terminal plasma half-life, determined from the 3 mg group, was 10 days.

There was no plasma accumulation with administration of repeated doses.

In addition, no serum antibodies against pegaptanib were detected(64, 65).

In monkeys’ eyes, biologically active pegaptanib could be detected in the vitreous humor for at least 28 days, following a single 0.5 mg intravitreous injection dose(66).

 

Clinical trials with pegaptanib

Phases I and II studies

A phase IA safety study with 15 patients with exudative AMD(67), as well as a phase II study with 21 patients treated with pegaptanib associated or not to photodynamic treatment (PDT)(68), with a follow up of 3 months, have shown that the intravitreous administration of pegaptanib with 6 week intervals was well tolerated and had anatomic and visual benefits(69).

 

Phase III study

The study VISION (VEGF Inhibition Study in Ocular Neovascularization) consists of two multicentric, randomized, prospective, controlled, dose-ranging and double-blinded phase III clinical trials, used for testing the safety and efficiency of pegaptanib sodium in the treatment of choroidal neovascularization secondary to AMD(70).

There were 1208 patients in this study, distributed by 117 centers and the main criteria for inclusion were: 50-year old or above with any kind of angiographic subtype of subfoveal choroidal neovascularization in the study eye secondary to AMD, with a lesion of 12 or below disc areas (including blood, scarring, atrophy and neovascularization).

The best-corrected visual acuity varied between 20/320 and 20/40.

Patients were randomized in four branches of the study: a group for simulation of pegaptanib intravitreous injections and one of three groups for administration of pegaptanib sodium intravitreous injections (with doses of 0,3 mg, 1mg or 3 mg).

The injections (or simulations) were performed with 6-week intervals for 48 weeks, in a maximum of 8 injections per patient.

All patients underwent the same procedures with exception of the scleral penetration performed in the group of intravitreous injection simulation.

The ophthalmologist performing the injections was not authorized to undertake the patients’ follow up in order to guarantee the researcher’s concealment.

For ethical reasons, treatment with PDT (Visudyne®) was allowed in some clinical centers in patients with mainly classic lesions, in all branches of the study and according to the researcher’s criteria.

The primary study outcome measure was the proportion of patients who lost <15 letters of VA at the end of week 54.

Additional efficacy end-points included: proportion of patients maintaining or gaining > 0, 5, 10, or 15 letters, or losing > 30 letters (severe vision loss); mean changes in VA from baseline to week 54, and the proportion of patients with VA of 20/200 or worse in the study eye at week 54.

In total, 1186 subjects received at least one study treatment (mean, 8.5 of 9 possible injections)(61).

All pegaptanib doses were superior to sham with regard to loss of < 15 letters of VA: 70, 71 and 65% for 0.3 mg (p < 0.001), 1 mg (p < 0.001) and 3 mg (p < 0.03) groups, respectively, versus 55% for sham.

Overall, the 0.3 mg dose was found to be most effective and further discussion is limited to the 0.3 mg (approved) dose.

Pegaptanib was significant superior to sham in the percentage of subjects maintaining or gaining 0, 5, 10 or 15 lines of vision (p <0.05)(71).

Pegaptanib treated subjects were less likely to have severe vision loss (10 versus 22%, p < 0.001) or progress to VA < 20/200 (38 versus 56%; < 0.001).

Mean VA loss at week 54 was 7.95 letters for pegaptanib compared with 15.05 letters for sham
(p < 0.05; 47% relative difference).

Treatment effect was independent of angiographic subtype, baseline VA and lesion size, sex, age, race or iris color(71). VISION trial had an extension for 48 additional weeks.

Those patients receiving pegaptanib were randomized to either continue their pegaptanib dose or discontinue treatment.

Subjects initially receiving sham were rerandomized to continue or discontinue sham or to receive one of the three pegaptanib doses.

Overall, 1053 subjects were rerandomized; 941 (89%) were assessed at week 102 (mean, 15.7 of 17 possible total injections).

Compared with sham (sham over 2 years or randomized to discontinue sham in year 2), more of those receiving pegaptanib 0.3 mg during 2 years lost < 15 letters (45 versus 59%; p < 0.05).

Subjects continuing pegaptanib had the greatest benefits(72).

An exploratory analysis was conducted to assess the vision benefit of treating early subfoveal choroidal neovascularization secondary to AMD with pegaptanib in the VISION trials.

Subjects were grouped according to two different definitions of early disease.

Group 1 included those with lesions < 2 disc areas and a baseline VA of ≥ 54 letters, no prior PDT or laser photocoagulation and scarring or atrophy (n = 34 for pegaptanib 0.3 mg and n = 28 for sham).

Group 2 included those with occult with no classic CNV, with an absence of lipid and worse VA in the study eye versus the fellow eye (n = 30 for pegaptanib 0.3 mg and n = 35 for sham)(70).

At week 54, the responder rates (lost < 15 letters) were significantly higher for pegaptanib versus sham (group 1: 76 versus 50%; p = 0.03; group 2: 80 versus 57%; p = 0.05).

Pegaptanib-treated subjects in group 1 were approximately 10-times less likely to have severe vision loss than those receiving sham (3 versus 29%; p < 0.01); differences for group 2 were not as large (10 versus 17%; p = 0.17).

On average, subjects in both pegaptanib-treated groups lost less VA (group 1: -5.6 versus -16.6 letters; p < 0.01; group 2: -4.0 versus -16.7 letters; p < 0.006).

Notably, among those receiving pegaptanib 0.3 mg 12% of subjects in group 1 and 20% in group 2 gained ≥ 3 lines of vision, compared with 6% in the VISION study.

These findings suggest that pegaptanib treatment early in the course of wet AMD may improve visual outcomes(65, 70).

 

Safety

During the VISION study and the second and third year extension no increased risk of systemic adverse events was identified, but patients with high risk of cardiovascular and cerebrovascular events were excluded from the clinical trials.

Most adverse events reported in the study eyes were attributed to the injection procedure.

The low risk of serious injection-related adverse events, such as endophthalmitis, traumatic cataract and retinal detachment were found to be modifiable with injection protocols changes during the study
(Table 1).

Event
Year 1
n = 7545 injections
Year 2
n = 4091 injections

Year 3
n = 3227 injections

Endophthalmitis 0.16 0.10 0.06
Traumatic cataract 0.07 0.02 0
Retinal detachment 0.08 0.17 0.03

Table 1 - VISION study serious ocular adverse events rates (% per injection)61. Adapted from Rajendra S Apte, 2008.

 

Because VEGF is involved in a wide range of physiological processes, inhibition of this factor raises many safety concerns particularly in the context of extended treatment regimens(75-77).

The pegaptanib sodium selectively inhibits the most biologically active isoform of VEGF (VEGF 165), and according to some authors this quality allows a theoretical advantage in terms of safety comparing to the non-selective anti-VEGF like ranibizumab and bevacizumab.

The systemic risks of non-selective VEGF inhibition have been illustrated with the use of intravenous injection of bevacizumab for the treatment of metastatic colorectal and non-small-cell lung cancer, both approved indications for this agent.

Nevertheless, the intravitreous administration of anti-VEGF agents for the treatment of exudative AMD results in much lower systemic exposures(65).

Although the theoretical superior safety of pegaptanip in comparison to other non‐selective anti‐VEGFs this has not been confirmed yet.

 

Bevacizumab (Avastin®)

Bevacizumab (Avastin®, Genentech, Roche) is a recombinant, humanized, monoclonal immunoglobulin G1 antibody (149 kD) that binds to and inhibits the biologic activity of all isoforms of human VEGF.

This molecule has 2 antigen-binding domains (ranibizumab has 1).

In 2004, the FDA approved bevacizumab for use in patients with metastatic colorectal cancer.

It has received additional approval for use in patients with non–small-cell lung cancer and those with metastatic breast cancer(78-81).

Though not formally studied or approved for any intraocular disease, Rosenfeld’s pioneering work and the unavailability of a related ocular drug, ranibizumab, led to rapid and wide use of bevacizumab all over the world(82,83).

Using bevacizumab as an intravitreal injection to treat neovascular AMD is off-label at this time, however many ophthalmologists, appropriately offer intravitreal bevacizumab to AMD patients based on multiple forms of evidence: results from several retrospective case series, extrapolation from the magnitude of the outcomes reported with ranibizumab, the structural similarity between ranibizumab and bevacizumab, the individual, and the natural history of the disease if left untreated(84).

In the human retina, it is unclear if the molecule of bevacizumab fully distributes within the retinal layers or if localized inhibition of VEGF in the vitreous and inner retina is responsible for the clinical effects associated with administration(85-87).

There are also theories that the larger size of bevacizumab relative to ranibizumab may result in bevacizumab not clearing as quickly from the eye, potentially resulting in longer duration of activity.

To the knowledge of this author, this claim has not been confirmed(84).

Full antibodies generally have longer systemic half-lives than antibody fragments.

Therefore, it is assumed that the half-life of bevacizumab in the eye and in the circulation is longer than that of ranibizumab after intravitreal injection.

Different half-lives for these 2 drugs may have implications for different dosing frequencies and different systemic toxicities(78,86-91).

 

Experimental and clinical studies

Following the initial successful administration of this drug in the management of exudative AMD in May 2005, numerous case series were published illustrating the effectiveness of this treatment in a high proportion of patients(92).

Almost all of the evidence supporting the use on neovascular AMD comes from off-label usage in short-term uncontrolled clinical case series, which suggests that intravitreal administration is apparently locally and systemically well tolerated and is associated with vision stabilization or improvement in most treated eyes(85,86,87,91,94).

One of the earlier large retrospective case series in the literature included 81 consecutive eyes
(79 patients) with subfoveal choroidal neovascularization treated with 1.25 mg (0.05 cc) intravitreal bevacizumab, at baseline and 1 month later if morphologic changes attributable to the CNV persisted (subretinal fluid, pigment epithelial detachment, retinal thickening).

Seventy-eight percent had prior treatment with pegaptanib, photodynamic therapy (PDT), or both.

After one IVB injection, 30 of 81 eyes had resolution of their subretinal fluid. At 2 months, 50% demonstrated resolution of leakage.

The mean best corrected visual acuity (BCVA) improved from 20/200 to 20/125 at week 8
(p < 0.0001)(91).

Spaide et al. in a subsequent study evaluated 266 eyes, 70% of which had prior treatment for exudative AMD (PDT or pegaptanib).

At the 3-month follow-up (data available for 141 patients) 38.3% patients improved by 2 or more Snellen lines.

Mean BCVA improved from 20/184 to 20/109 at 3 months (p<0.001).

Central retinal thickness measured by OCT improved over 3 months from a mean of 340 microns to a mean of 213 microns (p<0.001)(95).

A greater visual acuity effect has been reported in naïve eyes compared to those that have received previous treatment, for example in a study of 50 eyes (48 patients) treated with bevacizumab for exudative AMD found that naïve eyes responded more favorably than previously treated eyes.

Six of the 14 (43%) of naïve eyes gained 3 lines or more of vision versus 17% of eyes that had undergone prior treatment.

The naïve group’s mean visual acuity improved from 20/160 at baseline to 20/63 (p<0.001) at week 24(96).

Such visual acuity gains were not reported with PDT or pegaptanib treatment and were comparable to the results of the phase III studies of ranibizumab.

However, those with longstanding exudative AMD have also been shown to improve with treatment.

One retrospective study in 48 eyes with exudative AMD for 5 months or longer (mean 17.9 months) showed that 25% of those improved at least 3 lines with bevacizumab intravitreous injection after a mean follow-up of 27 weeks(96).

In another prospective case series, Bashshur et al. injected 2.5 mg (0.1ml) of bevacizumab (twice the dose most frequently used) into the vitreous in 17 eyes with wet AMD patients and followed by two additional injections at four-week intervals.

Mean best-corrected visual acuity was 20/252 at baseline and 20/76 at week 12 (P < 0.001).

Mean central subfield retinal thickness also improved between baseline and week 12 in all 17 patients.

No systemic or ocular side effects were noted(85).

 

Safety

Data on the safety of intravitreal bevacizumab are more limited than data on ranibizumab or pegaptanib safety because there are no large, prospective, controlled safety studies with this treatment.

Local side-effects are similar to those found for the other anti-VEGF agents(98).

A safety retrospective study evaluating the side effects of intravitreal bevacizumab reviewed 1265 patients for 12 months, with 92 lost to follow-up.

Ocular complications included seven (0.16%) bacterial endophthalmitis, seven (0.16%) tractional retinal detachments, four (0.09%) uveitis, and a case (0.02%) of rhegmatogenous retinal detachment and another case (0.02%) of retinal detachment and vitreous hemorrhage(99).

In electrophysiological studies no negative side-effects were seen on the retina. In contrast, the results showed a recovery effect on photoreceptors even at the site of the CNV(100).

Most of the in vitro, ex vivo and in vivo experiments excluded short-term negative effects on ocular cells and histology(101,102,103,104,105).

A paper, however, discloses mitochondrial disruption in the inner segment of photoreceptors and apoptosis after high doses of intravitreal bevacizumab in the rabbit eye.

The electrophysiological investigation and light microscopy, in contrast appeared unaltered.

This suggests that potential side-effects on the cellular level cannot be detected with the present diagnostic tools in clinical practice(98,106).

Intravenous use of bevacizumab in patients with colorectal cancer is associated with severe systemic side effects including arterial thromboembolism, gastrointestinal perforation, hemorrhage, hypertensive crisis and nephrotic syndrome.

Initial studies using this therapy intravenously for ocular disease in a healthier population did not find nearly the same risks(107,108).

The dose of intravitreal bevacizumab is much lower (1/400th) of the dose used for intravenous treatment and has not been found to result in unexpected systemic side effects(92).

There are no studies adequately undertaken to identify rare systemic events.

In a 3-month retrospective study of bevacizumab treatment in 266 patients, 1 (0.4%) developed a nonfatal myocardial infarction after the third injection.

Two patients (0.8%) had apparent transient ischemic attacks (diagnosis was not definitive).

There were 2 deaths, one from myocardial infarction.

Nevertheless, that patient was a smoker with a history of emphysema.

It is important to consider, however, that this population (mean age, 80.3 years) is at risk for myocardial infraction regardless of treatment.

Any potential safety concerns remain unknown and waiting for randomized and controlled clinical trials.

 

Discussion

The initial results of intravitreal bevacizumab for exudative AMD led to the acceptance of this off-label therapy by ophthalmologists around the world, assuming, based on case series evidence, that bevacizumab is at least almost as good as ranibizumab with respect to efficacy and safety.

Some ophthalmologists might recommend bevacizumab instead of ranibizumab, even when it is available and affordable to the patient, because of the concerns regarding the treatment costs(84,92).

Intravitreal bevacizumab accounts for more than 50% of all anti-VEGF therapy delivered for exudative AMD in the United States(109).

The National Eye Institute is sponsoring a clinical trial to compare the safety and efficacy betwen bevacizumab and ranibizumab for the treatment of exudative AMD – CATT study.

This study and other prospective, controlled and randomized trials in several countries (IVAN-UK, VIBERA-Germany, MANTA-Austria, LUCAS-Norway, GEFAL-France) will provide the best level of evidence regarding the efficacy and safety of bevacizumab.

Some of these ongoing studies can give consistent information about the necessary dose-ranging and dosing-frequency to control AMD neovascularization.

 

Nice recommendations(57) (National Institute for Health and Clinical Excellence; April 2008)

According to NICE, ranibizumab is the only anti-VEGF recommended for the treatment of Age-related Macular Degeneration (as per the NICE Guidelines, published in 2008).

Differences are clear when comparing the outcomes of clinical programs for both drugs (ranibizumab and pegaptanib).

In clinical trials with ranibizumab, the percentage of patients who gained 15 letters or more was substantially higher, whereas in clinical trials with pegaptanib few patients gained 15 letters or more compared to the control group.

Regarding visual acuity outcomes (expressed as the average number of letters lost or gained by both treatment groups versus the control group), the observed results revealed that ranibizumab leads to statistically significant average gains, whereas pegaptanib only leads to a decrease in the average loss, i.e., ranibizumab is more effective than pegaptanib regarding improvements in visual acuity.

Additionally, no benefits were observed in patients whose treatment with pegaptanib was discontinued after the first year, when compared to patients in the placebo group (VISION study results, published in 2006).

According to NICE, both drugs (ranibizumab and pegaptanib) have demonstrated clinical efficacy in the treatment of exudative AMD, although ranibizumab leads to increased clinical benefits and pegaptanib fails to represent a cost-effective example of healthcare resource use, thus not being recommended in the treatment of AMD.

On the contrary, ranibizumab is referred as an option in the treatment of this condition, providing the following are observed for the treated eye:

  • visual acuity between 6/12 and 6/96
  • no permanent structural damage to the central fovea
  • lesion size less than or equal to 12 disc areas in its greatest linear dimension
  • evidence of recent disease progression (vessel pro liferation, observed in fluorescein angiography, or recent changes in visual acuity).

 

>> References

References - Anti-VEGF in the treatment of AMD

References - Anti-VEGF in the treatment of AMD

English
  1. Lewis WH. Endothelium in tissue cultures. Am J Anat 1922; 30:39-60.
     
  2. Ide A. G. Baker N H. Warren S L. Vascularization of the Brown Pearce rabbit epithelioma transplant as seen in the transparent ear chamber. Am J Roentgenol 1939; 42; 891–899.
     
  3. Michaelson IC. The mode of development of the vascular system of the retina with some observations on its significance for certain retinal disorders. Trans Ophthalmol Soc UK 1948; 68: 137–180.
     
  4. Folkman J. Tumor angiogenesis: therapeutic implications. New England Journal of Medicine 1971; 285: 1182-1186.
     
  5. Senger DR, Galli SJ, Dvorak AM, et al. Tumor cells secrete a vascular permeability factor that promotes accumulation of ascites fluid. Science 1983; 219: 983-5.
     
  6. Ferrara N, Henzel WJ. Pituitary follicular cells secrete a novel heparin-binding growth factor specific for vascular endothelial cells. Biochem Biophys Res Commun 1989; 161: 851-8.
     
  7. Connolly DT, Heuvelman D, Nelson R, et al. Tumor vascular permeability factor stimulates endothelial cell growth and angiogenesis. J Clin Invest 1989; 84: 1470-8.
     
  8. Plouet J, Schilling J, Gospodarowicz D. Isolation and characterization of a newly identified endothelial cell mitogen produced by AtT-20 cells. EMBO J 1989; 8: 3801-6.
     
  9. Plate KH,Warnke PC. Vascular endothelial growth factor. J Neurooncol 1997; 35 (3): 365-72.
     
  10. Ferrara N, Gerber HP, LeCouter J. The biology of VEGF and its receptors. Nat Med 2003; 9 (6): 669-76.
     
  11. Houck KA, Ferrara N, Winer J, Cachianes G, Li B, Leung DW. The vascular endothelial growth factor family: identification of a fourth molecular species and characterization of alternative splicing of RNA. Mol. Endocrinol 1991; 5: 1806–1814.
     
  12. Olsson AK, Dimberg A, Kreuger J, Claesson-Welsh L. VEGF receptor signalling - in control of vascular function. Nat Rev Mol Cell Biol 2006; 7 (5): 359-71.
     
  13. Sheikpranbabu S, Kalishwaralal K, Venkataraman D, Eom SH, Park J, Gurunathan S. Silver nanoparticles inhibit VEGF-and IL-1beta-induced vascular permeability via Src dependent pathway in porcine retinal endothelial cells. J Nanobiotechnology 2009; 7: 8.
     
  14. Allen WR, Gower S, Wilsher S. Immunohistochemical localization of vascular endothelial growth factor (VEGF) and its two receptors (Flt-I and KDR) in the endometrium and placenta of the mare during the oestrous cycle and pregnancy. Reprod Domest Anim 2007; 42 (5): 516-26.
     
  15. Witmer AN, Vrensen GF, Van Noorden CJ, Schlingemann RO. Vascular endothelial growth factors and angiogenesis in eye disease. Prog Retin Eye Res 2003; 22 (1): 1-29.
     
  16. Nishijima K, Ng YS, Zhong L, Bradley J, Schubert W, Jo N, Akita J, Samuelsson SJ, Robinson GS, Adamis AP, Shima DT. Vascular endothelial growth factor-A is a survival factor for retinal neurons and a critical neuroprotectant during the adaptive response to ischemic injury. Am J Pathol 2007; 171 (1): 53-67.
     
  17. Ablonczy Z, Crosson CE. VEGF modulation of retinal pigment epithelium resistance. Exp Eye Res 2007; 85 (6): 762-71. Epub 2007 Aug 24.
     
  18. Takahashi H, Shibuya M. The vascular endothelial growth factor (VEGF)/VEGF receptor system and its role under physiological and pathological conditions. Clin Sci (Lond) 2005; 109 (3): 227-41.
     
  19. Slomiany MG, Rosenzweig SA. Autocrine effects of IGF-I-induced VEGF and IGFBP-3 secretion in retinal pigment epithelial cell line ARPE-19. Am J Physiol Cell Physiol 2004; 287 (3): C746-53. Epub 2004 May 12.
     
  20. Funatsu H, Noma H, Mimura T, Eguchi S, Hori S. Association of vitreous inflammatory factors with diabetic macular edema. Ophthalmology 2009; 116 (1): 73-9.
     
  21. Patel JI, Tombran-Tink J, Hykin PG, Gregor ZJ, Cree IA. Vitreous and aqueous concentrations of proangiogenic, antiangiogenic factors and other cytokines in diabetic retinopathy patients with macular edema: Implications for structural differences in macular profiles. Exp Eye Res 2006; 82 (5): 798-806. Epub 2005 Dec 1.
     
  22. Aiello LP, Avery RL, Arrigg PG, Keyt BA, Jampel HD, Shah ST, Pasquale LR, Thieme H, Iwamoto MA, Park JE, et al. Vascular endothelial growth factor in ocular fluid of patients with diabetic retinopathy and other retinal disorders. N Engl J Med 1994; 331 (22): 1480-7.
     
  23. Adamis, AP, Miller JW, Bernal MT, D’Amico DJ, Folkman J, Yeo TK, and Yeo KT. Increased vascular endothelial growth factor levels in the vitreous of eyes with proliferative diabetic retinopathy. Am J Ophthalmol 1994; 118: 445-450.
     
  24. Tsai DC, Charng MJ, Lee FL, Hsu WM, Chen SJ. Different plasma levels of vascular endothelial growth factor and nitric oxide between patients with choroidal and retinal neovascularisation. Ophthalmologica 2006; 220 (4): 246-51.
     
  25. Matsuoka M, Ogata N, Otsuji T, Nishimura T, Takahashi K, Matsumura M. Expression of pigment epithelium derived factor and vascular endothelial growth factor in choroidal neovascular membranes and polypoidal choroidal vasculopathy. Br J Ophthalmol 2004; 88 (6): 809-15.
     
  26. Tatar O, Adam A, Shinoda K, Stalmans P, Eckardt C, Lüke M, Bartz-Schmidt KU, Grisanti S. Expression of VEGF and PEDF in choroidal neovascular membranes following verteporfin photodynamic therapy. Am J Ophthalmol 2006; 142 (1): 95-104.
     
  27. Lopez PF, Sippy BD, Lambert HM, Thach AB, Hinton DR. Transdifferentiated retinal pigment epithelial cells are immunoreactive for vascular endothelial growth factor in surgically excised age-related macular degeneration-related choroidal neovascular membranes. Invest Ophthalmol Vis Sci 1996; 37 (5): 855-68.
     
  28. Ogata N, Matsushima M, Takada Y, et al. Expression of basic fibroblast growth factor mRNA in developing choroidal neovascularization. Curr Eye Res 1996;15:1008–18.
     
  29. Kvanta A, Algvere PV, Berglin L, Seregard S. Subfoveal fibrovascular membranes in age-related macular degeneration express vascular endothelial growth factor. Invest Ophthalmol Vis Sci 1996; 37 (9): 1929-34.
     
  30. Rudolf M, Winkler B, Aherrahou Z, Doehring L C, P Kaczmarek, Schmidt-Erfurth U. Increased expression of vascular endothelial growth factor associated with accumulation of lipids in Bruch’s membrane of LDL receptor knockout mice. Br J Ophthalmol 2005; 89 (12): 1627–1630.
     
  31. Boulton ME, Cai J, Grant MB. gamma-Secretase: a multifaceted regulator of angiogenesis. J Cell Mol Med 2008; 12 (3): 781-95. Epub 2008 Feb 8.
     
  32. Ferrara N. Role of vascular endothelial growth factor in regulation of physiological angiogenesis. Am J Physiol Cell Physiol 2001; 280 (6): C1358-66.
     
  33. Chappelow AV, Kaiser PK. Neovascular age-related macular degeneration: potential therapies. Drugs 2008; 68 (8): 1029-36.
     
  34. Raig ET, Jones NB, Varker KA, Benniger K, Go MR, Biber JL, Lesinski GB, Carson WE 3rd. VEGF secretion is inhibited by interferon-alpha in several melanoma cell lines. J Interferon Cytokine Res 2008; 28 (9): 553-61.
     
  35. Michels S, Schmidt-Erfurth U, Rosenfeld PJ. Promising new treatments for neovascular age-related macular degeneration. Expert Opin Investig Drugs 2006; 15 (7): 779-93.
     
  36. Hayek S, Scherrer M, Barthelmes D, Fleischhauer JC, Kurz-Levin MM, Menghini M, Helbig H, Sutter FK. First clinical experience with anecortave acetate (Retaane). Klin Monbl Augenheilkd 2007; 224 (4): 279-81.
     
  37. Emerson MV, Lauer AK. Emerging therapies for the treatment of neovascular age-related macular degeneration and diabetic macular edema. BioDrugs 2007; 21 (4): 245-57.
     
  38. Emerson MV, Lauer AK Current and emerging therapies for the treatment of age-related macular degeneration. Clin Ophthalmol 2008; 2 (2): 377-88.
     
  39. Steinbrook R. The price of sight--ranibizumab, bevacizumab, and the treatment of macular degeneration. N Engl J Med 2006; 355 (14): 1409-12.
     
  40. Rosenfeld PJ, Brown DM, Heier JS, et al. Ranibizumab for neovascular age-related macular degeneration. N Engl J Med. 2006;355:1419-1431. Abstract
     
  41. Brown DM, Kaiser PK, Michels M, et al, for the Anchor Study Group. Comparison of ranibizumab and verteporfin photodynamic therapy for neovascular age-related macular degeneration. N Engl J Med. 2006;355:1432-1444.
     
  42. Ferrara N, Damico L, Shams N, Lowman H, Kim R. Development of ranibizumab, an anti-vascular endothelial growth factor antigen binding fragment, as therapy for neovascular age-related macular degeneration. Retina 2006; 26 (8): 859-70.
     
  43. Bakri SJ, Snyder MR, Reid JM, Pulido JS, Ezzat MK, Singh RJ. Pharmacokinetics of intravitreal ranibizumab (Lucentis). Ophthalmology 2007; 114 (12): 2179-82.
     
  44. Rouvas AA, Papakostas TD, Vavvas D, Vergados I, Moschos MM, Kotsolis A, Ladas ID. Intravitreal ranibizumab, intravitreal ranibizumab with PDT, and intravitreal triamcinolone with PDT for the treatment of retinal angiomatous proliferation: a prospective study. Retina 2009; 29 (4): 536-44.
     
  45. Rosenfeld PJ, Rich RM, Lalwani GA. Ranibizumab: Phase III clinical trial results. Ophthalmol Clin North Am 2006; 19 (3): 361-72.
     
  46. Bressler NM, Chang TS, Fine JT, Dolan CM, Ward J; Anti-VEGF Antibody for the Treatment of Predominantly Classic Choroidal Neovascularization in Age-Related Macular Degeneration (ANCHOR) Research Group. Improved vision-related function after ranibizumab vs photodynamic therapy: a randomized clinical trial. Arch Ophthalmol 2009; 127 (1): 13-21.
     
  47. Brown DM, Kaiser PK, Michels M, et al. Ranibizumab versus verteporfin for neovascular age-related macular degeneration. N Engl J Med. 2006;355:1432-44.
     
  48. Rosenfeld PJ, Brown DM, Heier JS, et al. Ranibizumab for neovascular age-related macular degeneration. N Engl J Med. 2006;355:1419-31.
     
  49. Kaiser PK, Blodi BA, Shapiro H, Acharya NR; MARINA Study Group. Angiographic and optical coherence tomographic results of the MARINA study of ranibizumab in neovascular age-related macular degeneration. Ophthalmology 2007; 114 (10): 1868-75.
     
  50. Regillo CD, Brown DM, Abraham P, Yue H, Ianchulev T, Schneider S, Shams N. Randomized, double-masked, sham-controlled trial of ranibizumab for neovascular age-related macular degeneration: PIER Study year 1. Am J Ophthalmol 2008; 145 (2): 239-248.
     
  51. Holz FG, Korobelnik JF, Lanzetta P, Mitchell P, Schmidt-Erfurth U, Wolf S, Markabi S, Schmidli H, Weichselberger A. The effects of a flexible visual acuity-driven ranibizumab treatment regimen in age-related macular degeneration: outcomes of a drug and disease model. Invest Ophthalmol Vis Sci 2010; 51 (1): 405-12.
     
  52. Bolz M, Schmidt-Erfurth U. Ranibizumab EXCITE study: Exploring the value of optical coherence tomography for the management of ranibizumab therapy in age-related macular degeneration. Presented at the 8th EURETINA Congress, Vienna, Austria 22–25 May 2008.
     
  53. Suñer IJ, Kokame GT, Yu E, Ward J, Dolan C, Bressler NM. Responsiveness of NEI VFQ-25 to changes in visual acuity in neovascular AMD: validation studies from two phase 3 clinical trials. Invest Ophthalmol Vis Sci 2009; 50 (8): 3629-35.
     
  54. Lalwani GA, Rosenfeld PJ, Fung AE, Dubovy SR, Michels S, Feuer W, Davis JL, Flynn HW Jr, Esquiabro M. A variable-dosing regimen with intravitreal ranibizumab for neovascular age-related macular degeneration: year 2 of the PrONTO Study. Am J Ophthalmol 2009; 148 (1): 43-58.e1.
     
  55. Dafer RM, Schneck M, Friberg TR, Jay WM. Intravitreal ranibizumab and bevacizumab: a review of risk. Semin Ophthalmol 2007; 22 (3): 201-4.
     
  56. NCT00331864. SUSTAIN: Study of Ranibizumab in Patients with Subfoveal Choroidal Neovascularization Secondary to Age-Related Macular Degeneration. Available at http://www.clinicaltrials.gov/ct2/show/NCT00331864 . Accessed January 18, 2010.
     
  57. NICE Final Guidance. Ranibizumab and pegaptanib for age-related macular degeneration. At www.nice.org.uk.
     
  58. Sadda S. HORIZON extension trial of ranibizumab for neovascular age-related macular degeneration: First-year safety and efficacy results. Presented at the Annual Meeting of the Retina Society, Scottsdale, Ariz., Sept 28, 2008.
     
  59. Antoszyk AN, Tuomi L, Chung CY, Singh A; FOCUS Study Group. Ranibizumab combined with verteporfin photodynamic therapy in neovascular age-related macular degeneration (FOCUS): year 2 results. Am J Ophthalmol 2008; 145 (5): 862-74.
     
  60. Krzystolik MG, Afshari MA, Adamis AP, et al. Prevention of experimental choroidal neovascularization with intravitreal anti-vascular endothelial growth factor antibody fragment. Arch Ophthalmol 2002; 120 (3): 338 -46
     
  61. Gragoudas ES, Adamis AP, Cunningham ET Jr, Feinsod M, Guyer DR. Pegaptanib for neovascular age-related macular degeneration. N Engl J Med 2004; 351 (27): 2805 -16
     
  62. Sivaprasad S. Role of pegaptanib sodium in the treatment of neovascular age-related macular degeneration. Clin Ophthalmol. 2008 Jun; 2(2):339-46
     
  63. Waheed NK, Miller JW. Aptamers, intramers, and vascular endothelial growth factor. Int Ophthalmol Clin. 2004 Summer; 44(3):11-22.
     
  64. Apte RS, Modi M, Masonson H, et al. Pegaptanib 1-year systemic safety results Expert Opin. Pharmacother. (2008) 9(3) 507 from a safety-pharmacokinetic trial in patients with neovascular age-related macular degeneration. Ophthalmology 2007; 114 (9): 702 -12.
     
  65. Apte RS. Pegaptanib sodium for the treatment of age-related macular degeneration. Expert Opin Pharmacother. 2008 Feb; 9(3):499-508.
     
  66. Drolet DW, Nelson J, Tucker CE, et al. Pharmacokinetics and safety of an anti-vascular endothelial growth factor aptamer (NX1838) following injection into the vitreous humor of rhesus monkeys. Pharm Res 2000; 17 (12): 1503 -10.
     
  67. Eyetech Study Group. Preclinical and Phase 1A clinical evaluation of an anti-VEGF pegylated aptamer (EYE001) for the treatment of exudative age-related macular degeneration. Retina 2002; 22 (2): 143 -52.
     
  68. Eyetech Study Group. Anti-vascular endothelial growth factor therapy for subfoveal choroidal neovascularisation secondary to age-related macular degeneration: Phase II study results. Ophthalmology 2003; 110 (5): 979 -86
     
  69. Ruckman J, Green LS, Beeson J, et al. 2 -Fluoropyrimidine RNA-based aptamers to the 165-amino acid form of vascular endothelial growth factor (VEGF165). Inhibition of receptor binding and VEGF-induced vascular permeability through interactions requiring the exon 7-encoded domain. J Biol Chem 1998; 273 (32): 20556 -67.
     
  70. Gonzales CR; VEGF Inhibition Study in Ocular Neovascularization (V.I.S.I.O.N.) Clinical Trial Group. Enhanced efficacy associated with early treatment of neovascular age-related macular degeneration with pegaptanib sodium: an exploratory analysis. Retina. 2005 Oct-Nov; 25(7):815-27.
     
  71. Ng EW, Adamis AP. Targeting angiogenesis, the underlying disorder in neovascular age-related macular degeneration. Can J Ophthalmol 2005 ; 40 (3): 352 -68.
     
  72. Chakravarthy U, Adamis AP, Cunningham ET Jr, et al. Year 2 efficacy results of 2 randomized controlled clinical trials of pegaptanib for neovascular age-related macular degeneration. Ophthalmology 2006; 113 (9): 1508, e1501-25.
     
  73. D’Amico DJ, Masonson HN, Patel M, et al. Pegaptanib sodium for neovascular age-related macular degeneration: two-year safety results of the two prospective, multicenter, controlled clinical trials. Ophthalmology 2006; 113 (6): 992 -1001.
     
  74. Suner IJ. Safety of pegaptanib sodium in age-related macular degeneration (AMD): 3-year results of the VISION trial. Annual Meeting of the American Academy of Ophthalmology, Las Vegas, Nevada; 2006.
     
  75. Ferrara N, Mass RD, Campa C, Kim R. Targeting VEGF-A to treat cancer and age-related macular degeneration. Annu Rev Med 2007; 58: 491 -504.
     
  76. Kamba T, McDonald DM. Mechanisms of adverse effects of anti-VEGF therapy for cancer. Br J Cancer 2007; 96 (12): 1788 -95.
     
  77. Verheul HM, Pinedo HM. Possible molecular mechanisms involved in the toxicity of angiogenesis inhibition. Nat Rev Cancer 2007; 7 (6): 475 -85.
     
  78. Ferrara N, Hillan KJ, Nowotny W. Bevacizumab (Avastin), a humanized anti-VEGF monoclonal antibody for cancer therapy. Biochem Biophys Res Commun. 2005; 333:328–35.
     
  79. Presta LG, Chen H, O’Connor SJ, Chisholm V, Meng YG, Krummen L, et al. Humanization of an anti-vascular endothelial growth factor monoclonal antibody for the therapy of solid tumors and other disorders. Cancer Res. 1997; 57:4593–9.
     
  80. Hurwitz HI, Fehrenbacher L, Hainsworth JD, Heim W, Berlin J, Holmgren E, et al. Bevacizumab plus irnotecan, fluorouracil and leucovorin for metastatic colorectal cancer. N Engl J Med. 2004; 350:2335–42.
     
  81. Yang JC, Harworth L, Sherry RM, Hwu P, Schwartzentruber DJ, Topalian SL, et al. A randomized trial of bevacizumab, an anti-vascular endothelial growth factor antibody, for mewztastatic renal cancer. N Engl J Med. 2003; 349:427–34.
     
  82. Rosenfeld PJ, Mosfeghi AA, Puliafito CA. Optical coherence tomography findings after an intravitreal injection of bevacizumab (avastin) for neovascular age-related macular degeneration. Ophthalmic Surg Lasers Imaging. 2005;36:331–5.
     
  83. Rosenfeld PJ, Fung AE, Puliafito CA. Optical coherence tomography findings after an intravitreal injection of bevacizumab (avastin) for macular edema from central retinal vein occlusion. Ophthalmic Surg Lasers Imaging. 2005;36:336–9.
     
  84. Bressler NM. Antiangiogenic approaches to age-related macular degeneration today. Ophthalmology. 2009 Oct;116(10 Suppl):S15-23.
     
  85. Bashshur ZF, Bazarbachi A, Schakal A, et al. Intravitreal bevacizumab for the management of choroidal neovascularization in age-related macular degeneration. Am J Ophthalmol 2006;142:1–9.
     
  86. Rich RM, Rosenfeld PJ, Puliafito CA, et al. Short-term safety and efficacy of intravitreal bevacizumab (Avastin) for neovascular age-related macular degeneration. Retina 2006;26:495– 511.
     
  87. Bressler SB. Introduction: Understanding the role of angiogenesis and antiangiogenic agents in age-related macular degeneration. Ophthalmology. 2009 Oct;116(10 Suppl):S1-7.
     
  88. Hudson PJ, Souriau C. Engineered antibodies. Nat Med 2003;9:129 –34.
     
  89. Mordenti J, Cuthbertson RA, Ferrara N, et al. Comparisons of the intraocular tissue distribution, pharmacokinetics, and safety of 125I-labeled full-length and Fab antibodies in rhesus monkeys following intravitreal administration. Toxicol Pathol 1999;27:536–44.
     
  90. Mordenti J, Thomsen K, Licko V, et al. Intraocular pharmacokinetics and safety of a humanized monoclonal anti-body in rabbits after intravitreal administration of a solution or a PLGA microsphere formulation. Toxicol Sci 1999;52:101– 6.
     
  91. Avery RL, Pieramici DJ, Rabena MD, et al.  Intravitreal bevacizumab (Avastin) for neovascular age-related macular degeneration. Ophthalmology 2006;113:363–72.
     
  92. Gunther JB, Altaweel MM. Bevacizumab (Avastin) for the treatment of ocular disease. Surv Ophthalmol. 2009 May-Jun;54(3):372-400.
     
  93. Emerson MV, Lauer AK, Flaxel CJ, et al. Intravitreal bevacizumab (Avastin) treatment of neovascular age-related macular degeneration. Retina 2007;27:439–44.
     
  94. Goff MJ, Johnson RN, McDonald HR, et al. Intravitreal bevacizumab for previously treated choroidal neovascularization from age-related macular degeneration. Retina 2007; 27:432–8.
     
  95. Spaide RF, Laud K, Fine JFF, et al. Intravitreal bevacizumab treatment of choroidal neovascularization secondary to age-related macular degeneration. Retina. 2006; 26:383—90
     
  96. Yoganathan P, Deramo VA, Lai JC, et al.  Visual improvement following intravitreal bevacizumab (Avastin) in exudative age-related macular degeneration. Retina. 2006; 26:994—8.
     
  97. Ehrlich R, Weinberger D, Priel E, Axer-Siegel R. Outcome of bevacizumab (avastin) injection in patients with agerelated macular degeneration and low visual acuity. Retina. 2008; 28(9):1302—7.
     
  98. Grisanti S, Ziemssen F. Bevacizumab: off-label use in ophthalmology. Indian J Ophthalmol. 2007 Nov-Dec;55(6):417-20.
     
  99. Wu L, Martinez-Castellanos MA, Quiroz-Mercado H, et al. for the Pan American Collaborative Retina Group (PACORES). Twelve-month safety of intravitreal injections of bevacizumab (Avastin): results of the Pan-American Collaborative Retina Study Group (PACORES). Graefes Arch Clin Exp Ophthalmol. 2008; 246(1):81—7.
     
  100. Moschos MM, Brouzas D, Apostolopoulos M, Koutsandrea C, Loukianou E, Moschos M. Intravitreal use of bevacizumab (Avastin) for choroidal neovascularization due to ARMD: A preliminary multifocal-ERG and OCT study: Multifocal-ERG after use of bevacizumab in ARMD. Doc Ophthalmol. 2007;114:37–44.
     
  101. Spitzer MS, Yoeruek E, Sierra A, Wallenfels-Thilo B, Schraermeyer U, Spitzer B, Bartz-Schmidt KU, Szurman P. Comparative antiproliferative and cytotoxic profile of bevacizumab (Avastin), pegaptanib (Macugen) and ranibizumab (Lucentis) on different ocular cells. Graefes Arch Clin Exp Ophthalmol. 2007 Dec; 245(12):1837-42. Epub 2007 Mar 9.
     
  102. Luke M, Warga M, Ziemssen F, Gelisken F, Grisanti S, Schneider T, et al. Effects of bevacizumab on retinal function in isolated vertebrate retina. Br J Ophthalmol. 2006;90:1178–82.
     
  103. Manzano RP, Peyman GA, Khan P, Kivilcim M. Testing intravitreal toxicity of bevacizumab (Avastin). Retina. 2006;26:257–61.
     
  104. Luthra S, Narayanan R, Marques LE, Chwa M, Kim DW, Dong J, et al. Evaluation of in vitro effects of bevacizumab (Avastin) on retinal pigment epithelial, neurosensory retinal and microvascular endothelial cells. Retina. 2006;26:512–8.
     
  105. Feiner L, Barr EE, Shui YB, Holekamp NM, Brantley MA., Jr Safety of intravitreal injection of bevacizumab in rabbit eyes. Retina. 2006;26:882–8.
     
  106. Inan UU, Avci B, Kusbeci T, Kaderli B, Avci R, Temel SG. Preclinical safety evaluation of Intravitreal injection of full-length humanized vascular endothelial growth factor antibody in rabbit eyes. Invest Ophthalmol Vis Sci. 2007; 48:1773–81.
     
  107. Michels S, Rosenfeld PJ, Puliafito CA, et al. Systemic bevacizumab (Avastin) therapy for neovascular age-related macular degeneration twelve-week results of an uncontrolled open-label clinical study. Ophthalmology. 2005; 112(6):1035—47.
     
  108. Moshfeghi AA, Rosenfeld PJ, Puliafito CA, et al. Systemic bevacizumab (Avastin) therapy for neovascular age-related macular degeneration: twenty-four-week results of an uncontrolled open-label clinical study. Ophthalmology. 2006;113(11):1—12.
     
  109. American Society of Retina Specialists. Preferences And Trends Survey. Palm Springs, CA 2007.

Combined Treatment

Combined Treatment

English

Authors:

Ana Fernandes Fonseca
Instituto Oftalmológico Gama Pinto, Lisbon, Portugal 
ALM, Lisbon, Portugal

Mário Guitana, MD
Portuguese Red Cross Hospital, Lisbon, Portugal
 
Victor Ágoas, MD
Instituto Oftalmológico Gama Pinto, Lisbon, Portugal 
 
Teresa Luísa Quintão, MD
Santa Casa da Misericórdia de Lisboa, Lisbon, Portugal
IRL-Instituto de Retina de Lisboa, Lisnon, Portugal
 
José Henriques, MD
Instituto Oftalmológico Gama Pinto, Lisbon, Portugal
IRL-Instituto de Retina de Lisboa, Lisnon, Portugal

 

Updated/reviewed by the authors, July 2017.

1. Introduction

Choroidal neovascularization in AMD has become a serious medical and social problem. One of the reasons for this is ageing of the population. However, a better understanding of this disease and the emergence of new treatment options have been witnessed in recent years.

In clinical practice, combination treatments are necessary when a disease is not well properly controlled with a single therapeutic modality. Using the anti-VEGF gold standard therapy to control exsudative AMD is not enough because it does not structurally change the neovascular membrane, i.e., it does not result in its regression. Combined therapy is the logical step to counteract disease progression mechanisms which are self-supportive once initiated.

We can combine(1,2) anti-VEGF drugs with corticosteroids, verteporfin photodynamic therapy and anti-pericyte agents (targeting vascular endothelial growth factors, angiogenesis-regulating cytokines, proliferating endothelial cells and pericytes, respectively), with the common goal to achieve a synergistic action, with improved outcomes, reduced retreatment frequency and more sustained effects. This additive effect allows patients to be treated with lower doses, entailing added value through increased tolerability and decreased costs.(3).

 

2. Health economics in AMD treatment: efficacy versus efficiency and the importance of equity

2.1 Seeking efficacy and efficiency – resource saving

The efficacy of a given drug or technique is assessed when comparative studies of visual outcomes are performed. 

If the observed outcomes are identical to those observed in previous studies it is concluded that no apparent advantages result from using the technique or drug in question. 

However, if the number of treatment sessions decreases, a smaller number of medicine vials is used or patients visit the hospital less frequently, it is concluded that the efficiency of the drug or method is greater. 

Drugs and methods that are equivalent in terms of efficacy may vary widely in terms of efficiency. We are thus faced with efficiency gains and better use of resources – more patients are treated with the same budget.

Ophthalmologists have been researching for the more efficacious and efficient treatment regime that allows reducing the therapeutic burden and customize the treatment.(80)

This theoretical improvement in clinical efficiency has the advantage of reducing the number of treatment sessions, with a consequent decrease in the drugs cost used to treat each patient. 

Since each patient will make less visits to the hospital, the number of medical, nursing and technical staff hours required will decrease, the same occurring for equipment operation times and time spent at hospital premises.(4)

Different studies refer that the cost of drugs represents the greatest percentage of AMD treatment costs, as opposed to usual health cost distribution, where the largest percentage, of approximately 40%, corresponds to staff costs. In economic studies about the costs of AMD treatment with the ranibizumab protocol, 83% of costs were associated to the drug.(5,6).

 

2.2 Combined treatments – visual outcomes and efficiency results

It has been proved in a randomized prospective trial that some combination therapies for wet macular degeneration produced better visual results (statistically significant superior efficacy) than monotherapy with anti-VEGF agent.(45,46).

In some other studies, it has been showed that some combination therapies are comparatively effective and cost-effective, with considerably less treatments needed than in monotherapy studies.(7,64)

Calculated costs of 1 year treatment per line of visual acuity were $84 for a regimen of treatment when necessary (PRN) with bevacizumab compared to $766 for the gold standard protocol treatment with ranibizumab. Combined treatment costs varied between $71 and $269.(5,6)

Due to their synergistic effect, combined treatments potentially lead to a decrease in the number of retreatment sessions, as well as sustained long-term visual benefits(1), and better clinical efficiency.

 

3. Synergistic action and increased treatment effect

How can we explain the fact that a synergistic effect is theoretically achieved by using various mechanisms of action, sometimes more effective than the sum of their separate effects?(Figure 1) 

4.3.8 Combination therapy with ICON-1 and ranibizumab

ICON-1 is an anti-Tissue Factor (TF) immunoconjugate protein that binds to pathologic vessels overexpressing TF and acts via a new mechanism of action that can eliminate abnormal CNV as well as inhibit the exudation. 

EMERGE(76-77) is a phase 2, randomized, double masked, active control study in the United States that examines the hypothesis that ICON-1 eliminates abnormal CNV as well as inhibits the exudation either alone or in combination with ranibizumab as compared to ranibizumab alone.

A total of 90 patients with treatment naïve (in the study eye) CNV secondary to AMD are being enrolled.

Patients are randomized in a 1:1:1 ratio to receive intravitreal injections of ICON-1 (0.3 mg) as monotherapy (n=30) or in combination with ranibizumab 0.5 mg (n=30) or ranibizumab 0.5 mg monotherapy (n=30). Patients will receive 3 initial monthly injections followed by maximum 3 additional monthly injections based on protocol re-treatment criteria, for a total of 6 months of treatment.

The primary outcomes are mean change from baseline in BCVA letter score and in Central Retinal Thickness (CRT) at 3 months.

Figure 1. Angiogenesis - New blood vessels are formed in response to various physiological and/or pathological stimuli, of which hypoxia is one of the most relevant. Hypoxia activates multiple cellular response cascades, with special emphasis on activation of extracellular matrix metalloproteases and increased synthesis and release of growth factors, including VEGF. The latter acts on membrane receptors, activating intracellular enzyme pathways through intracellular signalling, which results in response amplification. This ultimately leads to cellular proliferation, migration and differentiation, with formation of new blood vessels. Different drug categories act on different stages of the neovascularization process. Joint action on various cascade levels should theoretically lead to an increase in treatment effect and/or a decrease in the effective dose and/or a more prolonged effect.

 

3.1 Anti-VEGFs

The primary need is to act on the key mechanism of the neovascularization process – VEGF. By acting on this mechanism not only do we inhibit neovascularization but we also act on oedema and inflammatory mechanism, to a certain extent.(9).

Several studies – CATT(10), SECURE(11), HORIZON(12), SEVEN YEAR update in ANCHOR/MARINA(13) – showed that antiangiogenic therapy does not cause neovascular network regression and that mean visual acuity tends to return to baseline when therapy is discontinued.

In fact, anti-VEGF therapy is effective at inhibiting tip cells because these are not protected by pericytes that envelop the already formed neovascular network.(14)

Tip cells and tip cell membrane projections called filopodia, "the fingers that do the walking", "lead the way"(15) aare crucial for neovascularization sprouting. Tip cells produce PDGF-B, a growth factor which stimulates pericyte recruitment to line the neovascular network(16), which means that angiogenesis progression does not depend solely on cell proliferation, but it also requires the formation of a pericytic reinforcement to stabilize the neovascular network. This summarily explains the anti-VEGF resistance described in the literature, because when therapy is discontinued the tip cells become active and grow again(87).

Therefore, additional therapy is needed.

 

3.2 Synergistic action of corticoids

When steroids are added, a synergistic action is achieved, since steroids act on various levels of the inflammatory process and angiogenesis regulation.

Various mechanisms of action are proposed for steroids in AMD treatment.(79).

It is known that steroids act on local inflammatory mediators, stabilizing blood-retinal barrier function by increasing gap junction density and activity in capillary endothelial cells. 

It is thought that triamcinolone decreases VEGF, which is a potent agent in increasing capillary permeability by increasing phosphorylation of proteins involved in intercellular tight-junctions, such as occludin and Zonula Occludens-1 (ZO-1). 

These agents also have an anti-inflammatory effect by inhibiting phospholipase A2, an enzyme that metabolizes cell membrane phospholipids to free arachidonic acid, which, in turn, originates thromboxane, leukotrienes and prostaglandins that cause vasodilatation, increased permeability and oedema. 

They also have an angiostatic effect by promoting a decrease in extracellular matrix (ECM) turnover through inhibition of plasmin activation.

Plasmin activates matrix collagenases and metalloproteinases (MMP’s) that dissolve the capillary basement membrane and trigger angiogenesis, with endothelial cell differentiation, migration and proliferation (Figure 2).

combined_treatmnet_image_2_0.png

Figure 2. Corticoids inhibit MMP’s (metalloproteinases) activity.

 

These agents also act on the interaction between ICAM-1 (Intercellular Adhesion Molecule-1) and leukocytes, inhibiting recruitment of the latter, thereby contributing to reduce the inflammatory component.
It is also thought steroids may act on SDF-1 (Stromal-cell Derived Factor-1), inhibiting its action (Figure 3)(8,17).

combined_treatmnet_image_3_0.png

Figure 3. Corticoids decrease VEGF, SDF - 1, ICAM - 1 and VCAM - 1 activity.

 

Steroids also decrease the expression of Major Histocompatibility Complex Class II (MHC-II) molecules involved in the inflammatory process.(18)

Therefore, we have scientific grounds supporting the combined action of treatment with corticoids (Figure 4).

combined_treatmnet_image_4_1.png

Figure 4. Corticosteroids inhibit angiogenesis and macular edema through multiple mechanisms of action. See text.

 

3.3 PDGF-B inhibition 

Inhibiting PDGF-B facilitates the action of anti-VEGF therapies by inducing pericyte stripping, thereby allowing anatomical regression of the neovascular membrane (see lower section 4.4.1 on Fovista).

 

3.4 Associated FGF-2 inhibition and PEDF action

It is also possible to inhibit other factors, such as FGF-2, or induce PEDF locally, which has antiangiogenic effects that counteract the angiogenic effect of VEGF.(15-17).

 

3.5 Acting on the structural level by damaging newly formed blood vessels – PDT

The actions already referred involve blocking or inhibiting neovascularization and inflammatory mediators.

However, it is also possible to act on a structural level, by damaging newly formed blood vessels.

This is achieved through cellular damage and death mediated by free radicals induced by the action of laser on a photosensitizing agent – verteporfin.(18,19).

Standard fluence values of 50-J/cm² or lower fluence values of 25-J/cm² or 12-J/cm² are normally used in combined treatment.

Photodynamic therapy causes vascular occlusion but is associated to an inflammatory response that may be minimized by using corticoids and an anti-VEGF agent. 

Both agents may also inhibit the angiogenic stimulus represented by a VEGF rebound effect following occlusion of new blood vessels.(27,78).

Capillary occlusion induced by PDT leads to hypoperfusion of the treated area, a condition that is theoretically worsened by concomitant use of anti-VEGF agents that prevent recapillarization.

This effect has not been shown as negative; on the contrary, it appears that this recapillarization delay promotes neuronal recovery by decreasing oxygen and free radical concentrations.(27,78).

 

3.6 Other sites of action – associated surgical therapy

We shall not elaborate on this treatment combination, as it will be referred in a chapter dedicated to surgery.

 

4. Main studies

4.1 Gold standard treatment

The efficacy and safety of combined treatments are evaluated in studies where drug combinations are compared with the gold standard treatment, which, according to the results of the MARINA(28) and ANCHOR(29) studies, consists of twelve consecutive monthly intravitreal injections of an antiangiogenic agent.

Use of this treatment regime led to outcomes of 90% in vision stabilization and approximately 30-40% of significant improvement after one year.

 

4.2 Combined treatments: double and triple treatments

Most combined treatments that have been used include reduced-fluence PDT associated to an anti-VEGF agent. PDT, antiangiogenic agents and steroids are often used in triple treatments.(31,82)

Combined therapy with PDT and anti-VEGF therapy seems to have a valid role, as suggested by the RADICAL(35) study results.

Dexamethasone seems to be preferable to triamcinolone since there is a decreased risk of ocular tension increase.(31) .

There are no precise guidelines on triple therapy indications, on dosage and type of drugs. 

Some combined-treatment studies should be referred for their relevance, albeit in a summarized manner.

 

4.2.1 SUMMIT Trial programme

This program includes three large randomized clinical trials, DENALI (USA), MONT BLANC (Europe) and EVEREST (Asia), whose objective is to evaluate the efficacy and safety of combining PDT (Visudyne®) and ranibizumab, compared to monotherapy with this antiangiogenic agent, in patients with neovascular AMD and polypoidal disease (EVEREST). 

Analysis results from the MONT BLANC(32) 

study at twelve months have shown no significant differences between the two groups, although the number of treatments required is slightly lower with the combined treatment. 

The DENALI(33) study had a similar result. The overall benefit for patients was a reduced frequency of Lucentis to at least three months during the study.

In the EVEREST(34) study, in the patients with polypoidal vasculopathy, the combined treatment was more effective than monotherapy to promote regression of polypoid formations.

 

4.2.2 RADICAL Phase 2 study (Reduced Fluence Visudyne-Anti-VEGF-Dexamethasone In Combination for AMD Lesions)

In this study, three Visudyne–Lucentis combination therapies were evaluated against Lucentis monotherapy.

Two-year, phase 2 final 24 months results(35). showed that the combination approach led to significantly fewer visits at which re-treatments were applied. Of the four treatment groups, the triple therapy half-fluence group had the fewest retreatment visits compared with Lucentis monotherapy.

Through 24 months, patients in the triple therapy half-fluence group had a mean of 4.2 retreatment visits compared with 8.9 for patients who received Lucentis monotherapy (P<.001).

More recently, Gallemore et al. conducted a phase II randomized trial and and verified that combination therapy resulted in significantly fewer retreatment visits than a ranibizumab monotherapy regimen at months 12 and 24.(81)

 

4.2.3 PDEX II study

This was a prospective, multicentric, randomized, non-inferiority study comparing the relative advantages of treatment with PDT in a reduced dose (reduced fluence), dexamethasone and ranibizumab versus monotherapy with ranibizumab.(37)

Trial results were presented at the American Society of Retina Specialists meeting 2008, which compared combination therapy using reduced fluence PDT, intravitreal dexamethasone 500 mcg and intravitreal ranibizumab 0.5 mg, to monthly ranibizumab.

Patients receiving the triple therapy needed fewer treatments than 12 per year.

 

4.2.4  LuceDex study

This study(36researched the role of dexamethasone in neovascular AMD treatment. 

This is a prospective, randomized clinical trial comparing two treatment groups, one treated with a combination of ranibizumab and dexamethasone and the other undergoing ranibizumab monotherapy. 

The LuceDex pilot study suggested a possible benefit of adding intravitreal dexamethasone to treatment of neovascular age-related macular degeneration with intravitreal ranibizumab. 

Choroidal neovascular membrane size decreased in Group 1 significantly compared with Group 2 (P < 0.05). 

A larger study is needed to further identify and define possible benefits of this combination therapy.

 

4.2.5 Cave study

In the multicenter CAVE study(38,39), reduced-fluence PDT plus bevacizumab plus triamcinolone therapy was compared with bevacizumab monotherapy, as well as with PDT plus bevacizumab dual therapy. Data from 103 patients were evaluated at one year. 

Mean retreatment rates were highest with bevacizumab monotherapy (4.57) and lowest with triple therapy (3.18). 

There were no significant between-group differences in mean VA gains. Interestingly, the greatest extension in treatment-free interval was with dual therapy, leading the investigators to conclude that addition of triamcinolone did not produce additional benefit.

 

4.2.6 Anti-Inflammatory treatment (Dexamethasone implant in combination with ranibizumab)

Six-month trial of 243 patients with CNV secondary to AMD, adjunctive therapy with the dexamethasone implant (Ozurdex) delayed the time to as-needed injections of 500 µg ranibizumab and reduced the need for repeated ranibizumab injections. 

DEX Implant delayed the time to as-needed injection of ranibizumaband reduced the need for repeated ranibizumab treatment in patients with CNV secondary to AMD.(40)

 

4.2.7 Anti-Inflammatory treatment. Pilot Study (Intravitreal ranibizumab and topical bromfenac)

This was a single-site, multiinvestigator, prospective, open-label, interventional, phase II study of patients with new or recurrent exudative/neovascular age-related macular degeneration. 

Thirty eyes were enrolled consecutively and were randomized in a ratio of 2:1 to combination therapy with intravitreal ranibizumab and topical bromfenac, and ranibizumab alone. 

This pilot study(41) was the first to prospectively identify a biologic signal that may indicate combination therapy with an easily administered well-tolerated eyedrop and ranibizumab is efficacious for the treatment of neovascular age-related macular degeneration.

 

4.2.8 CABERNET study

The phase 3, multicenter, prospective, randomized CABERNET(42) ((CNV secondary to AMD treated with beta radiation epiretinal therapy) study included a treatment arm of 302 patients receiving strontium-90 epimacular brachytherapy (NeoVistaand two mandatory ranibizumab injections, followed by as-needed ranibizumab injections. 

The control arm consisted of 155 patients receiving ranibizumab on a modified PIER protocol with 10 mandatory injections.

The CABERNET study did not achieve its endpoint with a 10% non-inferiority margin. 

 

4.2.9 Combination therapy with low-dose transpupillary thermotherapy and intravitreal ranibizumab

Treatment with low-dose TTT reduced the number or intravitreal injections of ranibizumab over 24 months. 

The results suggested that low-dose TTT can serve as an adjuvant in combination with intravitreal ranibizumab for neovascular AMD(43).

 

4.2.10 ATLANTIC Study

ATLANTIC study (Protocol ECR-AMD-2015-09) is a still running randomized, double-masked, sham-controlled phase 4 study of the efficacy, safety, and tolerability of intravitreal aflibercept monotherapy compared to aflibercept with adjunctive photodynamic therapy in patients with polypoidal choroidal vasculopathy.

 

4.2.11 VIA study

The objective of this study(7) was to determine whether a combination of a reduced dose of PDT and bevacizumab leads to a decrease in the number of treatment sessions required within a 6-month period, compared to monotherapy with bevacizumab.

This randomized, double-blind, controlled clinical trial revealed that a combination of bevacizumab and 25-J/cm² or 12-J/cm² PDT led to a decrease of approximately 50% in the number of treatment sessions required within a 6-month period. 

Favourable outcomes were also observed for visual acuity, although evaluation of this parameter was not the main objective of this study.

Other studies(83-85) confirm the favourable outcome and a consistent reduction of the number of injections, including a study using intravitreal pegaptanib and low-fluence PDT.

 

4.2.12 PDT combined with vitrectomy and dexamethasone

A combined pharmacosurgical intravitreal procedure was performed 24 to 36 hours after PDT, consisting of 23-gauge core vitrectomy and intravitreal substitution with BSS, dexamethasone and bevacizumab. 

The intravitreal retreatment rate was low (13/52) for this safe pharmaco-surgical regimen, corresponding to 25%.(44)

 

4.2.13 Triple-Combination Therapy and Zeaxanthin

Reports of triple-combination therapy for wet AMD suggest a benefit, as do reports for zeaxanthin.

An interventional comparative study(64) was undertaken to evaluate the efficacy of triple combination therapy with and without zeaxanthin, as well as the economic viability of the therapies.

Cases of 543 consecutive eyes with subfoveal CNV secondary to AMD were reviewed. 

All eyes were treated with triple-combination therapy consisting of reduced-fluence PDT, intravitreal bevacizumab and intravitreal dexamethasone. Therapy was repeated as necessary. 

One cohort was also given supplementation with 20 mg of oral zeaxanthin (Zx) daily. 

The triple-therapy group without Zx received a mean of 2.8 treatment cycles, and 87 percent of subjects had stable or improved vision at 24 months. 

In the triple-therapy group with Zx, the mean number of treatment cycles was 2.1, with 83 percent of people having stable or improved vision at 24 months. 

Also at 24 months, CNV developed in 12.5 percent of fellow eyes treated with triple therapy alone and in 6.25 percent of eyes undergoing triple therapy with Zx (p=0.03). 

Triple therapy was comparatively effective for neovascular AMD, since considerably less treatment was needed than they said has been reported in monotherapy studies. 

The addition of oral Zx appeared to further reduce treatment cycles required and risk of CNV development in fellow eyes.

 

4.3 Other promising forms of combined therapy

4.3.1 FovistaTM (pegpleranib)

In pre-clinical models, combination of anti-PDGF and anti-VEGF successfully induced neovascular regression (fig.5).(21)

Figure 5. Combination of anti-PDGF and anti-VEGF induced angiogenic vessel regression in choroidal neovascular model. Preclinical model of the combined treatment.

 

Fovista is a 50 kD DNA anti-platelet-derived growth factor aptamer (directed against PDGF-B subunit B) that strongly binds to PDGF-B, and strips pericytes from mature neovascular vessel, leaving them more susceptible to treatment with an anti-VEGF agent. 

In dec 2016 Ophthotech completed two large, prospective, randomized, controlled Phase 3 clinical trials of 1248 patients with wet AMD. 

The goal of the study - a superiority trial- was to assess the safety and efficacy of a combination of the FovistaTM plus ranibizumab compared with ranibizumab monotherapy in patients with wet AMD. 

Ophthotech announced that the "pre-specified primary endpoint of mean change in visual acuity at 12 months was not achieved in its two Phase 3 clinical trials"(88-90)

The study goes on the second year. While these results are disappointing, it remains to be seen if Fovista, in combination with the Anti-VEGF might be able to reduce the number of required injections and visits, or result in a better or more stable vision in the long term.

In the phase 3 trials, the anti-PDGF and anti-VEGF drugs were administered 30 minutes apart. However, in a small investigator-sponsored trial, Pravin Dugel administered the anti-PDGF a day before he injected the anti-VEGF. 

This protocol which involved 30 patients with AMD, 27 of whom were treatment-resistant and who had persistent or recurrent fluid and no improvement in visual acuity and demonstrated encouraging results.(92)  endpoint with a 10% non-inferiority margin. 

Two simultaneous clinical trials have been in progress with the drugs Avastin and Eylea(91), and the complete the results have yet to be released.

 

4.3.2 FGF-2 inhibition

RPE from CNV patients expresses angiogenic growth factors whose action is partly independent from VEGF. 

In a study(51), Sthal concluded that anti-VEGF treatment (bevacizumab) inactivated all RPE-derived VEGF in a 3D collagen matrix culture of RPE isolated from surgically excised CNV-membranes (CNV-RPE) used to stimulate sprouting of endothelial cell (EC) spheroids, but was unable to fully inhibit EC sprouting induced by CNV-RPE. Combined anti-VEGF/anti-FGF treatment inactivated both growth factors and reduced EC sprouting significantly.

In a comparison between the antiangiogenic effect of solitary anti-VEGF antibodies and combination treatment with anti-VEGF and anti-FGF-2 antibodies, greater inhibition was achieved for the latter. 

Targeted combined therapy can be better than anti-VEGF monotherapy.

 

4.3.3 The balancing effect of PEDF and its delivery

In AMD, pigment epithelium-derived factor (PEDF) is significantly lower in RPE cells, the RPE basal lamina, Bruch’s membrane and choroidal stroma. These data suggest the existence of a critical balance between PEDF and VEGF and the hypothesis that PEDF may be able to counteract the angiogenic potential of VEGF. 

A decrease in PEDF may disrupt this balance and CNV in AMD(52).

Results from a phase I clinical trial of intravitreal administration of an IE4-deleted adenoviral vector expressing human pigment epithelium-derived factor (AdPEDF.11) suggest that antiangiogenic activity may be sustained for several months after single intravitreal injection of AdPEDF.11(53).

This study provided evidence that adenoviral vector-mediated ocular gene transfer might be a viable approach in the treatment of ocular disorders. However, since the publication of phase I results, there is no evidence of further studies on AdPEDF.11 in the treatment of patients with wet AMD. 

 

4.3.4 Ranibizumab and tyrosine kinase inhibitors

In two cases of recurrent wet AMD in which intravitreal ranibizumab was used in combination with oral sorafenib (a tyrosine kinase inhibitor approved for the treatment of some kidney, live rand thyroid carcinomas) improvements were observed in optical coherence tomography(54), indicating this could also be a promising combination treatment.

Following promising preclinical experiments and a phase 1 trial, X-82 (Tyrogenex) is being investigated in a double-masked, placebo-controlled phase 2 study as an oral treatment for patients with wet AMD.(71)

X-82 is a small-molecule tyrosine kinase inhibitor derived from sunitinib (Sutent, Pfizer), a drug that is approved for the treatment of advanced renal cell carcinoma, gastrointestinal stromal tumors, and pancreatic neuroendocrine tumors. Compared with sunitinib, X-82 has a shorter half-life and accumulates less in tissues, it inhibits VEGF and PDGF receptors without the toxicities associated with sunitinib.

X-82 blocks kinase activity associated with all receptor subtypes for VEGF and PDGF, and its potency against all receptor subtypes is thought to provide targeted inhibition at a relatively low dose with reduced risk of side effects.(72) 

Oral X-82 provides dual inhibition of VEGF and PDGF via systemic distribution, which may allow it to efficiently target bilateral disease with fewer injections.

A phase 1 trial showed that treatment-refractory and treatment-naïve patients maintained or improved visual acuity and SD-OCT findings with few or no intravitreal injections, when on oral X-82.(73)

A phase 2 trial of oral X-82 at doses up to 200 mg daily (the APEX Study) is in progress.(74) It is a randomized, double-masked, comparative dose trial with placebo control intended to find the most appropriate dose of X-82 for treatment of wet AMD and to observe its safety and efficacy for a full year.

The trial seeks to enroll 132 patients, and it is running at 25 sites across the United States. 

Intravitreal anti-VEGF treatments will be administered as needed, based on SD-OCT evidence of increased macular fluid. 

The primary endpoint is visual acuity and secondary endpoints include the number of injections required and anatomic changes on SD-OCT.

 

4.3.5 DE-120 (anti-VEGF/anti-PDGF agent, Santen)

DE-120 (Santen) is a small molecule dual tyrosine kinase receptor inhibitor for both VEGF and PDGF signaling pathways.

DE-120 belongs to a class of active small molecules that inhibit the activity of receptor tyrosine kinases involved in the angiogenic process.

As a single anti-VEGF/anti-PDGF agent, DE-120 has the potential to provide a synergic antiangiogenic effect and reduce the potential side effects associated with intravitreal injection procedures.

Santen is conducting a phase 2 clinical trial(75) in the United States to evaluate the safety and efficacy of intravitreal DE-120 as monotherapy and along with a single injection of aflibercept (Eylea) in subjects with treatment-naïve active subfoveal CNV secondary to AMD.

 

4.3.6 Combination therapy of the angiopoietin2 antibody nesvacumab and aflibercept

Regeneron Pharmaceuticals and Bayer developed two separate phase 2 clinical studies to evaluate the combination therapy as a co-formulated single intravitreal injection in patients with wet AMD or diabetic macular edema.

Angiopoietins are a family of vascular growth factors. 

Preclinical data(66-68) demonstrated that angiopoietins act together with the VEGF family to promote the formation and maturation of blood and lymphatic vessels in the eye. 

Ang2 and VEGF together therefore have the potential to influence the pathological development of new blood vessels and the permeability of blood vessel walls in certain diseases of the eye.

 

4.3.7 Combination therapy of intravitreal anti-VEGF therapy and topical Squalamine

Squalamine is a small anti-angiogenic drug that acts against the development of aberrant neovascularization by inhibiting multiple growth factors, including VEGF, PDGF and basic fibroblast growth factor (bFGF), wich play a role in angiogenesis and ocular neovascular disease. 

The topical formulation of squalamine, called OHR-102, is late stage development for the treatment wet AMD. In the IMPACT studyl(69), a phase 2, prospective, randomized, double-masked, placebo-controlled, multicenter study completed in 2015 in treatment naïve patients with CNV due to AMD, all patients enrolled l(142) received ranibizumab at baseline and randomized 1:1 to topical OHR-102 BID (combination group) or placebo vehicle solution BID (monotherapy group). 

Patients were followed monthly for 9 months.

Retreatment with ranibizumab was performed by criteria-based PRN ranibizumab IVT injections monthly to month 9. OHR-102 demonstrated an improvement in visual function when used in combination with an anti-VEGF agent versus anti-VEGF monotherapy.

There were no safety issues identified.

An analysis of visual outcome as a function of lesion characteristics (presence and size of classic and occult CNV) was performedl(70) and demonstrated a strong correlation in Figure 5. 

Combination of anti-PDGF and anti-VEGF induced angiogenic vessel regression in choroidal neovascular model.

Preclinical model of the combined treatment. the combination therapy group (p=<0.0001) but not with ranibizumab monotherapy. 

Angiographic CNV lesion characteristics were predictive of visual acuity outcomes in combination therapy: baseline occult CNV area showed the highest correlation to visual outcome in combination treatment with IVT anti-VEGF and topical squalamine.

Patients with occult CNV <10 mm2 represented an optimal target population for this combination therapy in this clinical trial. 

OHR-102 used in combination with an anti-VEGF agent may provide several potential advantages over other combination therapy approaches currently being investigated in clinical studies. 

Based on the good results from the IMPACT study in wet AMD, a phase III program is being initiated. 

 

4.3.8 Combination therapy with ICON-1 and ranibizumab

ICON-1 is an anti-Tissue Factor (TF) immunoconjugate protein that binds to pathologic vessels overexpressing TF and acts via a new mechanism of action that can eliminate abnormal CNV as well as inhibit the exudation. 

EMERGE(76-77) is a phase 2, randomized, double masked, active control study in the United States that examines the hypothesis that ICON-1 eliminates abnormal CNV as well as inhibits the exudation either alone or in combination with ranibizumab as compared to ranibizumab alone.

A total of 90 patients with treatment naïve (in the study eye) CNV secondary to AMD are being enrolled.

Patients are randomized in a 1:1:1 ratio to receive intravitreal injections of ICON-1 (0.3 mg) as monotherapy (n=30) or in combination with ranibizumab 0.5 mg (n=30) or ranibizumab 0.5 mg monotherapy (n=30).

Patients will receive 3 initial monthly injections followed by maximum 3 additional monthly injections based on protocol re-treatment criteria, for a total of 6 months of treatment.

The primary outcomes are mean change from baseline in BCVA letter score and in Central Retinal Thickness (CRT) at 3 months.

 

4.3.9 Combination therapy with aflibercept coformulated with rinucumab (CAPELLA study)

CAPELLA(86) study is an ongoing phase 2 double-masked, randomized, controlled, multiple-dose, regimen-ranging study that has enrolled patients with wet AMD who were treated with aflibercept coformulated with rinucumab, an anti–platelet-derived growth factor (anti-PDGF) receptor beta antibody. 

Patients were randomly assigned to one of 3 groups: they received fixed doses every 4 weeks of 2 mg aflibercept monotherapy, 2 mg aflibercept/1 mg rinucumab combination therapy, or 2 mg aflibercept/3 mg rinucumab combination therapy. 

At week 12, two of the three treatment groups were rerandomized, resulting in five total dosing groups in the second phase of the study. 

This study failed to reach its primary endpoint, according to a company press release. At week 12, the combination therapy did not demonstrate an improvement in BCVA compared with aflibercept monotherapy.

The press release reported that patients in the combination therapy arms showed a 5.8-letter improvement in BCVA, compared with a 7.5-letter improvement in the monotherapy arm. 

The company reported that combination therapy showed no benefit over monotherapy on anatomic endpoints including reduction in retinal thickness or resolution of subretinal hyperreflective material. 

Ocular adverse events at week 12 were more common in the combination therapy groups (23.5% and 20%) than in the monotherapy group (16%), according to the company.

Researchers will reevaluate data at weeks 28 and 52.

 

5. The future of combined treatment

Only time will tell us whether double and triple treatments will be the gold standard treatment. 

In most combined studies, the number of re-treatments needed was lower than with monotherapy. It is known that some patients respond better to monotherapy, while others respond better to combined treatment; this is very likely due to individual patient and disease characteristics.

Intensive research is currently in course regarding alternative actions on crucial neovascularization cascade steps and mechanisms that trigger this process (signalling), since these have the potential to become alternative strategies for the combined treatment of AMD.

Emerging therapies will be described in a separate chapter. 

It shall be referred that the complexity of signalling pathways supports the concept of combined therapy as a way of achieving more adequate control of biological functions in general and neovascularization in particularr(62,63).

 

6. Practical aspects

Combined therapies have long been used in the treatment of oncological and numerous other systemic diseases. 

In neovascular AMD, the objective of combined treatments acting upon different stages of the physiopathological process or the signalling pathway that triggers its mechanisms of action is to achieve synergistic action; therefore, an increased treatment effect and/or a decrease in the number of retreatment sessions required to stabilize vision are to be expected, as well as a more prolonged effect, smaller doses and increased drug tolerance(1).

Treatment of this disease focuses on three main targets(1):

  1. neovascularization;
  2. angiogenic process; 
  3. inflammatory cicatricial and exudative process.

In clinical practice, the following therapies are currently used in combined treatment:

  1. Photodynamic therapy with verteporfin, with standard, low or very low fluence.
  2. Antiangiogenic agents ranibizumab 0.5mg (Lucentis®), bevacizumab 1.25mg (Avastin®) and VEGF Trap – aflibercept 2 mg (EYLEA®).
  3. Anti-inflammatory treatment with intravitreal dexamethasone with or without core vitrectomy. 

 

6.1 Particular cases of AMD

Due to their poor response to monotherapy, cases of Retinal Angiomatous Proliferation (RAP) and Polypoid Choroidal Vasculopathy should be primary candidates for anti-VEGF-based combined treatment. 

In the Everest Study,verteporfin PDT combined with ranibizumab 0.5 mg or alone was superior to ranibizumab monotherapy in achieving complete regression of polyps in this 6-month study in patients with symptomatic macular polypoidal choroidal vasculopathy.(34).

ATLANTIC study (Protocol ECR-AMD-2015-09), sponsored by AIBILI, Coordinating Centre of EVICR.net, is a still running randomized, double-masked, sham-controlled phase 4 study of the efficacy, safety, and tolerability of intravitreal aflibercept monotherapy compared to aflibercept with adjunctive photodynamic therapy in patients with polypoidal choroidal vasculopathy.

 

6.2 Improved healthcare and increased equity

If combined treatments are proved to lead to better outcomes (greater VA line gains) or more sustained gains, as referred in most studies, the superior clinical efficacy of this treatment approach will be established.

Although the costs of two or three different treatments need to be considered when calculating combined treatment costs, costs per patient will be reduced if fewer overall resources are used.

This is a more efficient strategy, as well as a principle to follow in health economics: to manage scarce resources so that health investments may benefit more patients, instead of necessarily making expense cuts.

It is also about increasing equity – increasing the number of patients benefiting from treatment(1).

Levels of clinical evidence:

Level I– At least one well-designed study – randomized, controlled studies.
Level II-1– High-quality, non-randomized, controlled studies.
Level II-2– Studies with a control group involving more than one research centre or group.
Level II-3– Studies with no control group; series studies, with or without intervention.
Level III– Opinion of respected authorities, based on clinical experience, descriptive studies or specialised committee reports.

Abbreviations

ACE- angiotensin converting enzyme
AR- angiotensin receptor
ICAM-1- Intercellular Adhesion Molecule-1
SDF-1-CRXR4 Axis– Stromal-Derived Factor-1 and its CRXR-4 receptor
PEDF- Pigment Epithelium-Derived Factor
VEGF- Vascular Endothelial Growth Factor

 

>> References

References - Combined Treatment

References - Combined Treatment

English
  1. Schmidt-Erfurth UM, Richard G, Augustin A, Aylward WG, Bandello F, Corcostegui B, Cunha-Vaz J, Gaudric A, Leys A, Schlingemann RO. Guidance for the treatment of neovascular age-related macular degeneration. Acta Ophthalmol Scand. 2007;85(5):486-494.
     
  2. Zuluaga MF, Mailhos C, Robinson G, Shima DT, Gurny R, Lange N. Synergies of VEGF inhibition and photodynamic therapy in the treatment of age-related macular degeneration. Invest Ophthalmol Vis Sci. 2007;48(4):1767-1772.
     
  3. Markabi S. Combination therapy from the regulatory perspective. Retina. 2009;29(6Suppl):S12-S14.
     
  4. Karel I. Moznosti a ekonomické ukazatele lécby exsudativní vekem podmínené makulární degenerace s choroidální neovaskulární membránou. Cesk Slov Oftalmol. 2007;63(5):311-319.
     
  5. Smiddy WE. Relative cost of a line of vision in age-related macular degeneration. Ophthalmology. 2007;114(5):847-854.
     
  6. Smiddy WE. Economic implications of current age-related macular degeneration treatments. Ophthalmology. 2009;116(3):481-487.
     
  7. Dugel PU. Anti-Platelet Derived Growth Factor: Where Do We Stand? Paper presented at: American Academy of Ophthalmology Annual Meeting; November 10, 2012; Chicago.
     
  8. Dugel PU. Anti-PDGF Combination Therapy in Neovascular Age-related Macular Degeneration: Results of a Phase 2b Study. RetinaToday March 2013. Available at: http://retinatoday.com/2013/03/anti-pdgf-combination-therapy-in-neovascular-age-related-macular-degeneration-results-of-a-phase-2b-study/ . Accessed April 30, 2017.
     
  9. Potter MJ, Claudio CC, Szabo SM. A randomised trial of bevacizumab and reduced light dose photodynamic therapy in age-related macular degeneration: the VIA study. Br J Ophthalmol. 2010;94(2):174-179.
     
  10. Olk RJ, Peralta E, Gierhart DL, Brown GC, Brown MM. Triple combination therapy and zeaxanthin for the treatment of neovascular age-related macular degeneration: an interventional comparative study and cost-effectiveness analysis. Int J Retina Vitreous. 2015 Nov;1:22.
     
  11. Ambati J, Fowler BJ. Mechanisms of age-related macular degeneration. Neuron. 2012;75(1):26–39.
     
  12. Martin DF, Maguire MG, Fine SL, Ying GS, Jaffe GJ, Grunwald JE, Toth C, Redford M, Ferris FL 3rd. Comparison of Age-related Macular Degeneration Treatments Trials (CATT) Research Group. Ranibizumab and bevacizumab for treatment of neovascular age-related macular degeneration. Ophthalmology. 2012;119(7):1388-1398.
     
  13. Silva R, Axer-Siegel R, Eldem B, Guymer R, Kirchhof B, Papp A, Seres A, Gekkieva M, Nieweg A, Pilz S; SECURE Study Group. The SECURE Study: Long-Term Safety of Ranibizumab 0.5 mg in Neovascular Age-related Macular Degeneration. Ophthalmology. 2013;120(1):130-139.
     
  14. Singer MA, Awh CC, Sadda S, Freeman WR, Antoszyk AN, Wong P, Tuomi L. HORIZON: an open-label extension trial of ranibizumab for choroidal neovascularization secondary to age-related macular degeneration. Ophthalmology. 2012;119(6):1175-1183.
     
  15. Bhisitkul RB, Rofaghaet S, Boyer DS, Sadda S, Zhang K. Year 7 outcomes for ranibizumab-treated subjects in ANCHOR/MARINA: a multicenter, prospective cohort study. Paper presented at: ARVO Annual Meeting; May 8, 2012; Fort Lauderdale,FL. Invest Ophthalmol Vis Sci. 2012; 53(14):3679.
     
  16. 16. Gerhardt H, Golding M, Fruttiger M, Ruhrberg C, Lundkvist A, Abramsson A, Jeltsch M, Mitchell C, Alitalo K, Shima D, Betsholtz C.VEGF guides angiogenic sprouting utilizing endothelial tip cell filopodia. J Cell Biol. 2003;161(6):1163-77.
     
  17. De Smet F, Segura I, De Bock K, Hohensinner PJ, Carmeliet P. Mechanisms of Vessel Branching Filopodia on Endothelial Tip Cells Lead the Way. Arterioscler Thromb Vasc Biol. 2009;29:639-649.
     
  18. Larrivée B, Freitas C, Suchting S, Brunet I, Eichmann A. Guidance of vascular development: lessons from the nervous system. Circ Res. 2009; 104(4):428-41.
     
  19. Wang Y, Wang VM, Chan CC. The role of anti-inflammatory agents in age-related macular degeneration (AMD) treatment. Eye (Lond). 2011 Feb;25(2):127–139.
     
  20. Funatsu H, Yamashita H, Sakata K, Noma H, Mimura T, Suzuki M, Eguchi S, Hori S.Vitreous levels of vascular endothelial growth factor and intercellular adhesion molecule 1 are related to diabetic macular edema.Ophthalmology. 2005 May;112(5):806-16.
     
  21. Butler JM, Guthrie SM, Koc M, Afzal A, Caballero S, Brooks HL, Mames RN, Segal MS, Grant MB, Scott EW.SDF-1 is both necessary and sufficient to promote proliferative retinopathy. J Clin Invest. 2005 Jan;115(1):86-93.
     
  22. Penfold PL, Wong JG, Gyory J, Billson FA. Effects of triamcinolone acetonide on microglial morphology and quantitative expression of MHC-II in exudative age-related macular degeneration. Clin Experiment Ophthalmol. 2001;29(3):188-192.
     
  23. Schmidt-Erfurth U, Hasan T. Mechanisms of action of photodynamic therapy with verteporfin for the treatment of age-related macular degeneration. Surv Ophthalmol. 2000 Nov-Dec;45(3):195-214.
     
  24. Schmidt-Erfurth U, Kiss C, Sacu S. The role of choroidal hypoperfusion associated with photodynamic therapy in neovascular age-related macular degeneration and the consequences for combination strategies. Prog Retin Eye Res. 2009;28(2):145-154.
     
  25. Tatar O, Adam A, Shinoda K, Stalmans P, Eckardt C, Lüke M, Bartz-Schmidt KU, Grisanti S. Expression of VEGF and PEDF in choroidal neovascular membranes following verteporfin photodynamic therapy. Am J Ophthalmol. 2006 Jul;142(1):95-104.
     
  26. Rosenfeld PJ, Brown DM, Heier JS, Boyer DS, Kaiser PK, Chung CY, Kim RY; MARINA Study Group. Ranibizumab for neovascular age-related macular degeneration. N Engl J Med. 2006;355(14):1419-1431.
     
  27. Brown DM, Kaiser PK, Michels M, Soubrane G, Heier JS, Kim RY, Sy JP, Schneider S; ANCHOR Study Group. Ranibizumab versus verteporfin for neovascular age-related macular degeneration. N Engl J Med. 2006;355(14):1432-1444.
     
  28. QLT announces final results from its radical study evaluating verteporfin PDT (Visudyne®) combination therapy in exudative AMD. Press Release, June 22, 2010. Available at: https://globenewswire.com/news-release/2010/06/22/423627/194795/en/QLT-Announces-Final-Results-From-Its-RADICAL-Study-Evaluating-Verteporfin-PDT-Visudyne-R-Combination-Therapy-in-Exudative-AMD.html . Accessed April 30, 2017.
     
  29. Augustin AJ, Puls S, Offermann I. Triple therapy for choroidal neovascularization due to age-related macular degeneration: verteporfin PDT, bevacizumab, and dexamethasone. Retina. 2007;27:133-140.
     
  30. Larsen M, Schmidt-Erfurth U, Lanzetta P, Wolf S, Simader C, Tokaji E, Pilz S, Weisberger A; MONT BLANC Study Group. Verteporfin plus ranibizumab for choroidal neovascularization in age-related macular degeneration: twelve-month MONT BLANC study results.Ophthalmology. 2012 May;119(5):992-1000.
     
  31. 12-Months results  from Denali Study evaluating verteporfin PDT (Visudyne®) combination therapy QLT news release, June 15, 2010. Available at: https://globenewswire.com/news-release/2010/06/15/423254/194371/en/12-Month-Results-From-DENALI-Study-Evaluating-verteporfin-PDT-Visudyne-R-Combination-Therapy.html . Accessed April 30, 2017.
     
  32. Koh A, Lee WK, Chen LJ, Chen SJ, Hashad Y, Kim H, Lai TY, Pilz S, Ruamviboonsuk P, Tokaji E, Weisberger A, Lim TH. THE EVEREST study: efficacy and safety of verteporfin photodynamic therapy in combination with ranibizumab or alone versus ranibizumab monotherapy in patients with symptomatic macular polypoidal choroidal vasculopathy. Retina. 2012 Sep;32(8):1453-64.
     
  33. Triple Therapy - PDT Plus IVD and Intravitreal Ranibizumab Versus Lucentis Monotherapy to Treat Age-Related Macular Degeneration (PDEX). https://clinicaltrials.gov/ct2/show/NCT00390208?term=NCT00390208&rank=1 . Accessed April 30, 2017.
     
  34. Ranchod TM, Ray SK, Daniels SA, Leong CJ, Ting TD, Verne AZ. LUCEDEX: A Prospective Study Comparing Ranibizumab plus Dexamethasone Combination Therapy Versus Ranibizumab Monotherapy for Neovascular Age-Related Macular Degeneration. Retina. 2013 Sep;33(8):1600-4.
     
  35. Buchholz P, Buchholz A, Kirchhof J, Augustin A. Combination Therapies for AMD: Latest Developments. Retinal Physician. 2010 Mar. Available at: http://www.retinalphysician.com/issues/2010/march-2010/combination-therapies-for-wet-amd-latest-developm . Accessed April 30, 2017.
     
  36. Sheidow T, Kertes PJ et al. The role of combination therapy in AMD: the Canadian study of Avastin and Visudyne in exudative AMD (CAVE STUDY), paper presented at Retina Congress 2009;sept 30-Oct. 4, 2009; New York, NY.
     
  37. Loewenstein A, Kuppermann BD, Weinberger D, Goldstein M, Singer M, Liu CC, Lou J, Li XY, Whitcup SM. Safety and Efficacy of OZURDEXTM (Dexamethasone Intravitreal Implant) as Adjunctive Therapy to Lucentis® in Patients With Choroidal Neovascularization (CNV) Secondary to Age-Related Macular Degeneration (AMD). Invest Ophthalmol Vis Sci. 2010; 51(13):1255.
     
  38. Flaxel C, Schain MB, Hamon SC, Francis PJ. Prospective randomized controlled trial of combination ranibizumab (Lucentis) and bromfenac (Xibrom) for neovascular age-related macular degeneration: a pilot study. Retina. 2012 Mar;32(3):417-23.
     
  39. Dugel P. Angiogenesis, Exudation, and Degeneration 2012 Meeting in Miami on Februay 4th. The 2-year results of the CABERNET study.
     
  40. Söderberg AC, Algvere PV, Hengstler JC, Söderberg P, Seregard S, Kvanta A. Combination therapy with low-dose transpupillary thermotherapy and intravitreal ranibizumab for neovascular age-related macular degeneration: a 24-month prospectiverandomised clinical study. Br J Ophthalmol. 2012;96(5):714-8
     
  41. Koch F, Scholtz S, Singh P, Koss MJ. Kombinierte intravitreale Therapie zur Behandlung der altersbedingten Makuladegeneration. Klin Monbl Augenheilkd. 2008;225(12):1003-1008.
     
  42. Jo N, Mailhos C, Ju M, Cheung E, Nishijima K, Robinson GS, Adamis AP, and Shima DT. Inhibition of platelet-derived growth factor B signaling enhances the efficacy of anti-vascular endothelial growth factor therapy in multiple models of ocular neovascularization. Am J Pathol. 2006;168(6):2036-2053.
     
  43. Ciulla TA, Jaffe GJ, Patel S. Assessment of Retinal Pigment Epithelium (RPE) Atrophy in a Phase 2b Study of a Platelet Derived Growth Factor inhibitor (Fovista®), in combination with a Vascular Endothelial Growth Factor inhibitor (Ranibizumab) for Neovascular Age-Related Macular Degeneration (NAMD). Invest Ophthalmol Vis Sci. 2016; 57(12):4418.
     
  44. Stahl A, Paschek L, Martin G, Feltgen N, Hansen LL, Agostini HT. Combinatory inhibition of VEGF and FGF2 is superior to solitary VEGF inhibition in an in vitro model of RPE-induced angiogenesis. Graefes Arch Clin Exp Ophthalmol. 2009;247(6):767-773.
     
  45. Bhutto IA, McLeod DS, Hasegawa T, Kim SY, Merges C, Tong P, Lutty GA. Pigment epithelium-derived factor (PEDF) and vascular endothelial growth factor (VEGF) in aged human choroid and eyes with age-related macular degeneration. Exp Eye Res. 2006; 82(1):99-110.
     
  46. Campochiaro PA, Nguyen QD, Shah SM, Klein ML, Holz E, Frank RN, Saperstein DA, Gupta A, Stout JT, Macko J, DiBartolomeo R, Wei LL. Adenoviral vector-delivered pigment epithelium-derived factor for neovascular age-related macular degeneration: results of a phase I clinical trial. Hum Gene Ther. 2006;17(2):167-176.
     
  47. Diago T, Pulido JS, Molina JR, Collett LC, Link TP, Ryan EH Jr. Ranibizumab combined with low-dose sorafenib for exudative age-related macular degeneration. Mayo Clin Proc. 2008;83(2):231-234.
     
  48. Chaudhry N. Oral VEGF receptor/PDGF receptor inhibitor X-82. Paper presented at: the American Academy of Ophthalmology Annual Meeting; November 14-17, 2015; Las Vegas, NV.
     
  49. Jackson T. A phase 1 study of oral tyrosine kinase inhibitor (X-82) in previously treated wet age-related macular degeneration. Paper presented at: 15th Annual Euretina Congress; September 17-20, 2015; Nice, France.
     
  50. Jackson T. A phase 1 study of oral tyrosine kinase inhibitor (X-82) in previously treated wet age-related macular degeneration. Paper presented at: 15th Annual Euretina Congress; September 17-20, 2015; Nice, France.
     
  51. X-82 to Treat Age-related Macular Degeneration. https://clinicaltrials.gov/ct2/show/NCT02348359. Accessed November 28, 2016.
     
  52. A Safety and Efficacy Study of DE-120 Injectable Solution for Age-related Macular Degeneration (VAPOR1). https://clinicaltrials.gov/ct2/show/NCT02401945. Accessed November 28, 2016.
     
  53. Hanahan D. Signaling vascular morphogenesis and maintenance. Science. 1997;277:48–50.
     
  54. Maisonpierre PC, Suri C, Jones PF, Bartunkova S, Wiegand SJ, Radziejewski C, Compton D, McClain J, Aldrich TH, Papadopoulos N, Daly TJ, Davis S, Sato TN, Yancopoulos GD. Angiopoietin-2, a natural antagonist for Tie2 that disrupts in vivo angiogenesis. Science. 1997;277:55–60.
     
  55. Bergers G, Song S. The role of pericytes in blood-vessel formation and maintenance. Neuro Oncol. 2005 Oct;7(4):452–464.
     
  56. Slakter JS, Ciulla TA, et al. Final Results from a Phase 2 study of Squalamine Lactate Ophthalmic Solution 0.2% (OHR-102) in the treatment of neovascular age-related macular degeneration (AMD). Invest Ophthalmol Vis Sci. 2015;56(7):4805.
     
  57. Brown D, Ingerman A, Shearn SP, Slakter JS. CNV lesion characteristics as a predictor of visual outcome in wet AMD patients receiving combination therapy of intravitreal anti-VEGF therapy and topical Squalamine lactate ophthalmic solution. Invest Ophthalmol Vis Sci. 2016; 57(12):4419.
     
  58. A Phase 2 Study (EMERGE) Evaluating Repeated Intravitreal Administration of ICON-1 in Patients With Choroidal Neovascularization (CNV) Secondary to Age-related Macular Degeneration (AMD). https://clinicaltrials.gov/ct2/show/NCT02358889. Accessed November 28, 2016.
     
  59. Christmas NJ. A Phase 2 Study (EMERGE) Evaluating Repeated Intravitreal Administration of ICON-1 in Patients With Choroidal Neovascularization (CNV) Secondary to Age-related Macular Degeneration (AMD). Invest Ophthalmol Vis Sci. 2016; 57(12):4434.
  60. Kaiser PK. Strategies for inhibiting vascular endothelial growth factor. Retina. 2009;29(6Suppl):S15-S17.
     
  61. Geltzer A, Turalba A, Vedula SS.Surgical implantation of steroids with antiangiogenic characteristics for treating neovascular age-related macular degeneration. Cochrane Database Syst Rev. 2007 Oct;(4):CD005022.

Surgery in AMD

Surgery in AMD

English

Authors:

Angelina Meireles, MD
Hospital Santo António-CHP - Porto, Portugal

Bernardete Pessoa, MD
Hospital Santo António-CHP - Porto, Portugal

Natália Ferreira, MD
Hospital Santo António-CHP - Porto, Portugal

Rui Martinho, MD
Hospital Lusíadas Porto - Porto, Portugal

Updated/reviewed by the authors, July 2017.

Introduction

In recent years, new treatment modalities such as photodynamic therapy and intravitreal anti-VEGF injections have been added to the treatment armamentarium for age-related macular degeneration (AMD).

Prior to the introduction of these therapies, only laser photocoagulation had been shown in a large randomized controlled trial to confer a statistically significant benefit in the treatment of subfoveal choroidal neovascularization (CNV) secondary to AMD regarding to long-term visual acuity when compared to the natural history of the condition(1-3).

Unfortunately, the Macular Photocoagulation Study also showed that laser photocoagulation was associated with immediate average reduction of visual acuity, with benefits over no treatment apparent only six months after treatment(1-3).

In fact, recovery of good vision rarely occurred in these patients.

As a result of the limited treatment options, alternative therapies, such as submacular surgery for CNV removal, were pursued with limited or no success.

Meanwhile, a different management paradigm for AMD was established with macular translocation.

However, because of its higher risk of complications, its popularity has waned with the wider availability of photodynamic therapy and the introduction of intravitreal anti-VEGF agents.

Nonetheless, they remain potentially useful treatment options, even if their role in the management of AMD is neither established nor consensual.

In the meantime, vitreoretinal surgery and vitrectomy have had important developments, mostly due to advances in sutureless transconjunctival vitrectomy, use of dyes, tamponade agents and new equipments.

Novel surgical approaches for AMD are under scrutiny: the association of submacular surgery with pigment cell transplantation, the use of adjuncts, such as recombinant tissue plasminogen activator (r-TPA) for subretinal hemorrhages displacement combined with anti-VEGF treatment.

Recently, a new insight involving the vitreoretinal surface has been proposed for the pathophysiologic mechanisms underlying the development of CNV in AMD.

Accordingly, posterior vitreomacular adhesion may be another risk factor in a subpopulation of patients with wet AMD, opening another path for a surgical approach in the treatment algorithm for this condition(4-7).

Future studies are needed to define the role of vitreoretinal surgery in such cases.

 

Macular translocation

The first experiments on retinal relocation were conducted and reported in the beginning of the 80’s.

Their aim was to study the anatomic dependency of the foveal retina on foveal retinal pigment epithelium (RPE) and choroid.

The assumption that visual function could be preserved with foveal displacement has originated from cases of ectopic macula from retinal traction or after surgery in patients with retinopathy of prematurity and retinal detachments with giant tears(8).

They ended up showing the feasibility of rotating the macula around the optic disc with reattachment of the fovea in animal eyes(9).

One might say that this was the starting point for the idea of rotating the macula of eyes with subfoveal CNV to a new area of underlying RPE-Bruch’s membrane-choriocapillaris complex – macular translocation  – as a treatment for the condition.

Although the exact pathogenesis of CNV secondary to AMD is not known, the natural history of this condition is a progressive loss of central vision over time.

The initial retinal dysfunction responsible for impaired vision in eyes with subfoveal CNV may be attributable to factors such as subretinal fluid, subretinal hemorrhage and retinal edema.

When fibrous proliferation and degeneration of the overlying photoreceptors occur during the later stages of the disease, the visual loss becomes irreversible.

The rationale of macular translocation is that moving the neurosensory retina of the fovea in one eye with recent-onset subfoveal CNV to a new location before the occurrence of permanent retinal damage, may allow it to recover or to maintain its visual function over a healthier bed of RPE-Bruch’s membrane-choriocapillaris complex.

In addition, relocating the fovea to an area outside the CNV allows the ablation of the later, by laser photocoagulation without destroying the fovea, in an attempt to preserve central vision.

On the other hand, some surgeons have combined macular translocation with CNV removal, allowing the fovea to be relocated to an area outside the RPE defect created during submacular surgery.
Several different surgical techniques for macular translocation have been described and were currently used.

These techniques produce different degrees of postoperative foveal displacement, and can broadly be classified into two categories:

Full macular translocation, and limited macular translocation.

 

Full macular translocation

After developing their surgical techniques in rabbit eyes(9), Machemer and Steinhorst became, in 1993, the first surgeons to demonstrate the feasibility of macular translocation in humans(10,11).

Their technique involved lensectomy, complete vitrectomy, planned total retinal detachment by transretinal infusion of fluid under the retina, 360º peripheral circumferential retinotomy, rotation of the retina around the optic disc, and reattachment of the retina with silicone oil tamponade.

Besides allowing retinal rotation to occur, the retinotomy also made way to the subretinal space for syncronous blood and CNV removal.

Since then, a number of investigators have subsequently modified this technique(12-15).

Corrective extraocular muscle surgery for globe counter-rotation, due to frequent postoperative cyclovertical diplopia or awareness of a tilted image, may be done during the primary surgery or at a later stage(13,16).

While some surgeons have found the results of macular translocation encouraging(13,17-20), others found the surgery unpredictable(21,22).

In a series of 50 consecutive eyes with subfoveal CNV from AMD that underwent full macular translocation followed-up for a median period of 21 months (12-36 months), Pertile and Claes reported an improvement of ≥ 2 Snellen lines in 66%, while 28% remained stable (±1 line) and 6% decreased by ≥ 2 lines(17).

Van Romunde et al. published long-term results of full macular translocation, in a retrospective study, of 158 patients with subfoveal CNV from AMD followed-up for a mean period of 45 months (12-108 months). A clinically and statistically significant improvement of best-corrected visual acuity (BCVA) was detected and sustained for up to 5 years after full macular translocation. They found that 45% had a 3-line or more gain at last visit. In a subgroup of 56 patients followed up for 5 years or more, there was a gain of 1.5 lines compared with preoperative BCVA(19).

In a series of 40 consecutive patients with subfoveal CNV from AMD, who undergone the mentioned surgical technique and with a mean follow-up period of 37.6 months (12.4-67.4 months), Chen et al. reported at the last observation that 68% achieved a BCVA of 6/60 or better, 43% achieved 6/30 and 15% 6/12 or better. A 3-line gain in BCVA was obtained in 30% at 1 year and in 25% at the most recent follow-up. However, in 25% of patients achieving 6/60 or better at 1 year, BCVA declined by greater than 2 lines primarily as a result of recurrent CNV(20). In another prospective, interventional, consecutive non-comparative case series of 61 AMD patients who underwent the same procedure and followed-up for 12 months, the visual acuity improved by ≥ 1 Snellen lines in 52%(23).

Toth et al. also showed improvements in distance and near visual acuity, contrast sensitivity and reading speed in a series of 25 consecutive AMD patientss(24).

In addition to exceptional surgical technique, avoiding intra and postoperative complications, the key for success after macular translocation seems to be patient selection(25,26).

If this procedure is performed on a patient without viable foveal photoreceptors, there is no chance for visual improvement.

Still, the main drawback of macular translocation lies on its high rate of complications. Rhegmatogenous retinal detachment, with or without proliferative vitreoretinopathy development, is the most common serious complication of macular translocation. Rates up to 19% have been reported(27).

Persistent or recurrent subfoveal CNV has been described in up to 30% of patients undergoing this procedure(19,27). Almost all reported cases of recurrent CNV occurred at the edge of the RPE defect created by the CNV excision. In order to reduce the recurrence of CNV, the fovea needs to be reconstructed by using healthier RPE such as those derived from stem cells (28). Chen et al. recommend close monitoring of the edge of excision site, clinically and with SD-OCT, to detect early signs of CNV. Prompt initiation of a course of anti-VEGF therapy can maintain visual acuity gained following translocation(20).With a longer follow-up, cystoid macular edema (CME) and subfoveal RPE atrophy may be limiting factors for improved postoperative visual acuity. Terasaki and colleagues found a 70% incidence rate of CME(29). Chen et al. described CME in 38 of patients (95%), at some stage postoperatively(20).

In a long-term follow-up, Aisenbrey et al. (14 to 79 months – mean 38) and von Romunde et al. (12 to 108 months – mean 45) reported 28% versus 47% of patients with subfoveal atrophy by the third and fifth year postoperatively respectively(18,19).

In figures 1 and 2, we present two anecdotal cases in which full macular translocation was performed. They presented with extensive submacular hemorrhages that were considered inadequate for anti-VEGF therapy. Visual rehabilitation was achieved, enabling the patient’s independence for daily activities.

Figure 1. Case I – Patient with severe visual acuity loss in his left eye 3 days before admission (right eye was lost to advanced AMD). He presented with a submacular hemorrhage complicating wet AMD. Full macular translocation was performed (António Travassos – Coimbra Surgical Centre, Coimbra, Portugal). Three years after surgery, left eye visual acuity was 20/200.

 

​Figure 2. Case II – A 85-year-old man with an extensive submacular hemorrhage associated with subfoveal CNV (A). Four months after a full macular translocation (Claus Eckardt-Staedtische Kliniken Frankfurt A.M., Germany) his visual acuity was 20/80 (B). A recurrent subfoveal CNV developed after the first year of follow-up, and the patient was started with intravitreal ranibizumab injections. After 4 treatments, a visual acuity of 20/63 remained stable (C).

 

Limited macular translocation

In an effort to overcome the major complication following macular translocation with large retinotomy, proliferative vitreoretinopathy, de Juan developed a new technique in 1998(30).

His technique involved transretinal hydrodissection using small posterior retinotomies to induce a subtotal retinal detachment, a complete vitrectomy, anterior-posterior shortening near the equator and retinal reattachment.

As no large retinal break was created, the likelihood of developing proliferative vitreoretinopathy was thought to be lower.

As more experience was gained with this surgery, additional modifications were made to the original technique.

This is essentially a five-step procedure. It starts with the placement of scleral imbricating sutures (either in the superotemporal or inferotemporal quadrants for inferior and superior translocations, respectively) for antero-posterior shortening of the eyewall, followed by a pars plana vitrectomy with posterior vitreous detachment induction.

The third step should be an internal subretinal hydrodissection through small self-sealing retinotomies for creation of a partial neurosensory retinal detachment.

After tightening the scleral imbricating sutures, a redundancy of the neurosensory retina relative to the eyewall is created that should allow the desired foveal displacement, after retinal reattachment with fluid-air exchange.

Limited macular translocation may be either inferior or superior, depending on the movement of the neurosensory macula relative to the underlying tissues.

The largest series of limited macular translocation by Pieramici et al., in a retrospective review, analysed the outcome of 102 consecutive eyes of 101 patients that underwent inferior limited macular translocation for AMD-related subfoveal CNV. At three and six months postoperatively, 37% and 48% of the eyes, respectively, experienced ≥ 2 Snellen lines of visual improvement(31).

After one year of follow-up, 39.5% still maintained a ≥ 2 lines of improvement, while 29.0% remained unchanged and 31.4% lost ≥ 2 lines of visual acuity(32).

Recurrent CNV developed in 34.6% of the eyes, being subfoveal in most of the cases (65%)(32).

Besides the usual risks inherent to pars plana vitrectomy, limited macular translocation is associated with additional risks similar to those of scleral buckling surgery.

As with full macular translocation, the rate of complications won its popularity.

Although conceived in an effort to reduce the rate of proliferative vitreoretinopathy development, this has remained the most common serious complication after limited macular translocation.

Rates of retinal detachment up to 17.4% have been reported(33).

Insufficient macular translocation is another important limitation of this procedure, restricting its indications to smaller CNV.

 

Is there a role for macular translocation in the current era of anti-VEGF therapy and combined treatments?

IIn the latest edition of Vitreous Microsurgery (5th Edition, 2011 - Lippincott Williams & Wilkins)22),

Steve Charles et al. stated again that do not believe that this procedure is indicated due to the high incidence of complications.

In the review from Eandi et al. to access the functional outcomes of macular translocation for wet-AMD was concluded that there wasn’t enough evidence in randomized trials to support the benefit of surgery(34).

Future studies should involve patients with small neovascular membranes that are non-responsive to the present medical armamentarium, and that accept the risks of surgery in an effort to improve their visual function.

This is a procedure that demands a high level of surgical skill and experience, with a rather flat learning curve. The number of complications is unacceptably high. The widely use and quite effective pharmacologic therapy has limited this surgical modality to patients with large submacular haemorrhages, or patients unresponsive to VEGF inhibitors(19).

At the present time, the very small incidence of cases that could benefit from this procedure does not allow every Ophthalmology department with vitreoretinal surgical capability to gain expertise on macular translocation.

Referral this few cases to selected surgical units with the proper human and technical resources and experience seems the best choice.

 

Submacular surgery

As already has been said, a variety of surgical treatments have been developed for exudative AMD disease and one of those are the different modalities of submacular surgery.

We will review these surgical approaches and their actual role in the AMD treatment.

 

Surgery with subfoveal neovascular membrane removal

In 1988, Juan and Machemer(35) 

published the first results of vitreous surgical techniques on removal of blood or fibrous submacular complications in four AMD patients.

Countless publications of retrospective small sample size studies ensued, with no control group, describing the benefits of this technique in stabilizing the disease, albeit displaying reduced functional benefits(36-38).

The need to establish the benefits of submacular surgery in the treatment of CNV led to the Submacular Surgery Trials (SSTs). One of the aims of this study was to find out if surgery, for the different types of AMD lesions, more than vision stabilization, was able to increase vision with real impact in the life quality of these patients(39).

The pilot trial used to pre-test the research procedures of this multicentre, randomized, controlled study found no reason to prefer submacular surgery to laser photocoagulation in AMD patients with similar lesions to those displayed by study patients(40).

Several surgical techniques are described in the literature. In summary, these techniques include standard pars plana vitrectomy, with or without posterior hyaloid membrane removal; posterior retinotomy followed by infusion of subretinal saline solution or r-TPA into lesions with a large haemorrhagic component, membrane mobilization and its removal with surgical forceps, followed by eventual aspiration of fresh blood or clot, depending on the situation.

Possible intraocular haemorrhages during surgery may be controlled by increasing the intraocular pressure, either by raising the infusion bottle or using heavy perfluorocarbon liquids.

The procedure is finished with a fluid-air exchange, followed by gas buffering, maintaining the patient in the prone position until gas reabsorption(41-44).

Due to the physiopathology of this disease, it was observed in histopathological studies that inadvertent and undesired removal of the pigment epithelium often occurs during membrane removal, especially for type 1 membranes.

The irreversible absence of the pigment epithelium leads to loss or atrophy of photoreceptors and choriocapillaries, inducing an unfavourable visual recovery(45).

The percentage of removed epithelium is variable but may reach significantly high values, as observed in the SSTs, where the pigment epithelium was involved in 84% of removed membranes(46). ther complications like regmatogeneous retinal detachment were also observed, with higher incidence in eyes with larger hemorrhagic lesions and relatively poor visual acuity(47).

Therefore, both the functional results and the impact on life quality observed for the various subgroups considered in the SSTs, compared to natural disease progression, led the authors not to recommend submacular surgery as a treatment option(48-51).

 

Retinal pigment epithelium transplants

As previously referred, the poor results achieved with surgical removal of subfoveal membranes, largely due to resulting atrophy or rupture of the photoreceptor-RPE complex, led some groups to combine neovascular membrane removal with simultaneous autologous transplant of iris or RPE, in order to restore normal subretinal conditions(52-54).

The surgical maneuvers were similar to those used for subfoveal CNV removal with additional aspiration and subsequent pigment epithelial cell transplant procedures.

Although good tolerability has been demonstrated for pigment cells in the subretinal space, as well as absence of membrane recurrence, the poor functional results obtained in most studies have been attributed to several factors, including lack of differentiation in pigment cell reproliferation, failure in adhesion to Bruch’s membrane and failure to form a regular pigment epithelial cell monolayer(52-56).

Also according to other authors, the behavior of transplanted cells depends essentially on the type of environment found at the seeding location(57).

Meanwhile, Peyman et al. described a technique in which an autologous pedicle graft or homologous RPE cells and Bruch’s membrane replaced the submacular RPE(58). At mid-90s, Algvere et al. transplanted fetal RPE as a monolayer patch or in a suspension form(59,60).

LLater on, Aylward et al., from Moorfields Eye Hospital, were the first team to describe the concept of total patch translocation of the RPE-choroid complex to the subfoveal area(61). The patches collected in the macular area near the RPE lesion were small and even so they survived in the subfoveal space (assessed by OCT, FA and ICG angiography); after 5 years of follow-up, none of the nine transplanted patients displayed any function, what might be explained by chronic photoreceptor apoptosis(62).

In Rotterdam, Van Meurs et al., described a modification to this technique in which a healthier full-thickness patch of RPE-choroid of approximately 1.5x2 mm is harvested from within a circular zone isolated by heavy diathermia in the superior midperiphery where cells were thought to be less damaged(63). he group reported long-term results (up to 7 years) of at least 20/200 final BCVA in 15% of patients and the main complications were recurrence of CNV (10%) and proliferative vitreoretinopathy (10%)(64).

Subsequently, other groups have been publishing small case series presenting not only expressive and sustained improvement in visual acuity and reading ability, but also recovery of central fixation, evidence for grafts revascularization (also demonstrated histologically in animals experiments) and a normal autofluorescence over the patch in patients with exudative or dry AMD(63-72).

According to Parolini et al., those eyes whose vision is not seriously impaired (equal or higher than 20/200) and with a relatively well-preserved retinal layers, especially the external limiting membrane, are the ones with better outcomes. In this series, long-lasting beneficial effect of the graft for the function of the retina with improvement of vision was observed in most cases(73). Another critical issue for a successful outcome is the blood supply to the graft. Cereda et al. reported partial revascularization of the patch with choroidal vessels running parallel to one another unlike the radially choroidal vascular system in the macula area. In some cases localized feeder vessels, usually at the graft margin, were seen. These vessels are probably a large choroidal artery arising from the connection between the patch and the original choroid or choriocapillaris(74,75).

During CNV removal, choroidal structure should be left unharmed and the dimension of the graft must be large enough to connect to an area of preserved choroid. Revascularisation is a progressive event and time to completion is variable (over 1 month). Before revascularisation, patch survival may be accomplished through plasmatic diffusion, as in skin transplantation(76).

RPE-choroid transplant offers an interesting alternative to macular translocation, especially in patients with good vision in their fellow eye. This surgery is challenging and not free from intraoperative and postoperative complications. Improvements in microsurgical techniques can lead to less traumatic CNV removal and insertion of the RPE-choroid graft. This is crucial for promising functional results, as well as an adequate patient selection.

 

Management of submacular haemorrhage

In addition to the poor prognosis, exudative forms of AMD evolved into a major submacular haemorrhage  that are difficult to diagnose (concerning membrane location and extension) and to treat(77,78).

n the era of antiangiogenic agents, this is the most frequent form of AMD for which surgery is indicated, aiming to avoid further blood damages (mechanical, metabolic and toxic) to the photoreceptor-RPE complex(41,79,80).

As previously referred, blood was initially removed by aspiration or mechanically extracted.

In 1991, Lewis et al. were the first researchers to report the fibrinolysis properties of r-TPA, which made this agent useful in removing blood clots(42-44,81-84).

In order to avoid surgical manipulation of the macular retina, Herriot described a procedure involving displacement of submacular blood by intravitreal injection of r-TPA and gas, followed by prone position, for the first time in 1997.(84)

Blood is normally displaced temporally or infero-temporally, with a significant increase in visual acuity occurring immediately after the aforementioned procedure, as described in countless published outcomes.

Some authors have pointed out the duration of hemorrhage (the shorter the better) as the main predictive factor for the aforementioned improvement(85-92).

On the other hand, the usefulness of intravitreal r-TPA as an adjuvant to this technique has been questioned, not only because r-TPA diffusion to the subretinal space has not been proved in experimental studies, but also because some studies demonstrated the success of pneumatic displacement of subretinal blood without concomitant injection of r-TPA(90,91,93).

Therefore, Haupert introduced a hybrid technique in 2001, combining the concepts of pneumatic displacement with subretinal r-Tpa(94).

After a few changes, this technique is currently used in some centers, as described: pars plana vitrectomy, removal of the posterior hyaloid, injection of r-TPA (12.5 or 25 μg/0.1 mL) into the subretinal clot using a 39 or 41-gauge flexible translocation cannula and fluid-air exchange followed by prone position(95).

The advantages of this technique over the surgical blood removal are less intra and post-operative complications, which is probably due to the smaller extent of tissue manipulation and consequent reduction in retinal injury(83,94,95). Usually after successful blood displacement, the patient experience a short gain in visual acuity that frequently drop down because of recurrent and massive haemorrhage or progression of the underlying disease. Such cases might benefit from more invasive procedures as vitrectomy and 180º retinectomy for direct access to subretinal space allowing the complete removal of the underlying blood and CNV(96,97).

Since the advent of anti-VEGF therapy, a combined intraocular injection of bevacizumab and r-TPA seems to be a promising treatment for AMD with submacular haemorrhage.  Whatever the chosen procedure, close observation and additional intra- and/or postoperative anti-VEGF treatments might potentially prevent CNV related complications(98-100). n selected cases, namely patients with multiple diseases and little subretinal blood, monotherapy with intravitreal anti-VEGF could offer some chance of visual improvement with a minimally invasive approach(101).

More recently, there has been a trend towards more complex associated procedures:  vitrectomy and gas endotamponade with subretinal injection of r-TPA and anti-VEGF agents (1.25 mg bevacizumab). The thinking is that vitrectomy may relieve any vitreomacular traction, preventing wet-AMD from getting worse. Regarding subretinal anti-VEGF application, it ensures the direct delivery of the drug at the site of the CNV, which may enhance its effect; the short half-life of r-TPA (around 5 minutes) may reduce the risk of drugs interaction when given alongside anti-VEGF agents, particularly when injected into the confined subretinal space(98,102-104). In 2013, Martel et al. advocated a combined subretinal injection of r-TPA, bevacizumab and filtered air (0,2 mL), followed by intravitreal fluid-air or hexafluoride gas exchange and up-right positioning in order to get an easier hemorrhage displacement due to the lowering haemorrhage’s buoyancy obtained by the subretinal air(105).

As occurs in other macular diseases and also previous referred, the detection on OCT of an ellipsoid layer and external limiting membrane line beneath the fovea are strongly correlated with good visual prognosis(106). A study that used SD-OCT to evaluate the preoperative features of submacular haemorrhage and PED in nAMD, found out that the presence of an intact ellipsoid layer and lower height of submacular haemorrhage were predictors of good postoperative BCVA. In contrast, the preoperative characteristics of PED had no effects on visual prognosis(107). In the recent Stanesc-Segall literature review the greatest improvement is obtained when a combined approach with vitrectomy, subretinal r-TPA, intravitreal gas and anti-VEGF is performed(104). However, according to others, subretinal r-TPA, intravitreal gas and anti-VEGF treatment without vitrectomy, which is a less invasive procedure, had also good functional results with acceptable safety and fewer complications(100).

Concomitant systemic anticoagulation or platelet anti-aggregant medication should be discontinued if possible, especially in monocular patients. With the improvement of vitreoretinal surgery and instrumentation, described techniques are reasonably effective, although with widely variable outcome, depending on the severity of the underlying disease. Post-surgical addition of anti-VEGF medications appears to help preserve vision over time.

Regardless treatment modality, the benefits of surgery diminish with an increasing duration, extension and thickness of submacular haemorrhage(108).

Submacular haemorrhage associated with CNV due to AMD is a potentially visually devastating condition and to date, there is no consensus regarding an optimal treatment.

 

>> References

References - Surgery in AMD

References - Surgery in AMD

English
  1. Macular Photocoagulation Study Group. Laser photocoagulation of subfoveal neovascular lesions in age-related macular degeneration. Results of a randomized clinical trial. Arch Ophthalmol 1991;109(9):1220-1231.
     
  2. Macular Photocoagulation Study Group. Laser photocoagulation of subfoveal recurrent neovascular lesions in age-related macular degeneration. Results of a randomized clinical trial. Arch Ophthalmol 1991;109(9):1232-1241.
     
  3. Macular Photocoagulation Study Group. Laser photocoagulation of subfoveal neovascular lesions of age-related macular degeneration. Updated findings from two clinical trials. Arch Ophthalmol 1993;111(9):1200-1209.
     
  4. Krebs I, Brannath W, Glittenberg C, Zeiler F, Sebag J, Binder S. Posterior vitreomacular adhesion: a potential risk factor for exudative age-related macular degeneration? Am J Ophthalmol 2007;144(5):741-746.
     
  5. Schulze S, Hoerle S, Mennel S, Kroll P. Vitreomacular traction and exudative age-related macular degeneration. Acta Ophthalmol 2008;86(5):470-481.
     
  6. Lee SJ, Lee CS, Koh HJ. Posterior vitreomacular adhesion and risk of exudative age-related macular degeneration: paired eye study. Am J Ophthalmol 2009;147(4):621-626.
     
  7. Mojana F, Cheng L, Bartsch DU, Silva GA, Kozak I, Nigam N, Freeman WR. The role of abnormal vitreomacular adhesion in age-related macular degeneration: spectral optical coherence tomography and surgical results. Am J Ophthalmol 2008;146(2):218-227.
     
  8. Foulds WS. Factors influencing visual recovery in retinal detachment surgery. Trans Ophthalmol Soc U K 1980;100(Pt 1):72-77.
     
  9. Machemer R, Steinhorst UH. Retinal separation, retinotomy, and macular relocation: I. Experimental studies in the rabbit eye. Graefes Arch Clin Exp Ophthalmol 1993;231(11):629-634.
     
  10. Machemer R, Steinhorst UH. Retinal separation, retinotomy, and macular relocation: II. A surgical approach for age-related macular degeneration? Graefes Arch Clin Exp Ophthalmol 1993;231(11):635-641.
     
  11. Machemer R. Macular translocation. Am J Ophthalmol 1998;125(5):698-700.
     
  12. Wolf S, Lappas A, Weinberger AW, Kirchhof B. Macular translocation for surgical management of subfoveal choroidal neovascularizations in patients with AMD: first results. Graefes Arch Clin Exp Ophthalmol 1999;237(1):51-57.
     
  13. Eckardt C, Eckardt U, Conrad HG. Macular rotation with and without counter-rotation of the globe in patients with age-related macular degeneration. Graefes Arch Clin Exp Ophthalmol 1999;237(4):313-325.
     
  14. Ninomiya Y, Lewis JM, Hasegawa T, Tano Y. Retinotomy and foveal translocation for surgical management of subfoveal choroidal neovascular membranes. Am J Ophthalmol 1996;122(5):613-621.
     
  15. Toth CA, Machemer R. Macular translocation. In: Berger JW, Fine SL, Maguire MG, eds. Age-related macular degeneration. Saint-Louis, USA. Mosby. 1999;353-362.
     
  16. Holgado S, Enyedi LB, Toth CA, Freedman SF. Extraocular muscle surgery for extorsion after macular translocation surgery new surgical technique and clinical management. Ophthalmology 2006;113(1):63-69.
     
  17. Pertile G, Claes C. Macular translocation with 360 degree retinotomy for management of age-related macular degeneration with subfoveal choroidal neovascularization. Am J Ophthalmol 2002;134(4):560-565.
     
  18. Aisenbrey S, Bartz-Schmidt KU, Walter P, Hilgers RD, Ayertey H, Szurman P, Thumann G. Long-term follow-up of macular translocation with 360 degrees retinotomy for exudative age-related macular degeneration. Arch Ophthalmol 2007;125(10):1367-1372.
     
  19. van Romunde SHM, Polito A, Bertazzi L, Guerriero M, Pertile G.. Long-term results of full macular translocation for choriodal neovascularization in age-related macular degeneration. Ophthalmology 2015;122(7):1366-1374.
     
  20. Chen FK, Patel PJ, Uppal GS, Tufail A, Coffey PJ, Da Cruz L.. Long-term outcomes following full macular translocation surgery in neovascular age-related macular degeneration. Br J Ophthalmol 2010;94(10):1337-1343.
     
  21. American Academy of Ophthalmology. Macular translocation. Ophthalmology 2000;107(5):1015-1018.
     
  22. Charles S, Calzada J, Wood B. Submacular surgery and macular translocation. In: Charles S, Calzada J, Wood B, eds. Vitreous microsurgery. Fourth Edition. Philadelphia, USA. Lippincott Williams & Wilkins. 2007;14:163-171.
     
  23. Mruthyunjaya P, Stinnett SS, Toth CA. Change in visual function after macular translocation with 360 degrees retinectomy for neovascular age-related macular degeneration. Ophthalmology 2004;111(9):1715-1724.
     
  24. Toth CA, Lapolice DJ, Banks AD, Stinnett SS. Improvement in near visual function after macular translocation surgery with 360-degree peripheral retinectomy. Graefes Arch Clin Exp Ophthalmol 2004;242(7):541-548.
     
  25. Wong D, Stanga P, Briggs M, Lenfestey P, Lancaster E, Li KK, Lim KS, Groenewald C. Case selection in macular relocation surgery for age related macular degeneration. Br J Ophthalmol 2004;88(2):186-190.
     
  26. Uppal G, Milliken A, Lee J, Acheson J, Hykin P, Tufail A, da CL. New algorithm for assessing patient suitability for macular translocation surgery. Clin Experiment Ophthalmol 2007;35(5):448-457.
     
  27. Aisenbrey S, Lafaut BA, Szurman P, Grisanti S, Luke C, Krott R, Thumann G, Fricke J, Neugebauer A, Hilgers RD, Esser P, Walter P, Bartz-Schmidt KU. Macular translocation with 360 degrees retinotomy for exudative age-related macular degeneration. Arch Ophthalmol 2002;120(4):451-459.
     
  28. Vugler A, Carr AJ, Lawrence J, Chen LL, Burrell K, Wright A, Lundh P, Semo M, Ahmado A, Gias C, da Cruz L, Moore H, Andrews P, Walsh J, Coffey P. Elucidating the phenomenon of HESC-derived RPE: anatomy of cell genesis, expansion and retinal transplantation. Exp Neurol 2008;214(2):347-61.
     
  29. Terasaki H, Ishikawa K, Suzuki T, Nakamura M, Miyake K, Miyake Y. Morphologic and angiographic assessment of the macula after macular translocation surgery with 360 degrees retinotomy. Ophthalmology 2003;110(12):2403-8.
     
  30. De Juan E Jr, Loewenstein A, Bressler NM, Alexander J. Translocation of the retina for management of subfoveal choroidal neovascularization II: a preliminary report in humans. Am J Ophthalmol 1998;125(5):635-646.
     
  31. Pieramici DJ, de Juan E Jr, Fujii GY, Reynolds SM, Melia M, Humayun MS, Schachat AP, Hartranft CD. Limited inferior macular translocation for the treatment of subfoveal choroidal neovascularization secondary to age-related macular degeneration. Am J Ophthalmol 2000;130(4):419-428.
     
  32. Fujii GY, de Juan E Jr, Pieramici DJ, Humayun MS, Phillips S, Reynolds SM, Melia M, Schachat AP. Inferior limited macular translocation for subfoveal choroidal neovascularization secondary to age-related macular degeneration: 1-year visual outcome and recurrence report. Am J Ophthalmol 2002;134(1):69-74.
     
  33. Fujii GY, Pieramici DJ, Humayun MS, Schachat AP, Reynolds SM, Melia M, de Juan E Jr. Complications associated with limited macular translocation. Am J Ophthalmol 2000;130(6):751-762.
     
  34. Eandi CM, Giansanti F, Virgili G. Macular translocation for neovascular age-related macular degeneration. Cochrane Database Syst Rev 2008;(4):CD 006928.
     
  35. De Juan E Jr, Machemer R. Vitreous surgery for hemorrhagic and fibrous complications of age-related macular degeneration. Am J Ophthalmol 1988;105(1):25-29.
     
  36. Lambert HM, Capone A Jr, Aaberg TM, Sternberg P Jr, Mandell BA, Lopez PF. Surgical excision of subfoveal neovascular membranes in age-related macular degeneration. Am J Ophthalmol 1992;113(3):257-262.
     
  37. Berger AS, Kaplan HJ. Clinical experience with the surgical removal of subfoveal neovascular membranes. Short-term postoperative results. Ophthalmology 1992;99(6):969-975.
     
  38. Thomas MA, Grand MG, Williams DF, Lee CM, Pesin SR, Lowe MA. Surgical management of subfoveal choroidal neovascularization. Ophthalmology 1992;99(6):952-968.
     
  39. Bressler NM. Submacular surgery. Are randomized trials necessary? Arch Ophthalmol 1995;113(12):1557-1560.
     
  40. Bressler NM, Bressler SB, Hawkins BS, Marsh MJ, Sternberg P Jr, Thomas MA. Submacular surgery trials randomized pilot trial of laser photocoagulation versus surgery for recurrent choroidal neovascularization secondary to age-related macular degeneration: I. Ophthalmic outcomes submacular surgery trials pilot study report number 1. Am J Ophthalmol 2000;130(4):387-407.
     
  41. Conti SM, Kertes PJ. Surgical management of age-related macular degeneration. Can J Ophthalmol 2005;40(3):341-351.
     
  42. Ibanez HE, Williams DF, Thomas MA, Ruby AJ, Meredith TA, Boniuk I, Grand MG. Surgical management of submacular hemorrhage. A series of 47 consecutive cases. Arch Ophthalmol 1995;113(1):62-69.
     
  43. Wade EC, Flynn HW Jr, Olsen KR, Blumenkranz MS, Nicholson DH. Subretinal hemorrhage management by pars plana vitrectomy and internal drainage. Arch Ophthalmol 1990;108(7):973-978.
     
  44. Lewis H. Intraoperative fibrinolysis of submacular hemorrhage with tissue plasminogen activator and surgical drainage. Am J Ophthalmol 1994;118(5):559-568.
     
  45. Grossniklaus HE, Gass JD. Clinicopathologic correlations of surgically excised type 1 and type 2 submacular choroidal neovascular membranes. Am J Ophthalmol 1998;126(1):59-69.
     
  46. Grossniklaus HE, Green WR. Histopathologic and ultrastructural findings of surgically excised choroidal neovascularization. Submacular Surgery Trials Research Group. Arch Ophthalmol 1998;116(6):745-749.
     
  47. Soloman S, Dong LM, Haller JA, Gilson MM, Hawkins BS, Bressler NM; SST research group and the SST adverse event review committee. Risk factors for rhegmatogenous retinal detachment in the submacualr surgery trials. Retina 2009;29(6):819-824. 
     
  48. Hawkins BS, Bressler NM, Miskala PH, Bressler SB, Holekamp NM, Marsh MJ, Redford M, Schwartz SD, Sternberg P Jr, Thomas MA, Wilson DJ. Surgery for subfoveal choroidal neovascularization in age-related macular degeneration: ophthalmic findings: SST report no. 11. Ophthalmology 2004;111(11):1967-1980.
     
  49. Miskala PH, Bass EB, Bressler NM, Childs AL, Hawkins BS, Mangione CM, Marsh MJ. Surgery for subfoveal choroidal neovascularization in age-related macular degeneration: quality-of-life findings: SST report no. 12. Ophthalmology 2004;111(11):1981-1992.
     
  50. Bressler NM, Bressler SB, Childs AL, Haller JA, Hawkins BS, Lewis H, MacCumber MW, Marsh MJ, Redford M, Sternberg P Jr, Thomas MA, Williams GA. Surgery for hemorrhagic choroidal neovascular lesions of age-related macular degeneration: ophthalmic findings: SST report no. 13. Ophthalmology 2004;111(11):1993-2006.
     
  51. Childs AL, Bressler NM, Bass EB, Hawkins BS, Mangione CM, Marsh MJ, Miskala PH. Surgery for hemorrhagic choroidal neovascular lesions of age-related macular degeneration: quality-of-life findings: SST report no. 14. Ophthalmology 2004;111(11):2007-2014.
     
  52. Thumann G, Aisenbrey S, Schraermeyer U, Lafaut B, Esser P, Walter P, Bartz-Schmidt KU. Transplantation of autologous iris pigment epithelium after removal of choroidal neovascular membranes. Arch Ophthalmol 2000;118(10):1350-1355.
     
  53. Binder S, Stolba U, Krebs I, Kellner L, Jahn C, Feichtinger H, Povelka M, Frohner U, Kruger A, Hilgers RD, Krugluger W. Transplantation of autologous retinal pigment epithelium in eyes with foveal neovascularization resulting from age-related macular degeneration: a pilot study. Am J Ophthalmol 2002;133(2):215-225.
     
  54. Van Meurs JC, ter AE, Hofland LJ, van Hagen PM, Mooy CM, Baarsma GS, Kuijpers RW, Boks T, Stalmans P. Autologous peripheral retinal pigment epithelium translocation in patients with subfoveal neovascular membranes. Br J Ophthalmol 2004;88(1):110-113.
     
  55. Bindewald A, Roth F, Van MJ, Holz FG. Transplantation von retinalem Pigmentepithel (RPE) nach CNV-Exzision bei altersabhängiger Makuladegeneration. Techniken, Ergebnisse und Perspektiven. Ophthalmologe 2004;101(9):886-894.
     
  56. Aisenbrey S, Lafaut BA, Szurman P, Hilgers RD, Esser P, Walter P, Bartz-Schmidt KU, Thumann G. Iris pigment epithelial translocation in the treatment of exudative macular degeneration: a 3-year follow-up. Arch Ophthalmol 2006;124(2):183-188.
     
  57. Binder S, Krebs I, Hilgers RD, Abri A, Stolba U, Assadoulina A, Kellner L, Stanzel BV, Jahn C, Feichtinger H. Outcome of transplantation of autologous retinal pigment epithelium in age-related macular degeneration: a prospective trial. Invest Ophthalmol Vis Sci 2004;45(11):4151-4160.
     
  58. Peyman GA, Blinder KJ, Paris CL, Alturki W, Nelson NC Jr, Desai U. A technique for retinal pigment epithelium to extensive subfoveal scarring. Ophthalmic Surg 1991;22(2):102-108.
     
  59. Algvere PV, Berglin L, Gouras P, Sheng Y. Transplantation of fetal retinal pigment epithelium in age-related macular degeneration with subfoveal neovascularization. Graefes Arch Clin Exp Ophthalmol 1994;232(12):707–716.
     
  60. Algvere PV, Berglin L, Gouras P, Sheng Y, Kopp ED. Transplantation of RPE in age-related macular degenera- tion: Observations in disciform lesions and dry RPE atrophy. Graefes Arch Clin Exp Ophthalmol 1997;235(3):149–158.
     
  61. Stanga PE, Kychenthal A, Fitzke FW, Halfyard AS, Chan R, Bird AC, Aylward GW. Retinal pigment epithelium translocation after choroidal neovascular membrane removal in age-related macular degeneration. Ophthalmology 2002;109(8):1492-1498.
     
  62. MacLaren RE, Bird AC, Sathia PJ, Aylward GW. Long-term results of submacular surgery combined with macular translocation of the retinal pigment epithelium in neovascular age-related macular degeneration. Ophthalmology 2005;112(12):2081-2087.
     
  63. van Meurs JC, Van Den Biesen PR. Autologous retinal pigment epithelium and choroid translocation in patients with exudative age-related macular degeneration: short-term follow-up. Am J Ophthalmol 2003;136(4):688-695.
     
  64. Van Zeeburg EJ, Maaijwee KJ, Missotten TO, Heimann H, van Meurs JC. A free retinal pigment epithelium-choroid graft in patients with exudative age-related macular degeneration: Results up to 7 years. Am J Ophthalmol. 2012;153(1):120e7.e2.
     
  65. MacLaren RE, Uppal GS, Balaggan KS, Tufail AD, Ali RR, Aylward GW, DaCruz L. RPE Patch Graft Auto-Transplantation in Macular Degeneration: A Prospective Cohort Study. ARVO. Fort Lauderdale, USA, Apr 30-May 4, 2006. Invest Ophthalmol Vis Sci 2006; 47(13):2693.
      
  66. Joussen AM, Heussen FM, Joeres S, Llacer H, Prinz B, Rohrschneider K, Maaijwee KJ, Van MJ, Kirchhof B. Autologous translocation of the choroid and retinal pigment epithelium in age-related macular degeneration. Am J Ophthalmol 2006;142(1):17-30.
     
  67. Joussen AM, Joeres S, Fawzy N, Heussen FM, Llacer H, van Meurs JC, Kirchhof B. Autologous translocation of the choroid and retinal pigment epithelium in patients with geographic atrophy. Ophthalmology 2007;114(3):551-560.
     
  68. MacLaren RE, Uppal GS, Balaggan KS, Tufail A, Munro PM, Milliken AB, Ali RR, Rubin GS, Aylward GW, da CL. Autologous transplantation of the retinal pigment epithelium and choroid in the treatment of neovascular age-related macular degeneration. Ophthalmology 2007;114(3):561-570.
     
  69. Treumer F, Bunse A, Klatt C, Roider J. Autologous retinal pigment epithelium-choroid sheet transplantation in age related macular degeneration: morphological and functional results. Br J Ophthalmol 2007;91(3):349-353.
  70. Treumer F, Klatt C, Roider J. Autologe RPE-Chorioidea-Translokation bei exsudativer AMD. Eine Falldemonstration 10 konsekutiver Patienten. Ophthalmologe 2007;104(9):795-802.
  71. Maaijwee K, Heimann H, Missotten T, Mulder P, Joussen A, Van MJ. Retinal pigment epithelium and choroid translocation in patients with exudative age-related macular degeneration: long-term results. Graefes Arch Clin Exp Ophthalmol 2007;245(11):1681-1689.
  72. Maaijwee KJ, van Meurs JC, Kirchhof B, Mooij CM, Fischer JH, Mackiewicz J, Kobuch K, Joussen AM. Histological evidence for revascularisation of an autologous retinal pigment epithelium--choroid graft in the pig. Br J Ophthalmol 2007;91(4):546-550.
     
  73. Parolini, B, Alkabes, M, Baldi, A, Pinackatt, S. Visual recovery after autologous retinal pigment epithelium and choroidal patch in a patient with choroidal neovascularization secondary to angioid streaks: long-term results. Retin Cases Brief Rep. 2016 Fall;10(4):368-72.
     
  74. Maaijwee K, Van Den Biesen PR, Missotten T, Van Meurs JC. Angiographic evidence for revascularization of an rpe–choroid graft in patients with age-related macular degeneration. Retina 2008;28(3):498–503.
     
  75. Cereda MG, Parolini B, Bellesini E, Pertile G. Surgery for CNV and autologous choroidal RPE patch transplantation: exposing the submacular space. Graefes Arch Clin Exp Opththalmol 2010;248:37–47.
     
  76. Converse JM, Rapaport FT. The vascularisation of skin autografts and homografts. Ann Surg 1956;120:306–316.
     
  77. Bennett SR, Folk JC, Blodi CF, Klugman M. Factors prognostic of visual outcome in patients with subretinal hemorrhage. Am J Ophthalmol 1990;109(1):33-37.
     
  78. Berrocal MH, Lewis ML, Flynn HW Jr. Variations in the clinical course of submacular hemorrhage. Am J Ophthalmol 1996;122(4):486-493.
     
  79. Glatt H, Machemer R. Experimental subretinal hemorrhage in rabbits. Am J Ophthalmol 1982;94(6):762-773.
     
  80. Toth CA, Morse LS, Hjelmeland LM, Landers MB, III. Fibrin directs early retinal damage after experimental subretinal hemorrhage. Arch Ophthalmol 1991;109(5):723-729.
     
  81. Lewis H, Resuick SC, Flannery JG, Straatsma BR. Tissue plasminogen activator treatment of experimental subretinal hemorrhage. Am J Ophthalmol 1991;111:197-204.
     
  82. Peyman GA, Nelson NCJ, Alturki W, Blinder KJ, Paris CL, Desai UR, Harper CA 3rd. Tissue plasminogen activating factor assisted removal of subretinal hemorrhage. Ophthalmic Surg 1991;22(10):575-82.
     
  83. Kamei M, Tano Y. Tissue plasminogen activator-assisted vitrectomy: surgical drainage of submacular hemorrhage. Dev Ophthalmol 2009;44:82–88.
     
  84. Herriot WJ. Further experience in management of submacular hemorrhage with intravitreal tPA. Proceedings of the 1997 Update on Macular Surgery, American Academy of Ophthalmology. San Francisco, USA, October. 1997;82-84.
     
  85. Hesse L, Schmidt J, Kroll P. Management of acute submacular hemorrhage using recombinant tissue plasminogen activator and gas. Graefes Arch Clin Exp Ophthalmol 1999;237(4):273-277.
     
  86. Karlsson E, Carlsson J, Crafoord S, Jemt M, Martensson PA, Stenkula S. Tissue Plasminogen Activator and Expanding Gas Intravitreally in Treatment of Submacular Hemorrhage. Transactions of the Swedish Society of Ophthalmology 1997. Annual Meeting. Sundsvall, Sweden, August 27-30, 1997. Acta Ophthalmol Scand 1999;77(1): 119.
     
  87. Hassan AS, Johnson MW, Schneiderman TE, Regillo CD, Tornambe PE, Poliner LS, Blodi BA, Elner SG. Management of submacular hemorrhage with intravitreous tissue plasminogen activator injection and pneumatic displacement. Ophthalmology 1999;106(10):1900-1906.
     
  88. Handwerger BA, Blodi BA, Chandra SR, Olsen TW, Stevens TS. Treatment of submacular hemorrhage with low-dose intravitreal tissue plasminogen activator injection and pneumatic displacement. Arch Ophthalmol 2001;119(1):28-32.
     
  89. Hattenbach LO, Klais C, Koch FH, Gumbel HO. Intravitreous injection of tissue plasminogen activator and gas in the treatment of submacular hemorrhage under various conditions. Ophthalmology 2001;108(8):1485-1492.
     
  90. Ohji M, Saito Y, Hayashi A, Lewis JM, Tano Y. Pneumatic displacement of subretinal hemorrhage without tissue plasminogen activator. Arch Ophthalmol 1998;116(10):1326-1332.
     
  91. Daneshvar H, Kertes PJ, Leonard BC, Peyman GA. Management of submacular hemorrhage with intravitreal sulfur hexafluoride: a pilot study. Can J Ophthalmol 1999;34(7):385-388.
     
  92. Silva SR, Bindra MS. Early treatment of acute submacular haemorrhage secondary to wet AMD using intravitreal tissue plasminogen activator, C3F8, and an anti-VEGF agent. Eye (Lond) 2016;30(7):952-7.
     
  93. Kamei M, Misono K, Lewis H. A study of the ability of tissue plasminogen activator to diffuse into the subretinal space after intravitreal injection in rabbits. Am J Ophthalmol 1999;128(6):739-746.
     
  94. Haupert CL, McCuen BW, Jaffe GJ, Steuer ER, Cox TA, Toth CA, Fekrat S, Postel EA. Pars plana vitrectomy, subretinal injection of tissue plasminogen activator, and fluid-gas exchange for displacement of thick submacular hemorrhage in age-related macular degeneration. Am J Ophthalmol 2001;131(2):208-215.
     
  95. Olivier S, Chow DR, Packo KH, MacCumber MW, Awh CC. Subretinal recombinant tissue plasminogen activator injection and pneumatic displacement of thick submacular hemorrhage in Age-Related macular degeneration. Ophthalmology 2004; 111 (6): 1201-1208.
     
  96. Thompson JT, Sjaarda RN. Vitrectomy for the treatment of submacular hemorrhages from macular degeneration: a comparison of submacular hemorrhage/membrane removal and submacular tissue plasminogen activator-assisted pneumatic displacement. Trans Am Ophthalmol Soc. 2005;103:98-107, discussion 107.
     
  97. Fine HF, Iranmanesh R, Del Priore LV, Barile GR, Chang LK, Chang S, Schiff WM. Surgical outcomes after massive subretinal hemorrhage secondary to age-related macular degeneration. Retina. 2010;30(10):1588-94.
     
  98. Treumer F, Klatt C, Roider J, Hillenkamp J. Subretinal co-application of recombinant tissue plasminogen activator and bevacizumab for neovascular age-related macular degeneration with submacular hemorrhage. Br J Ophthalmol 2010;94(1):48-53. 
     
  99. Kamei M, Tano Y, Maeno T, Ikuno Y, Mitsuda H, Yuasa T. Surgical removal of submacular hemorrhage using tissue plasminogen activator and perfluorocarbon liquid. Am J Ophthalmol. 1996;121(3):267-75.
     
  100. Van Zeeburg EJT, van Meurs JC. Literature review of recombinant tissue plasminogen activator used for recent-onset submacular haemorrhage displacement in age-related macular degeneration. Ophthalmologica 2013;229(1):1-14.
     
  101. Kim JH, Chang YS, Kim JW, Kim CG, Yoo SJ, Cho HJ.. Intravitreal anti–vascular endothelial growth factor for submacular hemorrhage from choroidal neovascularization. Ophthalmology 2014;121(4):926–935.
     
  102. Simpson AR, Petrarca R, Jackson TL. Vitreomacular adhesion and neovascular age-related macular degeneration. Surv Ophthalmol. 2012;57(6):498-509.
     
  103. Treumer F, Roider J, Hillenkamp J. Long-term outcome of subretinal co-application of recombinant tissue plasminogen activator and bevacizumab followed by repeated intravitreal anti-VEGF injections for neovascular age-related macular degeneration with submacular hemorrhage. Br J Ophthalmol 2012;96(5):708-713.
     
  104. Stanescu-Segall D, Balta F, Jackson TL. Submacular hemorrhage in neovascular age-related macular degeneration: A synthesis of the literature. Surv Ophthalmol 2016;61(1):18-32.
     
  105. Martel JN, Mahmoud TH. Subretinal pneumatic displacement of subretinal hemorrhage. JAMA Ophthalmol 2013;131(12):1632-5.
     
  106. Ueda-Arakawa N, Tsujikawa A, Yamashiro K, Ooto S, Tamura H, Yoshimura N. Visual prognosis of eyes with submacular hemorrhage associated with exudative age-related macular degeneration. Jpn J Ophthalmol 2012;56(6): 589–598.
     
  107. Hirashima T, Moriya T, Bun T, Utsumi T, Hirose M, Oh H. Optical coherence tomography findings and surgical outcomes of tissue plasminogen activator-assisted vitrectomy for submacular hemorrhage secondary to age-related macular degeneration. Retina 2015;35(10):1969–78.
    ​​​​​​​
  108. Kitagawa Y, Shimada H, Mori R, Tanaka K, Yuzawa M.Intravitreal Tissue Plasminogen Activator, Ranibizumab, and Gas Injection for Submacular Hemorrhage in Polypoidal Choroidal Vasculopathy. Ophthalmology 2016; 123(6):1278-86.

AMD Future Perspectives: New promising drugs

AMD Future Perspectives: New promising drugs

English

Authors:

José Costa, MD MSc

Department of Ophthalmology. Centro Hospitalar e Universitario de Coimbra (CHUC). Portugal.

 

João Nascimento, MD

Ophthalmology Department. Hospital de Loures. Portugal

 

Susana Teixeira, MD

Ophthalmology Department. Hospital Amadora Sintra. Portugal

 

Rufino Silva, MD, PhD

Department of Ophthalmology. Centro Hospitalar e Universitario de Coimbra (CHUC). Portugal.

Faculty of Medicine, Institute for Biomedical Imaging and Life Sciences (IBILI), University of Coimbra, Coimbra, Portugal

Association for Innovation and Biomedical Research on Light and Image (AIBILI). Coimbra. Portugal

 

Updated/reviewed by the authors, July 2017.

Introduction

As researchers and clinicians begin to understand that wet age-related macular degeneration (wAMD) is a complex disease with angiogenic, vascular and inflammatory components, new pharmacological agents addressing multiple targets in its pathophysiology keep being introduced.

While some drugs are already available in human trials or even FDA and EMA-approved, most are still under laboratory investigation and little is known about them.

In this review, we will discuss promising emerging therapies for wAMD that aim to improve outcomes, safety and treatment burden through novel mechanisms of action. Due to its historical importance, we will also include the most significant drugs that did not achieve the desired clinical effects and never left the development pipeline.

For treatments for dry AMD, please refer to the chapter “Geographic Atrophy”.

 

VASCULAR ENDOTHELIAL GROWTH FACTOR INHIBITORS

Brolucizumab (RTH258, ESBA1008)

Mechanism of action: brolucizumab is a humanized single chain antibody fragment with high affinity to all VEGF-A isoforms. Its small molecular weight of 26 kDa (compared with 48 kDa for ranibizumab or 115 kDa for aflibercept) makes it possible to prepare higher molar concentrations and administer up to 6 mg of the drug in a single 0.05mL intravitreal injection(1), resulting in longer duration of effect, rapid systemic clearance, and better ocular tissue penetration(2)

Developer: Novartis®

Posology: 3.0-6.0 mg intravitreal injection q4w-q12w

Highest development phase: Phase III (completed)

Status: The OSPREY (NCT01796964) was a phase II trial that compared the safety and efficacy of 6mg brolucizumab to 2mg aflibercept in subjects with untreated active choroidal neovascularization due to AMD(3). Both groups received 3 monthly loading doses, were then treated every 8 weeks and assessed up to week 40. In the brolucizumab group, the final q8w cycle was extended to enable 2 cycles of treatment every 12 weeks until week 56. The mean BCVA change from baseline with brolucizumab was noninferior to aflibercept. During the q8w treatment cycles, the brolucizumab group had fewer unscheduled treatments than aflibercept (6 vs. 15) and more stable central subfield fluid thickness reductions. In addition, a greater proportion of brolucizumab treated eyes had resolved intraretinal and subretinal fluid compared with the aflibercept group. Approximately 50% of brolucizumab treated eyes had stable BCVA during the 12 week cycles and did not require unscheduled treatments(4).

Based on these results, more than 1800 patients were enrolled in two phase III studies: the HAWK (NCT02307682) and HARRIER (NCT02434328) trials. After three monthly loading doses, the brolucizumab groups were treated every 12 weeks, with the option of switching to 8-week dosing in case of disease activity. The aflibercept group was treated every 8 weeks. According to a press release from June 2017, brolucizumab was non-inferior to aflibercept through week 48, and 57% (HAWK) and 52% (HARRIER) of patients were maintained exclusively on a every 12 week interval immediately following the loading phase through week 48(5). Further results should be presented in the upcoming months.

 

Abicipar Pegol (AGN-150998, MP0112)

Mechanism of action: Abicipar is an anti-VEGF-A designed ankyrin repeat protein (DARPin), genetically modified antibody mimetic proteins that exhibit highly specific and high-affinity target protein binding proprieties.

Developer: Molecular Partners®, licensed to Allergan®

Posology: 2mg intravitreal injection q8w-q12w

Highest development phase: Phase III (active)

Status: Phase II data suggests that abicipar might have better efficacy and longer duration of effect than ranibizumab: 3 monthly doses of 1 or 2 mg abicipar compared with 5 monthly doses of ranibizumab led to an improvement of BCVA at least as significant as ranibizumab and possibly with a longer duration of action after 20 weeks(6). Two ongoing phase III trials (NCT02462928 and NCT02462486) are expected to be completed in 2019. These will compare 3 groups through 96 weeks: ranibizumab 0.5mg 4qw, abicipar 2mg q8w and abicipar 2mg q12w.

 

OPT-302 (soluble VEGFR-3; VGX 300)

Mechanism of action: OPT-302 is a soluble form of VEGFR-3 comprising the extracellular domains 1-3 of human vascular endothelial growth factor receptor (VEGFR)-3 and the Fc fragment of human IgG1. It binds and neutralizes the activity of VEGF-C and VEGF-D on endogenous VEGFR-2 and VEGFR-3, possibly resulting in a more complete blockade of the VEGF family when combined with anti-VEGF-A agents(7).

Developer: Ophtheaâ

Posology: intravitreal injection

Highest development phase: Phase IIA (completed)

Status: A phase I/IIA dose escalation study (NCT02543229) that included 20 patients met its primary objective, demonstrating OPT-302 safety and tolerability as monotherapy and in combination with ranibizumab. The combination group had better visual results and previous poor responders to anti-VEGF-A therapy fared better with the addition of OPT-302(8). The company is planning a phase IIB study for the 2nd semester of 2017.

 

PF-655 (PF-04523655, REDD14NP, RTP801i)

Mechanism of action: A siRNA is a protein that binds to and activates the RNA-induced silencing complex (RISC) that then degrades mRNA molecules. A single activated RISC can destroy hundreds of mRNAs, thus preventing translation and protein synthesis. Therefore, siRNA might be an efficient mechanism for preventing the expression and synthesis specific harmful proteins(9). PF-655 is a siRNA that inhibits expression of the hypoxia-inducible gene RTP801, which in turn inhibits the mammalian target of rapamycin (mTOR) pathway and reduces VEGF-A production(10).

Developer: Quark Pharmaceuticalsâ, licensed to Pfizer®

Posology: intravitreal injection q4w

Highest development phase: Phase II (completed)

Status: PF-655 showed efficacy in halting angiogenesis in animal models and its effect appears to be complementary to the anti-VEGF mechanisms of ranibizumab and bevacizumab.(11) The Phase II MONET trial (NCT00713518) evaluated the efficacy of different dosing paradigms of PF-655 versus ranibizumab in 151 subjects with wAMD(12). The BCVA change from baseline in the combination group was greater than in the ranibizumab group (9.5 vs 6.8 letters, respectively), but the difference was not statistically significant. The combination and ranibizumab groups had similar mean reductions in central subfield retinal thickness and total CNV area. A phase II clinical trial focused on Diabetic Macular Edema, the MATISSE Trial (NCT01445899), has been completed but no results have been published(13).

 

Rabinizumab Port Delivery System

Mechanism of action: The Ranibizumab Port Delivery System (RPDS) is a refillable reservoir system with a diffusion-control mechanism designed to gradually release ranibizumab. It is placed under the conjunctiva, fixed to the pars plana, and no sutures are needed. The port is then refilled as an in-office procedure that uses a proprietary refill needle system that simultaneously introduces drug into the reservoir and removes any remaining contents.

Developer: ForSight Vision4®, licensed to Genentech®

Posology: subconjunctival implant refilled with ranibizumab as needed (estimated every 4-6 months)

Highest development phase: Phase II (recruiting)

Status: A proof-of-concept phase I trial enrolled 20 patients. There were 3 patients with serious adverse effects, all associated with the placement of the RPDS – a case of endophthalmitis and two cases of persistent vitreous haemorrhage. Over 12 months, the average gain of BCVA was 15 letters if the 2 nonevaluable patients with persistent vitreous hemorrhage were excluded. The average number of refills for the full 20-patient cohort was 4.8(14). A phase II LADDER trial (NCT02510794) is currently recruiting and planning to enrol 220 patients. It will evaluate the efficacy, safety and pharmacokinetics of three different formulations of ranibizumab delivered via the RPDS.

 

Aflibercept Hydrogel Depot

Mechanism of action: hydrogel-based drug delivery depot that gradually dissolves and releases aflibercept for up to 6 months(15).

Developer: Ocular Therapeutixâ, licensed to Regeneronâ

Posology: intravitreal injection

Highest development phase: Preclinical

Status: Phase I studies are in the planning stages(16).

 

PAN-90806

Mechanism of action: topical VEGF-A inhibitor

Developer: PanOptica

Posology: daily topical administration

Highest development phase: Phase I/II (completed)

Status: A phase I/II trial (NCT02022540) included 50 treatment-naïve patients with wAMD that received topical PAN-90806 for 8 weeks. According to a press release from the study sponsor, there was positive biological response to topical PAN-90806 in 45-50% of treated patients, including outcomes such as vascular leakage, lesion morphology and vision. The most common adverse effect was punctate keratopathy(17). Another phase I/II clinical trial, using a new drug formulation aiming to reduce the corneal toxicity, is planned for 2017.

 

LHA-510

Mechanism of action: topical VEGF inhibitor

Developer: Novartis®

Posology: daily topical administration

Highest development phase: Phase II (active)

Status: A phase II trial (NCT02355028) was designed to evaluate the efficacy and tolerability of LHA-510 as maintenance therapy for 84 days after ranibizumab injection. The study has reached its primary completion date in September 2016 but no results have yet been published.

 

ANTI-PLATELET DERIVED GROWTH FACTOR INHIBITORS

Anti-VEGF monotherapy in wAMD markedly reduces the hyperpermeability of the neovascular membranes but does not lead to long term regression of the neovascular complexes(18). It has been proposed that pericytes may play an important role in this anti-VEGF resistance as they protect the endothelial cells when exposed to anti-VEGF. Pericytes share a common basement membrane with endothelial cells and provide them with VEGF-A and other growth factors via paracrine and juxtacrine signals(19). Additionally, pericytes are a major source of myofibroblasts and appear to be involved in the development of subretinal fibrosis associated with CNVs(20).

Platelet derived growth factor (PDGF) is a protein that binds to a tyrosine kinase receptor on pericytes and is critical for their survival, recruitment, and maturation(21). Therefore, inhibiting PDGF in order to strip pericytes from the neovacular complexes might be an adequate strategy to improve long-term visual outcomes in patients with wAMD, as it might leave the underlying endothelial cells more sensitive to the effects of VEGF blockade(22).

 

Pegpleranib (Fovista®, E-10030; E01AJ; OAP-030; X01E)

Mechanism of action: pegylated DNA aptamer that selectively binds to PDGF-BB and PDGF-AB, thereby disrupting the interaction with their tyrosine kinase receptors in the perycites and leading to perycite stripping from the underlying neovascular complex(23).

Developer: Ophthotech®, licensed to Novartis®

Posology: 1.5mg intravitreal injection

Highest development phase: Phase III (active)

Status: The OPH1002 (NCT01944839) and OPH1003 (NCT01940900) were two randomized, double blind, Phase III clinical trial studies designed to evaluate the safety and efficacy of pegpleranib in combination with ranibizumab versus ranibizumab monotherapy in patients with subfoveal wAMD. A total of 1,248 patients were enrolled across both studies. The studies did not meet the primary endpoint of superiority for the pegpleranib and ranibizumab combination therapy, measured as additional letter gains over ranibizumab monotherapy(24).
The results of these trials have decreased interest on pegpleranib. Still, a third phase III trial, the OPH1004 (NCT01940887), is underway. It is evaluating the safety and efficacy of 1.5mg pegpleranib administered in combination with 2.0mg aflibercept or 1.25mg bevacizumab anti-VEGF therapy compared to both in monotherapy for the treatment of wAMD. Results are expected in the second half of 2017.

 
Rinucumab (REGN2176-3)

Mechanism of action: monoclonal anti-PDGF receptor beta antibody, formulated with aflibercept in a single injection

Developer: Regeneron Pharmaceuticals®

Posology: 1.0-3.0mg monthly intravitreal injection with 3mg of aflibercept

Highest development phase: Phase II (completed)

Status: The Phase II CAPELLA study (NCT02418754), completed in August 2016, evaluated aflibercept co-formulated with rinucumab in 500 patients with wet AMD. Adding rinucumab to aflibercept showed no benefit on BCVA or on anatomic endpoints, including reduction in retinal thickness or in resolution of subretinal hyper-reflective material(25).

 

DE-120

Mechanism of action: double kinase receptor inhibitor that blocks both VEGF-A and PDGF.

Developer: Santen Pharmaceuticals®

Posology: intravitreal injection 

Highest development phase: Phase II (completed)

Status: VAPOR1 (NCT02401945), a phase II study assessing the efficacy, safety and duration of effect of DE-120 as monotherapy and after induction with aflibercept as reached its primary completion date in June 2016. No results have yet been published.

 

Vorolanib (X-82, CM-082)

Mechanism of action: oral tyrosine kinase inhibitor (derived from anti-cancer agent sunitibib) that blocks the kinase activity of all receptor subtypes for VEGF and PDGF(26)

Developer: Tyrogenex®

Posology: 50-200mg daily oral administration

Highest development phase: Phase II (active)

Status: A phase I trial (NCT02452385) enrolled 35 patients, most (n=29) with heavily treated refractory wAMD, that were given 4 dose levels of vorolanib and were studied for up to 6 months.  Anti-VEGF rescue treatment was administered if their macular thickness increased by 50 µm or if they experienced a 1-line decrease in BCVA. Twenty-five patients completed the follow-up period and 60% of these did not require anti-VEGF recue treatment - the mean increase in BCVA was 5.3 letters. Of the 10 patients that failed to complete the 24 weeks, 5 were due to adverse effects (transient transaminase elevations and gastrointestinal symptoms)(26). Based on these results, a phase II APEX trial (NCT02348359) comparing daily voloranib with PRN aflibercept vs. PRN aflibercept monotherapy is underway and is estimated to be completed in January 2018.

 

ANGIOPOIETIN PATHWAY INHIBITORS

In the past years, the angiopoietin pathway has been evaluated as a target to improve outcomes of ocular vascular pathologies. Angiopoietin-2 (ANG2) is essential for retinal vascular development and causes disruption of the blood-retinal barrier, increasing vascular permeability. Similar to what occurs with VEGF, ANG2 is expressed in retinal and choroidal neovascularization, and ischemic/hypoxic retina increases its production(27). ANG2 is significantly elevated in the vitreous of patients with retinal vascular pathologies, such as wAMD, diabetic retinopathy and retinal vein occlusion(28), and high-levels of VEGF at the inner surface of the retina do not cause retinal new vessels unless accompanied by increased expression of ANG2(29)

ANG2’s effect is due to inhibition of the Tie2 tyrosine kinase, a receptor for the angiopoietin family of proteins(30). Conversely, angiopoietin 1 (ANG1) binds to and activates Tie2, which stabilizes blood vessels(31). ANG2 competes with ANG1 for Tie2 binding, consequently increasing the responsiveness of retinal vessels to VEGF and promoting vascular leakage and neovascularization. On the contrary, ANG1 reduces responsiveness to VEGF and reduces vascular leakage(32). Therefore, inhibiting ANG2 and/or activating Tie2 are potential strategies to treat wAMD.

 

Nesvacumab (REGN-910-3)

Mechanism of action:  human IgG1 monoclonal antibody directed against ANG2 that blocks its interaction with the Tie2 receptor(33).

Developer: Regeneron®, licensed to Bayer® outside the USA

Posology: intravitreal injection q4w

Highest development phase: Phase II (active)

Status: A phase I study (NCT01997164) including 20 patients with wAMD and diabetic macular oedema showed no significant adverse effects and good functional and anatomical outcomes(34). Phase II trials for wAMD (ONYX, NCT02713204) and diabetic macular oedema (RUBY, NCT02712008) using nesvacumab combined with aflibercept are ongoing and should be completed in the 2nd semester of 2017.

 

RG-7716 (RO-6867461)

Mechanism of action: bispecific human IgG1 monoclonal antibody that binds to both ANG2 and VEGF-A(35).

Developer: Chugai Pharmaceutical®, licensed to Genentech®

Posology: 0.5-6.0mg intravitreal injection starting at q4w

Highest development phase: Phase II (active)

Status: A Phase I (NCT01941082) included 26 patients with persistent CNV activity despite =3 anti-VEGF injections. The drug was generally well tolerated and improvements in BCVA and OCT parameters were observed(36). The AVENUE (NCT02484690) and STAIRWAY (NCT03038880) studies are ongoing phase II trials that will evaluate the efficacy and tolerability of different dosages and treatment intervals of RG-7716 in patients with treatment-naïve wAMD.

 

ARP-1536

Mechanism of action: monoclonal antibody that inhibits the catalytic activity of the VE-PTP enzyme (vascular endothelial protein tyrosine phosphatase), a negative regulator of the Tie2 receptor, therefore decreasing vascular permeability(32).

Developer: Akebia?, licensed to Aerpio Pharmaceuticals®

Posology: intravitreal injection 

Highest development phase: Preclinical

Status: Phase I trials are projected to start in the 1st quarter of 2018.

 

AKB-9778

Mechanism of action: subcutaneously administered VE-PTP enzyme inhibitor(32).

Developer: Akebia®, licensed to Aerpio Pharmaceuticals®

Posology: daily subcutaneous administration

Highest development phase: Preclinical stages

Status: AKB-9778 for wAMD is currently in preclinical stages. Phase II trials to evaluate its efficacy were first started for DMO (NCT03197870, recruiting) and RVO (NCT02387788, completed). 

 

ALTERNATIVE TARGETS

ICON-1 (hI-con1)

Mechanism of action: chimeric protein designed to bind to and inhibit Tissue Factor (TF) with higher affinity than monoclonal antibodies. This, in turn, triggers NK-cells to selectively destroy overexpressing TF tissues, such as pathologic neovascularization(37).

Developer: Iconic Therapeutics®

Posology: 0.3mg intravitreal injection

Highest development phase: Phase II

Status: The EMERGE study was a phase II trial (NCT02358889) that randomized patients to 3 groups: ICON-1 monotherapy, ranibizumab monotherapy or combination treatment. Patients received 3 initial monthly injections, then remained masked in their respective randomized group for additional 3 possible injections based on protocol retreatment criteria. The study results were presented in February 2017: mean BCVA increased from baseline to month 3 by 0.3 letters with ICON-1 monotherapy, 6.8 letters with ICON-1 in combination with ranibizumab, and 7.6 letters with ranibizumab monotherapy and was associated with central retinal thickness reduction in all treatment groups. BCVA gain was comparable in the ICON-1 combination and ranibizumab groups, although it was maintained with fewer treatments from month 3 to 6 with ICON-1 combination(38).

 

Carotuximab (DE-122, TRC105)

Mechanism of action: chimeric antibody to endoglin/CD105, a transmembrane glycoprotein that is expressed on proliferating endothelial cells and functions as a co-receptor of the transforming growth factor ß (TGF-ß) superfamily, which makes it essential for angiogenesis(39).

Developer: TRACON Pharmaceuticals®, licensed to Santen®

Posology: intravitreal injection

Highest development phase: Phase II (recruiting)

Status: A phase II trial has begun recruiting patients in July 2017 (NCT03211234).

 

Conclusion

Despite all the efforts that the scientific community has directed towards neovascular AMD research, the treatment options currently available remain suboptimal. Intravitreal injections, the current standard of care, are invasive and expensive procedures that require frequent patient monitoring. We have presented the most important drugs currently in development and, based on the data that we presented, Tie2 receptor modulation appears to be the most promising strategy to improve the visual outcomes of patients with AMD. Still, drug research & development is an ever-evolving field and several more therapies, such as complement pathway modulators, gene therapy and cell implants, are in the pipeline, promising future breakthroughs.

 

>> References

References - AMD Future Perspectives: New promising drugs

References - AMD Future Perspectives: New promising drugs

English

  1. Holz FG, Dugel PU, Weissgerber G, et al. Single-Chain Antibody Fragment VEGF Inhibitor RTH258 for Neovascular Age-Related Macular Degeneration: A Randomized Controlled Study. Ophthalmology. 2016 May;123(5):1080-9.
  2. Hussain RM, Ciulla TA. Emerging vascular endothelial growth factor antagonists to treat neovascular age-related macular degeneration. Expert Opin Emerg Drugs. 2017 Jul 31.
  3. Singerman LJ WA, Sallstig P. OSPREY trial: randomized, active-controlled, phase II study to evaluate safety and efficacy of RTH258, a humanized single-chain anti-VEGF antibody fragment, in patients with neovascular AMD. Association for Research in Vision and Ophthalmology Annual Meeting; May 3-7, 2015; Denver, CO.
  4. Dugel PU, Jaffe GJ, Sallstig P, et al. Brolucizumab Versus Aflibercept in Participants with Neovascular Age-Related Macular Degeneration: A Randomized Trial. Ophthalmology. 2017 May 24.
  5. Novartis RTH258 (brolucizumab) demonstrates robust visual gains in nAMD patients with a majority on a 12-week injection interval. 2017 [05/07/2017]; Available from: https://www.novartis.com/news/media-releases/novartis-rth258-brolucizumab-demonstrates-robust-visual-gains-namd-patients.
  6. Positive Top-Line Data Reported from Phase 2 Study for DARPin abicipar pegol in wet AMD. 2014 [10/07/2017]; Available from: https://www.molecularpartners.com/positive-top-line-data-reported-from-phase-2-study-for-darpin-abicipar-pegol-in-wet-amd/.
  7. Study Evaluating the Safety, Pharmacokinetics and Pharmacodynamics of OPT-302 With or Without Lucentis™ in Patients With Wet AMD. [updated 04/04/201710/07/2017]; Available from: https://clinicaltrials.gov/show/NCT02543229.
  8. Opthea Reports Positive Data from wet AMD Clinical Trial 2016 [updated 26/07/2016]; Available from: http://www.opthea.com/pub/pdf/ASX_OPT_OPT302 Ph 1 Secondary Endpoints_FINAL_16 07 26_Aust.pdf.
  9. Garba AO, Mousa SA. Bevasiranib for the treatment of wet, age-related macular degeneration. Ophthalmol Eye Dis. 2010;2:75-83.
  10. Lee M, Bikram M, Oh S, Bull DA, Kim SW. Sp1-dependent regulation of the RTP801 promoter and its application to hypoxia-inducible VEGF plasmid for ischemic disease. Pharm Res. 2004 May;21(5):736-41.
  11. Brafman A, Mett I, Shafir M, et al. Inhibition of oxygen-induced retinopathy in RTP801-deficient mice. Invest Ophthalmol Vis Sci. 2004 Oct;45(10):3796-805.
  12. Nguyen QD, Schachar RA, Nduaka CI, et al. Evaluation of the siRNA PF-04523655 versus ranibizumab for the treatment of neovascular age-related macular degeneration (MONET Study). Ophthalmology. 2012 Sep;119(9):1867-73.
  13. Titze-de-Almeida R, David C, Titze-de-Almeida SS. The Race of 10 Synthetic RNAi-Based Drugs to the Pharmaceutical Market. Pharm Res. 2017 Jul;34(7):1339-63.
  14. Laganovska G. First-in-Human Results of a Refillable Drug delivery Implant Providing Release of Ranibizumab in Patients with Neovascular AMD. Anual Meeting AAO; 09/12/2012; Chicago, IL2012.
  15. 2017 [16/07/2017]; Available from: http://www.ocutx.com/pipeline/intravitreal-depots.
  16. Ocular Therapeutix™ and Regeneron Enter Into Strategic Collaboration to Develop Sustained Release Formulation of Aflibercept for the Treatment of Wet AMD and Other Serious Retinal Diseases. 2016 [16/07/2017]; Available from: http://investors.ocutx.com/phoenix.zhtml?c=253650&p=irol-newsArticle&ID=2211419.
  17. PanOptica Reports Positive Results from Phase 1/2 Clinical Trial of PAN-90806, a Novel Topical Anti-VEGF Eye Drop. [cited 2016 29/06/2017]; Available from: http://www.panopticapharma.com/docs/news/Positive-Results-from-Phase-1-2-Clinical-Trial-of-PAN-90806.pdf.
  18. Jaffe GJ, Eliott D, Wells JA, Prenner JL, Papp A, Patel S. A Phase 1 Study of Intravitreous E10030 in Combination with Ranibizumab in Neovascular Age-Related Macular Degeneration. Ophthalmology. 2016 Jan;123(1):78-85.
  19. Armulik A, Abramsson A, Betsholtz C. Endothelial/pericyte interactions. Circ Res. 2005 Sep 16;97(6):512-23.
  20. Humphreys BD, Lin SL, Kobayashi A, et al. Fate tracing reveals the pericyte and not epithelial origin of myofibroblasts in kidney fibrosis. Am J Pathol. 2010 Jan;176(1):85-97.
  21. Andrae J, Gallini R, Betsholtz C. Role of platelet-derived growth factors in physiology and medicine. Genes Dev. 2008 May 15;22(10):1276-312.
  22. Erber R, Thurnher A, Katsen AD, et al. Combined inhibition of VEGF and PDGF signaling enforces tumor vessel regression by interfering with pericyte-mediated endothelial cell survival mechanisms. FASEB J. 2004 Feb;18(2):338-40.
  23. Mitchell TS, Bradley J, Robinson GS, Shima DT, Ng YS. RGS5 expression is a quantitative measure of pericyte coverage of blood vessels. Angiogenesis. 2008;11(2):141-51.
  24. Ophthotech Announces Results from Pivotal Phase 3 Trials of Fovista® in Wet Age-Related Macular Degeneration. 2016 [02/07/2017]; Available from: http://investors.ophthotech.com/releasedetail.cfm?ReleaseID=1003651.
  25. Regeneron Announces Phase 2 Study of Aflibercept Co-Formulated with Rinucumab (anti-PDGFR-beta) Shows No Benefit Over Aflibercept Alone in Neovascular Age-Related Macular Degeneration. 2016 [04/07/2017]; Available from: http://www.prnewswire.com/news-releases/regeneron-announces-phase-2-study-of-aflibercept-co-formulated-with-rinucumab-anti-pdgfr-beta-shows-no-benefit-over-aflibercept-alone-in-neovascular-age-related-macular-degeneration-300337055.html.
  26. T. J. A phase 1 study of oral tyrosine kinase inhibitor (X-82) in previously treated wet age-related macular degeneration. 15th EURETINA Congress; 17-20/09/2015; Nice, France2015.
  27. Shen J, Frye M, Lee BL, et al. Targeting VE-PTP activates TIE2 and stabilizes the ocular vasculature. J Clin Invest. 2014 Oct;124(10):4564-76.
  28. Koss MJ, Pfister M, Koch FH. Inflammatory and angiogenic protein detection in the human vitreous: cytometric bead assay. J Ophthalmol. 2011;2011:459251.
  29. Oshima Y, Takahashi K, Oshima S, et al. Intraocular gutless adenoviral-vectored VEGF stimulates anterior segment but not retinal neovascularization. J Cell Physiol. 2004 Jun;199(3):399-411.
  30. Maisonpierre PC, Suri C, Jones PF, et al. Angiopoietin-2, a natural antagonist for Tie2 that disrupts in vivo angiogenesis. Science. 1997 Jul 04;277(5322):55-60.
  31. Thurston G, Suri C, Smith K, et al. Leakage-resistant blood vessels in mice transgenically overexpressing angiopoietin-1. Science. 1999 Dec 24;286(5449):2511-4.
  32. Campochiaro PA, Sophie R, Tolentino M, et al. Treatment of diabetic macular edema with an inhibitor of vascular endothelial-protein tyrosine phosphatase that activates Tie2. Ophthalmology. 2015 Mar;122(3):545-54.
  33. Papadopoulos KP, Kelley RK, Tolcher AW, et al. A Phase I First-in-Human Study of Nesvacumab (REGN910), a Fully Human Anti-Angiopoietin-2 (Ang2) Monoclonal Antibody, in Patients with Advanced Solid Tumors. Clin Cancer Res. 2016 Mar 15;22(6):1348-55.
  34. Brown DM. Phase 1 study of combination therapy with nesvacumab and aflibercept for neovascular AMD and diabetic macular edema. AAO Annual Meeting; October 2016; Chicago, IL2016.
  35. Regula JT, Lundh von Leithner P, Foxton R, et al. Targeting key angiogenic pathways with a bispecific CrossMAb optimized for neovascular eye diseases. EMBO Mol Med. 2016 Nov;8(11):1265-88.
  36. P. Dugel DS, P. Cech, M. Giraudon, A. Boulay, P. Szczesny, J. Sahni. A novel anti-VEGF/anti-angiopoietin2 bispecific monoclonal antibody for wet age-related macular degeneration and diabetic macular edema. 16th EURETINA Congress; 8-11/09/20162016.
  37. Cocco E, Hu Z, Richter CE, et al. hI-con1, a factor VII-IgGFc chimeric protein targeting tissue factor for immunotherapy of uterine serous papillary carcinoma. Br J Cancer. 2010 Sep 07;103(6):812-9.
  38. Christine R. Gonzales GB. A Phase 2 Study (EMERGE) Evaluating Repeated Intravitreal Administration of ICON-1 in Patients With Choroidal Neovascularization (CNV) Secondary to Age-related Macular Degeneration (AMD). ARVO 2017; Baltimore2017.
  39. Saito T, Bokhove M, Croci R, et al. Structural Basis of the Human Endoglin-BMP9 Interaction: Insights into BMP Signaling and HHT1. Cell Rep. 2017 May 30;19(9):1917-28.

Optical Coherence Tomography in Age-related Macular Degeneration

Optical Coherence Tomography in Age-related Macular Degeneration

English

Authors:

Garcia-Layana, Alfredo1

Ciuffo, Gianfranco1

Javier Zarranz-Ventura2

Alvarez-Vidal, Aurora1
 

1 Ophthalmology Department. Clínica Universidad de Navarra.
Pamplona. Spain

2 Institut Clínic d´Oftalmologia (ICOF); Hospital Clinic Barcelona; Barcelona; Spain

Updated/reviewed by the authors, July 2017.

Introduction

Optical coherence tomography (OCT) technology allows the acquisition of cross-sectional images of the retina with semihistologic resolution.

It permits to define the location and nature of the changes in the retina and adjacent structures and objectively evaluates the thickness of the retina and surrounding structures.

These capabilities allow detection of newly emerging fluid and/or intraretinal or subretinal tissue and tissue below the retinal pigment epithelium
c. Neurosensory Retinal Detachment (RPE).

OCT provides important information on serous retinal detachments, hemorrhages and subretinal neovascular membranes that are components of exudative macular degeneration and allows a more precise and detailed analysis of anatomic structures and neovascular membrane lesions subtypes.

These capabilities facilitate an understanding of the differences between the classic membranes, occult membranes, retinal angiomatous proliferation (RAP) and disciform scars in the natural course of the disease and assess the response to anti-vascular endothelial growth factor (VEGF) drugs.

The images are presented in a color (or gray) scale based on the different reflectivity of the tissue structures.

More concisely, tissues that reflect more light or disperse more light are shown in red and white, respectively, while the ones that reflect or disperse less light are shown in blue and black.

Tissues that moderately reflect light are shown as green or yellow. It should be noted that the color shown in the images represents the optical properties of the tissues and not the tissues themselves.

Therefore, the image is not real but represents the true dimensions of the measured structures(1-10)

 

Normal eye

The vitreous transmits light without reflecting it and is depicted in black in OCT images (Figure 1). The posterior hyaloid usually is indistinguishable from the retinal surface except when the posterior vitreous is detached and appears as a weakly reflective band. The choriocapillaris and choroid are highly reflective layers, because they are vascular and limit light penetration into the deeper layers. Blood vessels are identified by their high reflectivity and the masking effect generated on adjacent tissues (Table 1) (7,8,11)

Figure 1

Figure 1. A normal eye.

 

 
Table 1. Reflectivity Patterns of the Retinal Structures in Normal Eyes
 
 
Hyperreflectivity pattern
 
 
Nerve fiber layer, thicker in the nasal portion (red), RPE and choriocapillaris (red)
 
Normalreflectivity pattern
 
 
Inner and outer plexiform layer (yellow-green), inner and outer nuclear layers (blue)
 
Hyporeflectivity pattern  
 
 
Photoreceptor layer

 

Age-Related Maculopathy

Age-related maculopathy, which is currently considered a previous stage of age-related macular degeneration (AMD), is defined as the presence of areas of hyperpigmentation or hypopigmentation of the RPE and/or confluent or soft drusen. When soft drusen in the macular region are associated with focal areas of pigmentary changes (hypopigmentation and hyperpigmentation), there is an increased risk of progression to AMD. Drusen, which are degenera
c. Neurosensory Retinal Detachmenttive nodular formations located mainly in Bruch’s membrane, are accumulations of proteins, lipids, mucopolysaccharides and other components that appear in adulthood and tend to increase in size and number over time. Drusen is the earliest AMD sign that is detected clinically in fundus examinations. On OCT, drusen appear as RPE deformation or thickening that may form irregularities and undulations (Figure 2).

Drusen are classified histologically as hard formations, defined as small hyaline deposits with delimited margins that are considered age-related low-risk changes, and soft drusen, which are deposits of granular or amorphous material considered to be precursors of AMD. Drusen can progress to an atrophic form (dry) or exudative form (wet) of AMD. The risk is based on the lesion number, size and confluence of drusen. (4,6,8,12-15)

Figure 2. Drusen are seen as undulations and elevations in the hyperreflective band of the RPE with less reflective material beneath them, while the inner retinal layers remain generally intact.

 

Atrophic Macular Degeneration

In the macular region, progressive atrophy of the RPE, the outer retinal layers, the choriocapillaris, and dense clusters of In the macular region, progressive atrophy of the RPE, the outer retinal layers, the choriocapillaris, and dense clusters of drusen can be seen.

OCT can detect decreases in retinal thickness and increases in the reflectivity of the RPE detachment, which results from the decreased ability of the atrophic retinal tissue to attenuate light and allows greater penetration of the laser beam to deeper structures as the choroid (Figure 3). The reduced retinal thickness and volume can be determined by the retinal maps that display the areas of greatest atrophy, identify the extent of the atrophy, and monitor progression. Geographic atrophy (GA) represents the final stage of dry AMD. (6,8,10,16,17)

Figure 3

Figure 3. In a case of retinal atrophy, the OCT shows a highly reflective choroidal signal due to retinal thinning and RPE hypopigmentation, which allows greater beam penetration into the choroid and greater reflectivity. The retinal map allows quantification of the decreased retinal thickness.

 

Exudative Macular Degeneration

OCT, a fundamental tool in the diagnosis and management of patients with choroidal neovascularization (CNV), allows identification of active neovascular membranes and determination of the extent of the membranes in many cases. The technology is useful for the assessment of the subfoveal region and also helps diagnose occult choroidal neovascular (CNV) membranes, in which fluorescein angiography (FA) often shows confusing patterns. OCT has also been proven useful for monitoring treatment response to anti-VEGF agents and is a reliable tool to determine the need for retreatment in pro re nata (PRN) treatment regimes. (3,8,15,18-23)

Well-defined classic CNV appears on OCT as hyperreflective areas in contact with or in front of the RPE; the pathology may be dome-shaped or appear as a thin formation (fusiform or nodular) (Figure 4). Retinal edema is always present to some extent in front of the active membrane; if the retina is thinner than normal, new vessels may be latent. CNV is less evident a few weeks after onset, and only interruption, breakdown and pronounced thickening of the RPE can be seen.

Figure 4. OCT visualizes the components of the neovascular membrane, a RPE detachment (PED), neurosensorial detachment, intraretinal fluid and subretinal haemorrhage. D: Drusenoid PEDs (∆), subretinal hemorrhage (*), dense particles present in the subretinal fluid (↓), intraretinal migration of RPE cells (>).

 

RPE Detachment

a. Hemorrhagic Detachment

An RPE detachment can be hemorrhagic in the presence of neovascular membranes. Retinal hemorrhages are often seen under the RPE. Detachment of the RPE forms a steep angle to the choriocapillaris, and the accumulated blood cells block penetration of the light rays from OCT in the area of the detachment, so the penetration is minimal and forms a shadow that hides the choriocapillaris and other subsequent layers (Figure 5).
 
 

Figure 5. Hemorrhagic detachment of the RPE.

 

b. Serous Detachment

During exudative AMD development, OCT shows serous detachments of the RPE (pigment epithelium detachments, PEDs) as optically clear areas between the RPE and the choriocapillaris (Figure 6). The PEDs form a steep angle with the underlying choriocapillaris, which may be initially clear in an early stage and becomes blurred at a later stage as a consequence of the different cells and aggregates that are released to the virtual space created under the detached RPE. A PED also can be single, multiple, dome-shaped or bilobulated.(6,8,24)
 

Figure 6. Serous detachment of the RPE associated with a neurosensory retinal detachment.

 

c. Neurosensory Retinal Detachment

Active CNV membranes cause edema and small serous detachments of the neurosensory retina. Eventually, CNV membranes with large PEDs may evolve to pigment epithelial rips (Figure 7).
 
 
Figure 7. Neurosensory retinal detachment.
 
 

Intraretinal Fluid

Intraretinal fluid can occur either diffusely, creating increased retinal thickness and reduced retinal reflectivity, or appear localized in non-reflective well defined cysts (cystic macular edema) (Figure 8).

Figure 8. Intraretinal cysts.

 

RPE Rips

RPE rips are serious complications of AMD and appear during the evolution of serous or serohemorrhagic PEDs (Figure 9). RPE rips also can be secondary to photodynamic therapy (PDT) and laser photocoagulation. Patients with occult CNV membranes and PEDs are at high risk of developing RPE rips, being the PED base diameter and the PED height the main risk factors for this complication, which is estimated to appear in 17% of the cases.(6,8,25)

Figure 9. RPE tears, sudden disruptions in the continuity of the hyperreflectivity layer of the pigment epithelium, show strong reflectivity from the choroid in the absence of the pigment epithelium, followed by sudden interruption of the band’s pigment epithelial hyperreflectivity. C: (↓) Intraretinal fluid. D: (↓) RPE interruption site, (*) RPE tear bent over itself.

 

Polypoidal Choroidal Vasculopathy

Polypoidal choroidal vasculopathy (PCV) is considered a variant of neovascular AMD with serous or hemorrhagic detachment of the retina and the RPE in the posterior pole. The pathology is seen on FA and especially on indocyanine green angiography (ICGA). Both tests show dilated vessels in the choroid layer with the characteristic presence of polypoidal structures which usually occur at the termini of vessels at the edge of the vascular network (Figure 10).(8,24-27)

Figure 10. PCV case. A: Color photography shows hemorrhagic PED. B: FA shows the PED and signs of occult CNV. C: ICGA shows the presence of hot spots corresponding to polypoidal structures. D: OCT shows small cup-shaped RPE elevations corresponding to PCV; subretinal fluid (*) and blood under the RPED (^).

 

Retinal Angiomatous Proliferation (RAP)

Retinal angiomatous proliferation (RAP) is a particular type of intraretinal neovascularization consisting of vascular anastomosis of aberrant retinal vessels. In choroidal neovascular membranes, new vessels proliferate through the RPE and infiltrate the retina and eventually communicate with the retinal circulation to produce retinochoroidal anastomosis. In RAP, the neovascularization originates in the same retinal vessels. The diagnosis is based on the presence of hyperfluorescent points corresponding to the areas of neovascularization (hot spot) in ICGA images, and the presence of dilated retinal vessels associated with multiple intraretinal, preretinal or subretinal hemorrhages, with varying degrees of intraretinal edema in the fundus examination. A proposed three-stage classification is seen in Table 2.

Table 2. RAP Stages

 
1
 
 
Proliferation of intraretinal capillaries originating from the deep capillary network in the macular area
 
 
2
 
 
Presence of retinal-retinal anastomosis and intraretinal edema and a possible association with a PED and/or subretinal neovascularization
 
 
3               
 
Retinal-choroidal anastomosis and emergence of new choroidal vessels
 
 
Although the initial stages are only visible by fundus examination and angiographic examinations due to the small size of the initial injury, OCT is extremely useful for identifying RAP and its associated manifestations (Figure 11)(6-8,19,28-30)
 
Figure 11
 
Figure 11. A (FA) and B (OCT) of a RAP lesion. OCT shows RAP (*) associated with edema, intraretinal cysts (↘) and subretinal fluid (∆). C (OCT), OCT shows the RPE and the retinal neurosensory layer detachments; Intraretinal cysts (∆), dense retinal particles (↘), the posterior hyaloid membrane (v) and a vessel (*) can be appreciated too.
 
 
The initial signs that correspond to stage 1 (intraretinal vascularization) consist of a focal area, usually extrafoveal, with increased retinal reflectivity that are not associated with epiretinal, intraretinal, or subretinal changes or changes in the retinal thickness. Progression of RAP on OCT is characterized by the presence of intraretinal or subretinal fluid, the former characterized by well-defined confluent hyporeflective spaces and the latter (neurosensory detachment) characterized by a well-defined hyporeflective space between the neurosensory foveal retina and other highly reflective bands corresponding to the RPE. When RAP reaches the subretinal space and merges with the RPE, a serous detachment of the RPE usually develops (stage 2). In well-developed cases, there may be retinal choroidal anastomoses (stage 3) and it is impossible to clearly differentiate stage 2 from stage 3 on OCT.
 

Disciform scars

Disciform scars are the end-stage of exudative AMD in which the retinal tissue is replaced by scar tissue (usually vascularized). Disciform scars develop with regression of subretinal hemorrhages and retinal edema and hyperplasic elements of the RPE. Disciform scars are usually dry and show marked destruction of all retinal layers. It is less common to find diffuse exudation with elevation of the entire sensory retina at the posterior pole (Figure 12).

Figure 12

Figure 12. OCT shows a hyperreflective fibrotic nodular area that corresponds to the fibrotic area and easily detects diffuse or cystoid edema. In the wet form, OCT shows diffuse or cystoid edema involving the RPE detachment.

 

 

Time-Domain vs. Spectral-Domain OCT in AMD 

The new spectral-domain (SD) OCT devices include a spectrometer in the receiver that analyzes the spectrum of reflected light on the retina and transforms it into information about the depth of the structures according to the Fourier principle. This technology eliminates the need to mechanically move the reference arm with the consequent increase in the speed with which images are received and axial resolution in time-domain (TD) OCT.(31) In addition to greater speed in capturing images and higher definition, the algorithms used by SD-OCT differ from those of TD-OCT, and the retinal thickness measurements are not comparable between the two OCT systems. While TD-OCT determines the total retinal thickness by measuring the distance from the internal limiting membrane to the highest hyperreflective band, i.e., that combining the inner and outer segments of the photoreceptors, SD-OCT set this threshold in the RPE hyperreflective band, so the retinal thickness values are higher than those obtained by TD-OCT (Figure 13)(31,32)
 
 
Figure 13. SD- vs. TD-OCT. A, TD-OCT (Stratus). B, SD-OCT (Cirrus). Foveal cysts are barely distinguishable with the TD-OCT, but are clearly identified by SD-OCT.
 
 
These differences justify the results of studies that compared the two devices in patients with wet AMD. Thus, Mylonas and colleagues found that in a number of patients with wet AMD, the retinal thickness measurements obtained by three SD-OCT devices were higher than those obtained by TD-OCT instruments. The authors also emphasized the importance of segmentation analysis as the main source of errors with both devices.(33
 
The effect of this parameter had already been studied widely for Stratus OCT (Carl Zeiss Meditec Inc., Dublin, CA), and the proposed solution to avoid these errors was the manual correction of segmentation lines in each image, especially in clinical trials of neovascular AMD, in which the difficulty establishing the limits in cases with subretinal and intraretinal fluid was automatically higher.(22,34,35)
 
To assess the incidence of errors at this level in the SD-OCT apparatus, Krebs and colleagues assessed 104 patients with neovascular AMD with both TD-OCT and SD-OCT and analyzed the position of the lines drawn automatically by segmentation analysis in each case(36). The results showed differences between the devices, with TD-OCT committing errors in 69.2% of cases and SD-OCT in 25%. These data suggested that SD-OCT makes fewer errors in automatic segmentation analysis, and these can be corrected manually identical to TD-OCT and therefore constitute a marked improvement in the main source of measurement errors.
 
Finally, some studies have compared both devices in a series of patients treated with antiangiogenic agents. Sayanagi and colleagues compared tomographic findings with TD-OCT and SD-OCT in 58 patients with wet AMD treated with ranibizumab (Lucentis, Novartis©, Basel, Switzerland) and concluded that SD-OCT is better than the TD linear mode B and mode 3D cube for detecting intraretinal cysts and intraretinal and subretinal fluid or fluid under the RPE, making it a more effective tool for managing these patients.(37)
 

Correlation with Other Techniques in AMD

OCT retinal studies in patients with AMD can complement information obtained with other conventional examinations, such as FA, or other more modern techniques such as autofluorescence. We discuss the added value of combining these techniques in recent studies.

 

OCT and Fluorescein and Indocyanine Green Angiography

Because of the widespread use of OCT, the use of FA and ICG has decreased considerably in routine consultations, but it is still an essential examination for studying CNV in patients with AMD, especially in the identification of occult CNV. When comparing both techniques, due to the improved definition of the latest OCT devices some studies have reported similar results for both examinations. Krebs and colleagues reported that OCT can detect signs of activity similarly to FA in two different studies directed to evaluate the agreement between two independent graders. The authors highlighted that OCT was slightly more sensitive than FA for determining the activity of the membranes, while other authors stated that it was necessary to manually adjust the segmentation analysis of images to achieve similar results(22,38). More recently, Malamos and colleagues used a high-definition OCT device to analyze patients diagnosed by FA in different types of AMD (early, classic, minimally classic and occult). They reported a significant correlation between different angiographic patterns and various OCT parameters. The authors could discriminate different types of lesions using OCT to the extent that they could identify their precise anatomic location in retinal maps types of AMD (early, neovascular classic, minimally classic, and occult). They reported a significant correlation between different angiographic patterns and various parameters with OCT. The authors could discriminate different types of damage using OCT to the extent that they could identify their precise anatomic retinal maps.(39) In cases with occult membranes and some cases of RAP, the relevance of FA is greater because of the difficulty to delineate the precise location of the pathology in the tomographic images (Figure 14). However, some authors have described qualitative OCT changes associated with CNV membranes, such as the presence of exudation signs and PEDs, and changes within the polymorphic features of these lesions.(3)

Figure 14. The retinography corresponds to a PCV (A), where you can observe the presence of intraretinal hemorrhages. In the red-free image (B) you can observe the tomographic cut line. The OCT shows a marked elevation of the cup-shaped RPE (C). By ICGA you can observe a hot spot that confirms the diagnosis (D, E).

 
Until the development of new techniques to visualize the choroid with OCT such as the Enhanced Depth Imaging OCT technique (EDI-OCT) proposed by Spaide and colleagues, the most informative method to explore these cases is FA/ICG.(40-42) However, OCT seems to override FA in the identification of signs of activity in patients previously treated with PDT, in which interobserver agreement is much higher for OCT than for FA.(43-44) Finally, the sequential use of both techniques for screening of wet AMD in patients with impaired visual acuity (VA) has been proposed. The screening was first performed with OCT, and then based on the OCT results, FA and ICG were performed when required, combination which obtained adequate sensitivity and specificity for the early detection of AMD in these patients.(45)
 

OCT and Autofluorescence

Autofluorescence allows gathering of information about the metabolic status of the retina. This technique is based on the ability of tissues to emit more or less fluorescence depending on the amount of accumulated lipofuscin after being excited by light at specific wavelengths, providing the opportunity to evaluate the retinal functional status. Since OCT obtains a detailed picture of the retinal anatomy, several studies have evaluated the structure/function in patients with AMD and correlated the two techniques. Brar and colleagues(46) reported that there was a significant association between OCT and autofluorescence findings, and highlighted that increased hyperfluorescence at the edges of the GA plaques correlated with hyperreflective changes in the outer retinal layers identified by OCT, finding that did not occur in healthy retinas with normal autofluorescence and OCT images (Figure 15).

Figure 15. (A) The retinography shows a juxtafoveal atrophic plaque. (B) The autofluorescence image delineates the atrophic plaque as an area of complete absence of fluorescence. (C) The OCT shows an increased choroidal reflectance due to the window effect produced by the overlying atrophic plaque. (D) Infrared image captured by the OCT that shows the tomographic section of figure C.

 

Therefore, the combination of both techniques allowed the determination of the patterns of progression of GA plaques in patients with AMD.(47)

Whereas autofluorescence is currently the most widely employed test to assess disease progression in dry AMD, OCT provides additional information and some authors have also proposed alternative OCT-based methods to evaluate progression of atrophic plaques, as described in the next section.(17)
 
 

Clinical and Therapeutic Implications of OCT in AMD

OCT is now a routine examination for patients with AMD. Although some studies have advocated a role for OCT in the dry forms of the disease, its use is greater in patients with the neovascular forms, and it is now the main method employed for lesion monitoring in this condition in routine clinical care. The role of OCT as a reliable tool for assessment of treatment response has been described since the PDT era. In recent years, the development of antiangiogenic therapies have made OCT a fundamental tool in routine patient management, and several landmark studies have based the need for retreatment in OCT.

 

Dry AMD

The better definition achieved by OCT devices that include SD technology allows a more detailed study of the outer retinal layers and RPE, which are important structures in the development of dry AMD. Thus, new SD-OCT instruments can accurately distinguish the presence and size of drusen and RPE changes, making it possible to differentiate the different retinal layers to identify changes at that level(45) Thus, in a group of patients with dry AMD, Schuman and colleagues(48) found localized thinning of the photoreceptor layer immediately above the drusen compared to healthy controls, suggesting a degenerative process with cell loss to explain the decreased visual function in this group (Figure 16).

Figure 16. (A) Central soft drusen may be seen in the retinography. (B) Drusen are observed as undulations and elevations of RPE hyperreflective band with less reflective material below them.

 

Bearelly and colleagues(17studied the thinning at the edges of the plaques of GA to establish a gradient in the thickness of the photoreceptors layer from the healthy retina to the atrophic plaque. Although they had a small sample size in their study (n = 17), they concluded that SD-OCT allows quantitative measurement of disease progression and postponed for future study the application of the technique. Moreover, to study progression of the dry forms, other studies have been designed to correlate the SD-OCT findings with other techniques such as autofluorescence. To this end, Stopa and colleagues(47) analyzed a series of patients with dry AMD and correlated SD-OCT images of areas of GA, isolated hard drusen and soft coalescing drusen with retinographies and autofluorescence images. Moreover, in addition to different patterns of reflectivity for each type of drusen, they observed that certain patterns of hyperreflectivity in some drusen and the overlying retina corresponded to increased or decreased autofluorescence at these points. This certainly established a morphology-function relationship between the images obtained with both techniques (Figure 17).

Figure 17. (A) Central hard drusen surrounded by atrophic areas of the RPE are seen in the retinography. (B) In the autofluorescence image, a hypofluorescence area corresponding to atrophic areas of the RPE may be appreciated. (C) In the OCT, two hyperreflectant points corresponding to hard drusen observed in the retinography are seen. (D) Infrared image captured by the OCT which shows the tomographic section line of figure C.

Finally, in some cases Stopa and colleagues identified small hyporeflective spots in retinas with soft coalescent drusen and drusenoid PEDs corresponding to subretinal fluid, which were not found in other examinations. In conclusion, the study of dry AMD need a combined approach with both autofluorescence and OCT in a multimodal imaging approach. Whereas OCT provides anatomic information on retinal structures, autofluorescence gives information of the metabolic status of the retinal tissues and allows accurate delineation of RPE atrophy lesions.(49)

 

WET AMD

Retinal study using OCT is now a fundamental tool to manage patients with wet AMD. Since the introduction of PDT as an antiangiogenic therapy a decade ago, OCT has been proven a useful tool to provide information about the activity of neovascular membranes, and has determined the need for treatment, the treatment response, and identified the early signs of recurrence or resolution. Several multicenter studies were designed to evaluate treatment protocols with antiangiogenic therapies that rely on OCT as a key tool to guide retreatment, as a way to optimize treatment regimens and decrease the number of injections in these neovascular AMD patients.

The PrONTO study (Prospective Optical Coherence Tomography Imaging of Patients with Neovascular Age-Related Macular Degeneration Treated with intraOcular Ranibizumab): The injection protocols established in multicenter phase III studies of ranibizumab for neovascular AMD, the ANCHOR and MARINA trials, consisted of monthly intravitreal injections for 2 years for all study eyes that resulted in substantial gains in VA in various intermediate controls and at the final examination(50-52) However, the dose of 24 injections carried a human and economic burden unaffordable in routine clinical care. Before these studies, in the phase I and II studies of ranibizumab, an extension study was conducted in which administration of new additional doses were left to the discretion of the investigator based on the presence of diffusion on FA or intraretinal or subretinal fluid on OCT. This subanalysis showed that the presence of retinal fluid could often be detected much earlier by OCT than FA, suggesting a greater role for OCT in the assessment and monitoring of these patients.

Based in this idea, Fung and colleagues at the Bascom Palmer Eye Institute in Miami, Florida, proposed the PrONTO Study. In that study, after three monthly injections (loading dose), monthly retreatments of ranibizumab were administered for 2 years according to the criteria in Table 3 in a PRN basis.

 
Table 3. PrONTO Study (1st year). Criteria for Reinjection after Three Loading Doses
 

1. Decrease of 5 letters of VA (ETDRS) and fluid on OCT
2. Increase of 100 microns in central retinal thickness on OCT
3. Macular hemorrhage
4. Classic CNV in emerging FA
5. Persistent fluid on OCT 1 month after the last injection

*Corrections for the second year of study: Qualitative changes on OCT images suggestive of recurrent macular fluid, including: intraretinal cysts, subretinal fluid, PEDs.

 

Of note, three out of five criteria were based on OCT analysis (VA decreases exceeding five ETDRS letters with fluid in the macula detected by OCT, central retinal thickness increases exceeding 100 microns compared to previous lowest value measured by OCT or fluid recurrence on OCT in a previously dry eye) (Figure 18).(53)

Figure 18. In the PrONTO clinical trial, three criteria of retreatment are based in the analysis by OCT: Decrease of more than 5 letters of VA (ETDRS) with fluid detected in the macula by OCT (A), increase of more than 100 µm in the central retinal thickness respect to the lowest previous value measured by OCT (B1, B2) or fluid reappearance in the OCT in a previous dry lesion (C1, C2).

 

Interestingly, the VA results of this study were similar to those obtained in the previous phase III trials, but required significantly fewer injections (an average of 5.6 vs. 12 injections during the first year of the study and 9.9 vs. 24 injections to complete the second year).
The good VA results obtained with the lower number of injections based on OCT led to the inclusion of OCT as a primary criteria in new multicenter studies such as the SAILOR or SUSTAIN studies. Those studies included a larger number of patients than in the original PrONTO study (n = 40) and confirmed the important role of OCT in the monitoring of these patients.

Figure 19. In the second year of the study, the presence of qualitative changes in the OCT images that suggested fluid recurrence was included, such as intraretinal cysts (A), subretinal fluid (B), or progression of a PED (C1 and C2)(54).

 

OCT and Reviews: “Treat & Extend.”

In patients with wet AMD treated with antiangiogenic drugs, the two more important parameters are the injection frequency and the number of visits. The latter represents a substantial burden for both patients and families, and healthcare system. Several studies (PIER, EXCITE) proposed fixed revisions, but the results were worse than those obtained with monthly revisions, with both fixed reinjection protocols and those based on OCT(55-57). Therefore, it is necessary to establish specific criteria to control AMD and allow more time between retreatments without losing the treatment effectiveness. Once again, OCT had a fundamental role in this approach. In this sense, Spaide proposed a revision scheme called Treat and Extend; after three monthly loading injections, patients were reevaluated in 6 weeks, at which time the physicians determine whether signs of activity by funduscopy and edema or intraretinal or subretinal fluid by OCT are found (Figure 20).(58)

Figure 20. Treatment schedule and revisions of Treat & Extend.

 

The VA results depend not only on the number of injections but also the frequency of subsequent visits by the protocol presented in this proposal, which again gives OCT a leading role, as it indicates the detection of activity and therefore the frequency of visits. The use of this treatment regimen is being evaluated in a study conducted at Wills Eye Institute, Philadelphia, Pennsylvania. The preliminary results are promising, with a lower frequency of examinations, i.e., an average of 7.4 injections annually and similar VA outcomes to those obtained in the PrONTO study in a greater number of patients (n=92)(59)

 

Prognostic Factors in OCT

Several OCT parameters are associated with decreased VA and different responses to other antiangiogenic treatments regarding VA, central retinal thickness and prognosis after treatment. Singh and colleagues retrospectively analyzed a group of patients with wet AMD treated with bevacizumab (Avastin, Genentech, south San Francisco, CA) and concluded that in addition to VA pre-treatment, the total retinal thickness measured by OCT affected the final VA and the total decrease in the post-treatment retinal thickness, which was lower in patients who had received previous treatments.(60) Keane and colleagues also found that the main factor associated with decreased VA was the volume of subretinal tissue, and in fewer cases, thickening of the neurosensory retina, without a significant association with the total volume of subretinal fluid and RPE detachment and VA.(22However, these parameters did not justify the variability found in the VA for similar values of subretinal tissue volume or thickening of the neurosensory retina. The authors pointed to the complex pathophysiology of the CNV membranes and the limitations of the TD-OCT devices employed to explain the results. Sayanagi and colleagues reported that SD-OCT is a superior generation of TD-OCT than its predecessors for assessing the activity of the neovascular membranes and changes in AMD after ranibizumab treatment, and Kiss and colleagues pointed to the RPE status of neovascular membranes as main predictor of VA in patients treated with ranibizumab in addition to conventional parameters (i.e. central retinal thickness).(37-61) Several studies also reported a significant correlation between a hyperreflective band indicating the integrity of the junction between the inner and outer segments of the photoreceptors and higher VA in patients treated with ranibizumab.(62-63) It appears that parallel to the development of new devices, structures assume special importance, such as the junction of the inner and outer segments of the photoreceptors, which until now was only identified, as shown by studies with ultrahigh resolution OCT (UHR-OCT) prototypes with a high-definition axial resolution of 3.5 microns.(64-65Lee and colleagues reported that the presence of posterior vitreomacular adhesion on OCT was associated with CNV in a large series of patients with AMD (n = 251). Those authors suggested that this finding is a possible risk factor for subretinal membrane development because of chronic vitreomacular traction on the retina, opening the door to a possible surgical approach in patients not responding or resistant to drug treatment (Figure 21).(66-67)

Figure 21. The chronic retinal vitreomacular traction has been considered as a possible risk factor for development of subretinal membranes. In the image, a subretinal membrane associated with a vitreomacular traction in the presence of a PED and subretinal fluid may be appreciated.

 

Ahlers and colleagues at the Medical University of Vienna pointed to a new parameter as a prognostic factor in patients with AMD, the optical density ratio of the subretinal fluid detected by OCT(68). The authors suggested that this ratio may be an indirect way to measure the integrity of the blood-retinal barrier and, therefore, be useful in the differential diagnosis of different exudative macular diseases, such as central serous chorioretinopathy, as well as to assess the response to antiangiogenic drugs (Table 4).

 
Table 4. Prognosis Factors According to OCT as Identified by Various Authors.
 

Parameter

Author and year

 
Total retinal thickness
 
 
Singh 2009
 
Subretinal tissue volume
Neurosensory retina thickening
 
 
Keane 2008
 
RPE status
 
 
Kiss 2009

Integrity of the union of outer and inner segments of photoreceptors

Sayanagi 2009, Kaluzny 2009, Chen 2009, Witkins 2009

Posterior vitreomacular adhesion

Lee 2009, Mojana 2008

Optical density ratio

Ahlers 2009

 

OCT and Wet AMD in Clinical Practice

To reach a consensus on the criteria of treatment and frequency of revisions, daily management of these patients is individualized, and the decisions are based on clinical examination and qualitative analysis of OCT images(2). The main signs of CNV activity in OCT are the presence of intraretinal or subretinal fluid, PEDs and/or RPE rips. These OCT findings should be evaluated biomicroscopically for the presence of fibrosis in disciform scars, which are final and irreversible stages of the disease that sometimes can be shown on OCT. The presence of any of these tomographic signs, the patient’s VA, and the ophthalmoscopic and angiographic appearance of the lesions should be evaluated by ophthalmologists to reach treatment decisions and the frequency of follow up visits in each individual case. It is also important to identify other OCT signs that may be confounded with signs of CNV activity, such as retinal tubulations and viteliform detachments, that do not represent actual exudation of fluid in the subretinal space (Figures 22-26).

Figure 22. Massive subretinal hemorrhage secondary to wet AMD. Blood (yellow arrows) is collected between the RPE detachment and the neurosensory retina.

 

optical-coherence-tomography-age-related-macular-degeneration-image23.png

Figure 23. Adult-onset foveomacular vitelliform lesion in a patient with drusen (A); multiple hyperautofluorescent material deposits are seen (B). OCT shows the material over the RPE (C) that must not be confounded with a neovascular membrane related to wet AMD.

 

optical-coherence-tomography-age-related-macular-degeneration-image24.png

Figure 24. Adult-onset foveomacular vitelliform lesion in a patient with drusen (A); multiple hyperautofluorescent deposits are seen (B) together to hypoautofluorescent RPE atrophy. OCT shows the material over the soft drusen (yellow arrow) (C) that must not be confounded with a neovascular membrane related to wet AMD.

 

optical-coherence-tomography-age-related-macular-degeneration-image25.png

Figure 25. Adult-onset foveomacular vitelliform lesion with hyperautofluorescent central deposit. OCT shows a vitelliform detachment due to reabsorption of the vitelliform material that must not be confounded with subretinal fluid related to wet AMD.

 

optical-coherence-tomography-age-related-macular-degeneration-image26.png

Figure 26. Outer retinal tubulation (yellow arrow) in cross-sectional SD-OCT. They must not be confounded with subretinal fluid in wet AMD.

 

New Techniques and Prototypes in AMD

OCT and Confocal Scanning Laser Ophthalmoscopy

Some new SD-OCT instruments (Spectralis HRA-OCT, Heidelberg Engineering, Inc., Heidelberg, Germany; Spectral OCT/SLO, Opko/OTI, Inc., Miami, FL) are equipped with this method, which simultaneously captures digital funduscopic images (infrared, red free autofluorescence, FA and ICGA) and OCT images from the same light source. This enables clinicians to correlate the two images pixel by pixel and allows a detailed topographic survey and tomographic retinal structures.(69) In AMD, these devices are used especially in studies attempting to correlate OCT and autofluorescence in GA.(70-71)

 

Displaying the Choroid with Depth-Enhanced Optical Imaging OCT

The principle of OCT requires penetration of the emitted laser beam through the retinal and choroidal structures before being received for analysis. Thus, with the standard technique, images of the full-thickness retina to the RPE, Bruch’s membrane and choroid immediately adjacent space can be obtained, but the remainder of the choroid is beyond the capabilities of the instrument. To study this layer, Spaide developed a system in which the focus of the retinal scan is displaced to the choroid in a way that multiple inverted images are processed to obtain a higher quality image in which the choroid can be seen in greater depth. Several diseases have been studied using this technique. This technique may be relevant for studying wet AMD associated with RPE detachment and especially in occult CNV, which often cannot be visualized with conventional OCT. (40,42,72-74)

 

New Prototypes: High-Speed Resolution Ultrahigh OCT

Spectral Domain (SD) instruments reach axial resolutions of 5 microns, using a laser source that emits light at a certain wavelength and a detection system based on the Fourier principle. To obtain even higher resolutions, several groups have designed devices that modify these components, mainly the light source. A group at Tufts Medical Center, Boston, developed a prototype that uses a superluminescent diode broadband laser combined with a Fourier detection system that can achieve axial resolutions up to 3.5 microns, allowing precise delineation and analysis of the outer retinal layers, such as the photoreceptors layer, the RPE and Bruch’s membrane(75). The investigators have published several studies highlighting the importance of these stratus in the pathogenesis of dry AMD, the ability for early detection of intraretinal and subretinal fluid and the status of the photoreceptors before and after ranibizumab treatment in a series of patients with wet AMD.(4,65,76,77)

 

En-Face OCT Imaging

En-face OCT is one of the OCT visualization approaches that has significantly benefited from technical advancements in OCT technology(78). En-face software reconstruct images on the coronal plane, from the scans obtained using a dense raster scan cube of the posterior pole. En-face OCT produces transverse images of retinal and choroidal layers at any specified depth, which offers additional benefit on top of longitudinal cross-sections. This technology is very useful for diagnosing and monitoring diseases with layer-specific anatomic abnormalities (Figures 27-29). It also has the ability to localize OCT images in other fundus imaging modalities, using retinal vessels as landmarks. 

 

optical-coherence-tomography-age-related-macular-degeneration-image27.png

Figure 27. Outer retinal tubulation (yellow arrows) in cross-sectional (upside) and en-face SD-OCT (downside, green arrows). Tubulations are hyporeflective areas representing a lumen surrounded by a hyperreflective border consisting of cone photoreceptor mitochondria and external limiting membrane, indicating that cones have lost their normal directionality due to loss of outer segments and subsequent retinal remodelling. They must not be confounded with subretinal fluid in wet AMD.

 

optical-coherence-tomography-age-related-macular-degeneration-image28.png

Figure 28. En-face OCT imaging: dome of the PED appears as hyporeflective area, predominantly circular with marked and well-defined hyperreflective margin. Polyps appears as round multiple heterogeneous hyperreflective structures.

 

optical-coherence-tomography-age-related-macular-degeneration-image29.png

Figure 29. SS-OCT and en-face OCT of a patients with late AMD with geography atrophy. En-face OCT is a reliable method to detect and quantify GA, particularly when analyzed at the outer retina and choriocapillaris level, where the photoreceptors' loss creates an abrupt transition in OCT reflectivity.

 

SS-OCT

Swept-source (SS) OCT utilizes a longer wavelength of 1060 nm, that increases depth penetration, thus improving penetration and enhanced resolution of the deeper layers of the eye, including the choroid and sclera(86). Furthermore, it is capable of faster scanning rates, allowing for image capture, doubling the scan acquisition speed of SD-OCT devices. SS-OCT has acquisition rates of up to 100,000–236,000 A-scans per second and a resolution of up to 11 μm, thus permitting to obtain widefield B-scans (12 mm vs 6–9 mm with conventional SD-OCT) and more accurate 3-D imaging of the vitreous, retina and choroid (Figures 30-39).

 

optical-coherence-tomography-age-related-macular-degeneration-image30.png

Figure 30. SS-OCT of a normal eye combined with color photography. Vitreoretinal interface with vitreous adhesions and choroid are clearly identified.

 

optical-coherence-tomography-age-related-macular-degeneration-image31.png

Figure 31. SS-OCT of soft drusen.

 

optical-coherence-tomography-age-related-macular-degeneration-image32.png

Figure 32. SS-OCT of wet AMD with subretinal fluid (yellow star) and RPE detachment (yellow arrow). Top, inverted image.

 

optical-coherence-tomography-age-related-macular-degeneration-image33.png

Figure 33. SS-OCT of a normal eye combined with autofluorescence. Vitreoretinal interface with vitreous adhesions and choroid are clearly identified. Retinal thickness graph after automatic measurement is also visualized.

 

optical-coherence-tomography-age-related-macular-degeneration-image34.png

Figure 34. SS-OCT of an eye with intermediate AMD with drusen and retinal pigment abnormalities combined with red-free fundus photography. Retinal thickness graph after automatic measurement is also visualized.

 

optical-coherence-tomography-age-related-macular-degeneration-image35.png

Figure 35. SS-OCT of an eye with late AMD with GA and retinal pigment abnormalities combined with autofluorescence fundus photography. Intraretinal cysts may be sometimes observed in areas of GA in absence of CNV.

 

optical-coherence-tomography-age-related-macular-degeneration-image36.png

Figure 36. SS-OCT of an eye with late AMD with GA combined with autofluorescence fundus photography. Increased transmission in the area of atrophy and reduced choroidal thickness.

 

optical-coherence-tomography-age-related-macular-degeneration-image37.png

Figure 37. SS-OCT of an eye with late wet AMD with subretinal fluid and RPE detachment in a case of active CNV (top) and after intravitreal anti-VEGF injection with disappearance of the subretinal fluid (bottom).

 

optical-coherence-tomography-age-related-macular-degeneration-image38.png

Figure 38. SS-OCT of an eye with AMD with drusen and retinal pigment areas of atrophy.

 

optical-coherence-tomography-age-related-macular-degeneration-image39.png

Figure 39. SS-OCT of an eye with late wet AMD with manual measurement of the choroidal thickness, measured each 1000 microns (bottom, horizontal numbers).

 

ANGIO OCT (OCTA)

En-face OCT is a valuable imaging strategy for anatomic and angiographic viewing of the fundus. Their use combined with Doppler OCT, that utilizes ultra-high-speed image acquisition using blood flow as a contrast medium, to evaluate retinal and choroidal vasculature has result in the development of optical coherence angiography (OCTA)(92). OCTA is a non-invasive technique, based on Doppler shift induced by red blood cell flow, to visualize vessels with active flow. The three-dimensional reconstruction of vascular mapping at the microcirculation vascular flow highlights the potential of this technology to identify important vascular changes in different retinal and choroidal diseases including CNV and GA(93,94), and it is a very promising technique to replace more invasive imaging techniques such as FA and ICGA.

However, image artefacts occur frequently in OCTA images, and these artefacts are more frequent in eyes with pathology(95). It is possible that SS-OCTA is able to better delineate the full extent of CNV vasculature, and yielded significantly larger CNV areas than SD-OCTA(96). Moreover, it should be noted that OCTA is able to detect asymptomatic neovascularization in eyes with intermediate AMD, lesions that are visible only as plaques in ICGA(97).

Figures 40 to 53 show several cases of OCTA exams of a normal eye and of diseased eyes in different stages of AMD.

 

optical-coherence-tomography-age-related-macular-degeneration-image40.png

Figure 40. OCTA of a normal eye. From left to right, superficial retinal vascular plexus, deep retinal vascular plexus, outer avascular retina and choriocapillaris, combined with SS-OCT, color representation of all the combined vascular zones and retinal color photography.

 

optical-coherence-tomography-age-related-macular-degeneration-image41.png

Figure 41. OCTA of an eye with late AMD with GA. An artefact simulating a CNV (blue arrows) is observed in the outer retina and choriocapillaris over the area of atrophy.

 

optical-coherence-tomography-age-related-macular-degeneration-image42.png

Figure 42. OCTA of an eye with late AMD with GA. An artefact simulating a CNV is observed in the outer retina and choriocapillaris over the area of atrophy.

 

optical-coherence-tomography-age-related-macular-degeneration-image43.png

Figure 43. OCTA of an eye with late AMD with type 1 CNV. SS-OCT shows subretinal fluid and drusen. Hyperreflective spots in the neurosensory retina and subretinal fluid are suspected to be inflammatory cells (yellow arrow). CNV vascular network is clearly observed in the outer retina and choriocapillaris of the OCTA.

 

optical-coherence-tomography-age-related-macular-degeneration-image44.png

Figure 44. OCTA of an eye with late AMD with type 2 CNV. SS-OCT shows intraretinal fluid and some subretinal fluid over a small RPE detachment. Vitreoretinal traction is also observed. CNV vascular network is clearly observed in the outer retina and choriocapillaris of the OCTA. The neovascular membrane appears as a medusa-shaped complex or a glomerulus-shaped lesion. The external borders of the lesion appear as a dark ring.

 

optical-coherence-tomography-age-related-macular-degeneration-image45.png

Figure 45. OCTA of an eye with late AMD with type 2 CNV. SS-OCT shows subretinal fluid. CNV vascular network is clearly observed in the outer retina and choriocapillaris of the OCTA. The neovascular membrane appears as a medusa-shaped complex or a glomerulus-shaped lesion. 

 

optical-coherence-tomography-age-related-macular-degeneration-image46.png

Figure 46. OCTA of an eye with late AMD with type 3 CNV. SS-OCT shows subretinal fluid with the presence of an intraretinal-subretinal continuous hyperreflective lesion (blue arrow). Small intraretinal neovascularization vascular network is observed in the deep retina plexus of the OCTA (yellow arrow). ICGA shows the hot spot corresponding to the RAP (green arrow).

 

optical-coherence-tomography-age-related-macular-degeneration-image47.png

Figure 47. OCTA of an eye with late AMD with PCV. SS-OCT shows a steeply-sloped dome-shaped PED in PCV. A hyporeflective space (cleft) between hyperreflective materials in PED and Bruch's membrane sometimes appears in neovascular AMD. This cleft may represent a space that results from fluid accumulation that originates from active CNV components in the materials in PED (yellow arrow). ICGA shows the hot spot corresponding to the PCV (green arrows). Branching vascular networks showed more clearly on OCTA than on ICGA. Polypoidal lesions had variable patterns on OCTA and were not always detected. The hypo-flow round appearance of the polyps on OCTA (orange arrows) is probably due to unusual blood flow inside of the polypoidal lesions, in contrast with the branching vascular network.

 

optical-coherence-tomography-age-related-macular-degeneration-image48.png

Figure 48. OCTA of an eye with late AMD with disciform lesion. SS-OCT shows no subretinal fluid with the presence of a subretinal hyperreflective fibrous scar. A dead tree image of an inactive neovasvascular network is observed in the OCTA (yellow arrow). 

 

optical-coherence-tomography-age-related-macular-degeneration-image49.png

Figure 49. FA is the gold standard for determining the presence of leakage (yellow arrow) or stain of the lesion (blue arrow). SS-OCT shows fluid accumulation (green arrow). Different CNV patterns are identified on OCTA even in the same patient. A lesion was assessed as active (Pattern I, large blue arrow), if it showed at least three of the following five features: 1) shape, a well-defined (lacy-wheel or sea-fan shaped) CNV lesion in contrast to one with long filamentous linear vessels; 2) branching, numerous tiny capillaries, typical of a recent lesion, in contrast to rare large mature vessels, typical of a mature one; 3) the presence of anastomoses and loops; 4) morphology of the vessel termini, assessing the presence of a peripheral arcade in contrast to a “dead tree” appearance (inactive, large yellow arrow); 5) presence of a perilesional hypointense halo, considered as regions of choriocapillaris alteration, either corresponding to flow impairment, steal or localized atrophy. A CNV lesion was considered as Pattern II if it showed less than three of the previously reported OCTA features(98). Horizontal lines are artefacts.

 

optical-coherence-tomography-age-related-macular-degeneration-image50.png

Figure 50. SS-OCT showed no signs of active CNV without intraretinal or subretinal fluid over the RPE detachment.  OCTA revealed the presence of a type 1 CNV. After each injection a pruning of smaller vessels is seen immediately that increases for one or two weeks when it reaches a maximum. Bottom, a reopening or new sprouting of the vessels one month later was observed still without the recurrence of the subretinal fluid (yellow arrows).

 

optical-coherence-tomography-age-related-macular-degeneration-image51.png

Figure 51. SS-OCT showed no signs of CNV and no intraretinal or subretinal fluid over the RPE changes that are clearly visible (yellow arrows).  OCTA revealed the presence of a quiescent type 1 CNV.
 

optical-coherence-tomography-age-related-macular-degeneration-image52.png

Figure 52. SS-OCT showed no signs of CNV and no intraretinal or subretinal fluid over the RPE changes that are clearly visible (yellow arrows).  OCTA revealed the presence of a quiescent type 1 CNV.

 

optical-coherence-tomography-age-related-macular-degeneration-image53.png

Figure 53. SS-OCT showed subtle subretinal fluid (green arrow) over the RPE changes (yellow arrow).  OCTA revealed the presence of a quiescent type 1 CNV.

 

>> References

References - Optical Coherence Tomography in Age-related Macular Degenera

References - Optical Coherence Tomography in Age-related Macular Degenera

English
  1. Sakata LM, Deleon-Ortega J, Sakata V, Girkin CA. Optical Coherence Tomography of the retina and the Optic Nerve. Clin Experiment Ophthalmol 2009;37:90-99.
     
  2. Brown DM, Regillo CD. Anti-VEGF agents in the treatment of neovascular age-related macular degeneration: applying clinical trial results to the treatment of everyday patients. Am J Ophthalmol 2007;144:627–37.
     
  3. Coscas F, Coscas G, Souied E, Tick S, Soubrane G. Optical coherence tomography identification of occult choroidal neovascularization in age-related macular degeneration. Am J Ophthalmol 2007;144:592–599.
     
  4. Pieroni CG, Witkin AJ, Ko TH, Fujimoto JG, Chan A, Schuman JS, Ishikawa H, Reichel E, Duker JS.. Ultrahigh resolution optical coherence tomography in non-exudative age related macular degeneration. Br J Ophthalmol 2006;90:191–197.
     
  5. Puliafito CA, Hee MR, Lin CP, Reichel E, Schuman JS, Duker JS, Izatt JA, Swanson EA, Fujimoto JG. Imaging of macular diseases with optical coherence tomography. Ophthalmology 1995;102:217–229.
     
  6. Brancato R, Tremoloada G. Tomografia óptica de coherencia en la degeneración macular asociada a la edad. En: Monés J, Gómez-Ulla F. Degeneración macular asociada a la edad. Proas Science 2005;73-84.
     
  7. Criado Muñoz P. Tomografía de Coherencia Óptica en el estudio de la patología macular. Boletín de la Sociedad Oftalmologica de Madrid - N.º 48, 2008.
     
  8. España Contreras M, Fernández-Baca Casares I. OCT en las enfermedades retinianas y del nervio óptico. E.U.R.O.M.E.D.I.C.E., Ediciones Médicas, S. L. 2007;15-30.
     
  9. Schmidt-Erfurth U, Leitgeb RA, Michels S, Povazay B, Sacu S, Hermann B, Ahlers C, Sattmann H, Scholda C, Fercher AF, Drexler W.. Three-dimensional ultrahigh-resolution optical coherence tomography of macular diseases. Invest Ophthalmol Vis Sci 2005;46:3393-3402.
     
  10. Sunness JS, Margalit E, Srikumaran D, Applegate CA, Tian Y, Perry D, Hawkins BS, Bressler NM.. The long-term natural history of geographic atrophy from age-related macular degeneration: enlargement of atrophy and implications for interventional clinical trials. Ophthalmology 2007;114:271-277.
     
  11. Chen TC, Cense B, Miller JW, Rubin PA, Deschler DG, Gragoudas ES, de Boer JF. Histologic correlation of in vivo optical coherence tomography images of the human retina. Am J Ophthalmol 2006;141:1165-8.
     
  12. Age-Related Eye Disease Study Research Group. The Age-Related Eye Disease Study system for classifying age-related macular degeneration from stereoscopic color fundus photographs: the Age-Related Eye Disease Study report nº 6. Am J Ophthalmol 2001;132:668–681.
     
  13. Hageman GS, Luthert PJ, Chong NH, Johnson LV, Anderson DH, Mullins RF. An integrated hypothesis that considers drusen as biomarkers of immune-mediated processes at the RPE-Bruch’s membrane interface in aging and age-related macular degeneration. Prog Retin Eye Res 2001;20:705-732.
     
  14. Yi K, Mujat M, Park BH, Sun W, Miller JW, Seddon JM, Young LH, de Boer JF, Chen TC. Spectral Domain Optical Coherence Tomography for quantitative evaluation of  drusen and associated structural changes in non-neovascular age-related macular degeneration. Br J Ophthalmol 2009;93:176-81.
     
  15. Khanifar AA, Koreishi AF, Izatt JA, Toth CA. Drusen ultrastructure imaging with spectral domain optical coherence tomography in age-related macular degeneration. Ophthalmology 2008;115:1883-90.
     
  16. Sandhu SS, Talks SJ. Correlation of optical coherence tomography, with or without additional colour fundus photography, with stereo fundus fluorescein angiography in diagnosing choroidal neovascular membranes. Br J Ophthalmol 2005;89:967-970.
     
  17. Bearelly S, Chau FY, Koreishi A, Stinnett SS, Izatt JA, Toth CA. Spectral domain optical coherence tomography imaging of geographic atrophy margins. Ophthalmology 2009;116(9):1762-9.
     
  18. Klein R, Klein BE, Knudtson MD, Meuer SM, Swift M, Gangnon RE. Fifteen-year cumulative incidence of age-related macular degeneration: the Beaver Dam Eye Study. Ophthalmology 2007;114:253-62.
     
  19. Liakopoulos S, Ongchin S, Bansal A, Msutta S, Walsh AC, Updike PG, Sadda SR. Quantitative Optical Coherence Tomography findings in varios subtypes of Neovascular Age-related Macular Degeneration. Invest Ophthalmol Vis Sci 2008;49:5048-54.
     
  20. Hughes EH, Khan J, Patel N, Kashani S, Chong NV. In vivo demonstration of the anatomic differences between classic and occult choroidal neovascularization using optical coherence tomography. Am J Ophthalmol 2005;139:344-346.
     
  21. Hee MR, Baumal CR, Puliafito CA, Duker JS, Reichel E, Wilkins JR, Coker JG, Schuman JS, Swanson EA, Fujimoto JG. Optical coherence tomography of age-related macular degeneration and choroidal neovascularization. Ophthalmology 1996;103:1260-1270.
     
  22. Keane PA, Liakopoulos S, Chang KT, Wang M, Dustin L, Walsh AC, Sadda SR. Relationship between optical coherence tomography retinal parameters and visual acuity in neovascular age-related macular degeneration. Ophthalmology 2008;115:2206-14.
     
  23. Sadda SR, Liakopoulos S, Keane PA, Ongchin SC, Msutta S, Chang KT, Walsh AC. Relationship between angiographic and optical coherence tomographic (OCT) parameters for quantifying choroidal neovascular lesions. Graefes Arch Clin Exp Ophthalmol 2009;247(8):1031-7.
     
  24. Iijima H, Iida T, Imai M, Gohdo T, Tsukahara S. Optical coherence tomography of orange-red subretinal lesions in eyes with idiopathic polypoidal choroidal vasculopathy. Am J Ophthalmol 2000;129:21–26.
     
  25. Chiang A, Chang LK, Yu F, Sarraf D. Predictors of anti-VEGF-associated retinal pigment epithelial tear using FA and OCT analysis. Retina 2008;28(9):1265-9.
     
  26. Okubo A, Sameshima M, Uemura A, Kanda S, Ohba N. Clinicopathological correlation of polypoidal choroidal vasculopathy revealed by ultrastructural study. Br J Ophthalmol 2002;86:1093–1098.
     
  27. Ozawa S, Ishikawa K, Ito Y, Nishihara H, Yamakoshi T, Hatta Y, Terasaki H. Differences in macular morphology between polypoidal choroidal vasculopathy and exudative age-related macular degeneration detected by optical coherence tomography. Retina 2009;29:793-802.
     
  28. Gross NE, Aizman A, Brucker A, Klancnik JM Jr, Yannuzzi LA. Nature and risk of neovascularization in the fellow eye of patients with unilateral retinal angiomatous proliferation. Retina 2005;25:713-718.
     
  29. Yannuzzi LA, Negrao S, Iida T, Carvalho C, Rodriguez-Coleman H, Slakter J, Freund KB, Sorenson J, Orlock D, Borodoker N. Retinal angiomatous proliferation in age-related macular degeneration. Retina 2001;21:416-434.
     
  30. Brancato R, Introini U, Pierro L, Setaccioli M, Forti M, Bolognesi G, Tremolada G. Optical coherence tomography (OCT) in retinal angiomatous proliferation (RAP). Eur J Ophthalmol 2002;12:467-72.
     
  31. Chen TC, Cense B, Pierce MC, Nassif N, Park BH, Yun SH, White BR, Bouma BE, Tearney GJ, de Boer JF. Spectral domain optical coherence tomography: ultra-high speed, ultra-high resolution ophthalmic imaging. Arch Ophthalmol 2005;123:1715-20.
     
  32. Han IC, Jaffe GJ. Comparison of spectral- and time-domain optical coherence tomography for retinal thickness measurements in healthy and diseased eyes. Am J Ophthalmol 2009;147:847-58.
     
  33. Mylonas G, Ahlers C, Malamos P, Golbaz I, Deak G, Schuetze C, Sacu S, Schmidt-Erfurth U. Comparison of retinal thickness measurements and segmentation performance of four different spectral and time domain OCT devices in neovascular age-related macular degeneration.  Br J Ophthalmol 2009;93:1453-60.
     
  34. Patel PJ, Chen FK, da Cruz L, Tufail A. Segmentation error in Stratus optical coherence tomography for neovascular age-related macular degeneration. Invest Ophthalmol Vis Sci 2009;50:399-404.
     
  35. Ghazi NG, Kirk T, Allam S, Yan G. Quantification of error in optical coherence tomography central macular thickness measurement in wet age-related macular degeneration. Am J Ophthalmol 2009;148:90-6.
     
  36. Krebs I, Falkner-Radler C, Hagen S, Haas P, Brannath W, Lie S, Ansari-Shahrezaei S, Binder S. Quality of the threshold algorithm in age-related macular degeneration: Stratus versus Cirrus OCT. Invest Ophthalmol Vis Sci 2009;50:995-1000.
     
  37. Sayanagi K, Sharma S, Yamamoto T, Kaiser PK. Comparison of spectral-domain versus time-domain optical coherence tomography in management of age-related macular degeneration with ranibizumab. Ophthalmology 2009;116:947-55.
     
  38. Krebs I, Ansari-Shahrezaei S, Goll A, Binder S. Activity of neovascular lesions treated with bevacizumab: comparison between optical coherence tomography and fluorescein angiography. Graefes Arch Clin Exp Ophthalmol 2008;246:811-5.
     
  39. Malamos P, Sacu S, Georgopoulos M, Kiss C, Pruente C, Schmidt-Erfurth U. Correlation of high-definition optical coherence tomography and fluorescein angiography imaging in neovascular macular degeneration. Invest Ophthalmol Vis Sci 2009;50:4926-33.
     
  40. Spaide RF, Koizumi H, Pozzoni MC. Enhanced depth imaging spectral-domain optical coherence tomography. Am J Ophthalmol. 2008;146:496-500.
     
  41. Spaide RF. Enhanced depth imaging optical coherence tomography of retinal pigment epithelial detachment in age-related macular degeneration. Am J Ophthalmol 2009;147:644-52.
     
  42. Margolis R, Spaide RF. A pilot study of enhanced depth imaging optical coherence tomography of the choroid in normal eyes. Am J Ophthalmol 2009;147:811-5.
     
  43. Salinas-Alamán A, García-Layana A, Maldonado MJ, Sainz-Gómez C, Alvárez-Vidal A. Using optical coherence tomography to monitor photodynamic therapy in age related macular degeneration. Am J Ophthalmol 2005;140:23-8.
     
  44. van Velthoven ME, de Smet MD, Schlingemann RO, Magnani M, Verbraak FD. Added value of OCT in evaluating the presence of leakage in patients with age-related macular degeneration treated with PDT. Graefes Arch Clin Exp Ophthalmol 2006;244:1119-23.
     
  45. Talks J, Koshy Z, Chatzinikolas K. Use of optical coherence tomography, fluorescein angiography and indocyanine green angiography in a screening clinic for wet age-related macular degeneration. Br J Ophthalmol 2007;91:600-1.
     
  46. Brar M, Kozak I, Cheng L, Bartsch DU, Yuson R, Nigam N, Oster SF, Mojana F, Freeman WR. Correlation between spectral-domain optical coherence tomography and fundus autofluorescence at the margins of geographic atrophy. Am J Ophthalmol 2009;148:439-44.
     
  47. Stopa M, Bower BA, Davies E, Izatt JA, Toth CA. Correlation of pathologic features in spectral domain optical coherence tomography with conventional retinal studies. Retina 2008;28:298-308.
     
  48. Schuman SG, Koreishi AF, Farsiu S, Jung SH, Izatt JA, Toth CA. Photoreceptor layer thinning over drusen in eyes with age-related macular degeneration imaged in vivo with spectral-domain optical coherence tomography. Ophthalmology 2009;116:488-496.
     
  49. Lujan BJ, Rosenfeld PJ, Gregori G, Wang F, Knighton RW, Feuer WJ, Puliafito CA. Spectral domain optical coherence tomographic imaging of geographic atrophy. Ophthalmic Surg Lasers Imaging 2009;40:96-101.
     
  50. Kaiser PK, Brown DM, Zhang K, Hudson HL, Holz FG, Shapiro H, Schneider S, Acharya NR. Ranibizumab for predominantly classic neovascular age-related macular degeneration: subgroup analysis of first-year ANCHOR results. Am J Ophthalmol 2007;144:850-857.
     
  51. Brown DM, Michels M, Kaiser PK, Heier JS, Sy JP, Ianchulev T,; ANCHOR Study Group. Ranibizumab versus verteporfin photodynamic therapy for neovascular age-related macular degeneration: Two-year results of the ANCHOR study. Ophthalmology 2009;116:57-65.
     
  52. Kaiser PK, Blodi BA, Shapiro H, Acharya NR; MARINA Study Group. Angiographic and optical coherence tomographic results of the MARINA study of ranibizumab in neovascular age-related macular degeneration. Ophthalmology 2007;114:1868-75.
     
  53. Fung AE, Lalwani GA, Rosenfeld PJ, Dubovy SR, Michels S, Feuer WJ, Puliafito CA, Davis JL, Flynn HW Jr, Esquiabro M. An optical coherence tomography-guided, variable dosing regimen with intravitreal ranibizumab (Lucentis) for neovascular age-related macular degeneration. Am J Ophthalmol 2007;143:566-83.
     
  54. Lalwani GA, Rosenfeld PJ, Fung AE, Dubovy SR, Michels S, Feuer W, Davis JL, Flynn HW Jr, Esquiabro M. A variable-dosing regimen with intravitreal ranibizumab for neovascular age-related macular degeneration: year 2 of the PrONTO Study. Am J Ophthalmol 2009;148:43-58.
     
  55. Regillo CD, Brown DM, Abraham P, Yue H, Ianchulev T, Schneider S, Shams N. Randomized, double-masked, sham-controlled trial of ranibizumab for neovascular age-related macular degeneration: PIER Study year 1. Am J Ophthalmol 2008;145:239-248.
     
  56. Schmidt-Erfurth U, Eldem B, Guymer R, Korobelnik JF, Schlingemann RO, Axer-Siegel R, Wiedemann P, Simader C, Gekkieva M, Weichselberger A; EXCITE Study Group. Efficacy and safety of monthly versus quarterly ranibizumab treatment in neovascular age-related macular degeneration: the EXCITE study. Ophthalmology 2011;118:831-9.
     
  57. Mitchell P, Korobelnik JF, Lanzetta P, Holz FG, Pruente C, Schmidt-Erfurth UM, Tano Y, Wolf S. Ranibizumab (Lucentis) in neovascular age-related macular degeneration: evidence from clinical trials. Br J Ophthalmol 2010;94(1):2-13.
     
  58. Spaide R. Ranibizumab according to need: a treatment for age-related macular degeneration. Am J Ophthalmol 2007;143:679-80.
     
  59. Regillo CD. Overview of Maintenance Phase Paradigms for Anti-VEGF Therapy for Wet AMD. Retina: The Cutting Edge. AAO 2009 Subspeciality Day, 23-24 October San Francisco, CA, USA. Section III, AMD Part III-Exudative AMD:40.
     
  60. Singh RP, Fu EX, Smith SD, Williams DR, Kaiser PK. Predictive factors of visual and anatomical outcome after intravitreal bevacizumab treatment of neovascular age-related macular degeneration: an optical coherence tomography study. Br J Ophthalmol 2009;93:1353-8.
     
  61. Kiss CG, Geitzenauer W, Simader C, Gregori G, Schmidt-Erfurth U. Evaluation of ranibizumab-induced changes in high-resolution optical coherence tomographic retinal morphology and their impact on visual function. Invest Ophthalmol Vis Sci 2009;50:2376-83.
     
  62. Sayanagi K, Sharma S, Kaiser PK. Photoreceptor status after antivascular endothelial growth factor therapy in exudative age-related macular degeneration. Br J Ophthalmol 2009;93(5):622-6.
     
  63. Kaluzny JJ, Wojtkowski M, Sikorski BL, Szkulmowski M, Szkulmowska A, Bajraszewski T, Fujimoto JG, Duker JS, Schuman JS, Kowalczyk A. Analysis of the outer retina reconstructed by high-resolution, three-dimensional spectral domain optical coherence tomography. Ophthalmic Surg Lasers Imaging 2009;40:102-8.
     
  64. Chen Y, Vuong LN, Liu J, Ho J, Srinivasan VJ, Gorczynska I, Witkin AJ, Duker JS, Schuman J, Fujimoto JG. Three-dimensional ultrahigh resolution optical coherence tomography imaging of age-related macular degeneration. Opt Express 2009;17:4046-60.
     
  65. Witkin AJ, Vuong LN, Srinivasan VJ, Gorczynska I, Reichel E, Baumal CR, Rogers AH, Schuman JS, Fujimoto JG, Duker JS. High-speed ultrahigh resolution optical coherence tomography before and after ranibizumab for age-related macular degeneration. Ophthalmology 2009;116:956-63.
     
  66. Lee SJ, Lee CS, Koh HJ. Posterior vitreomacular adhesion and risk of exudative age-related macular degeneration: paired eye study. Am J Ophthalmol 2009;147:621-626.
     
  67. Mojana F, Cheng L, Bartsch DU, Silva GA, Kozak I, Nigam N, Freeman WR. The role of abnormal vitreomacular adhesion in age-related macular degeneration: spectral optical coherence tomography and surgical results. Am J Ophthalmol 2008;146:218-227.
     
  68. Ahlers C, Golbaz I, Einwallner E, Dunavölgyi R, Malamos P, Stock G, Pruente C, Schmidt-Erfurth U.  Identification of optical density ratios in subretinal fluid as a clinically relevant biomarker in exudative macular disease. Invest Ophthalmol Vis Sci 2009;50:3417-24.
     
  69. Brinkmann CK, Wolf S, Wolf-Schnurrbusch UE. Multimodal imaging in macular diagnostics: combined OCT-SLO improves therapeutical monitoring. Graefes Arch Clin Exp Ophthalmol 2008;246:9-16.
     
  70. Wolf-Schnurrbusch UE, Enzmann V, Brinkmann CK, Wolf S. Morphologic changes in patients with geographic atrophy assessed with a novel spectral OCT-SLO combination. Invest Ophthalmol Vis Sci 2008;49:3095-9.
     
  71. Fleckenstein M, Charbel Issa P, Helb HM, Schmitz-Valckenberg S, Finger RP, Scholl HP, Loeffler KU, Holz FG. High-resolution spectral domain-OCT imaging in geographic atrophy associated with age-related macular degeneration. Invest Ophthalmol Vis Sci 2008;49:4137-44.
     
  72. Imamura Y, Fujiwara T, Margolis R, Spaide RF. Enhanced depth imaging optical coherence tomography of the choroid in central serous chorioretinopathy. Retina 2009;29:1469-73.
     
  73. Spaide RF. Age-related choroidal atrophy. Am J Ophthalmol  2009 ;147:801-10.
     
  74. Fujiwara T, Imamura Y, Margolis R, Slakter JS, Spaide RF. Enhanced depth imaging optical coherence tomography of the choroid in highly myopic eyes. Am J Ophthalmol 2009;148:445-50.
     
  75. Wojtkowski M, Srinivasan V, Fujimoto JG, Ko T, Schuman JS, Kowalczyk A, Duker JS. Three-dimensional retinal imaging with high-speed ultrahigh-resolution optical coherence tomography. Ophthalmology 2005;112:1734-46.
     
  76. Gorczynska I, Srinivasan VJ, Vuong LN, Chen RW, Liu JJ, Reichel E, Wojtkowski M, Schuman JS, Duker JS, Fujimoto JG. Projection OCT fundus imaging for visualising outer retinal pathology in non-exudative age-related macular degeneration. Br J Ophthalmol 2009;93:603-9.
     
  77. Srinivasan VJ, Monson BK, Wojtkowski M, Bilonick RA, Gorczynska I, Chen R, Duker JS, Schuman JS, Fujimoto JG. Characterization of outer retinal morphology with high-speed, ultrahigh-resolution optical coherence tomography. Invest Ophthalmol Vis Sci 2008;49:1571-9.
     
  78. Heiferman M, Simonett J; Fawzi A. En Face OCT Imaging in Retinal Disorders An emerging technique offers innovative insights. Retinal Physician 2015;12:45-48, 50.
     
  79. Lau T, Wong IY, Iu L, Chhablani J, Yong T, Hideki K, Lee J, Wong R. En-face optical coherence tomography in the diagnosis and management of age-relate macular degeneration and polypoidal choroidal vasculopathy. Indian J Ophthalmol 2015;63:378-83.
  80. Nunes RP, Gregori G, Yehoshua Z, Stetson PF, Feuer W, Moshfeghi AA, Rosenfeld PJ. Predicting the progression of geographic atrophy in age-related macular degeneration with SD-OCT en face imaging of the outer retina. Ophthalmic Surg Lasers Imaging Retina 2013;44:344-359.
     
  81. Flores-Moreno I, Arias-Barquet L, Rubio-Caso MJ, Ruiz-Moreno JM, Duker JS, Caminal JM. En face swept-source optical coherence tomography in neovascular age-related macular degeneration. Br J Ophthalmol 2015;99:1260-7.
     
  82. Sayanagi K, Gomi F, Akiba M, Sawa M, Hara C, Nishida K. En-face high-penetration optical coherence tomography imaging in polypoidal choroidal vasculopathy. Br J Ophthalmol 2015;99:29-35.
     
  83. Kokame GT, Shantha JG, Hirai K, Ayabe J. En Face Spectral-Domain Optical Coherence Tomography for the Diagnosis and Evaluation of Polypoidal Choroidal Vasculopathy. Ophthalmic Surg Lasers Imaging Retina 2016;47:737-44.
     
  84. Semoun O, Coscas F, Coscas G, Lalloum F, Srour M, Souied EH. En face enhanced depth imaging optical coherence tomography of polypoidal choroidal vasculopathy. Br J Ophthalmol 2016;100:1028-34.
     
  85. Alasil T, Ferrara D, Adhi M, Brewer E, Kraus MF, Baumal CR, Hornegger J, Fujimoto JG, Witkin AJ, Reichel E, Duker JS, Waheed NK.En face imaging of the choroid in polypoidal choroidal vasculopathy using swept-source optical coherence tomography. Am J Ophthalmol 2015;159:634-643.
     
  86. Grulkowski I, Liu JJ, Potsaid B, Jayaraman V, Lu CD, Jiang J, Cable AE, Duker JS, Fujimoto JG. Retinal, anterior segment and full eye imaging using ultrahigh speed swept source OCT with vertical-cavity surface emitting lasers. Biomed Opt Express 2012;3:2733-51.
     
  87. Ruiz-Medrano J, Flores-Moreno I, Peña-García P, Montero JA, García-Feijóo J, Duker JS, Ruiz-Moreno JM. Analysis of age-related choroidal layers thinning in healthy eyes using swept-source coherence tomography. Retina 2016 Oct [Epub ahead of print].
     
  88. Uzun S, Pehlivan E. Choroidal and Retinal Thickness in Children With Different Refractive Status Measured by Swept-Source Optical Coherence Tomography. Am J Ophthalmol 2016;169:298-9. 
     
  89. Maruko I, Arakawa H, Koizumi H, Izumi R, Sunagawa H, Iida T. Age-Dependent Morphologic Alterations in the Outer Retinal and Choroidal Thicknesses Using Swept Source Optical Coherence Tomography. PLoS One 2016;11(7):e0159439. 
     
  90. Kuroda Y, Ooto S, Yamashiro K, Oishi A, Nakanishi H, Tamura H, Ueda-Arakawa N, Yoshimura N. Increased Choroidal Vascularity in Central Serous Chorioretinopathy Quantified Using Swept-Source Optical Coherence Tomography. Am J Ophthalmol 2016;169:199-207.
     
  91. Govetto A, Sarraf D, Figueroa MS, Pierro L, Ippolito M, Risser G, Bandello F, Hubschman JP. Choroidal thickness in non-neovascular versus neovascular age-related macular degeneration: a fellow eye comparative study. Br J Ophthalmol 2016 Sep 1. pii: bjophthalmol-2016-309281 [Epub ahead of print].
     
  92. Koustenis A Jr, Harris A, Gross J, Januleviciene I, Shah A, Siesky B. Optical coherence tomography angiography: an overview of the technology and an assessment of applications for clinical research. Br J Ophthalmol 2016 Oct 4. pii: bjophthalmol-2016-309389 [Epub ahead of print].
     
  93. Brijesh T, Shorya A. Assessment of Choroidal Topographic Changes by Swept Source Optical Coherence Tomography After Intravitreal Ranibizumab for Exudative Age-Related Macular Degeneration. Am J Ophthalmol 2016;162:202-3.
     
  94. Choi W, Moult EM, Waheed NK, Adhi M, Lee B, Lu CD, de Carlo TE, Jayaraman V,Rosenfeld PJ, Duker JS, Fujimoto JG. Ultrahigh-Speed, Swept-Source Optical Coherence Tomography Angiography in Nonexudative Age-Related Macular Degeneration with Geographic Atrophy. Ophthalmology 2015;122:2532-44.
     
  95. Ghasemi Falavarjani K, Al-Sheikh M, Akil H, Sadda SR. Image artefacts in swept-source optical coherence tomography angiography. Br J Ophthalmol. 2016 Jul 20. pii: bjophthalmol-2016-309104 [Epub ahead of print].
     
  96. Novais EA, Adhi M, Moult EM, Louzada RN, Cole ED, Husvogt L, Lee B, Dang S, Regatieri CV, Witkin AJ, Baumal CR, Hornegger J, Jayaraman V, Fujimoto JG, Duker  JS, Waheed NK. Choroidal Neovascularization Analyzed on Ultrahigh-Speed Swept-Source Optical Coherence Tomography Angiography Compared to Spectral-Domain Optical Coherence Tomography Angiography. Am J Ophthalmol 2016;164:80-8.
     
  97. Roisman L, Zhang Q, Wang RK, Gregori G, Zhang A, Chen CL, Durbin MK, An L, Stetson PF, Robbins G, Miller A, Zheng F, Rosenfeld PJ. Optical Coherence Tomography Angiography of Asymptomatic Neovascularization in Intermediate Age-Related Macular Degeneration. Ophthalmology 2016;123:1309-19. 
     
  98. Coscas GJ, Lupidi M, Coscas F, Cagini C, Souied EH. Optical Coherence Tomography Angiography versus Traditional Multimodal Imaging in Assessing the Activity of Exudative Age-Related Macular Degeneration: A New Diagnostic Challenge. Retina 2015;35:2219-28.

Optical Coherence Tomography Angiography in AMD

Optical Coherence Tomography Angiography in AMD

English

Authors

João Pedro Marques, MD, MSc

Rufino Silva MD, PhD

Affiliations

Department of Ophthalmology, Centro Hospitalar e Universitário de Coimbra (CHUC), Coimbra, Portugal

Association for Innovation and Biomedical Research on Light and Image (AIBILI), Coimbra, Portugal

Faculty of Medicine, University of Coimbra (FMUC), Coimbra, Portugal

Created in January 2018.

1. BACKGROUND

The advent of optical coherence tomography (OCT) revolutionized retinal imaging by providing a fast, simple, and noninvasive method to assess retinal structure at a microscopic level.(1)

Since its debut more than two decades ago, profound advances in the scanning speed, resolution and image depth led to an explosive growth of OCT in clinical practice, with a significant effect in clinical decision making.(2) Even though technical sophistication improved the OCT diagnostic yield for wet age-related macular degeneration (AMD), there are no agreed-on standards for diagnosing CNV based strictly on cross-sectional OCT(3) and it must ultimately be combined with angiography systems to diagnose CNV accurately. The recent introduction of optical coherence tomography angiography (OCTA), allowed us to redefine our understanding of AMD by conveying three-dimensional, depth-resolved scans of both the superficial and deep retinal capillary plexuses and the choriocapillaris (Video 1). The ability to accurately image CNV noninvasively (Figure 1) and to qualitatively and quantitatively appraise its changes over time (4, 5), turn OCTA into an attractive alternative to conventional dye angiography systems.

Video 1. OCTA allows a three-dimensional depth-resolved view of the retinal and choroidal circulations. This enables us to independently evaluate the vasculature of the inner and outer retina and the choriocapillaris, accurately locating the CNV net, as depicted in this video, recorded with the Optovue Angiovue OCTA system.

 

Figure 1

Figure 1. Example of an active CNV lesion imaged with the Optovue Angiovue OCTA system. The superficial and deep retinal plexuses are depicted, along with the outer retina (normally avascular) and the choriocapillaries. Note the presence of a densely packed CNV net composed of loops, peripheral anastomoses and surrounded by a hypointense halo. On SD-OCT, subretinal fluid can be seen.

 

CNV – choroidal neovasculzarization; SD-OCT – spectral domain optical coherence tomography

 

2. BRIEF OVERVIEW OF THE TECHNOLOGY

OCTA detects differences in amplitude, intensity or phase variance between a series of sequential B-scans taken at the same location of the retina.(6) The degree of decorrelation in signal is then calculated. This enables visualization of the moving part only, assumed to correspond to the endoluminal flow of erythrocytes within the bloodstream. The above procedure is then repeated for a series of different Y-positions in the retina to achieve the three-dimensional data set, from which proprietary algorithms such as split-spectrum amplitude-decorrelation angiography (SSADA), optical microangiography and OCT angiography ratio analysis are used to reconstruct en face angiograms.(7)

Compared to conventional angiography, OCTA offers several advantages, namely the noninvasive nature and fast acquisition time, thus allowing the exam to be repeated as needed without the added risk of dye injection. In addition, OCTA conveys high-resolution details of the retinal and choroidal microvasculature and a depth-resolved analysis within specific axial locations.(7)

 

3. OCTA IN EARLY/INTERMEDIATE AMD

The pathogenesis of AMD is known to be associated with inflammation, oxidative damage, ageing, genetic predisposition and environmental factors. However, there is a growing body of evidence implicating choroidal and retinal blood flow in the development of drusen, with hypoxia being associated with the progression of the disease.(8) The advent of OCTA opened doors into the noninvasive evaluation of early/intermediate AMD and has been contributing to a deeper understanding of the disease.

In a recent study, Toto et al(9) used OCTA to evaluate the superficial and deep vessel density of 37 eyes of patients with early (n=14) and intermediate (n=23) AMD. The results were compared to 21 eyes of 21 age- and sex-matched healthy subjects. The authors found a statistically significant difference in both the superficial and deep retinal plexuses between controls and the intermediate AMD group, although not between controls and the early AMD group. These preliminary results corroborate the hypothesis that ischemia plays an important role in the development of AMD. Further studies, preferably longitudinal and with larger populations, are needed to confirm this hypothesis.

Another hallmark of AMD are reticular pseudodrusen (RPD). RPD have been identified as a significant risk factor for late forms of AMD.(10, 11) A recent study used OCTA to evaluate choriocapillaris non-perfusion in early/intermediate AMD eyes with and without RPD.(12) The authors found that eyes with RPD have significantly larger areas of choriocapillaris non-perfusion compared to eyes with drusen and no RPD. Furthermore, the percent choriocapillaris area of non-perfusion (PCAN) strongly correlated with visual acuity, thus providing a potential mechanistic explanation for vision compromise in RPD compared to other forms of drusen in AMD.

 

4. OCTA IN EXUDATIVE AMD AND ITS NEOVASCULAR PHENOTYPES

Several studies have demonstrated that OCTA can accurately identify type 1 and type 2 neovascularization,(3, 4, 13-16) offering an unrivaled morphological characterization of the choroidal neovascularization (CNV) net that far exceeds FA, ICGA and structural OCT.

Since the OCTA image is not obscured by dye leakage, a clear depiction of the CNV net is provided (Figure 2). Findings from a study by de Carlo et al(3) where 30 eyes with OCTA and same-day FA were evaluated, reported a sensitivity of 50% (4/8) and a specificity of 91% (20/22) for CNV detection with OCTA.

Morphological changes have been reported as the CNV net evolves from active to inactive stages. Characteristic features suggestive of an active CNV net include the presence of a tangle of vessels in a well-defined shape (lacy-wheel, glomerular, sea fan, etc), the presence of anastomoses or loops branching in numerous tiny capillaries, the presence of a peripheral arcade and the presence of a hypointense halo (Figure 3). In contrast, inactive chronic CNVs have larger, more mature vessels, a 'dead tree' appearance with the absence of the anastomoses, loops and peripheral arcades.(13) (Figure 4)

The non-invasive nature of the technique allows comprehensive lesion monitoring after treatment with anti-VEGF compounds. Quantitative measurements of the CNV flow area and flow index reported by Huang et al(17) showed a rapid shutdown of flow over the initial 2 weeks after treatment with anti-VEGF, followed by reappearance of the CNV channel by the fourth week and fluid re-accumulation at 6 weeks. As demonstrated by Lumbroso et al(5) and Marques et al(4), sequential examinations can shadow morphological changes in the neovascular network, closely tracking the timings of vascular network remodeling in patients undergoing treatment (Figure 5). Ultimately, this may allow for the advent of tailored and customized treatment regimens for wet AMD patients, built upon the CNV net morphology at baseline and its response to treatment.(1)

Figure 2

Figure 2. OCTA provides a clear, depth-resolved visualization of the CNV net, not obscured by the staining or pooling effects seen in FA. The image on the right was obtained with the Optovue Angiovue OCTA system.

 

FA – Fluorescein Angiography; OCTA – Optical Coherence Tomography Angiography; CNV – choroidal neovascularization

Figure 3

Figure 3. Example of a chronic but still active CNV lesion imaged with the Optovue Angiovue OCTA system. The patient underwent treatment with 8 intravitreal injections of anti-VEGF so far. Note the presence of loops, peripheral arcades and a hypointense halo surrounding the lesion. On SD-OCT, both subretinal and intraretinal fluid can be seen.

 

CNV – choroidal neovasculzarization; VEGF – vascular endothelial growth factor; SD-OCT – spectral domain optical coherence tomography

Figure 4

Figure 4. Example of a chronic, inactive CNV lesion imaged with the Optovue Angiovue OCTA system. The larger, more mature vessels are depicted in a 'dead tree' appearance but no peripheral anastomosis or loops are seen. No exudation is visible on SD-OCT and macular atrophy is observed instead.

 

CNV – choroidal neovasculzarization; SD-OCT – spectral domain optical coherence tomography

Figure 5

Figure 5. The non-invasive nature of OCTA allows comprehensive lesion monitoring after treatment with anti-VEGF compounds. Sequential examinations can shadow morphological changes in the neovascular network, closely tracking the timings of vascular network remodelling in patients undergoing treatment with anti-VEGF. The image on the left depicts de status of the CNV net before initiating treatment with intravitreal aflibercept. One week after the IVI (middle), marked regression of the peripheral anastomosis and substantial capillary dropout can be seen, along with decreased new vessel density. The feeder trunk and some other central vessels are still visible. Note recanalization of the peripheral anastomosis, resumption of capillary sprouting and increased new vessel density 1 month after the IVI (right). Images acquired with the Optovue Angiovue OCTA system. Adapted with permission from Marques et al (2016).(4)

IVI – intravitreal injection; VEGF – vascular endothelial growth factor

 

4.1. NONEXUDATIVE CNV

OCTA may also be useful in the screening of eyes at risk for CNV. Unilateral CNV is an established risk factor for the development of neovascularization in the fellow eye. Screening this population for early detection of CNV may have both therapeutic and prognostic implications. A pilot study by Palejwala et al(18) showed that OCTA was able to identify foci of CNV that were not discernible on angiography systems and coined it ‘nonexudative CNV’. OCTA detects CNV by the presence of an abnormal pattern of vascular flow above the Bruch membrane, therefore being able to identify neovascular lesions that do not leak on FA.

A prospective study conducted to determine the natural history of nonexudative CNV found that 23/160 eyes with intermediate AMD had subclinical CNV at presentation (prevalence 14.4%).(19) After the detection of nonexudative CNV, the risk of exudation at 12 months was 15.2 times (95% confidence interval, 4.2-55.4) greater compared with eyes without subclinical CNV. For eyes with nonexudative CNV, recommendations include more frequent follow-up and home monitoring with an amsler grid. So far, intravitreal therapy is not recommended until prospective studies are performed.(19)

 

4.2. OCTA IN TYPE 3 NEOVASCULARIZATION/RETINAL ANGIOMATOUS PROLIFERATION (RAP)

Qualitative and quantitative analyses of type 3 neovascular complexes can be performed using OCTA.(20, 21) Querques et al(22) evaluated the features of retinal angiomatous proliferation (RAP) with this new imaging modality and found lesions emerging from the deep capillary plexus, forming a clear, tuft-shaped, high-flow network in the outer retinal segment in all eyes, abutting in the sub-RPE space (Figure 6). The authors also reported a small, clew-like lesion present in the choriocapillaris and that, in some cases, this clew-like lesion seemed to be connected to the choroid through a small-caliber vessel. Longitudinal imaging of type 3 neovascularization showed that OCTA could detect changes in the vascular complex even before the presence of exudation on structural OCT. It was also noted that OCTA enabled the distinction between hyper-reflective vascular structures of type 3 neovascularization from hyper-reflective foci devoid of flow, which may correlate to pigment migration.(23) In fact, compared with conventional imaging, OCTA may improve the detection and delineation of vascular changes occurring in type 3 neovascularization (Video 2).

Video 2. Type 3 neovascularization imaged with the Zeiss Angioplex OCTA system. The video shows the presence of the neovascular complex deep in the retina.

 

A recent study by Martins et al(24), qualitatively and quantitatively evaluated OCTA changes in fellow eyes of type 3 neovascularization. These eyes were shown to have a significantly inferior perfusion density and vessel density in the superficial retinal layer (SRL), deep retinal layer (DRL) and choriocapillaris layer (CL), when compared to both healthy controls and early/intermediate AMD eyes. The authors hypothesized that the compromised vascular perfusion in the CL may be associated with reduced oxygen transport to the RPE and neuro-sensory retina, thus inducing a sub ischemic state that may potentiate the development of RAP.(24). Further longitudinal studies with OCTA may bring additional insights on the pathophysiology of type 3 neovascularization and establish OCTA parameters as biomarkers of this neovascular phenotype of AMD.

Figure 6

Figure 6. Multimodal imaging of a 73-year old patient with early type 3 NV. OCTA angiogram obtained with the Optovue Angiovue OCTA system (top) shows a large anomalous vessel in the deep retinal plexus (green and red line intersection). The hyper-reflective foci observed on structural SD-OCT (middle) within this particular location have abnormal flow (yellow arrow). However, intraretinal fluid accumulation cannot be seen. Fluorescein angiography (bottom) confirms the presence of an initial type 3 NV lesion with leakage on late phases.

 

OCTA – optical coherence tomography angioraphy; SD-OCT – spectral domain optical coherence tomography angiography

4.3. OCTA IN POLYPOIDAL CHOROIDAL VASCULOPATHY (PCV)/ANEURISMATIC TYPE 1 NEOVASCULARIZATION

Due to a superior visualization of polyps and branching vascular network (BVN), ICGA has long been the gold standard for the diagnosis of polypoidal choroidal vasculopathy (PCV). Previous studies have shown that OCTA is comparable to ICGA for the detection of the BVN, often showing it with even greater detail than ICGA.(7) On the other hand, the rate of polyp detection by OCTA was much more variable, ranging from 17 to 85%.(25-27) A good example of PCV imaged with OCTA is shown on Video 3.

Video 3. Polypoidal Choroidal Vasculopathy imaged with the Zeiss Angioplex OCTA system. The video shows both the branching vascular network and a polyp. Depending on the position of the slab, the polyp may show either hyperflow or hypoflow. A multimodal evaluation of this patient is shown on Figure 7.

 

Cross-sectional OCTA of polyps (Figure 7) showed that the lumen is largely devoid of flow.(26, 27) and that the polyp area measured on OCTA was consistently smaller when compared to ICGA.(28) En face OCTA imaging of the polyps was reported to show a more common hypoflow round structure (75%) or less common (25%) hyperflow round structure surrounded by a hypointense halo.(29) It is now widely accepted that the appearance of the polyps on OCTA is highly dependent on the correct positioning of the slab. Some authors hypothesize that the slow or turbulent flow within the polyp may explain its frequently hypoflow appearance.

Figure 7

Figure 7.

PCV lesion with a large polyp imaged with OCTA, ICGA, FA and SD-OCT. Note a hypoflow round structure surrounded by a hypointense halo on OCTA. The details and extent of the BVN are better identified on OCTA than ICGA.

 

OCTA – optical coherence tomography angioraphy; ICGA – indocyanine green angiography; FA – fluorescein angiography; IR – infrared imaging; SD-OCT – spectral domain optical coherence tomography; PCV – polypoidal choroidal vasculopathy; BVN – branching vascular network.

 

5. OCTA IN GEOGRAPHIC ATROPHY

Geographic atrophy (GA) often follows the disappearance of drusen and is characterized by the well-delineated loss of the RPE, photoreceptors, and choriocapillaris. In fact, late atrophic AMD is an important cause of irreversible vision loss, even in patients with neovascular AMD, because the long-term use of anti-VEGF successfully suppresses neovascularization and results in the progression of these lesions to atrophy.(30)

With the advent of OCTA, it is now possible to visualize the vascular changes that occur in dry AMD (Figures 8 and 9). OCTA shows choriocapillaris flow impairment under the regions of GA.(31-33) In these areas, larger choroidal vessels may be displaced into the area ordinarily occupied by the choriocapillaris and may be seen on the en face OCTA image at the depth level where the choriocapillaris is ordinarily seen.(33) In many cases, the areas of choriocapillaris changes extend beyond the margins of GA in an asymmetric pattern. These changes outside of the margins of GA may be either quite extensive or very limited and subtle.(33) One of the debates regarding choriocapillaris changes in patients with GA is whether they truly represent the absence of flow or merely reduced flow. If the velocity of flow in the vessels is very slow, then OCTA may not be able to detect this slow flow versus the parasitic retinal motion.(33) With the continuous improvement in the software capabilities, this problem will most likely disappear in a near future.

The ability to image the choriocapillaris structure and flow impairments may be useful for detecting and monitoring the progression of dry AMD and for monitoring treatment responses in GA clinical trials.

Figure 8

Figure 8. Geographic atrophy (GA) imaged with the Optovue Angiovue OCTA system. The area of GA can be easily observed on structural SD-OCT (A) and on en face OCT (B). Its limits correlate well with the OCTA scan at the level of the choriocapillaris (C). Here, the loss of choriocapillaris allows visualization of the larger choroidal vessels that rest below.

 

OCTA – optical coherence tomography angioraphy; SD-OCT – spectral domain optical coherence tomography

 

Figure 9

Figure 9. Geographic atrophy (GA) imaged with the Spectralis OCTA system. On the OCTA scan (top right image), the window defect on the choriocapillaris allows visualization of the larger choroidal vessels.

 

OCTA – optical coherence tomography angioraphy

 

OCTA ARTIFACTS

Although OCTA merits all the credit for the recent reshaping of our understanding of AMD, attention should be drawn to the appearance of several doubtful findings, mainly related to segmentation errors, ‘ghost images’ due to RPE reflection and the prototype nature of many OCTA devices. The so-called ‘false-friend’ images can induce significant pitfalls in interpreting OCTA scans and lead to unnecessary treatment. (34, 35)

 

FUTURE PERSPECTIVES

The constant software updates and technical enhancements are making OCTA devices faster, more precise and ultimately more reliable. For instance, the introduction of 3D projection artifact removal and the incorporation of tools to quantitative evaluate vessel density, perfusion/non perfusion, etc increased the usefulness and trustworthiness of OCTA. We believe that this will further reduce the use of conventional angiographic systems, turning OCTA in the gold standard for diagnosis, managing and monitoring of AMD and other retinal diseases. Reading centres around the world are beginning to work with this technology and the inclusion of OCTA images in clinical trials will most likely lead to the advent of OCTA biomarkers of AMD progression and treatment response.

Furthermore, the combination of OCTA and OCT Leakage (OCT-L) seems particularly promising in order to evaluate both vessel parameters and fluid accumulation non-invasively.(36, 37)

 

CONCLUSION

We embark on a new era of retinal imaging with great expectations in mind, hoping that OCTA can point out new clinical coordinates to improve the everyday management of AMD patients.

 

>> References

References of Optical Coherence Tomography in Age-related Macular Degeneration

References of Optical Coherence Tomography in Age-related Macular Degeneration

English

REFERENCES

  1. Marques JP, Silva R. Optical coherence tomography angiography in wet age-related macular degeneration (AMD). Eye Sci. 2016.
     
  2. Puliafito CA. OCT angiography: the next era of OCT technology emerges. Ophthalmic surgery, lasers & imaging retina. 2014;45(5):360.
     
  3. de Carlo TE, Bonini Filho MA, Chin AT, Adhi M, Ferrara D, Baumal CR, et al. Spectral-domain optical coherence tomography angiography of choroidal neovascularization. Ophthalmology. 2015;122(6):1228-38.
     
  4. Marques JP, Costa JF, Marques M, Cachulo ML, Figueira J, Silva R. Sequential Morphological Changes in the CNV Net after Intravitreal Anti-VEGF Evaluated with OCT Angiography. Ophthalmic research. 2016;55(3):145-51.
     
  5. Lumbroso B, Rispoli M, Savastano MC. Longitudinal Optical Coherence Tomography-Angiography Study of Type 2 Naive Choroidal Neovascularization Early Response after Treatment. Retina. 2015;35(11):2242-51.
     
  6. Gao SS, Jia Y, Zhang M, Su JP, Liu G, Hwang TS, et al. Optical Coherence Tomography Angiography. Investigative ophthalmology & visual science. 2016;57(9):OCT27-36.
     
  7. Tan ACS, Tan GS, Denniston AK, Keane PA, Ang M, Milea D, et al. An overview of the clinical applications of optical coherence tomography angiography. Eye. 2017.
     
  8. Remsch H, Spraul CW, Lang GK, Lang GE. Changes of retinal capillary blood flow in age-related maculopathy. Graefe's archive for clinical and experimental ophthalmology = Albrecht von Graefes Archiv fur klinische und experimentelle Ophthalmologie. 2000;238(12):960-4.
     
  9. Toto L, Borrelli E, Di Antonio L, Carpineto P, Mastropasqua R. Retinal Vascular Plexuses' Changes in Dry Age-Related Macular Degeneration, Evaluated by Means of Optical Coherence Tomography Angiography. Retina. 2016;36(8):1566-72.
     
  10. Gil JQ, Marques JP, Hogg R, Rosina C, Cachulo ML, Santos A, et al. Clinical features and long-term progression of reticular pseudodrusen in age-related macular degeneration: findings from a multicenter cohort. Eye. 2017;31(3):364-71.
     
  11. Hogg RE, Silva R, Staurenghi G, Murphy G, Santos AR, Rosina C, et al. Clinical characteristics of reticular pseudodrusen in the fellow eye of patients with unilateral neovascular age-related macular degeneration. Ophthalmology. 2014;121(9):1748-55.
     
  12. Nesper PL, Soetikno BT, Fawzi AA. Choriocapillaris Nonperfusion is Associated With Poor Visual Acuity in Eyes With Reticular Pseudodrusen. American journal of ophthalmology. 2017;174:42-55.
     
  13. Coscas GJ, Lupidi M, Coscas F, Cagini C, Souied EH. OPTICAL COHERENCE TOMOGRAPHY ANGIOGRAPHY VERSUS TRADITIONAL MULTIMODAL IMAGING IN ASSESSING THE ACTIVITY OF EXUDATIVE AGE-RELATED MACULAR DEGENERATION: A New Diagnostic Challenge. Retina. 2015;35(11):2219-28.
     
  14. Moult E, Choi W, Waheed NK, Adhi M, Lee B, Lu CD, et al. Ultrahigh-speed swept-source OCT angiography in exudative AMD. Ophthalmic surgery, lasers & imaging retina. 2014;45(6):496-505.
     
  15. Jia Y, Bailey ST, Wilson DJ, Tan O, Klein ML, Flaxel CJ, et al. Quantitative optical coherence tomography angiography of choroidal neovascularization in age-related macular degeneration. Ophthalmology. 2014;121(7):1435-44.
     
  16. Kuehlewein L, Bansal M, Lenis TL, Iafe NA, Sadda SR, Bonini Filho MA, et al. Optical Coherence Tomography Angiography of Type 1 Neovascularization in Age-Related Macular Degeneration. American journal of ophthalmology. 2015;160(4):739-48 e2.
     
  17. Huang D, Jia Y, Rispoli M, Tan O, Lumbroso B. Optical Coherence Tomography Angiography of Time Course of Choroidal Neovascularization in Response to Anti-Angiogenic Treatment. Retina. 2015;35(11):2260-4.
     
  18. Palejwala NV, Jia Y, Gao SS, Liu L, Flaxel CJ, Hwang TS, et al. Detection of Nonexudative Choroidal Neovascularization in Age-Related Macular Degeneration with Optical Coherence Tomography Angiography. Retina. 2015.
     
  19. de Oliveira Dias JR, Zhang Q, Garcia JMB, Zheng F, Motulsky EH, Roisman L, et al. Natural History of Subclinical Neovascularization in Nonexudative Age-Related Macular Degeneration Using Swept-Source OCT Angiography. Ophthalmology. 2017.
     
  20. Kuehlewein L, Dansingani KK, de Carlo TE, Bonini Filho MA, Iafe NA, Lenis TL, et al. Optical Coherence Tomography Angiography of Type 3 Neovascularization Secondary to Age-Related Macular Degeneration. Retina. 2015;35(11):2229-35.
     
  21. Tan AC, Dansingani KK, Yannuzzi LA, Sarraf D, Freund KB. Type 3 Neovascularization Imaged with Cross-Sectional and En Face Optical Coherence Tomography Angiography. Retina. 2016.
     
  22. Querques G, Miere A, Souied EH. Optical Coherence Tomography Angiography Features of Type 3 Neovascularization in Age-Related Macular Degeneration. Dev Ophthalmol. 2016;56:57-61.
     
  23. Tan AC, Dansingani KK, Yannuzzi LA, Sarraf D, Freund KB. Type 3 Neovascularization Imaged with Cross-Sectional and En Face Optical Coherence Tomography Angiography. Retina. 2017;37(2):234-46.
     
  24. Martins A, Farinha C, Raimundo M, Lopes M, Santos AR, Melo P, et al. Multimodal Evaluation of the Fellow Eye of Patients with Retinal Angiomatous Proliferation. Ophthalmic research. 2017.
     
  25. Inoue M, Balaratnasingam C, Freund KB. Optical Coherence Tomography Angiography of Polypoidal Choroidal Vasculopathy and Polypoidal Choroidal Neovascularization. Retina. 2015.
     
  26. Tanaka K, Mori R, Kawamura A, Nakashizuka H, Wakatsuki Y, Yuzawa M. Comparison of OCT angiography and indocyanine green angiographic findings with subtypes of polypoidal choroidal vasculopathy. The British journal of ophthalmology. 2017;101(1):51-5.
     
  27. Tomiyasu T, Nozaki M, Yoshida M, Ogura Y. Characteristics of Polypoidal Choroidal Vasculopathy Evaluated by Optical Coherence Tomography Angiography. Investigative ophthalmology & visual science. 2016;57(9):OCT324-30.
     
  28. Takayama K, Ito Y, Kaneko H, Kataoka K, Sugita T, Maruko R, et al. Comparison of indocyanine green angiography and optical coherence tomographic angiography in polypoidal choroidal vasculopathy. Eye. 2017;31(1):45-52.
     
  29. Srour M, Querques G, Semoun O, El Ameen A, Miere A, Sikorav A, et al. Optical coherence tomography angiography characteristics of polypoidal choroidal vasculopathy. The British journal of ophthalmology. 2016;100(11):1489-93.
     
  30. Velez-Montoya R, Oliver SC, Olson JL, Fine SL, Mandava N, Quiroz-Mercado H. Current knowledge and trends in age-related macular degeneration: today's and future treatments. Retina. 2013;33(8):1487-502.
     
  31. Moult EM, Waheed NK, Novais EA, Choi W, Lee B, Ploner SB, et al. Swept-Source Optical Coherence Tomography Angiography Reveals Choriocapillaris Alterations in Eyes with Nascent Geographic Atrophy and Drusen-Associated Geographic Atrophy. Retina. 2016;36 Suppl 1:S2-S11.
     
  32. Kvanta A, Casselholm de Salles M, Amren U, Bartuma H. Optical Coherence Tomography Angiography of the Foveal Microvasculature in Geographic Atrophy. Retina. 2017;37(5):936-42.
     
  33. Waheed NK, Moult EM, Fujimoto JG, Rosenfeld PJ. Optical Coherence Tomography Angiography of Dry Age-Related Macular Degeneration. Dev Ophthalmol. 2016;56:91-100.
     
  34. Spaide RF, Fujimoto JG, Waheed NK. Image Artifacts in Optical Coherence Tomography Angiography. Retina. 2015.
     
  35. Coscas G, Lupidi M, Cagini C, Coscas F. 'False-friend' images on optical coherence tomography angiography: early choroidal neovascularization or artefact? Acta ophthalmologica. 2016.
     
  36. Farinha C, Santos T, Marques IP, Marques JP, Ribeiro L, Figueira J, et al. OCT-Leakage Mapping. Ophthalmology Retina.1(6):486-96.
     
  37. Cunha-Vaz J, Santos T, Ribeiro L, Alves D, Marques I, Goldberg M. OCT-Leakage: A New Method to Identify and Locate Abnormal Fluid Accumulation in Diabetic Retinal Edema. Investigative ophthalmology & visual science. 2016;57(15):6776-83.