You are on page 1of 29

ARTICLE IN PRESS

Progress in Retinal and Eye Research 25 (2006) 325–353


www.elsevier.com/locate/prer

Retinal assessment using optical coherence tomography$


Rogério A. Costaa,!, Mirian Skafb, Luiz A.S. Melo Jr.b, Daniela Caluccia, Jose A. Cardilloa,
Jarbas C. Castroc, David Huangd, Maciej Wojtkowskie
a
U.D.A.T.—Retina Diagnostic and Treatment Division, Hospital de Olhos de Araraquara, Rua Padre Duarte 989 ap 172, Araraquara, SP 14801 310, Brazil
b
Glaucoma Section, Hospital de Olhos de Araraquara, Araraquara, SP, Brazil
c
Instituto de Fı´sica de São Carlos—USP, São Carlos-SP, Brazil
d
Department of Ophthalmology, University of Southern California, Los Angeles, CA, USA
e
Institute of Physics, Nicolaus Copernicus University, Torun, Poland

Abstract

Over the 15 years since the original description, optical coherence tomography (OCT) has become one of the key diagnostic
technologies in the ophthalmic subspecialty areas of retinal diseases and glaucoma. The reason for the widespread adoption of this
technology originates from at least two properties of the OCT results: on the one hand, the results are accessible to the non-specialist
where microscopic retinal abnormalities are grossly and easily noticeable; on the other hand, results are reproducible and exceedingly
quantitative in the hands of the specialist. However, as in any other imaging technique in ophthalmology, some artifacts are expected to
occur. Understanding of the basic principles of image acquisition and data processing as well as recognition of OCT limitations are
crucial issues to using this equipment with cleverness.
Herein, we took a brief look in the past of OCT and have explained the key basic physical principles of this imaging technology. In
addition, each of the several steps encompassing a third generation OCT evaluation of retinal tissues has been addressed in details.
A comprehensive explanation about next generation OCT systems has also been provided and, to conclude, we have commented on the
future directions of this exceptional technique.
r 2006 Elsevier Ltd. All rights reserved.

Keywords: Artifacts; Cross-sectional; Fourier domain; Glaucoma; Interferometer; Macula; Macular map; Measurement; Nerve fiber layer; Optic disc;
Optical coherence tomography (OCT); Photoreceptor; Retinal boundary; Retinal thickness; Spectral

Contents

1. History of optical coherence tomography (OCT): conception of the idea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326


2. Basic physical principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326
3. OCT in clinical setting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
3.1. Retina (and retinal diseases) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
3.1.1. Image acquisition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
3.1.2. Qualitative analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
3.1.3. Quantitative analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
3.2. RNFL and glaucoma . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
3.2.1. RNFL thickness protocols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337

Abbreviations: A-scan(s), axial scan(s); HRL, highly reflective layer; OCT, optical coherence tomography; RNFL, retinal nerve fiber layer; RPE, retinal
pigment epithelium; RTA, retinal thickness analyzer; SLD, superluminescent diode
$
Supported in part by Fundac- ão de Amparo à Pesquisa do Estado de São Paulo, FAPESP Grant no.: 98/14270-8, and by Grant no.: KBN
4T11E02322.
!Corresponding author. Tel./fax: +55 16 3331 1001.
E-mail address: roger.retina@globo.com (R.A. Costa).

1350-9462/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.preteyeres.2006.03.001
ARTICLE IN PRESS
326 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

3.2.2. Reproducibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338


3.2.3. Diagnostic capability and progression evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
3.2.4. Normative database . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
3.2.5. Future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
3.2.6. Additional considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
4. Next generation OCT devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
4.1. Spectral OCT instrument using Fourier domain detection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
4.1.1. Standard-resolution retinal imaging with high-speed spectral OCT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
4.1.2. High-resolution retinal imaging with high-speed spectral OCT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
4.2. Additional considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
5. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349

1. History of optical coherence tomography (OCT): 2. Basic physical principles


conception of the idea
Clinical examination using the slit lamp has been used
OCT was first developed by David Huang and colleagues for several years as the main instrument for retinal
in James Fujimoto’s laboratory at the Massachusetts structural assessment. Meanwhile, many other imaging
Institute of Technology (MIT) and published in a 1991 techniques have been developed to examine cross-sectional
Science article (Huang et al., 1991). The Fujimoto’s retinal morphology. The confocal scanning laser ophthal-
laboratory was specialized in femtosecond lasers at the moscope (cSLO) forms retinal images by sequentially
time. These are lasers that emit pulses of only several tens collecting reflections from laterally and longitudinally
of femtosecond (million billionth of a second) and can be well-defined retinal volumes. Several cSLO images taken
used to measure the delay of light reflected from tissue with sequential focal depths can generate three-dimen-
structures with near micron precision. Because femtose- sional information on the distribution of retinal reflectivity
cond lasers were too bulky and expensive for routine for topographic and tomographic assessments (Huang,
clinical use, Huang worked on an interferometer system 1999). The longitudinal resolution of a cSLO, however, is
that could use a cheap and compact diode light source to limited to !300 mm due to the available numerical aperture
measure the time-of-flight of light with the same precision. through the pupil and ocular aberrations (Bartsch and
He realized then that this technique, called optical Freeman, 1994). Cross-sectional measurements of the
coherence domain reflectometry, could be the basis of a retina can be achieved in the clinical setting as well by
new imaging technology with unprecedented potential for the instrument coined retinal thickness analyzer (RTA).
non-invasive imaging of retina and other tissues with The RTA employs the principle of optical triangulation to
micron resolution. This new technique was coined optical provide direct measurement of the retinal thickness with an
coherence tomography because it relied on measuring the estimated accuracy of 20–30 mm (Zeimer et al., 1989). The
coherence of light reflected from tissue structures and instrument projects a narrow slit of 543 nm He–Ne laser
generates cross-sectional images, or tomographs. light onto the retina and calculates the distance between the
The initial retinal OCT experiment that Huang con- reflections that correspond to the vitreoretinal and
ducted with Joel Schuman, an ophthalmologist then at chorioretinal interfaces. Although recent advances in this
Harvard, took several hours to acquire a single image. To instrumentation have enabled rapid multiple optical
improve the imaging speed, Fujimoto recruited Eric sectioning of neighboring retinal regions to generate a
Swanson, then working on optical communications at retinal thickness map (Zeimer et al., 1996), information is
MIT Lincoln Laboratory. With Swanson’s crucial assis- restricted to fundus (macular) regions of 2 " 2 mm and
tance, it was first developed an efficient fiber-optic OCT limited qualitative data can be extracted from such imaging
system that was fast enough for clinical testing (Swanson methodology.
et al., 1993). The first clinical tests of retinal scanning were Optical coherence tomography (OCT) is based on the
conducted by Carmen Puliafito’s group that was then at imaging of reflected light. But unlike a simple camera
the Massachusetts Eye and Ear Infirmary, Harvard image that only has transverse dimensions (left/right,
Medical School. The encouraging results lead to commer- up/down), it resolves depth. The depth resolution of OCT
cialization of the technology in the mid-1990s by Hum- is extremely fine, typically on the order of 0.01 mm or 0.4
phrey Instruments, Inc. (since acquired by Carl Zeiss thousandth of an inch. This provides cross-sectional views
Meditec, Inc). The latest Zeiss Stratus OCT system (third (tomography) of internal tissue structures similar to tissue
generation OCT or OCT3) is now used by thousands of sections under a microscope, without disturbing the tissue
ophthalmologists for the management of macular diseases as in histology. Thus, OCT has been described as a method
and glaucoma. for non-invasive tissue ‘‘biopsy.’’
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 327

Fig. 1. The optical coherence tomography beam is scanned across the


Fig. 3. An optical coherence tomography cross-sectional image (gray-
retina (1). The delay of a superficial reflection (2) is shorter than that of
scale image) is built up from many A-scans (red plot lines). Reprinted by
a deeper reflection (3). Reprinted by courtesy of Elsevier (Huang et al.,
courtesy of Elsevier (Huang et al., in progress).
in progress).

Ultrasound imaging and RADAR are also reflectome-


try-based imaging methods. Because OCT employs light,
several advantages are gained. The wavelength of light
(!0.001 mm) is shorter than that of ultrasound (!0.1 mm)
and radio wave (410 mm). Therefore the spatial resolution
of OCT is much higher. And unlike ultrasound imaging,
OCT does not require probe-tissue contact or an immer-
sion fluid since light passes through the air–tissue interface
easily.
Because light travels very rapidly (3 " 108 m/s), it is not
possible to directly measure the time-of-flight delay on a
small spatial scale. The micron-scale resolution of OCT is
achieved by comparing the delays of sample reflections
with the known delay of a reference reflection in an
interferometer. The classic OCT system (Fig. 4) employs a
‘‘low-coherence’’ fiber-optic Michelson interferometer
(Huang et al., 1991). Interferometry measures the effect
Fig. 2. An axial scan (A-scan) of the retina. The amplitude of reflection on of combining 2 light waves. Low coherence means that the
a decibel (dB) scale is plotted against depth. Reprinted by courtesy of
system employs a wide range of wavelengths. We will
Elsevier (Huang et al., in progress).
explain the concepts of interferometry and coherence
separately.
The interferometer (Fig. 4) has source, sample, reference,
In OCT, a beam of light (typically 800–1400 nm and detector arms all centered on a 50/50 fiber coupler.
wavelength in the near infrared) is scanned across Output of the superluminescent diode (SLD) light source is
the tissue sample. The OCT system then collects the launched into the source arm fiber and split by the coupler
reflected light and measures its time-of-flight delay. Light into the sample and reference arms. Sample and reference
reflected from deeper layers have longer propagation reflections are recombined at the coupler and produce
delays than that reflected from more superficial layers interference. This interferometric signal is converted from
(Fig. 1). The amplitude of reflected light can be plotted light to electrical current by a photodetector, processed
against delay (Fig. 2) to demonstrate tissue reflectivity at electronically, and transferred to computer memory.
successively deeper levels of tissue penetration along the To understand how the Michelson interferometer works,
axis of beam propagation. This is called an axial scan let us start with the simple case (Fig. 5) where the sample
(A-scan). As the OCT probe beam is scanned across a arm reflection comes from a simple mirror surface and
sample, many A-scans are acquired to form an image the light source emits only one wavelength. Think of the
(Fig. 3). A color or gray-scale is used to represent the signal sample and reference reflections as 2 waves of light. The
amplitude. coupler partially transfers theses waves to the detector arm,
ARTICLE IN PRESS
328 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

Fig. 6. Combining interference signals from a range of wavelengths (left)


Fig. 4. Schematic of the classic optical coherence tomography system. produces a pulse (right). The width of this pulse determines the axial
Reprinted by courtesy of Elsevier (Huang et al., in progress). resolution of optical coherence tomography. Reprinted by courtesy of
Elsevier (Huang et al., in progress).

interference signals from all wavelength components have


the same phase (peaks and troughs lined up) or are
‘‘coherent.’’ This adds up to large interferometric modula-
tion (see large peaks and troughs in the center of the right-
side waveform). When the mismatch is large (away from
the center of the waveforms), the interference waveforms
vary widely in phase over the wavelength range (peaks and
troughs not lined up) and add up to near a flat line. The
summed interference signal (Fig. 6, right) forms a wave
Fig. 5. In a single-wavelength Michelson interferometer, varying the pulse. In an OCT system, the pulse is demodulated
reference delay produces a sinusoidal oscillation of optical power in the electronically to extract the pulse envelope (shape of the
detector arm. Reprinted by courtesy of Elsevier (Huang et al., in progress).
pulse without the sinusoidal oscillation). The width of the
pulse envelope is the coherence length, which determines
the axial resolution of the OCT system. The coherence
where they combine and produce an interference signal. length is inversely proportional with the wavelength range
Interference can be thought of as the addition of the or ‘‘bandwidth.’’
amplitude of two waves. When the two waves are in phase When the OCT system is used to image an actual tissue
(lined up peak to peak), they interfere constructively, sample, there are many reflections at different depths. As
forming a peak in the interference waveform. When the 2 the reference mirror is scanned, each sample reflection gives
waves are exactly out of phase (lined up peak to trough), rise to a signal pulse when the reference delay matches it.
they interfere destructively, forming a trough in the The plot of the demodulated interferometric signal is the
interference waveform. As the reference mirror is moved, axial scan (Fig. 2) waveform, which represents amplitude
the phase of the reference wave changes, producing a of reflection vs. depth.
sinusoidal interference signal. As the reference mirror All clinical commercial available OCT systems to date
moves through 12 cycle of the source wavelength, the employ SLD light sources. SLDs are similar to the diode
roundtrip delay of the reference wave varies by one lasers inside the common compact disc (CD) player, but
wavelength and the interference signal goes through one are made to emit over a wider range of wavelengths. SLDs
cycle of sinusoidal oscillation. are used because they are economical, compact, long-
To resolve the delay of sample reflections, the OCT lasting, and emit high quality beams that couples efficiently
system uses a light source that has a wide range of with an optical fiber. The resolution of clinical OCT is
wavelengths (low coherence). When the interference signals basically limited by the state of SLD technology. Early
are added together over the range of wavelengths, the retinal OCT systems typically employ SLDs emitting
interferometric oscillation fades as the delay mismatch around an 820 nm center wavelength over a bandwidth of
between the reference and sample reflections increases 20 nm full-width half-maximum (FWHM) (Swanson et al.,
(Fig. 6). To understand why this occurs, look carefully 1993; Hee et al., 1995a). This limits the axial resolution to
at the phase relationship between the wavelength compo- roughly 15 mm FWHM in air and 11 mm in tissue. The
nents shown schematically on the left panel of Fig. 6. When Stratus OCT System (Carl Zeiss Meditec, Inc., Dublin, CA,
the delay mismatch is near zero (center of waveforms), the USA) has 9–10 mm axial resolution in tissue.
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 329

3. OCT in clinical setting

3.1. Retina (and retinal diseases)

Definitive inclusion of OCT in the diagnostic arsenal of


the retina specialist occurred with the availability of third
generation devices. Indeed, OCT has turned out to be the
ancillary exam of choice to assist the diagnosis of several
chorioretinal diseases (Haouchine et al., 2004; Jorge et al.,
2004; Browning et al., 2004; Johnson, 2005; Niwa et al.,
2005; Sandhu and Talks, 2005; Catier et al., 2005; Costa Fig. 7. Quantitative assessment by third generation optical coherence
et al., 2005; Gaucher et al., 2005; Montero and Ruiz- tomography of morphologic retinal changes induced by treatments in a
Moreno, 2005; Meirelles et al., 2005). Additionally, the role patient with diffuse diabetic macular edema. (A) At baseline, retina is
diffusely thickened due to intra- and sub-retinal fluid. (B) Four weeks after
of OCT to monitor morphologic retinal changes over time high-density grid laser photocoagulation, favorable macular remodeling is
has turned out to be apparent, particularly given the latest disclosed by OCT. At the right side of the figure, quantitative analysis
concept of ‘‘disease modulation’’ associated with new (micrometer scale) of the changes in the average retinal thickness from
treatment modalities and the contemporary ‘‘pharmacolo- baseline in nine different macular sub-fields is disclosed.
gical’’ era of alternative management of chorioretinal
diseases (Costa et al., 2002, 2003a–c; Campochiaro et al.,
2004; Michels et al., 2005; Eter and Spaide, 2005; Gross, scholarship purposes, we have divided this section into
2005; Salinas-Alaman et al., 2005; Hillenkamp et al., 2005; 3 parts: (a) image acquisition; (b) qualitative analysis; and
Sahni et al., 2005; Cardillo et al., 2005; Bonini-Filho et al., (c) quantitative analysis.
2005).
Two representative examples of drugs fitting in such 3.1.1. Image acquisition
modern treatment concept might be triamcinolone and Part of the artifacts generated during OCT data
pegaptanib. It has been demonstrated that disease control processing may be attenuated given that optimal image
using intravitreal drug injections may be preferable to acquisition has been performed. For such, the examiner
photothermal destruction of pathological tissues in some should be aware of as well as should follow all guidelines
instances. Borne out from several studies over the past presented in the device’s user manual. In this segment,
decade, this premise has been to a great extent based on the some guidelines are emphasized and additional tips to
results of OCT evaluation, which has provided incon- minimize bias generated by the examiner are presented.
testable evidence of favorable macular remodeling after
treatment. While the initial reports about the alternative 3.1.1.1. Scanning centralization. When acquiring macular
use of intravitreal injection of triamcinolone for the scans for retinal thickness measurements (macular thickness
treatment of macular edema of a variety of causes may map and fast macular thickness map scan acquisition
represent the launch of the concept of disease modulation protocols), it is imperative that the examiner centers all 6
for the management of chorioretinal diseases (Jonas and scans at the fovea. This is relatively easy for cooperative
Sofker, 2001; Antcliff et al., 2001; Martidis et al., 2001, patients with relative good vision. However, the profile of
2002; Jonas et al., 2002; Greenberg et al., 2002), the recent patients with chorioretinal diseases submitted to OCT
FDA approval of pegaptanib (Macugens, Pfizer) for the evaluation at our institution is quite different. In general,
treatment of neovascular age-related macular degeneration patients are aged 50 or more, and have poor fixation due to
serves as the ultimate frontier that a new treatment era has impaired visual acuity. Additionally, the examiner fre-
definitively began (Gragoudas et al., 2004; Doggrell, 2005). quently faces a situation in which precise identification of
Presently, several drugs and alternative laser treatments the foveal center is problematic because of distortions
for the management of macular diseases are under of the macular architecture associated with the disease.
evaluation in randomized clinical trials, and there is Therefore, the examiner not only must be capable to
growing common sense that OCT may be preferable to identify the location of the fovea in abnormal maculas but
angiographic studies to monitor the retinal response in also has to be extremely serene to correctly place all 6 scans
such setting, particularly because of the theoretical at the presumed foveal center. While the former will be sole
possibility of quantitative analysis of the induced changes dependent on examiner expertize, the latter task may be
in a micrometer scale (Fig. 7). However, as commonly seen facilitated by switching on the center line tool of the OCT3
in other techniques of retinal imaging, one must bear in software. Right-click anywhere in the scan acquisition
mind that artifacts are expected to occur in the several steps window and a vertical line appears in the center of the scan
encompassing an OCT evaluation. Understanding of the image. This feature facilitates macular scans centralization
basic principles of image acquisition and data processing and shortens the acquisition time of optimal scans for
as well as recognition of OCT limitations may help macular maps. To deactivate the center line, right-click
us to use this imaging technique with cleverness. For another time in the scan acquisition window (Fig. 8).
ARTICLE IN PRESS
330 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

3.1.1.2. Data verification and validation. Immediately at thickness/volume (OU) analyze protocols, the software
the end of one scanning session for macular thickness maps automatically calculates the average (7SD) retinal thick-
using either standard (512 A-scans/image) or fast (128 ness at the fovea, named ‘‘foveal thickness’’ or ‘‘foveal
A-scans/image) acquisition protocols, some actions should height’’ (Fig. 10). The more central A-scan of each one of
be performed by the examiner prior to data processing to the 6 B-scans acquired is used to calculate foveal height.
generate the macular maps. Initially, the examiner should Since all 6 scans are to be centered at the same point
verify possible artifacts in the delineation of the retinal (fovea), in theory, in a perfect scan acquisition session, the
boundaries. For such, each image (B-scan) should be central A-scan should be the same for all 6 B-scans
processed in separate using the retinal thickness (single eye) (intersecting point). Therefore, in this hypothetical situa-
analysis protocol and accuracy of automatic delineation tion, the SD of the average foveal height is to be equal to
confirmed. If delineation errors were verified in any of the 6 zero. Depending on the macular status, SD values higher
B-scans, a new complete scanning session should be than 30 mm are highly suggestive that at least one of the 6
performed (Fig. 9). Once adequate images have been scans is not correctly centered at the fovea, and a new
acquired, the examiner should then verify centralization of complete scan acquisition session should be performed.
the 6 scans (in respect to the foveal center). By means of
data processing using the retinal map (single eye) or retinal 3.1.1.3. Manual raster scanning. At the end of a scanning
session for macular map, it is highly advisable to the
examiner to perform a manual raster scanning throughout
the macular area to minimize the chance of missing
morphological details in the adjacencies of the macular

Fig. 10. Checking of scans displacement in respect to the foveal center


(after verification of the automatic delineation of retinal boundaries
[Fig. 9]). (A) By using the retinal map (single eye) or retinal thickness/
volume (OU) analyze protocols, the software automatically calculates the
average (7SD) retinal thickness at the fovea, named ‘‘foveal thickness’’,
Fig. 8. (A) During fundus scanning, right-click anywhere in the scan ‘‘center’’ or ‘‘foveal height’’. In spite of delineation flaws observed in first
acquisition window (arrowhead), and a vertical line appears in the center scanning (Scan 5—Fig. 9(A)), which induced a !10% difference in
of the scan image (B). This feature facilitates macular scans centralization average RT in two subfields (between red lines) (B), note that the SD of
(C and D), and shortens the acquisition time of adequate scans for the foveal thickness was o30 mm in both scanning acquisition sessions
macular maps. (A and B), indicating good centralization of the 6 scans in both settings.

Fig. 9. Verifying in separate the automatic delineation of the retinal boundaries in each image (B-scan) to be used to macular maps. (A) At the end of one
scanning session for macular thickness maps, analysis of the automatic retinal boundaries delineation using the retinal thickness (single eye) analysis
protocol revealed one major delineation error in scan 5 (red dashed line/asterisk). (B) A new complete scanning session has solved the issue (red dashed
line/asterisk).
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 331

Fig. 11. Manual raster scanning. (A) Perifoveal detachment of the


posterior hyaloid and presumed foveal split were seen (impending macular
hole); detailed raster scanning revealed that in fact the pseudocyst had
already extended posteriorly (A0 ), and there was an eccentric opening of its
roof (arrowhead) (stage 2 macular hole). (B) Incomplete vitreofoveal
separation and a tomographic appearance that gives the impression that a
lamellar hole is to be formed; however, detailed scanning demonstrates an
already established full-thickness macular hole (B0 ).

center. This maneuver enables an enhanced ‘‘view’’ of the


macular status. It is not rare to find new findings such as
abnormal vitreoretinal adhesions. The use of 8 mm in Fig. 12. Review software tool. After activating the scanning mode (A), the
length scans for manual raster scanning may also facilitate examiner may freeze the image at any desired time (B). By pressing the
the detection of vitreomacular adhesions (Fig. 11). review button (C), the examiner is allowed to retrieve all tomograms
acquired in that session (scan review window). If an optimal tomogram
exists amongst recovered images, it can be selected (D) and saved by the
3.1.1.4. Scan review software tool. Occasionally during examiner (E).
scanning of ocular fundus, the examiner misses the exact
timing to freeze the desired image or some computer delay
occurs after the freeze button has been pressed, and the
subsequent tomogram is now frozen in the scan acquisition
window computer screen. By pressing the review button,
the examiner is allowed to retrieve all tomograms acquired
in that session (scan review window). If an optimal
tomogram exists amongst recovered images, it can be
selected and saved by the examiner (Fig. 12).

3.1.1.5. Bimanual technique. A simple change proposed


by one of the authors (D.C.) in the disposition of the
mouse within the device’s table may also facilitate
immensely OCT evaluation in patients with poor colla-
boration. By changing the mouse position to the left side of
the keyboard, the examiner will be able to control
simultaneously the alignment of the scanner unit (patient Fig. 13. Alternative mouse disposition within the table of third generation
module) with the right hand (joystick), and all the screen OCT. The examiner may keep one of freeze (with or without flash) buttons
settings (such as Z-offset, polarization, and fixation LED) of the scan acquisition window pressed during scanning to speed up image
with the left hand (mouse). In addition, to shorten acquirement.
acquisition time, the examiner may keep the freeze (with
or without flash) button of the scan acquisition window
pressed during scanning; when an optimal tomogram is while in standard mode each tomogram consists of 512
displayed in the screen, the image is rapidly frozen by just A-scans. In both modes, a total of 6 tomograms oriented at
releasing the button (Fig. 13). A short period of adaptation 301 intervals are acquired. Topographic macular maps are
may be needed until the examiner get used to this ultimately derived from each individual A-scan per OCT
alternative mouse disposition. study (768 A-scans in fast mode; 3072 A-scans in standard
mode) (Fig. 14). The equivalence of retinal maps generated
3.1.1.6. Additional considerations. The third generation by standard and fast modes in clinical scenarios other
of OCT instruments offers basically two modes of image than diabetic macular edema remains to be determined.
acquisition for macular thickness measurements. In the fast Additionally one must bear in mind that, although
mode, each tomogram (B-scan) consists of 128 A-scans quite convenient for the patient and presumably less
ARTICLE IN PRESS
332 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

time-consuming for the examiner, the fast mode acquires ment of one or more of the tomograms in relation to the
all six tomograms in sequence at one single scanning foveal center (Fig. 8(D)). Since macular maps are generally
session, increasing the chance for an undesirable misalign- used to evaluate possible changes in the macular status
over time, it is highly advisable to verify the capability of
the examiner to generate reproducible maps (Fig. 15).

3.1.2. Qualitative analysis


Third generation OCT is an established strategy that
improves first generations OCT technology and achieves
extraordinary in vivo visualization of retinal features and
disease. OCT3 exceeds first generations OCT imaging by
obtaining superior axial-image resolution and higher pixel-
density images, and therefore offers better recognition of
the intraretinal layers. Improved acquisition speed (400
A-scans/s) allows, within 1–2 s, the visualization of the
retinal morphology approaching a level of structural
differentiation obtainable only with histopathology (Fig. 16).

3.1.2.1. Defining inner and outer HRL. Ever since the


initial images obtained using first generation OCT became
available, several studies have focused on the precise
interpretation of retinal reflective signals and their correla-
tion with retinal morphology. From the first generation
tomograms of normal retinas it was demonstrated that a
clear highly reflective layer (HRL) exists at the outer aspect
of the neuro-sensory retina. Based on clinical and
tomographical correlation studies, it had been suggested
Fig. 14. Fast and standard macular thickness maps. (A) In the fast mode,
each tomogram (B-scan) consists of 128 A-scans while in standard mode that the outer most reflective layer might correspond
(B) each tomogram consists of 512 A-scans. In both modes, a total of 6 to a complex formed by the RPE and choriocapillaris,
tomograms oriented at 301 intervals are acquired. and such layer was then coined ‘‘RPE-choriocapillaris

Fig. 15. (A–D) Four scanning sessions performed in 5-min intervals using the fast macular thickness map acquisition protocol in a patient with central
serous chorioretinopathy. (A0 –D0 ) Data processing revealed optimal centralization of the scans in all four scanning sessions and good correspondence of
topographic macular maps.
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 333

Fig. 17. Third generation optical coherence tomography appearance of


one normal macula. (A) Two, well-defined, linear, highly reflective layers
(HRL) are seen at the outer aspect of the neural retina. (B) The inner HRL
(corresponding to the photoreceptors IS/OS junction) is thinner than the
outer, and is characterized by mild forward bow-shaped configuration at
the foveal center. The outer HRL may correspond to a hyper-reflective
Fig. 16. Differences in image quality related to A-scan density (normal complex formed by RPE [and choriocapillaris].
macula). Note that image ‘‘quality’’ is increases as the number of A-scans/
image multiplies (128 [A], 256 [B], and 512 [C]).

3.1.2.2. Outer retina. Particular analysis of the inner


HRL tomographic appearance has provided new insights
hyper-reflective complex’’. In 1998, Huang et al. first for better comprehension of diseases affecting the macula.
proposed an alternative interpretation of such hyper- The integrity of the inner HRL has been correlated to some
reflective complex (Huang et al., 1998). Based on an extent with visual acuity. Eyes with vitreomacular traction
OCT study with histological and pathological correlation syndrome or macular edema and extraordinary good
in normal and rd chickens, they suggested that the outer visual acuity levels represent a good example of the
most hyper-reflective complex might correspond to photo- value of such particular analysis (Fig. 18A,B). In unilateral
receptors’ inner and outer segments, RPE, and anterior resolved central serous chorioretinopathy as well as in eyes
choroidal pigmentation (they termed it outer retina–chor- presenting visual acuity deterioration due to photic
oid complex). At that time, they have also demonstrated maculopathy the same rationale may be used (Eandi
that in the most severely affected rd chickens, which lacked et al., 2005; Jorge et al., 2004). Third generation OCT
photoreceptor layer, the total thickness of the outer most evaluation has demonstrated abnormal reflectivity at the
hyper-reflective complex is reduced, suggesting missing the level of the outer foveal retina, such as intense fragmenta-
highly reflective peak derived from the photoreceptor tion or complete interruption of the inner HRL. It has been
IS/OS (Huang et al., 1998). Supportive data to Huang’s suggested that visual acuity may be more severely distorted
hypothesis occurred as soon as third generation and in eyes presenting full-thickness involvement of the
ultrahigh-resolution OCT clinical studies became available photoreceptors reflective layer (Jorge et al., 2004). Analysis
(Montero et al., 2003; Drexler et al., 2003; Jorge et al., of the inner HRL has also demonstrated usefulness in the
2004; Costa et al., 2004). The single outer hyper-reflective evaluation and better understanding of the outcomes of
layer seen by first generations OCT was then visualized as operated macular holes. It has been demonstrated that in
two parallel, highly reflective (red/white) layers separated spite of anatomical success and recovery of the macular
from each other by one thin layer of moderate reflectivity shape, the postoperative visual acuity and improvement of
(green/yellow) at the level of the posterior boundary of the visual acuity were not directly related to the morphological
retina (Fig. 17). The inner HRL, which most likely results. Outer retinal features appear to be more important
corresponds to the junction between the inner and outer to determine postoperative visual function as irregularities
segments of the photoreceptors, assumes a forward bow- at the level of the inner HRL after macular hole surgery
shaped configuration in the center of the macula consistent may prevent visual acuity improvement (Uemoto et al.,
with the well-known increase in length of the outer 2002; Kitaya et al., 2004; Villate et al., 2005) (Fig. 18C,D).
segments of the cones in such region. The outer HRL, New and unexpected findings have also been demon-
which appeared approximately two times thicker than the strated by studies using OCT in several hereditary
inner HRL, have been described to be equivalent to the retinopathies (Aleman et al., 2002; Jacobson et al., 1998,
RPE-choriocapillaris hyper-reflective complex seen on first 2000, 2003, 2004, 2005; Milam et al., 2003; Pianta et al.,
generations OCT. However, the precise interpretation of 2003). These scientific results include: localization of the
the outer HRL remains to be established. Is choriocapil- missing photoreceptor component in retinitis pigmentosa
laris signaling contributing to its formation? Such question (Jacobson et al., 1998, 2000); prediction of sub-retinal
is to be answered as soon as additional ultrahigh-resolution and sub-RPE deposits in cone-rod dystrophy and in Best
OCT data becomes available. macular dystrophy (Aleman et al., 2002; Pianta et al.,
ARTICLE IN PRESS
334 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

Fig. 18. Tomographic appearance of outer foveal retina and visual acuity (VA). (A) Typical inner HRL appearance in normal macula (VA ¼ 20/20). (B)
Macular edema in branch retinal vein occlusion and VA ¼ 20/25; note at the fovea, the inner HRL is well preserved. Intense fragmentation of the inner
HRL in one patient with outer macular hole and VA ¼ 20/50$1 (C), and in patient with photic maculopathy and VA ¼ 20/60 (D).

Fig. 20. Third generation optical coherence tomography evaluation of the


Fig. 19. Vitreomacular status and diabetic macular edema. From
left eye of a 57-year-old woman with idiopathic full-thickness macular
perifoveal posterior vitreous detachment (A and B) to vitreofoveal
hole in the right eye. (A) At presentation a full-thickness macular hole was
traction (and increased foveal high) (C), and finally to spontaneous
seen in patient’s right eye, and OCT3 scan of the left eye revealed complete
complete vitreofoveal separation (and favorable macular remodeling).
detachment of the posterior hyaloid in the macular region with hyper-
reflective signals in the plane of the detached posterior hyaloid over the
foveal region (arrow). Despite minimal irregularities of the inner foveal
2003); unexpected retinal lamination abnormalities (Jacob- contour, macular architecture was relatively well preserved at presenta-
tion, and visual acuity was 20/20. (B) Five months later, visual acuity
son et al., 2003, 2004); and, definition of the relationship decreased to 20/25$2 and an epiretinal membrane (grade 1) was seen in her
between outer nuclear layer thickness and visual function left eye associated with the development of a full-thickness macular hole.
(Jacobson et al., 2005). Oblique scans revealed complete interruption of the foveal retina of
approximately 99 mm associated with perifoveal intraretinal fluid accu-
3.1.2.3. Vitreoretinal interface. Unparalleled characteri- mulation and attenuation of the inner HRL suggesting edema of the outer
perifoveal neural retina. Tornambe’s ‘‘retinal hydration’’ theory may
zation of the vitreoretinal interface has been possible with explain macular hole formation in eyes presenting complete vitreofoveal
the advent of third generation OCT. Studies focused on the separation and associated disruption of the inner fovea.
analysis of the vitreomacular interface have led to new
insights about our understanding and management of
several maculopathies, such as idiopathic macular hole and Presently, there are more than 80 published OCT studies
diabetic macular edema. For example, a high prevalence of about macular holes. Old concepts have been revisited and
perifoveal posterior vitreous detachment with incomplete new findings and conjectures, such as the influence of
vitreofoveal separation has been demonstrated in diabetic oblique vitreofoveal tracional forces, the demonstration of
patients with macular edema, suggesting a possible ‘‘intrafoveal split’’ as well as the retinal ‘‘hydration’’
additional role of the vitreomacular interface status in the theory, have been possible due to enhanced appreciation
pathogenesis of diabetic macular edema (Gaucher et al., of the vitreomacular interface status and macular morpho-
2005). Anecdotal reports of favorable macular remodeling logic features in eyes with macular holes (Hee et al., 1995b;
in eyes with diabetic macular edema after spontaneous Gaudric et al., 1999; Haouchine et al., 2001; Tornambe,
vitreofoveal separation, as well as in the early post- 2003) (Fig. 20).
Nd:YAG laser capsulotomy period, may provide addi-
tional supportive data of the possible influence of the 3.1.2.4. Additional considerations. An extraordinary qua-
vitreomacular interface status in such scenario (Watanabe litative analysis of the macular morphologic features has
et al., 2000; Yamaguchi et al., 2003) (Fig. 19). been enabled by third generation OCT technology. However,
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 335

et al., 1995; Puliafito et al., 1996; Zeimer et al., 1996;


Podoleanu et al., 1998; Bartsch and Freeman, 1994).
Indubitably, one of the most attractive features of third
generation OCT is the theoretical possibility of attaining
reproducible and accurate measurements of ocular fundus
tissues. In clinical settings, it has significant potential both
as a diagnostic tool and particularly as a way to monitor
objectively subtle retinal changes induced by therapeutic
interventions. Understanding of the basic principles of
automatic retinal thickness measurements by OCT3 as well
as recognition of the software limitations are crucial steps
to facilitating users to extract tomographic data as ‘‘real’’
as possible.
Fig. 21. Young patient with pathologic myopia complaining of distortion
and acute loss of vision. (A) At presentation OCT evaluation suggested the 3.1.3.1. Fundamentals of OCT automatic retinal thickness
presence of a ‘‘true’’ abnormal hyper-reflective formation (arrowheads) measurement. At first moment, it is very difficult to
such as choroidal neovascularization. (B) Six weeks latter, the macula
understand why the OCT3 software erroneously delineates
contour was practically normal. The patient had serohemorrhagic
complications related to one small lacquer crack located juxtafoveally, the retinal boundaries in an optimal tomogram as such
without associated neovascularization. exemplified in Fig. 22. Initially, one should bear in mind
that automatic retinal thickness measurements are gener-
ated in essence by means of a ‘‘mathematical calculation’’
one should bear in mind that optimal interpretation of (algorithm). The algorithm identifies differences in the
OCT3 findings in clinical setting is highly dependent on image reflectance patterns in each A-scan (up to 512
adequate clinical correlation and, whenever suitable, A-scans in OCT 3) that compose one tomogram (B-scan),
clinical and angiographical correlation may be preferable. and assumes that the distance between two relatively high
Third generation OCT evaluation in eyes presenting early reflective structures represents the retinal thickness at that
stages (grade 0 and 1) epiretinal membranes and in patients A-scan. As a result, the OCT software locates the presumed
with serohemorrhagic macular complications due to inner retina boundary at the vitreoretinal interface (first
choroidal neovascularization are good examples of entities high reflective structure) and the presumed outer retina
in which interpretation bias may occur in the absence of boundary at the retinal pigment epithelial-photoreceptor
clinical/clinical and angiographical correlation. The occur- outer segment interface (second high reflective structure).
rence of ‘‘normal’’ cross sectional images (with no apparent The algorithm also compares the ‘‘shape’’ of one A-scan to
abnormal finding) is not rare in patients with early adjacent A-scans, once great differences in shape are not
stage epiretinal membranes. In the same view, one should
be caution in the interpretation of tomographic findings
in exudative maculopathies, given that outer retina edema
as well as blood may eventually simulate the tomo-
graphic appearance of one ‘‘virtually’’ abnormal formation
(Fig. 21).
Finally, an important aspect of OCT data that has been
available since the first generation technology, but is very
rarely studied, is the intensity changes resulting from intra-
retinal abnormalities in backscatter. In terms of future
work, this rich data source, which is available in each scan
performed nowadays, requires careful attention in the
future. Quantitative analysis of intra-retinal OCT intensity
changes have been demonstrated experimentally by Cide-
ciyan et al. in dogs (Cideciyan et al., 2005). Also relevant to
this subject, is one of the very few clinical examples of
changes in intra-retinal light scattering, reported by Lerche
et al. (2001), including 20 patients with retinal venous
occlusive disease.

3.1.3. Quantitative analysis


Fig. 22. Automatic retinal thickness measurement. Although quite good
Cross sectional retinal imaging provides a unique scans were acquired (blue-dashed box), the OCT 3 software algorithm was
opportunity to quantify the overall thickness of the retina not able to delineate correctly the outer retinal boundaries (red dashed
in vivo (Huang et al., 1991; Hee et al., 1995a; Schuman boxes, arrows).
ARTICLE IN PRESS
336 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

expected to occur in side-by-side A-scans. The software The use of the OCT3 automated retinal thickness
then places a line on the inner vitreoretinal interface and measurement tool (software versions 1.0, 2.0, and 3.0)
another on the retinal pigment epithelium (RPE)-outer has generating erroneous values due to incorrect inter-
retinal interface and determines retinal thickness as the pretation of the inner HRL as the outer neural retina
distance between these lines at each measurement point boundary (Costa et al., 2004; Pons and Garcia-Valenzuela,
along the scan’s x-axis. Therefore, even in ‘‘fine-looking’’ 2005) (Fig. 23). Measurements using the automated tool of
B-scans, errors in retinal boundaries delineation may the OCT3 software (version 3.0) in comparison to manual
occur. caliper-assisted technique, in which the outer HRL was
interpreted as the outer boundary, demonstrated that a
3.1.3.2. Software delineation of outer neuro-sensory retinal significant difference existed in the generated values for
boundary. It was believed in the past that scans of normal retinal thickness at specific macular regions in healthy
eyes did not have inner and outer retina boundaries subjects caused by such misalignment. Manual caliper-
misidentification artifacts and only had artifact related to assisted retinal thickness measurements at specific macular
examiner error. Therefore, under optimal scan acquisition, regions differed from those automatically generated by
measurement of the retinal thickness is expected to be 9.9% from up to 38% (Costa et al., 2004).
perfect, basically depending on the ability of the OCT 3
software to recognize both interfaces at each A-scan that 3.1.3.3. Topographic macular maps. Built-in software of
composes the tomogram. However, it has been recently third generation OCT performs measurements of macular
demonstrated that built-in OCT3 software has encounter- thickness using 6 intersecting 6-mm-long OCT images
ing severe problems in recognizing the outer boundaries of oriented in a radial pattern centered on the fovea (Hee
the neurosensory retina in optimal OCT3 scans (Costa et al., 1998). Six images of 512 A-scans (transverse pixels)
et al., 2004). Of particular concern was the finding that each can be acquired in approximately 8–10 s using
such recognition was invariably incorrect in normal eyes macular thickness map or radial lines acquisition proto-
(Costa et al., 2004). As explained above, the so-called cols, or 6 images of 128 A-scans each can be acquired
‘‘RPE-choriocapillaris reflective complex’’ seen in first in approximately 2 s using the fast macular thickness
generations OCT now is disclosed as two well-defined map. The radial scanning protocol was designed to
HRL at the level of the outer retina in the macular region concentrate measurements in the central fovea, where high
of healthy subjects in third generation tomograms; the sampling density is most important. The 6 OCT images
inner HRL corresponding to the junction of the inner and are segmented to detect the retinal thickness, which is
outer segments of the photoreceptors while the outer one displayed as a false-color topographic map divided into 9
most likely corresponding to the retinal pigment epithelium regions: one central (central macular thickness [CMT])
or a reflective complex formed by RPE and choriocapil- plus 8 Early Treatment Diabetic Retinopathy Study-
laris. A similar tomographic appearance has been also fashion sub-fields, and the average thickness value for
evidenced in non-affected macular regions of patients with each region is displayed in separate. Because the radial
selected eye diseases (idiopathic macular hole, central pattern of 6 OCT images samples the macular thickness
serous chorioretinopathy, and macular edema), whereas along clock hours, the retinal thickness in the wedges
affected regions generally demonstrated a single-layer high between each image is interpolated. Therefore, this imaging
reflective appearance with disappearance of the inner HRL protocol may miss focal peculiarities in a span of o1 clock
(Costa et al., 2004). hour, or 301. In addition, misidentification of retinal

Fig. 23. Automated retinal thickness measurement was obtained from corresponding A-scan at the fovea (left). Manual caliper-assisted measurement of
retinal thickness at the fovea using the automated delineation for the inner retinal boundary by the software as one point (inner caliper cross) and
positioning the outer caliper cross just above the outer HRL demonstrated a difference of 51 mm.
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 337

boundaries in the 6 B-scans that are used to generate 3.1.3.5. Additional considerations. Summing up, auto-
retinal maps may interfere significantly in the average matic OCT macular thickness is calculated by computer
retinal thickness value displayed for each sub-field and image-processing algorithms with several notable flaws.
CMT, as well as in values estimations for macular volume. The most important flaw is that the software does not truly
Of concern, is the fact that CMT values have been used as measure the anatomic macular thickness due to its inability
the main tomographic outcome to monitor retinal changes to reliably discriminate the junction between the inner and
by therapeutic interventions. Because of particular mor- outer segments of the photoreceptors (inner HRL) from
phologic foveal features, retinal thickness measurements at the RPE, because both produce a high backscattering
this region are more sensible to the OCT3 software signal. Therefore, macular thickness is actually measured
measurement flaws (Costa et al., 2004). Recent data from using the hyper-reflective layer corresponding to the
next generation OCT prototypes have been supportive to junction between the photoreceptor inner and outer
our concerns about the automatic delineation of the ‘‘true’’ segment (inner HRL) as the outer retinal boundary,
retinal boundaries by OCT software. Comparison of effectively truncating the outer segments in most subjects.
retinal thickness maps obtained using OCT3 and 3D Obviously, this is not an incapacitating limitation of the
OCT data from one high speed ultrahigh-resolution OCT methodology, but this issue should be stressed because one
prototype, which enables differentiation of the junction may assume that these measurements are more anatomi-
between the inner and outer segments of the photorecep- cally meaningful than they truly are. By addressing these
tors (inner HRL) as a distinct feature from the RPE (outer issues, we intent to promote a better comprehension of the
HRL), showed a 8–9% difference in retinal thickness actual limitations of the OCT3 software, and to assure
values for the 8 macular map sub-fields and up to 16% researchers’ as well as manufacture’s best efforts to
for CMT, when the thickness map that measures the overcome them as fast as possible. Recognition of the
retinal thickness as the distance from the inner interface of limitations of any particular device is the basic principle for
the hyporeflective band corresponding to the RPE to its use with cleverness.
the vitreal retinal interface was disclosed (Wojtkowski
et al., 2005). 3.2. RNFL and glaucoma

3.1.3.4. Current third generation OCT software versions. Interest in retinal nerve fiber layer (RNFL) analysis in
As recently clarified (Hee, 2005), the original macular glaucoma has been recorded as early as 1972, when Hoyt
thickness algorithms (upon which the current OCT3 and Newman initially reported RNFL atrophy in patients
algorithms are based) were designed to determine the with glaucoma (Hoyt and Newman, 1972), thus suggesting
inner and outer retinal boundaries in diabetic macular RNFL thinning as a possible sensitive indicator of
edema, a condition leading to retinal thickening in glaucomatous damage. When glaucomatous damage be-
which intraretinal fluid accumulation and hard exudates gins, ganglion cell degeneration can occur in either diffuse
occur but the retinal pigment epithelium and inner limiting or focal forms. Diffuse atrophy of the nerve fibers is more
membrane remain intact (Hee et al., 1998); in such difficult to assess especially in early stages of the disease,
scenario, the inner HRL is frequently attenuated or but focal axonal degeneration causes characteristic changes
absent due to intraretinal edema, and the OCT3 software in the appearance of the RNFL and is more easily
correctly delineates the outer HRL as the outer boundary recognized. Dark slits or grooves appear among the
(Costa et al., 2004; Costa, 2005). Additionally, we should arcuate bundles approaching the optic disc superiorly
remember that the inner HRL was not so evident in first or inferiorly (Hoyt et al., 1973; Sommer et al., 1991).
generation OCT. The enhanced resolution offered by Abnormal RNFL appearance may be sufficient evidence to
OCT3 compared with first generations OCT provides initiate glaucoma therapy since at least 25–35% retinal
images displaying more complex internal features that, ganglion cell loss is lost before detection of abnormalities in
paradoxically, require more refined boundary detection automated visual field testing (Kerrigan-Baumrind et al.,
algorithms. 2000). Therefore, evaluation of RNFL has gained growing
Improved versions of the OCT3 software are constantly interest amongst glaucoma specialists in the past few
under development. The new software version (4.0) decades.
including innovative features to minimize problems during OCT is a relatively new technology that provides high-
image acquisition has been just recently released. Norma- resolution cross-sectional imaging of the RNFL. A built-in
tive reference values for retinal thickness measurements of algorithm automatically calculates the RNFL thickness
the macula have also been included in such version. when interpreting data acquired using the several RNFL
However, one must bear in mind that misidentification of scan acquisition protocols.
the outer retinal boundary is still occurring in version 4.0,
and, of particular concern, that normative reference values 3.2.1. RNFL thickness protocols
have been established according to OCT3 data in which the There are innumerous ways to study the RNFL with
outer retinal boundary was considered to be the inner OCT, but a fixed diameter circular scan around the optic
HRL, and not the outer HRL. disc has been used as standard for most of the investigators.
ARTICLE IN PRESS
338 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

The most used glaucoma scan acquisition protocols are the variability of the optic disc size by multiplying the optic
RNFL thickness (3.4) and the fast RNFL thickness (3.4). disc radius by a factor to determine the final diameter
The former enables the acquirement of three circular scans of measurement scanning circle (Fig. 25), thus RNFL
of 3.46 mm diameter around the optic disc, which can then measurements taking place further away from disc margin
be averaged. The fast protocol acquires three circle images in larger discs. Most investigators prefer fixed circles rather
of lower resolution sequentially in one single-scan acquisi- than proportional circles to analyze the RNFL thickness.
tion session (Fig. 24). While the RNFL thickness acquisi- RNFL thickness measurements are displayed by quadrant,
tion protocol is composed of one circle with 512 clock hour and overall mean.
A-scans/image and requires 1.28 s of scanning time, the Compared with first generations OCT systems, the third
fast RNFL thickness comprises three circles of 256 A-scans/ generation OCT allows high-density scanning protocols,
image and requires 1.92 s. which produces images with high transverse pixel density,
Proportional circle and the RNFL thickness (2.27 " disc) thus resulting in better image quality. However, increase in
acquisition protocols enable one to account for the the image acquisition time may lead to measurement
artifacts due to eye motion. In a study comparing fast
RNFL scan acquisition protocol (256 A-scans/image) with
RNFL measurement using images of 512 A-scans, Leung et
al. (2004) concluded that the latter (high-density protocol)
has provided better sensitivity and stronger correlation
with visual function.

3.2.2. Reproducibility
Schuman et al. (1996) have suggested the use of 3.4-mm
diameter circle as the standard for RNFL OCT evaluation
after a reproducibility study involving 11 normal volun-
teers and 10 glaucomatous patients. Each subject under-
went five repetitions of a series of scans on five separate
occasions. Each series consisted of three circular scans
around the optic nerve head (diameters of 2.9, 3.4 and
4.5 mm). Reproducibility was better in a given eye on a
given visit than from visit to visit. In addition, the internal
Fig. 24. (A) Fundus image and corresponding B-scan of the RNFL fixation was superior to external fixation regarding
thickness 3.4-mm (circle diameter) acquisition protocol in the regular
mode (512 A-scans/image). (B) Fundus image and corresponding B-scans
reproducibility. The 3.4-mm circle was then suggested for
of the fast RNFL thickness 3.4-mm (circle diameter) acquisition protocol. future studies because reproducibility was significantly
Three images of 256 A-scans each are captured ‘‘simultaneously’’ better at this circle diameter than at 2.9 mm. Additionally,
(consecutively, at same scanning session). the 3.4-mm circle allowed measurement of NFL in a

Fig. 25. Four different RNFL thickness circle acquisition protocols. (A) Fixed 3.4 mm diameter circle. (B–D) After determining the disc radius (!1 mm)
three additional protocols enable to tailor de measurement circle accordingly: (B) nerve head circle (measurement circle was chosen to be 400 mm after the
disc margin), (C) RNFL thickness (2.27 " disc), and (D) proportional circle (1.5 was chosen as the multiplication factor).
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 339

thicker area than 4.5 mm, what potentially permits a higher normal and 29 glaucomatous eyes using 3.4-mm diameter
sensitivity to subtle NFL defects. scans (Bowd et al., 2000). Average RNFL thickness in
Other reproducibility studies have demonstrated that temporal, superior, nasal, inferior quadrants and total
OCT RNFL measurements are reproducible for both average were obtained. Mean RNFL was significantly
normal and glaucomatous eyes (Blumenthal et al., 2000; thinner in ocular hypertensive eyes than in normal eyes,
Carpineto et al., 2003; Paunescu et al., 2004; Budenz et al., more specifically in the inferior and nasal quadrants.
2005). Blumenthal et al. (2000) evaluated the reproduci- RNFL was significantly thinner in glaucomatous eyes than
bility of OCT RNFL measurements in normal and in ocular hypertensive and normal eyes (total RNFL and
glaucomatous patients for a prospective instrument valida- all quadrants).
tion study. Only a modest contribution to variability was In a study published in 2001, Bowd et al. (2001)
found for session (1%), visit (5%), and operator (2%). compared the abilities of scanning laser polarimetry
Budenz et al. (2005) studying the reproducibility of (SLP), OCT, SWAP, and frequency-doubling technology
standard and fast RNLF thickness scans found that both perimetry (FDT) to discriminate between healthy eyes and
scan acquisition protocols yields reproducible and compar- those with early glaucoma, classified based on standard
able measurements in glaucoma as well as in healthy automated perimetry and optic disc appearance. In general,
individuals. The nasal quadrant showed more variations in areas under the receiver operating characteristics (ROC)
the measurements than other sectors. curve were largest for OCT parameters, followed by FDT,
According to Paunescu et al. (2004), the best reprodu- SLP, and SWAP, regardless of the definition of glaucoma
cibility for RNFL measurements was found for dilated eyes used. The most sensitive OCT and FDT parameters tended
and scanning rate of 256 A-scans per image acquired in the to be more sensitive than the most sensitive SWAP and
fast mode when compared to high-density scanning (512 SLP parameters at the specificities investigated, regardless
A-scans per image) acquired in the regular acquisition of diagnostic criteria. Medeiros et al. (2004) using the
mode. Although increased density may cause a reproduci- current commercial available versions of SLP (GDx-VCC),
bility problem, higher-density scanning protocols have third generation OCT, and Heidelberg retina tomograph
provided better diagnostic sensitivity in glaucoma detection (HRT II), have demonstrated similar sensitivity results
and a stronger correlation with visual function according among the best parameters of each equipment.
to Leung et al. (2004). OCT RNFL thickness decreases with increasing RNFL
damage detected with red-free photography and visual field
3.2.3. Diagnostic capability and progression evaluation (Soliman et al., 2002). The global average OCT RNFL
According to Schuman et al. (1995), RNFL measure- thickness correlated significantly with the photographic
ments by OCT demonstrate a high degree of correlation total RNFL score. This study suggests the validity of OCT
with functional status, as measured by visual field measurements and its potential advantage for detection of
examination. Neither cupping of the optic disc nor neural early cases of glaucoma. Leung et al. (2005a) published a
rim area were as strongly associated with visual field loss as study evaluating the relationship between structure and
was RNFL thickness in that study. RNFL, especially in the function in glaucoma. In this study, the Advanced
inferior quadrant, was significantly thinner in glaucomatous Glaucoma Intervention Study and the Collaborative Initial
eyes than in normal eyes. Finally, it was found a decrease in Glaucoma Treatment Study scores as well as, the mean
RNFL thickness with aging, even when controlling for deviation in decibel an unlogged 1/Lambert were used as
factors associated with the diagnosis of glaucoma. measures of visual function. Better correlations were
In a retrospective observational case series study that demonstrated between OCT RNLF measurements and
included 29 glaucoma patients, RNFL thickness measured visual function than between GDx-VCC measurements
with OCT was topographically correlated with glaucoma- and visual function.
tous visual field defects measured with short-wavelength The RNFL thickness has also been measured in children
automated perimetry (SWAP) (Sanchez-Galeana et al., (Mrugacz and Bakunowicz-Lazarczyk, 2005; Hess et al.,
2004). In a paper by Pieroth et al. (1999), OCT-enabled 2005). In a study including 26 normal eyes and 26
focal defects detection with a sensitivity of 65% and a glaucomatous eyes, Mrugacz and Bakunowicz-Lazarczyk
specificity of 81%. OCT analysis of RNFL thickness in (2005) found that the mean RNFL thickness as well as the
eyes with focal defects showed good structural and inferior quadrant measurements was statistically thinner in
functional correlation with clinical parameters and allowed children with glaucoma than in healthy ones. In another
identification of focal defects in the RNFL in early stages study, Hess et al. (2005) showed that both macular and
of glaucoma (Fig. 26). By evaluating 64 eyes of glaucoma RNFL thickness were thinner in glaucomatous in compar-
or glaucoma-suspect individuals, Wollstein et al. (2005a) ison with healthy children.
have shown a greater likelihood of detection of glaucoma- OCT was used to measure macular and nerve fiber
tous progression by the OCT in comparison to standard layer thickness and to analyze their correlation with
automated perimetry. each other and with glaucoma status (Guedes et al.,
In another study, the mean RNFL thickness of 28 ocular 2003). Both macular and RNFL thickness as measured by
hypertensive eyes was compared with age-matched 30 OCT showed statistically significant correlations with
ARTICLE IN PRESS
340 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

Fig. 26. Fundus photography showing localized RNFL defects (arrows) in both eyes, which could also be identified by OCT using the RNFL thickness
average protocol (with normative data analysis). Glaucomatous damage was mild in the right eye, and moderate in the left eye.

glaucoma, although RNFL thickness showed a stronger macular symmetry testing (Bagga et al., 2005), are under
association than macular thickness. development.
Ishikawa et al. (2005) developed a software algorithm to More recently, investigators have been trying to increase
obtain automated segmentation measurements of retinal the sensitivity and specificity of the OCT by adding
layers in the macula. Four retinal layers were obtained and measurement information that this device provides rather
the algorithm was capable to discriminate between than analyzing isolated parameters (Medeiros et al., 2005;
glaucomatous and normal eyes. In three layers (macular Huang and Chen, 2005; Chen and Huang, 2005; Burgans-
nerve fiber layer, inner retinal complex, and outer retinal ky-Eliash et al., 2005). Medeiros et al. (2005) combining
complex), the measurements were statistically significantly selected optic nerve head and RNFL parameters obtained
thinner in glaucomatous eyes than in normal eyes. larger area under ROC curve than using single parameters.
In another study involving macular measurements, Huang and Chen (2005) compared several automated
Leung et al. (2005b) found macular thickness measure- classifiers to differentiate normal from glaucomatous eyes.
ments significantly reduced in glaucomatous patients. In this study, the Mahalanobis space showed better results
However, peripapillary RNFL thickness measurements than linear discriminant analysis and artificial neural
provided greater power to discriminate between normal, network. In another study by Chen and Huang (2005), 21
glaucoma-suspect and glaucoma eyes than macular parameters (optic nerve head and RNFL) were combined
measurements. Wollstein et al. (2005b) also found the to obtain a linear discriminant function. The use of linear
RNFL measurement to provide better discrimination discriminant analysis increased the discrimination power
between glaucomatous and normal individuals than to differentiate glaucomatous and healthy individuals
macular measurement. In another report, Wollstein et al. in that study. Five classifier methods to discriminate
(2004) show that the peripapillary RNFL measurements between glaucoma and healthy subjects were studied by
have higher sensitivity and specificity than macular Burgansky-Eliash et al. (2005). In this study, the classifier
measurements. New strategies to evaluate macular methods studied were linear discriminant analysis, support
thickness in relation to glaucoma detection, including vector machine, recursive partitioning and regression tree,
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 341

generalized linear model and generalized additive model. performed after positioning an aiming circle whose
The largest area under ROC curve was obtained using dimension can be adjusted to the optic disc size, and the
support vector machine, and the best discrimination actual scan radius is greater than the aiming circle radius
between advanced and early glaucoma was provided by by the designated R). For each option, variance compo-
the generalized additive model. nents and intraclass correlation coefficients were deter-
mined. The total variance increased with circle diameter
3.2.4. Normative database and the intersubject standard deviation showed a tendency
In spite of the good reproducibility and potential to to increase with radius in both groups. The RNFL
detect glaucomatous damage, the format of data presenta- thickness decreased with increasing circle radius. Multiple
tion by initial versions of third generation OCT software regression for intraclass correlation coefficient of RNFL
and absence of a normative reference restricted OCT thickness showed that intraclass correlation coefficient was
RNFL data interpretation and its use in clinical setting as a higher for normal eyes and for scan protocol 1.5R than for
routinely diagnostic tool. To address such particular issues, R ¼ 1:73 and 2.0R. The 1.5R option allowed RNFL
a normative RNFL database to analyze patient’s data has measurements in a thicker area than R ¼ 1:73 and 2.0R.
been made available within the last software package
(version 4.0) of third generation OCT systems. Never- 3.2.5. Future directions
theless, normative reference analysis software for glaucoma By the use of the 3.4-mm circle scan acquisition protocol,
applications is not fully developed, and there is a scarcity of Williams et al. (2002) have defined a parameter called NFL
age, refractive error, and mainly, race-specific normative (50), which is the RNFL thickness value at which there is a
data upon which to compare eyes. The normative database 50% likelihood of a visual field defect with either
is based on 3.4-mm diameter circular scan measurements. automated standard or FDT perimetries. RNFL layer
This might represent a potential source of error since a thickness analysis using this parameter demonstrated that
fixed diameter does not account for the distance between OCT might be clinically useful in identifying subjects who
the measurement circle and the optic disc margin. have visual field loss. However, the positive predictive
Considering the progressive decrease of the RNFL thick- value suggested that OCT might need higher resolution and
ness with increasing distance from disc margin (Varma better reproducibility to enhance its sensitivity and
et al, 1996), the disc size may interfere in RNFL thickness specificity for population screening.
measurements (Fig. 27). By using third generation OCT, Increase of the axial resolution may be needed to
Savini et al. (2005) have demonstrated that RNFL improve OCT efficacy in detecting and following glauco-
thickness increased significantly with an increase in optic matous loss, but it is quite likely that refinements in the
disc size, and this can be due to a shorter distance between actual software algorithm may be sufficient to increase the
the scan and optic disc margin. New strategies have to be specificity and sensitivity of this technology. According to
developed to better evaluate RNFL thickness. Jones et al. (2001), OCT measurements close to disc margin
These ideas are in accordance with the findings of underestimate RNFL thickness in approximately 37%. In
Carpineto et al. (2003). These authors studied glaucoma- a study performed by Skaf et al. (2005), the OCT algorithm
tous patients and a gender- and age-matched group of to determine RNFL thickness is incorrect close to disc
normal subjects with three different circle diameter nerve margin up to approximately 400 mm resulting in under-
head programs: R ¼ 1.73 mm (3.4-mm diameter circle) estimated RNFL measurements. This might happen
as well as 1.5R and 2.0R (optic nerve head scanning is because retinal nerve fibers have a different orientation

Fig. 27. Fixed 3.4-mm circle and large optic discs. The measurement is performed closer to disc margin and values for RNFL thickness tend to be
overestimated.
ARTICLE IN PRESS
342 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

close to disc margin (fibers curve to form the optic nerve) tomographic methods including CT and MRI. However,
and/or the software may not be prepared with proper the speeds of commercially available third generation OCT
landmarks for this region. and combined SLO-OCT OTI instruments are insufficient
to measure full sets of three-dimensional data having a
3.2.6. Additional considerations large number of pixels per image in vivo. The highest
OCT has demonstrated good reproducibility of measure- reported speeds for retinal OCT systems based on standard
ments and capability to detect early glaucomatous damage. OCT detection have been achieved by transverse scanning
A normative database has been incorporated to the last with an acousto-optic modulator generating a highly stable
version of the commercial available third generation OCT carrier frequency (Hitzenberger et al., 2003). This system
system. Nevertheless, while the 3.4-mm fixed circle around (Hitzenberger et al., 2003) can acquire cross-sectional
the optic disc remains as the standard protocol to analyze images of the retina almost five times faster than OCT 3.
RNFL thickness, improvement of the database is demand- The demonstrated system enables three-dimensional data
ing. Some other new features such as disc size compensa- acquisition with a fundus field of view up to 151, 64 points
tion as well as new measurement landmarks and scan per axial scan, and 256 lines per cross-sectional image.
positions may be needed. The belief that optimized Such a low number of pixels prevents exact analysis of
acquisition protocols and an improved algorithm for cross-sectional information. In order to design three-
RNFL measurements, coupled with the current axial dimensional OCT instruments capable of collecting cross-
resolution of third generation OCT systems, can offer an sectional images with pixel counts similar to third
extraordinary tool to glaucoma diagnostic and follow-up generation OCT, the acquisition speed should be increased
seems quite reasonable. by at least 50 times compared to the commercial unit. This
Third generation OCT has superior resolution (o10 mm) can be realized only by a significant redesign of the OCT
compared with other instruments currently available for system. A recently demonstrated development is the novel
the same purpose. Some next generation ultrahigh-resolu- application of Fourier domain detection to OCT technol-
tion OCT prototypes have even more axial resolution ogy (Hausler and Linduer, 1998; Wojtkowski et al., 2002b).
(!2–3 mm) (Wollstein et al., 2005c). In addition, the This new method significantly improves the speed and
possibility to associate the ultrahigh-resolution with a sensitivity of OCT instruments (Choma et al., 2003a;
spectral/Fourier domain detection enables a dramatic de Boer et al., 2003; Leitgeb et al., 2003a).
increase in image acquisition speed (Wojtkowski et al., One of the most important considerations for OCT
2005) and potentially eliminates the motion problem instruments imaging the laminar structure of the retina is
previously associated to high-density images. the axial image resolution. This parameter is determined by
the spectral bandwidth of the light source used in the OCT
4. Next generation OCT devices instrument (Drexler, 2004; Fercher et al., 2003). In order to
improve the axial resolution, new broad bandwidth light
Significant progress in the field of OCT retinal imaging sources have been constructed and applied to OCT systems
has been made since the third generation of OCT (Drexler, 2004; Drexler et al., 1999). Application of these
instruments was introduced by Zeiss Meditec in 2002. light sources to the novel high speed OCT instruments
Since then a new instrument, which combines OCT with based on Fourier domain detection can provide a very
scanning laser ophthalmoscopy has been introduced by powerful tool for ophthalmic diagnostics in the future.
OTI in 2004 (Podoleanu et al., 2004). Also a new way of
scanning has been used in OCT instrumentation for three- 4.1. Spectral OCT instrument using Fourier domain
dimensional imaging presented by Laser Diagnostic detection
Technologies (Hitzenberger et al., 2003). Novel tools and
OCT measurement techniques have been developed in Recently developed Fourier domain OCT imaging
research laboratories. Some of them may have significant techniques dramatically improve the sensitivity and ima-
impact on the OCT retinal-imaging field in the future. The ging speed of OCT (Choma et al., 2003a; de Boer et al.,
most important developments are probably new light 2003; Leitgeb et al., 2003a). In Fourier domain OCT the
sources enabling imaging with sub-micron axial resolution axial structure of an object (optical A-scan) is retrieved
(Drexler et al., 1999, 2001, 2003; Kowalevicz et al., 2002; from the interferometric signal detected as a function of the
Fujimoto, 2003) and a novel high-speed OCT technique optical frequency (spectral fringe pattern). Fourier domain
called spectral OCT (Fercher et al., 1995; Hausler and OCT detection can be performed in two complementary
Linduer, 1998; Wojtkowski et al., 2002b). Spectral OCT is ways: Spectral OCT (SOCT) using a spectrometer with a
based on ‘‘Fourier domain’’ detection, which allows multi-channel detector (Fercher et al., 1995) or Swept
increasing the measurement speed more than 50 times Source OCT using a rapidly tunable laser source (Chinn
comparing to commercial OCT instrument (Nassif et al., et al., 1997; Lexer et al., 1997; Yun et al., 2003). Spectral
2004; Wojtkowski et al., 2003). OCT and swept source OCT are especially promising for
In principle OCT is able to provide three-dimensional ultrahigh-resolution imaging because they overcome the
information about the retinal morphology similar to other imaging speed limitations of standard OCT. Therefore, it is
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 343

possible to use these techniques to form three-dimensional retinal imaging will be demonstrated by a spectral OCT
maps of the macula and the optic disc (Nassif et al., 2004; instrument.
Wojtkowski et al., 2004a). In addition, the advantage of
providing direct access to the spectral fringe pattern 4.1.1. Standard-resolution retinal imaging with high-speed
enables a wide range of novel applications. Direct access spectral OCT
to the spectral information and phase of the interference The measurement speed of a third generation OCT
fringes permits measurement of absorption (Leitgeb et al., instrument is less than 400 axial scans per second.
2000), numerical compensation of dispersion (Cense et al., Therefore, this instrument needs more than 1 s (approxi-
2004; Wojtkowski et al., 2004b), and numerical resolution mately 1.92 s) to acquire an OCT image with 512 optical
improvement (Szkulmowski et al., 2005). It was also A-scans. One-second measurements by regular OCT
demonstrated that SOCT measurements have the advan- suffer from motion artifacts in the obtained cross-sectional
tage of high phase stability, which causes minimum images (Fig. 28). These artifacts can be corrected
detectable flow velocity. This is 25 times less than what by automated numerical alignment of adjacent optical
has been measured using standard OCT and results in A-scans. This procedure can generate errors in the presence
100 mm/s–5 mm/s of measurable flow velocities (Leitgeb of discontinuities in the retinal structure caused by
et al., 2003b, 2004b; White et al., 2003). pathological changes, or that are naturally existent in the
The first ophthalmic application of single-scan spectral region of the optic disc. Also, these methods ‘‘flatten’’
OCT was the measurement of eye length (Fercher et al., cross-sectional images and information about the true
1995). Demonstration of biomedical OCT imaging of topography of the retina is automatically lost. Fig. 28
human skin in vivo using Fourier domain detection was shows a comparison of cross-sectional images of normal
presented in 1996 (Hausler and Linduer, 1998). The first macula taken by standard third generation OCT and the
retinal and anterior chamber imaging using spectral OCT new spectral OCT instrument based on Fourier domain
was reported in 2002 (Wojtkowski et al., 2002b). Con- detection. In both cases the axial resolution is 10 mm. The
tinuation of this work resulted in the demonstration of the motion artifacts in the presented third generation OCT
first high-speed retinal imaging obtained using a spectral image required numerical correction whereas the spectral
OCT system with 10 mm axial resolution and acquisition OCT cross-sectional image, measured in only 0.17 s, did
time of 8 ms for a 128 transverse pixel image (Wojtkowski not require any motion correction. The speed advantage of
et al., 2003). Application of the line scan CCD camera, the spectral OCT instrument enables the acquisition of
enabling high acquisition speeds of up to 29,000 axial scans cross-sectional images with many more optical A-scans.
per second (Nassif et al., 2004). This represents an The cross-sectional image of the macula presented in
approximately 70 " improvement in imaging speed com- Fig. 28(c) is reconstructed from 4000 A-scans. An increased
pared to third generation OCT, which only operates at 400 number of A-scans and a slightly increased transverse
axial scans per second. resolution can dramatically improve the quality of SOCT
One of the first experiments with combined Fourier images. Fig. 29(a) shows the cross sectional SOCT image of
domain and ultrahigh-resolution OCT was performed in the retinal ‘‘panorama’’ measured horizontally from the
2003, and examples of ophthalmic imaging with 3 mm axial fovea to the inferior part of the optic disc. Fig. 29(b) shows
resolution were published in 2004 (Wojtkowski et al., another cross-sectional image taken across the fovea of the
2004a). Other groups have also recently demonstrated same subject. Both of these images are reconstructed from
ultrahigh-resolution imaging using SOCT. A resolution of 2500 A-scans with an axial resolution of 10 mm. Here the
3.5 mm in the retina was achieved at acquisition rates of nerve fiber layer, ganglion cell layer, inner and outer
15,000 axial scans per second (Cense et al., 2004). An image plexiform layers, inner and outer nuclear layer, external
resolution of 2.5 mm was demonstrated at 10,000 axial limiting membrane, junction between the photoreceptor
scans per second (Leitgeb et al., 2004a). The best axial inner and outer segments, and retinal pigment epithelium
resolution in the retina to date, 2.1 mm, was achieved using are all well visualized and delineated more clearly than in
high-speed acquisition rates of 16,000 axial scans per standard third generation OCT.
second by the group at MIT (Wojtkowski et al., 2004b). Another important advantage of high-speed imaging is
The first demonstration of ophthalmic ranging by swept the possibility of real time observation of cross-sectional
source OCT was done in 1997 (Lexer et al., 1997). images measured in vivo. This makes it more comfortable
Recently, a group from Wellman Laboratories demon- for the operator and can shorten the total examination
strated a new high speed tunable laser operating with a time. It is also possible to choose the region of interest and
central wavelength of 1310 nm enabling imaging at the rate the scanning range during the examination, which can help
of 15,600 A-scans per second (Yun et al., 2003). The in imaging small focal pathologic changes present outside
absorption of water at 1310 nm prevents the application of the macula or optic disc regions. Furthermore, the real-
this laser to retinal imaging. A swept source OCT system time imaging performed by the spectral OCT instrument
using 800 nm wavelength suitable for retinal imaging has enables observation of dynamic changes present in the
not been reported to date. Therefore, in this contribution, retina. In Fig. 30 the set of real-time spectral OCT cross-
the potential of the new Fourier domain OCT detection for sectional images of the peripheral part of the human optic
ARTICLE IN PRESS
344 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

Fig. 28. Comparison of standard-resolution third generation optical coherence tomography (OCT) (Stratus OCT) and spectral OCT images. (A) Stratus
OCT image of macula contains 512 axial scans (A-scans) and was acquired in 1.3 s with axial resolution of 10 micrometers. (B) Stratus OCT image after
numerical correction of motion artifacts. (C) spectral OCT image of macula contains 4000 A-scans and was acquired in 0.17 s with axial resolution of
10 mm.

disc is presented. Each frame consists of 128 A-scans with OCT instruments enable the acquisition of all this
512 samples per A-scan. The exposure time is 32 ms per information with the use of one scanning protocol—the
A-scan. The transverse scan was performed superiorly to raster scan, which provides three-dimensional volumetric
the optic disc cup area in order to examine blood vessels. data of retinal structure (Nassif et al., 2004; Wojtkowski
The region indicated by the arrow in the last frame varies et al., 2004a). Additionally, raster scanning simplifies
during the measurement, whereas the rest of the image is data processing and reconstruction of cross-sectional
stable. This is most likely caused by a pulsation of the images rendered with arbitrary orientations. For spectral
blood vessels. OCT, raster scanning also can provide combined profilo-
The most important advantage of high-speed imaging metric and cross-sectional information that is important
with spectral OCT is that this technique enables three- for quantitative analysis of optic nerve head parameters
dimensional data collection similar to other tomographic or nerve fiber layer thickness. Fig. 31 shows examples of
techniques. Standard third generation OCT devices use the imaging and processing of three-dimensional retinal
specific imaging protocols in order to obtain quantitative data. In Fig. 31(a), the volume rendering of the optic
information about full retinal thickness, retinal nerve fiber disc region is shown. Using software similar to that used
layer thickness, and optic nerve head parameters. Spectral for MRI enables segmentation of specific intraretinal
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 345

Fig. 29. Standard-resolution spectral optical coherence tomography retinal imaging with high quality: (A) cross-sectional image of the retinal
‘‘panorama’’ measured horizontally from temporal to nasal along the fovea to the inferior part of the optic disc, (B) cross-sectional image of the fovea
measured in the same eye. Both images contain 2500 axial scans. Red square shows approximately the region where the bottom image was taken.

layers from the three-dimensional data set (Fig. 31(b)). visualization of retinal architectural morphology, and
Three-dimensional imaging also has the advantage of promise to improve the accuracy of quantitative morpho-
reconstructing fundus view (Wojtkowski et al., 2004a) metric measurements. Ultrahigh-resolution OCT enables
similar to this obtained by scanning laser ophthalmoscopy. the detection of individual retinal layers such as the ganglion
The OCT fundus image is generated by summing the cell layer, inner and outer nuclear and plexiform layers,
reflectivities of successive layers along the axial direction. as well as the photoreceptor and RPE morphology,
The SOCT fundus image enables precise registration of which are difficult to visualize with standard-resolution
OCT cross-sections with fundus photography. Fig. 32 OCT (Drexler et al., 1999, 2001, 2003; Ko et al., 2004). In
shows the fundus image created from a three-dimensional contrast to commercially available OCT instruments,
data set acquired using standard-resolution spectral OCT. ultrahigh-resolution OCT can reveal changes in retinal
The pattern created by the blood vessels is clearly visible morphology associated with retinal disease, such as photo-
and can be used to correlate the OCT data with the fundus receptor integrity or impairment (Drexler et al., 2003;
photograph. Ko et al., 2004).
Ultrahigh-resolution OCT has not yet been commercia-
4.1.2. High-resolution retinal imaging with high-speed lized because of the high cost of femtosecond lasers. The
spectral OCT recent introduction of compact broadband semiconductor
The first demonstration of ultrahigh-resolution retinal light sources will enable widespread use of ultrahigh-
imaging with 3-mm resolution using a standard OCT system resolution OCT instruments. The cost of the broadband
was demonstrated in 1999 by Drexler et al. This represents a superluminescent diode is approximately 5 times lower
factor of five- to ten- fold resolution improvement over than the cost of the full Ti:Sa laser system. These light
standard third generation OCT instruments, which has sources are based on two or more superluminescent diode
10 mm axial resolution. This technology has already been modules combined by single mode fiber couplers. The
implemented in two clinical systems constructed at MIT application of broadband semiconductor light sources to
(Drexler et al., 2001; Ko et al., 2005) and the University of OCT ophthalmic imaging has been demonstrated in 2004
Vienna (Drexler et al., 2003). Both of these systems use (Ko et al., 2004). The high-quality retinal images obtained
femtosecond titanium:sapphire (Ti:Sa) lasers as light sources. by a spectral OCT system using combined superlumines-
Clinical results obtained with these instruments demonstrate cent diode modules has been also presented in 2004 (Cense
that ultrahigh-resolution OCT can greatly improve the et al., 2004).
ARTICLE IN PRESS
346 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

Fig. 30. Real-time spectral optical coherence tomography observation of dynamic processes in the retina: (A) the set of cross-sectional images measured
with 8 frames per second; (B) cross-sectional image of optic disc region with indicated region of interest.

Fig. 33 shows the cross-sectional SOCT images of the including photoreceptor layer details such as the external
macula and optic disc measured with a high axial limiting membrane (ELM), junction between the inner and
resolution of 4.5 mm. The delineation of retinal layers in outer segments (IS/OS), and the RPE. Fig. 34 shows an
the presented image is much clearer than that of standard- example of the full thickness retinal map and a thickness
resolution imaging. The improvement is especially visible map of the part of the photoreceptor layer from IS/OS to
in the retinal pigment epithelium, ganglion cell layer, and RPE. This analysis has great potential in objective
photoreceptor layer. The high number of collected A-scans measurements of progression in various macular diseases.
helps to decrease the noise level and improves the Simultaneous increase of the coverage and resolution,
continuity of retinal layers. The performance of automated guaranteed by three-dimensional ultrahigh-resolution high-
segmentation algorithms for thickness measurements and speed SOCT, will enable analysis of small focal patholo-
layer identification can also be improved. The collection of gical changes that can be missed by standard OCT
three-dimensional ultrahigh-resolution SOCT data enables techniques. This can effectively improve the capability of
the quantitative analysis of all major retinal layers, OCT technology to diagnose early pathological changes
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 347

Fig. 31. Three-dimensional spectral optical coherence tomography imaging with standard resolution: (A) volume rendering of optic disc region,
(B) segmentation of intraretinal layers in macular region; NFL-nerve fiber layer, OPL-outer plexiform layer, IS/OS-junction between inner and outer
segments of photoreceptors, and RPE- retinal pigment epithelium.

Fig. 32. Three-dimensional spectral optical coherence tomography (OCT) imaging with standard resolution: (A) fundus view (400 horizontal points, 300
vertical points) reconstructed from the spectral OCT data contains 300 cross-sectional images; (B) two cross-sectional images from the three-dimensional
set of data. The location of each cross-sectional image is perfectly registered relative to the fundus view.

and can help in understanding the pathogenesis of retinal mology clinics. For example, images obtained by SOCT
diseases. can suffer from coherent noise artifacts and conjugate
images (Wojtkowski et al., 2002a, b), which decrease the
4.2. Additional considerations effective axial measurement range and can lead to
misinterpretation of resultant tomographic images. These
Combined ultrahigh resolution and Fourier domain artifacts cause ‘‘folding’’ of the OCT cross-sectional image
detection techniques can give unprecedented improvement if the optical distance of the structure to be imaged is
of the capability of OCT systems for retinal imaging. higher than the effective measurement range. It has been
However, the advantages of the new spectral OCT shown that these unwanted effects can be eliminated by
technique cannot be easily realized as a commercial clinical phase-shifting methods (Choma et al., 2003b; Targowski
instrument. Technical problems still exist, which impede et al., 2004, 2005; Wojtkowski et al., 2002a). These
the introduction of spectral OCT instruments to ophthal- techniques require the collection of multiple signals from
ARTICLE IN PRESS
348 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

Fig. 33. High-resolution spectral optical coherence tomography imaging: (A) cross-sectional image of macula, and (B) cross-sectional image of optic disc.
Both images contain 10.000 axial scans and were acquired in 0.43 s with axial resolution of 4.5 mm.

the same region taken with an additional selected phase technology, which will enable the application of swept
shift in the reference arm, with the condition that the object source OCT to retinal imaging.
is stationary between the measurements to within a fraction
of micrometer. The latter condition makes these phase- 5. Concluding remarks
shifting methods hard to apply for clinical practice.
Another unsolved problem in spectral OCT is the drop of The introduction of OCT in ophthalmology represents a
sensitivity with increase of imaging depth (Hausler and definitive change in the way doctors understand and treat
Linduer, 1998; Wojtkowski et al., 2002b). This effect is several diseases affecting the retina. It is quite likely as well,
especially significant in ultrahigh-resolution SOCT imaging that the role of OCT as a method to diagnose and manage
where the digitalization must be much finer than in glaucoma will be further defined in the near future.
standard-resolution OCT. This problem is fundamentally Understanding of the basic principles in which OCT relays
associated with the detection performed by a spectrometer on is essential to understand its actual limitations and to
(Hausler and Linduer, 1998) and it is much less severe in use this technology with wisdom. We have already learned
swept source OCT (Yun et al., 2003). Also the swept source a lot with data provided by first generations of OCT, and
OCT gives more flexibility in controlling of the axial scan there is much more to learn with forthcoming data from
parameters, which are fixed in spectral OCT (Wojtkowski numerous ongoing studies worldwide that, presently, are
et al., 2004b). The future of the clinical application of using third generation OCT systems. A huge leap forward
Fourier domain detection to retinal imaging will depend in improving OCT imaging performance is expected to
on how the problems mentioned above will be solved, occur within the next few years (or should we say months?)
as well as on the development of high-speed tunable laser with the commercial availability of next generation OCT
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 349

Fig. 34. Three-dimensional high-resolution, spectral optical coherence tomography imaging: (A) reconstruction of the OCT fundus view, (B) cross-
sectional image with delineated automatically segmented layers, (C) full retinal thickness map, and (D) thickness map of outer segments of photoreceptors;
ILM-inner limiting membrane, IS/OS-junction between inner and outer segments of photoreceptors, and RPE-retinal pigment epithelium.

devices. One should say that three-dimensional, high- Antcliff, R.J., Spalton, D.J., Stanford, M.R., Graham, E.M., Fytche, T.J.,
speed, ultrahigh-resolution retinal assessment is just Marshall, J., 2001. Intravitreal triamcinolone for uveitic cystoid
‘‘around the corner’’. macular edema: an optical coherence tomography study. Ophthalmol-
ogy 108, 765–772.
Bagga, H., Greenfield, D.S., Knighton, R.W., 2005. Macular symmetry
testing for glaucoma detection. J. Glaucoma 14, 358–363.
Acknowledgments Bartsch, D-U., Freeman, W.R., 1994. Axial intensity distribution analysis
of the human retina with a confocal scanning laser tomograph. Exp.
M.W. thanks Prof. Andrzej Kowalczyk and members of Eye Res. 58, 161–173.
Medical Physics Group from Nicolaus Copernicus Uni- Blumenthal, E.Z., Williams, J.M., Weinreb, R.N., Girkin, C.A., Berry,
versity in Torun especially to Iwona Gorczynska and Anna C.C., Zangwill, L.M., 2000. Reproducibility of nerve fiber layer
thickness measurements by use of optical coherence tomography.
Szkulmowska, and Prof. James G. Fujimoto from Massa- Ophthalmology 107, 2278–2282.
chusetts Institute of Technology, Cambridge, MA, USA Bonini-Filho, M.A., Jorge, R., Barbosa, J.C., Calucci, D., Cardillo, J.A.,
and members of his group: Vikas Sharma, Aurea Zare, Costa, R.A., 2005. Intravitreal injection versus sub-Tenon’s infusion
Vivek Srinivasan, Tony Ko and Mariana Carvalho. R.A.C. of triamcinolone acetonide for refractory diabetic macular edema:
and M.W. includes special thank to Yijun Huang, and to a randomized clinical trial. Invest. Ophthalmol. Vis. Sci. 46,
3845–3849.
Robert Huber and Robert Zawadzki, respectively, for
Bowd, C., Weinreb, R.N., Williams, J.M., Zangwill, L.M., 2000. The
creative discussions. retinal nerve fiber layer thickness in ocular hypertensive, normal, and
glaucomatous eyes with optical coherence tomography. Arch.
Ophthalmol. 118, 22–26.
References Bowd, C., Zangwill, L.M., Berry, C.C., Blumenthal, E.Z., Vasile, C.,
Sanchez-Galeana, C., Bosworth, C.F., Sample, P.A., Weinreb, R.N.,
Aleman, T.S., Cideciyan, A.V., Volpe, N.J., Stevanin, G., Brice, A., 2001. Detecting early glaucoma by assessment of retinal nerve fiber
Jacobson, S.G., 2002. Spinocerebellar ataxia type 7 (SCA7) shows a layer thickness and visual function. Invest. Ophthalmol. Vis. Sci. 42,
cone-rod dystrophy phenotype. Exp. Eye Res. 74, 737–745. 1993–2003.
ARTICLE IN PRESS
350 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

Browning, D.J., McOwen, M.D., Bowen Jr., R.M., O’Marah, T.L., 2004. Costa, R.A., Calucci, D., Skaf, M., Cardillo, J.A., Castro, J.C., Melo Jr.,
Comparison of the clinical diagnosis of diabetic macular edema with L.A., Martins, M.C., Kaiser, P.K., 2004. Optical coherence tomo-
diagnosis by optical coherence tomography. Ophthalmology 111, graphy 3: automatic delineation of the outer neural retinal boundary
712–715. and its influence on retinal thickness measurements. Invest. Ophthal-
Budenz, D.L., Chang, R.T., Huang, X., Knighton, R.W., Tielsch, J.M., mol. Vis. Sci. 45, 2399–2406.
2005. Reproducibility of retinal nerve fiber thickness measurements Costa, R.A., Calucci, D., Paccola, L., Jorge, R., Cardillo, J.A., Castro,
using the stratus OCT in normal and glaucomatous eyes. Invest. J.C., Scott, I.U., 2005. Occult chorioretinal anastomosis in age-related
Ophthalmol. Vis. Sci. 46, 2440–2443. macular degeneration: a prospective study by optical coherence
Burgansky-Eliash, Z., Wollstein, G., Chu, T., Ramsey, J.D., Glymour, C., tomography. Am. J. Ophthalmol. 140, 107–116.
Noecker, R.J., Ishikawa, H., Schuman, J.S., 2005. Optical coherence de Boer, J.F., Cense, B., Park, B.H., Pierce, M.C., Tearney, G.J., Bouma,
tomography machine learning classifiers for glaucoma detection: a B.E., 2003. Improved signal-to-noise ratio in spectral-domain com-
preliminary study. Invest. Ophthalmol. Vis. Sci. 46, 4147–4152. pared with time-domain optical coherence tomography. Opt. Lett. 28,
Campochiaro, P.A., C99-PKC412-003 Study Group, 2004. Reduction of 2067–2069.
diabetic macular edema by oral administration of the kinase inhibitor Doggrell, S.A., 2005. Pegaptanib: the first antiangiogenic agent approved
PKC412. Invest. Ophthalmol. Vis. Sci. 45, 922–931. for neovascular macular degeneration. Expert. Opin. Pharmacother. 6,
Cardillo, J.A., Melo Jr., L.A., Costa, R.A., Skaf, M., Belfort Jr., R., 1421–1423.
Souza-Filho, A.A., Farah, M.E., Kuppermann, B.D., 2005. Compar- Drexler, W., 2004. Ultrahigh-resolution optical coherence tomography.
ison of intravitreal versus posterior sub-Tenon’s capsule injection of J. Biomed. Opt. 9, 47–74.
triamcinolone acetonide for diffuse diabetic macular edema. Ophthal- Drexler, W., Morgner, U., Kartner, F.X., Pitris, C., Boppart, S.A., Li,
mology 112, 1557–1563. X.D., Ippen, E.P., Fujimoto, J.G., 1999. In vivo ultrahigh-resolution
Carpineto, P., Ciancaglini, M., Zuppardi, E., Falconio, G., Doronzo, E., optical coherence tomography. Opt. Lett. 24, 1221–1223.
Mastropasqua, L., 2003. Reliability of nerve fiber layer thickness Drexler, W., Morgner, U., Ghanta, R.K., Kärtner, F.X., Schuman, J.S.,
measurements using optical coherence tomography in normal and Fujimoto, J.G., 2001. Ultrahigh-resolution ophthalmic optical coherence
glaucomatous eyes. Ophthalmology 110, 190–195. tomography. Nat. Med. 7, 502–507 (Erratum in: 2001. Nat. Med. 7, 636).
Catier, A., Tadayoni, R., Paques, M., Erginay, A., Haouchine, B., Drexler, W., Sattmann, H., Hermann, B., Ko, T.H., Stur, M., Unterhuber,
Gaudric, A., Massin, P., 2005. Characterization of macular edema A., Scholda, C., Findl, O., Wirtitsch, M., Fujimoto, J.G., Fercher,
from various etiologies by optical coherence tomography. Am. J. A.F., 2003. Enhanced visualization of macular pathology with the use
Ophthalmol. 140, 200–206. of ultrahigh-resolution optical coherence tomography. Arch. Ophthal-
Cense, B., Nassif, N., Chen, T.C., Pierce, M.C., Yun, S., Park, B.H., mol. 121, 695–706.
Bouma, B., Tearney, G., de Boer, J.F., 2004. Ultrahigh-resolution Eandi, C.M., Chung, J.E., Cardillo-Piccolino, F., Spaide, R.F., 2005.
high-speed retinal imaging using spectral-domain optical coherence Optical coherence tomography in unilateral resolved central serous
tomography. Opt. Express 12, 2435–2447. chorioretinopathy. Retina 25, 417–421.
Chen, H.Y., Huang, M.L., 2005. Discrimination between normal and Eter, N., Spaide, R.F., 2005. Comparison of fluorescein angiography and
glaucomatous eyes using Stratus optical coherence tomography in optical coherence tomography for patients with choroidal neovascu-
Taiwan Chinese subjects. Graefes Arch. Clin. Exp. Ophthalmol. 243, larization after photodynamic therapy. Retina 25, 691–696.
894–902. Fercher, A.F., Hitzenberger, C.K., Kamp, G., Elzaiat, S.Y., 1995.
Chinn, S.R., Swanson, E.A., Fujimoto, J.G., 1997. Optical coherence Measurement of intraocular distances by backscattering spectral
tomography using a frequency-tunable optical source. Opt. Lett. 22, interferometry. Opt. Commun. 117, 43–48.
340–342. Fercher, A.F., Drexler, W., Hitzenberger, C.K., Lasser, T., 2003. Optical
Choma, M.A., Sarunic, M.V., Yang, C.H., Izatt, J.A., 2003a. Sensitivity coherence tomography-principles and applications. Reports Prog.
advantage of swept source and Fourier domain optical coherence Phys. 66, 239–303.
tomography. Opt. Express 11, 2183–2189. Fujimoto, J.G., 2003. Optical coherence tomography for ultrahigh
Choma, M.A., Yang, C.H., Izatt, J.A., 2003b. Instantaneous quadrature resolution in vivo imaging. Nat. Biotech. 21, 1361–1367.
low-coherence interferometry with 3 " 3 fiber-optic couplers. Opt. Gaucher, D., Tadayoni, R., Erginay, A., Haouchine, B., Gaudric, A.,
Lett. 28, 2162–2164. Massin, P., 2005. Optical coherence tomography assessment of the
Cideciyan, A.V., Jacobson, S.G., Aleman, T.S., et al., 2005. In vivo vitreoretinal relationship in diabetic macular edema. Am. J. Ophthal-
dynamics of retinal injury and repair in the rhodopsin mutant dog mol. 139, 807–813.
model of human retinitis pigmentosa. Proc. Natl. Acad. Sci. USA 102, Gaudric, A., Haouchine, B., Massin, P., Paques, M., Blain, P., Erginay,
5233–5238 (Epub 2005 Mar 22). A., 1999. Macular hole formation: new data provided by optical
Costa, R.A., 2005. Evaluation of image artifact produced by optical coherence tomography. Arch. Ophthalmol. 117, 744–751.
coherence tomography of retinal pathology. Am. J. Ophthalmol. 140, Gragoudas, E.S., Adamis, A.P., Cunningham Jr., E.T., Feinsod, M.,
349–350. Guyer, D.R., VEGF Inhibition Study in Ocular Neovascularization
Costa, R.A., Scapucin, L., Moraes, N.S., Calucci, D., Melo Jr., L.A., Clinical Trial Group, 2004. Pegaptanib for neovascular age-related
Cardillo, J.A., Farah, M.E., 2002. Indocyanine green-mediated macular degeneration. N. Engl. J. Med. 351, 2805–2816.
photothrombosis as a new technique of treatment for persistent Greenberg, P.B., Martidis, A., Rogers, A.H., Duker, J.S., Reichel, E.,
central serous chorioretinopathy. Curr. Eye Res. 25, 287–297. 2002. Intravitreal triamcinolone acetonide for macular oedema due to
Costa, R.A., Farah, M.E., Cardillo, J.A., Calucci, D., Williams, G.A., central retinal vein occlusion. Br. J. Ophthalmol. 86, 247–248.
2003a. Immediate indocyanine green angiography and optical coher- Gross, J.G., 2005. Late reopening and spontaneous closure of previously
ence tomography evaluation after photodynamic therapy for subfoveal repaired macular holes. Am. J. Ophthalmol. 140, 556–558.
choroidal neovascularization. Retina 23, 159–165. Guedes, V., Schuman, J.S., Hertzmark, E., Wollstein, G., Correnti, A.,
Costa, R.A., Calucci, D., Cardillo, J.A., Farah, M.E., 2003b. Selective Mancini, R., Lederer, D., Voskanian, S., Velazquez, L., Pakter, H.M.,
occlusion of subfoveal choroidal neovascularization in angioid streaks Pedut-Kloizman, T., Fujimoto, J.G., Mattox, C., 2003. Optical
by using a new technique of ingrowth site treatment. Ophthalmology coherence tomography measurement of macular and nerve fiber layer
110, 1192–1203. thickness in normal and glaucomatous human eyes. Ophthalmology
Costa, R.A., Calucci, D., Teixeira, L.F., Cardillo, J.A., Bonomo, P.P., 110, 177–189.
2003c. Selective occlusion of subfoveal choroidal neovascularization in Haouchine, B., Massin, P., Gaudric, A., 2001. Foveal pseudocyst as the
pathologic myopia using a new technique of ingrowth site treatment. first step in macular hole formation: a prospective study by optical
Am. J. Ophthalmol. 135, 857–866. coherence tomography. Ophthalmology 108, 15–22.
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 351

Haouchine, B., Massin, P., Tadayoni, R., Erginay, A., Gaudric, A., 2004. for human gene therapy success. Proc. Natl. Acad. Sci. USA 102,
Diagnosis of macular pseudoholes and lamellar macular holes by 6177–6182 (Epub 18 April 2005).
optical coherence tomography. Am. J. Ophthalmol. 138, 732–739. Johnson, M.W., 2005. Tractional cystoid macular edema: a subtle variant
Hausler, G., Linduer, M.W., 1998. ‘‘Coherence radar’’ and ‘‘spectral of the vitreomacular traction syndrome. Am. J. Ophthalmol. 140,
radar’’—new tools for dermatological diagnosis. J. Biomed. Opt. 3, 184–192.
21–31. Jones, A.L., Sheen, N.J., North, R.V., Morgan, J.E., 2001. The Humphrey
Hee, M.R., 2005. Artifacts in optical coherence tomography topographic optical coherence tomography scanner: quantitative analysis and
maps. Am. J. Ophthalmol. 139, 154–155. reproducibility study of the normal human retinal nerve fibre layer.
Hee, M.R., Izatt, J.A., Swanson, E.A., Huang, D., Schuman, J.S., Lin, Br. J. Ophthalmol. 85, 673–677.
C.P., Puliafito, C.A., Fujimoto, J.G., 1995a. Optical coherence Jonas, J.B., Sofker, A., 2001. Intraocular injection of crystalline cortisone
tomography of the human retina. Arch. Ophthalmol. 113, 325–332. as adjunctive treatment of diabetic macular edema. Am. J. Ophthal-
Hee, M.R., Puliafito, C.A., Wong, C., Duker, J.S., Reichel, E., Schuman, mol. 132, 425–427.
J.S., Swanson, E.A., Fujimoto, J.G., 1995b. Optical coherence Jonas, J.B., Kreissig, I., Degenring, R.F., 2002. Intravitreal triamcinolone
tomography of macular holes. Ophthalmology 102, 748–756. acetonide as treatment of macular edema in central retinal vein
Hee, M.R., Puliafito, C.A., Duker, J.S., Reichel, E., Coker, J.G., Wilkins, occlusion. Graefes Arch. Clin. Exp. Ophthalmol. 240, 782–783 (Epub
J.R., Schuman, J.S., Swanson, E.A., Fujimoto, J.G., 1998. Topogra- 13 August 2002).
phy of diabetic macular edema with optical coherence tomography. Jorge, R., Costa, R.A., Quirino, L.S., Paques, M.W., Calucci, D., Cardillo,
Ophthalmology 105, 360–370. J.A., Scott, I.U., 2004. Optical coherence tomography findings in
Hess, D.B., Asrani, S.G., Bhide, M.G., Enyedi, L.B., Stinnett, S.S., patients with late solar retinopathy. Am. J. Ophthalmol. 137, 1139–1143.
Freedman, S.F., 2005. Macular and retinal nerve fiber layer analysis of Kerrigan-Baumrind, L.A., Quigley, H.A., Pease, M.E., Kerrigan, D.F.,
normal and glaucomatous eyes in children using optical coherence Mitchell, R.S., 2000. Number of ganglion cells in glaucoma eyes
tomography. Am. J. Ophthalmol. 139, 509–517. compared with threshold visual field tests in the same persons. Invest.
Hillenkamp, J., Saikia, P., Gora, F., Sachs, H.G., Lohmann, C.P., Roider, Ophthalmol. Vis. Sci. 41, 741–748.
J., Baumler, W., Gabel, V.P., 2005. Macular function and morphology Kitaya, N., Hikichi, T., Kagokawa, H., Takamiya, A., Takahashi, A.,
after peeling of idiopathic epiretinal membrane with and without the Yoshida, A., 2004. Irregularity of photoreceptor layer after successful
assistance of indocyanine green. Br. J. Ophthalmol. 89, 437–443. macular hole surgery prevents visual acuity improvement. Am. J.
Hitzenberger, C.K., Trost, P., Lo, P.W., Zhou, Q.Y., 2003. Three- Ophthalmol. 138, 308–310.
dimensional imaging of the human retina by high-speed optical Ko, T.H., Adler, D.C., Fujimoto, J.G., Mamedov, D., Prokhorov, V.,
coherence tomography. Opt. Expr. 11, 2753–2761. Shidlovski, V., Yakubovich, S., 2004. Ultrahigh resolution optical
Hoyt, W.F., Newman, N.M., 1972. The earliest observable defect in coherence tomography imaging with a broadband superluminescent
glaucoma? Lancet 1, 692–693. diode light source. Opt. Express 12, 2112–2119.
Hoyt, W.F., Frisen, L., Newman, N.M., 1973. Fundoscopy of nerve fiber Ko, T.H., Fujimoto, J.G., Schuman, J.S., Paunescu, L.A., Kowalevicz,
layer defects in glaucoma. Invest. Ophthalmol. 12, 814–829. A.M., Hartl, I., Drexler, W., Wollstein, G., Duker, J.S., Ishikawa, H.,
Huang, D., Swanson, E.A., Lin, C.P., et al., 1991. Optical coherence 2005. Comparison of ultrahigh and standard resolution optical
tomography. Science 254, 1178–1181. coherence tomography for imaging of macular pathology. Ophthal-
Huang, D., Tan, O., Fujimoto, J.G., et al., Optical coherence tomography. mology (22 September 2005, Epub ahead of print).
In: Huang D, Kaiser PK, Lowder CY, Traboulsi E, (Eds.), Retinal Kowalevicz, A.M., Schibli, T.R., Kartner, F.X., Fujimoto, J.G., 2002.
Imaging, first ed., Elsevier; in progress. Ultralow-threshold Kerr-lens mode-locked Ti:Al2O3 laser. Opt. Lett.
Huang, M.L., Chen, H.Y., 2005. Development and comparison of 27, 2037–2039.
automated classifiers for glaucoma diagnosis using stratus optical Leitgeb, R., Wojtkowski, M., Kowalczyk, A., Hitzenberger, C.K., Sticker,
coherence tomography. Invest. Ophthalmol. Vis. Sci. 46, 4121–4129. M., Fercher, A.F., 2000. Spectral measurement of absorption by
Huang, Y., Cideciyan, A.V., Papastergiou, G.I., Banin, E., Semple- spectroscopic frequency-domain optical coherence tomography. Opt.
Rowland, S.L., Milam, A.H., Jacobson, S.G., 1998. Relation of optical Lett. 25, 820–822.
coherence tomography to microanatomy in normal and rd chickens. Leitgeb, R., Hitzenberger, C.K., Fercher, A.F., 2003a. Performance of
Invest. Ophthalmol. Vis. Sci. 39, 2405–2416. Fourier domain vs. time domain optical coherence tomography. Opt.
Huang, Y., 1999. Optical coherence tomography (OCT) in hereditary Express 11, 889–894.
retinal degenerations: layer-by-layer analyses in normal and diseased Leitgeb, R.A., Schmetterer, L., Drexler, W., Fercher, A.F., Zawadzki,
retinas. A dissertation in Bioengineering. Presented to the faculties of R.J., Bajraszewski, T., 2003b. Real-time assessment of retinal blood
the University of Pennsylvania in partial fulfillment of the require- flow with ultrafast acquisition by color Doppler Fourier domain
ments for the degree of Doctor of Philosophy. optical coherence tomography. Opt. Express 11, 3116–3121.
Ishikawa, H., Stein, D.M., Wollstein, G., Beaton, S., Fujimoto, J.G., Leitgeb, R.A., Drexler, W., Unterhuber, A., Hermann, B., Bajraszewski,
Schuman, J.S., 2005. Macular segmentation with optical coherence T., Le, T., Stingl, A., Fercher, A.F., 2004a. Ultrahigh resolution
tomography. Invest. Ophthalmol. Vis. Sci. 46, 2012–2017. Fourier domain optical coherence tomography. Opt. Express 12,
Jacobson, S.G., Cideciyan, A.V., Huang, Y., et al., 1998. Retinal 2156–2165.
degenerations with truncation mutations in the cone-rod homeobox Leitgeb, R.A., Schmetterer, L., Hitzenberger, C.K., Fercher, A.F.,
(CRX) gene. Invest. Ophthalmol. Vis. Sci. 39, 2417–2426. Berisha, F., Wojtkowski, M., Bajraszewski, T., 2004b. Real-time
Jacobson, S.G., Cideciyan, A.V., Iannaccone, A., et al., 2000. Disease measurement of in vitro flow by Fourier-domain color Doppler optical
expression of RP1 mutations causing autosomal dominant retinitis coherence tomography. Opt. Lett. 29, 171–173.
pigmentosa. Invest. Ophthalmol. Vis. Sci. 41, 1898–1908. Lerche, R.C., Schaudig, U., Scholz, F., Walter, A., Richard, G., 2001.
Jacobson, S.G., Cideciyan, A.V., Aleman, T.S., et al., 2003. Crumbs Structural changes of the retina in retinal vein occlusion: imaging and
homolog 1 (CRB1) mutations result in a thick human retina with quantification with optical coherence tomography. Ophthal. Surg.
abnormal lamination. Hum. Mol. Genet. 12, 1073–1078. Lasers 32, 272–280.
Jacobson, S.G., Sumaroka, A., Aleman, T.S., et al., 2004. Nuclear Leung, C.K., Yung, W.H., Ng, A.C., Woo, J., Tsang, M.K., Tse, K.K.,
receptor NR2E3 gene mutations distort human retinal laminar 2004. Evaluation of scanning resolution on retinal nerve fiber layer
architecture and cause an unusual degeneration. Hum. Mol. Genet. measurement using optical coherence tomography in normal and
13, 1893–1902 (Epub 30 June 2004). glaucomatous eyes. J. Glaucoma 13, 479–485.
Jacobson, S.G., Aleman, T.S., Cideciyan, A.V., et al., 2005. Identifying Leung, C.K., Chong, K.K., Chan, W.M., Yiu, C.K., Tso, M.Y., Woo, J.,
photoreceptors in blind eyes caused by RPE65 mutations: prerequisite Tsang, M.K., Tse, K.K., Yung, W.H., 2005a. Comparative study of
ARTICLE IN PRESS
352 R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353

retinal nerve fiber layer measurement by Stratus OCT and GDx VCC, Podoleanu, A.G., Seeger, M., Dobre, G.M., Webb, D.J., Jackson, D.A.,
II: structure/function regression analysis in glaucoma. Invest. Ophthal- Fitzke, F.W., 1998. Transversal and longitudinal images from the
mol. Vis. Sci. 46, 3702–3711. retina of the living eye using low coherence reflectometry. J. Biomed.
Leung, C.K., Chan, W.M., Yung, W.H., Ng, A.C., Woo, J., Tsang, M.K., Opt. 3, 12–20.
Tse, R.K., 2005b. Comparison of macular and peripapillary measure- Podoleanu, A.G., Dobre, G.M., Cucu, R.G., Rosen, R., Garcia, P., Nieto,
ments for the detection of glaucoma: an optical coherence tomography J., Will, D., Gentile, R., Muldoon, T., Walsh, J., Yannuzzi, L.A.,
study. Ophthalmology 112, 391–400. Fisher, Y., Orlock, D., Weitz, R., Rogers, J.A., Dunne, S., Boxer, A.,
Lexer, F., Hitzenberger, C.K., Fercher, A.F., Kulhavy, M., 1997. 2004. Combined multiplanar optical coherence tomography and
Wavelength-tuning interferometry of intraocular distances. Appl. confocal scanning ophthalmoscopy. J. Biomed. Opt. 9, 86–93.
Opt. 36, 6548–6553. Pons, M.E., Garcia-Valenzuela, E., 2005. Redefining the limit of the outer
Martidis, A., Duker, J.S., Puliafito, C.A., 2001. Intravitreal triamcinolone retina in optical coherence tomography scans. Ophthalmology 112,
for refractory cystoid macular edema secondary to birdshot retino- 1079–1085.
choroidopathy. Arch. Ophthalmol. 119, 1380–1383. Puliafito, C.A., Hee, M.R., Schuman, J.S., Fujimoto, J.G., 1996. Optical
Martidis, A., Duker, J.S., Greenberg, P.B., Rogers, A.H., Puliafito, C.A., Coherence Tomography of Ocular Diseases. Slack Inc., Thorafore, NJ.
Reichel, E., Baumal, C., 2002. Intravitreal triamcinolone for refractory Sahni, J., Stanga, P., Wong, D., Harding, S., 2005. Optical coherence
diabetic macular edema. Ophthalmology 109, 920–927. tomography in photodynamic therapy for subfoveal choroidal
Medeiros, F.A., Zangwill, L.M., Bowd, C., Weinreb, R.N., 2004. neovascularisation secondary to age related macular degeneration: a
Comparison of the GDx VCC scanning laser polarimeter, HRT II cross sectional study. Br. J. Ophthalmol. 89, 316–320.
confocal scanning laser ophthalmoscope, and stratus OCT optical Salinas-Alaman, A., Garcia-Layana, A., Maldonado, M.J., Sainz-Gomez,
coherence tomograph for the detection of glaucoma. Arch. Ophthal- C., Alvarez-Vidal, A., 2005. Using optical coherence tomography to
mol. 122, 827–837. monitor photodynamic therapy in age related macular degeneration.
Medeiros, F.A., Zangwill, L.M., Bowd, C., Vessani, R.M., Susanna Jr., Am. J. Ophthalmol. 140, 23–28.
R., Weinreb, R.N., 2005. Evaluation of retinal nerve fiber layer, optic Sanchez-Galeana, C.A., Bowd, C., Zangwill, L.M., Sample, P.A., Weinreb,
nerve head, and macular thickness measurements for glaucoma R.N., 2004. Short-wavelength automated perimetry results are correlated
detection using optical coherence tomography. Am. J. Ophthalmol. with optical coherence tomography retinal nerve fiber layer thickness
139, 44–55. measurements in glaucomatous eyes. Ophthalmology 111, 1866–1872.
Meirelles, R.L., Aggio, F.B., Costa, R.A., Farah, M.E., 2005. STRATUS Sandhu, S.S., Talks, S.J., 2005. Correlation of optical coherence
optical coherence tomography in unilateral colobomatous excavation tomography, with or without additional colour fundus photography,
of the optic disc and secondary retinoschisis. Graefes Arch. Clin. Exp. with stereo fundus fluorescein angiography in diagnosing choroidal
Ophthalmol. 243, 76–81 (Epub 4 August 2004). neovascular membranes. Br. J. Ophthalmol. 89, 967–970.
Michels, S., Rosenfeld, P.J., Puliafito, C.A., Marcus, E.N., Venkatraman, Savini, G., Zanini, M., Carelli, V., Sadun, A.A., Ross-Cisneros, F.N.,
A.S., 2005. Systemic bevacizumab (Avastin) therapy for neovascular Barboni, P., 2005. Correlation between retinal nerve fibre layer
age-related macular degeneration twelve-week results of an uncon- thickness and optic nerve head size: an optical coherence tomography
trolled open-label clinical study. Ophthalmology 112, 1035–1047. study. Br. J. Ophthalmol. 89, 489–492.
Milam, A.H., Barakat, M.R., Gupta, N., et al., 2003. Clinicopathologic Schuman, J.S., Hee, M.R., Puliafito, C.A., Wong, C., Pedut-Kloizman, T.,
effects of mutant GUCY2D in Leber congenital amaurosis. Ophthal- Lin, C.P., Hertzmark, E., Izatt, J.A., Swanson, E.A., Fujimoto, J.G.,
mology 110, 549–558. 1995. Quantification of nerve fiber layer thickness in normal and
Montero, J.A., Ruiz-Moreno, J.M., 2005. Optical coherence tomography glaucomatous eyes using optical coherence tomography. Arch.
characterisation of idiopathic central serous chorioretinopathy. Br. J. Ophthalmol. 113, 586–596.
Ophthalmol. 89, 562–564. Schuman, J.S., Pedut-Kloizman, T., Hertzmark, E., Hee, M.R., Wilkins,
Montero, J.A., Ruiz-Moreno, J.M., Tavolato, M., 2003. Follow-up of J.R., Coker, J.G., Puliafito, C.A., Fujimoto, J.G., Swanson, E.A.,
age-related macular degeneration patients treated by photodynamic 1996. Reproducibility of nerve fiber layer thickness measurements
therapy with optical coherence tomography 3. Graefes Arch. Clin. using optical coherence tomography. Ophthalmology 103, 1889–1898.
Exp. Ophthalmol. 241, 797–802. Skaf, M., Bernardes, A.B., Cardillo, J.A., Costa, R.A., Melo, L.A.,
Mrugacz, M., Bakunowicz-Lazarczyk, A., 2005. Optical coherence Castro, J.C., Varma, R., 2005. Retinal nerve fibre layer thickness
tomography measurement of the retinal nerve fiber layer in normal profile in normal eyes using third-generation optical coherence
and juvenile glaucomatous eyes. Ophthalmologica 219, 80–85. tomography. Eye (22 July 2005, Epub ahead of print).
Nassif, N.A., Cense, B., Park, B.H., Pierce, M.C., Yun, S.H., Soliman, M.A., Van Den Berg, T.J., Ismaeil, A.A., De Jong, L.A., De
Bouma, B.E., Tearney, G.J., Chen, T.C., de Boer, J.F., 2004. In vivo Smet, M.D., 2002. Retinal nerve fiber layer analysis: relationship
high-resolution video-rate spectral-domain optical coherence between optical coherence tomography and red-free photography. Am.
tomography of the human retina and optic nerve. Opt. Express 12, J. Ophthalmol. 133, 187–195.
367–376. Sommer, A., Katz, J., Quigley, H.A., Miller, N.R., Robin, A.L., Richter,
Niwa, H., Terasaki, H., Ito, Y., Miyake, Y., 2005. Macular hole R.C., Witt, K.A., 1991. Clinically detectable nerve fiber atrophy
development in fellow eyes of patients with unilateral macular hole. precedes the onset of glaucomatous field loss. Arch. Ophthalmol. 109,
Am. J. Ophthalmol. 140, 370–375. 77–83.
Paunescu, L.A., Schuman, J.S., Price, L.L., Stark, P.C., Beaton, S., Swanson, E.A., Izatt, J.A., Hee, M.R., et al., 1993. In vivo retinal imaging
Ishikawa, H., Wollstein, G., Fujimoto, J.G., 2004. Reproducibility of by optical coherence tomography. Opt. Lett. 18, 1864–1866.
nerve fiber thickness, macular thickness, and optic nerve head Szkulmowski, M., Wojtkowski, M., Bajraszewski, T., Gorczynska, I.,
measurements using Stratus OCT. Invest. Ophthalmol. Vis. Sci. 45, Targowski, P., Wasilewski, W., Kowalczyk, A., Radzewicz, C., 2005.
1716–1724. Quality improvement for high resolution in vivo images by Spectral
Pianta, M.J., Aleman, T.S., Cideciyan, A.V., et al., 2003. In vivo domain Optical Coherence Tomography with supercontinuum source.
micropathology of Best macular dystrophy with optical coherence Opt. Commun., forthcoming.
tomography. Exp. Eye. Res. 76, 203–211. Targowski, P., Gorczynska, I., Szkulmowski, M., Cyganek, M., Wojt-
Pieroth, L., Schuman, J.S., Hertzmark, E., Hee, M.R., Wilkins, J.R., kowski, M., Kowalczyk, A., 2005. Improved complex spectral domain
Coker, J., Mattox, C., Pedut-Kloizman, R., Puliafito, C.A., Fujimoto, OCT for in vivo eye imaging. Opt. Commun., forthcoming.
J.G., Swanson, E., 1999. Evaluation of focal defects of the nerve fiber Targowski, P., Wojtkowski, M., Kowalczyk, A., Bajraszewski, T.,
layer using optical coherence tomography. Ophthalmology 106, Szkulmowski, M., Gorczynska, W., 2004. Complex spectral OCT in
570–579. human eye imaging in vivo. Opt. Commun. 229, 79–84.
ARTICLE IN PRESS
R.A. Costa et al. / Progress in Retinal and Eye Research 25 (2006) 325–353 353

Tornambe, P.E., 2003. Macular hole genesis: the hydration theory. Retina Wojtkowski, M., Srinivasan, V.J., Ko, T.H., Fujimoto, J.G., Kowalevicz,
23, 421–424. A., Duker, J.S., 2004b. Ultrahigh resolution, high speed, Fourier
Uemoto, R., Yamamoto, S., Aoki, T., Tsukahara, I., Yamamoto, T., domain optical coherence tomography and methods for dispersion
Takeuchi, S., 2002. Macular configuration determined by optical coherence compensation. Optics Express 12, 2404–2422.
tomography after idiopathic macular hole surgery with or without internal Wojtkowski, M., Srinivasan, V., Fugimoto, J.G., Ko, T., Schuman, J.S.,
limiting membrane peeling. Br. J. Ophthalmol. 86, 1240–1242. Kowalczyk, A., Duker, J.S., 2005. Three-dimension retinal imaging
Varma, R., Skaf, M., Barron, E., 1996. Retinal nerve fiber layer thickness with high-speed ultrahigh-resolution optical coherence tomography.
in normal human eyes. Ophthalmology 103, 2114–2119. Ophthalmology 112, 1734–1746.
Villate, N., Lee, J.E., Venkatraman, A., Smiddy, W.E., 2005. Photo- Wollstein, G., Schuman, J.S., Price, L.L., Aydin, A., Beaton, S.A., Stark,
receptor layer features in eyes with closed macular holes: optical P.C., Fujimoto, J.G., Ishikawa, H., 2004. Optical coherence tomo-
coherence tomography findings and correlation with visual outcomes. graphy (OCT) macular and peripapillary retinal nerve fiber layer
Am. J. Ophthalmol. 139, 280–289. measurements and automated visual fields. Am. J. Ophthalmol. 138,
Watanabe, M., Oshima, Y., Emi, K., 2000. Optical cross-sectional 218–225.
observation of resolved diabetic macular edema associated with Wollstein, G., Schuman, J.S., Price, L.L., Aydin, A., Stark, P.C.,
vitreomacular separation. Am. J. Ophthalmol. 129, 264–267. Hertzmark, E., Lai, E., Ishikawa, H., Mattox, C., Fujimoto, J.G.,
White, B.R., Pierce, M.C., Nassif, N., Cense, B., Park, B.H., Tearney, Paunescu, L.A., 2005a. Optical coherence tomography longitudinal
G.J., Bouma, B.E., Chen, T.C., de Boer, J.F., 2003. In vivo dynamic evaluation of retinal nerve fiber layer thickness in glaucoma. Arch.
human retinal blood flow imaging using ultra-high-speed spectral Ophthalmol. 123, 464–470.
domain optical Doppler tomography. Opt. Express 11, 3490–3497. Wollstein, G., Ishikawa, H., Wang, J., Beaton, S.A., Schuman, J.S.,
Williams, Z.Y., Schuman, J.S., Gamell, L., Nemi, A., Hertzmark, E., 2005b. Comparison of three optical coherence tomography scanning
Fujimoto, J.G., Mattox, C., Simpson, J., Wollstein, G., 2002. Optical areas for detection of glaucomatous damage. Am. J. Ophthalmol. 139,
coherence tomography measurement of nerve fiber layer thickness and 39–43.
the likelihood of a visual field defect. Am. J. Ophthalmol. 134, Wollstein, G., Paunescu, L.A., Ko, T.H., Fujimoto, J.G., Kowalevicz, A.,
538–546. Hartl, I., Beaton, S., Ishikawa, H., Mattox, C., Singh, O., Duker, J.,
Wojtkowski, M., Kowalczyk, A., Leitgeb, R., Fercher, A.F., 2002a. Full Drexler, W., Schuman, J.S., 2005c. Ultrahigh-resolution optical
range complex spectral optical coherence tomography technique in eye coherence tomography in glaucoma. Ophthalmology 112, 229–237.
imaging. Opt. Lett. 27, 1415–1417. Yamaguchi, Y., Otani, T., Kishi, S., 2003. Resolution of diabetic cystoid
Wojtkowski, M., Leitgeb, R., Kowalczyk, A., Bajraszewski, T., Fercher, macular edema associated with spontaneous vitreofoveal separation.
A.F., 2002b. In vivo human retinal imaging by Fourier domain optical Am. J. Ophthalmol. 135, 116–118.
coherence tomography. J. Biomed. Opt. 7, 457–463. Yun, S., Tearney, G., de Boer, J.F., Iftima, N., Bouma, B., 2003. High-
Wojtkowski, M., Bajraszewski, T., Targowski, P., Kowalczyk, A., 2003. speed optical frequency-domain imaging. Opt. Express 11, 2953–2963.
Real-time in vivo imaging by high-speed spectral optical coherence Zeimer, R., Shahidi, M., Mori, M., Zou, S., Asrani, S., 1996. A new
tomography. Opt. Lett. 28, 1745–1747. method for rapid mapping of the retinal thickness at the posterior pole.
Wojtkowski, M., Bajraszewski, T., Gorczynska, I., Targowski, P., Invest. Ophthalmol. Vis. Sci. 37, 1994–2001.
Kowalczyk, A., Wasilewski, W., Radzewicz, C., 2004a. Ophthalmic Zeimer, R., Mori, M., Khoobehi, B., 1989. Feasibility test of a new
imaging by spectral optical coherence tomography. Am. J. Ophthal- method to measure retinal thickness noninvasively. Invest. Ophthal-
mol. 138, 412–419. mol. Vis. Sci. 30, 2099–2105.

You might also like