Directional light scanning laser ophthalmoscope

7
Directional light scanning laser ophthalmoscope Brian Vohnsen, Ignacio Iglesias, and Pablo Artal Laboratorio de Óptica, Universidad de Murcia, Campus de Espinardo (Edificio C), 30071 Murcia, Spain Received March 4, 2005; revised manuscript received May 10, 2005; accepted May 16, 2005 The cone photoreceptor mosaic of the living human eye has in a limited number of cases been imaged without the use of wavefront-correction techniques. To accomplish this, the directionality of the photoreceptors, as manifested by their waveguiding properties, may be used to advantage. In the present paper we provide a model of our recently proposed directional light scanning laser ophthalmoscope [Opt. Lett. 29, 968 (2004)] to- gether with a detailed numerical analysis of the device. The outcome is compared with experimental results. © 2005 Optical Society of America OCIS codes: 170.4470, 170.5810, 330.5310. 1. INTRODUCTION The optical performance of the normal human eye dete- riorates at large pupil diameters owing to the presence of aberrations. Aberrations also limit the resolution that can be attained with ocular imaging devices. Thus, to get be- yond this limitation the aberrations need to be compen- sated for. The introduction of adaptive optics in studies of the eye 1–4 has allowed for correction of even higher-order aberrations in essentially real time and thereby has led to the acquisition of high-resolution retinal images. Never- theless, these results are somewhat contradicted by re- ported cone photoreceptor mosaic images obtained in the living eye without the aid of wavefront-correction methods. 5–7 The Stiles–Crawford effect refers to the reduced visual sensitivity to light that is incident off the pupil center and therefore obliquely incident on the retina. 8,9 A similar phenomenon occurs for light reflected off the retina 10–13 where it is typically known as the optical Stiles–Crawford effect. Based on this observation, we recently introduced a confocal scanning laser ophthalmoscope (c-SLO) designed to make use of the directionality of reflected light in high- resolution retinal imaging. 7 This technique relies on an accurate alignment of the incident beam (chosen suffi- ciently narrow to reduce the influence of aberrations) at the pupil point where its visibility is maximized, i.e., at the peak of the Stiles–Crawford effect. In this way the coupling of light at the retina to the underlying photore- ceptors is maximized (allowing for the largest signal). 10,14 At the confocal detection a sufficiently small pinhole is chosen to highlight the contribution of directional backre- flected light produced by the waveguiding of the photoreceptors. 7 The photoreceptors permit waveguiding with negligible absorption if bleached or if near-IR illumi- nation is used, and a major fraction of the returning light can therefore be directional. 11,15 Generally, this light has suffered less from aberrations than broadly scattered light since it exits the eye mostly through a fraction of the entire pupil area. 11 A valid alter- native would be a reduced exit pupil at the eye although this would presumably require additional transverse alignment to be properly centered at the peak of the (op- tical) Stiles–Crawford effect (which is often displaced from the apparent pupil center). 8,12,13 The use of a small confocal pinhole has the additional advantage of high- lighting the contribution of scattering sources (both guided and nonguided light) located in the plane of the retina conjugate to the pinhole while discriminating against scattered light from elsewhere that could deterio- rate image quality. In this paper we make a numerical analysis of the im- aging capabilities of the technique in order to gain further insight and, in particular, to examine under what circum- stances the c-SLO may be used to directly access param- eters of the photoreceptor mosaic even in the absence of wavefront correction. The model will be used on light that originates in photoreceptor waveguiding (as from bleached or nonabsorbing cones), but it should be stressed that the same approach can easily be generalized to handle imaging of arbitrary scattering sources (not only photoreceptor waveguides) by a simple modification of their pupil field. The paper is organized as follows. In Sec- tion 2 a model of the directional light c-SLO is described. Section 3 contains a numerical analysis in which the im- portance of various physical parameters (pinhole size, photoreceptor diameter, etc.) is studied in order to eluci- date their influence on high-resolution retinal images, and the outcome is compared with experimental results. In Section 4 we present our conclusions. 2. MODEL OF THE DIRECTIONAL LIGHT SCANNING LASER OPHTHALMOSCOPE In this section the model constructed to analyze our re- cently proposed directional light c-SLO is discussed. 7 The description includes only directional light produced by cone photoreceptor waveguiding; consequently scattering from all other retinal structures has been omitted. This assumption is justified by the often large contribution of directional light compared with broadly scattered light 11,15 and by the selectivity of the confocal pinhole. 7 On the basis of a simulation of the optical Stiles– Crawford effect, and in agreement with the findings of others, 16 we have recently found that the light reflected 2606 J. Opt. Soc. Am. A/Vol. 22, No. 12/December 2005 Vohnsen et al. 1084-7529/05/122606-7/$15.00 © 2005 Optical Society of America

Transcript of Directional light scanning laser ophthalmoscope

1Trabystattplm

stpwectracttccAcflpwnc

ttnta

2606 J. Opt. Soc. Am. A/Vol. 22, No. 12 /December 2005 Vohnsen et al.

Directional light scanning laser ophthalmoscope

Brian Vohnsen, Ignacio Iglesias, and Pablo Artal

Laboratorio de Óptica, Universidad de Murcia, Campus de Espinardo (Edificio C), 30071 Murcia, Spain

Received March 4, 2005; revised manuscript received May 10, 2005; accepted May 16, 2005

The cone photoreceptor mosaic of the living human eye has in a limited number of cases been imaged withoutthe use of wavefront-correction techniques. To accomplish this, the directionality of the photoreceptors, asmanifested by their waveguiding properties, may be used to advantage. In the present paper we provide amodel of our recently proposed directional light scanning laser ophthalmoscope [Opt. Lett. 29, 968 (2004)] to-gether with a detailed numerical analysis of the device. The outcome is compared with experimental results.© 2005 Optical Society of America

OCIS codes: 170.4470, 170.5810, 330.5310.

tfclgrar

aisewobthpttSppdaI

2SIcdcfadl

Co

. INTRODUCTIONhe optical performance of the normal human eye dete-iorates at large pupil diameters owing to the presence ofberrations. Aberrations also limit the resolution that cane attained with ocular imaging devices. Thus, to get be-ond this limitation the aberrations need to be compen-ated for. The introduction of adaptive optics in studies ofhe eye1–4 has allowed for correction of even higher-orderberrations in essentially real time and thereby has led tohe acquisition of high-resolution retinal images. Never-heless, these results are somewhat contradicted by re-orted cone photoreceptor mosaic images obtained in theiving eye without the aid of wavefront-correction

ethods.5–7

The Stiles–Crawford effect refers to the reduced visualensitivity to light that is incident off the pupil center andherefore obliquely incident on the retina.8,9 A similarhenomenon occurs for light reflected off the retina10–13

here it is typically known as the optical Stiles–Crawfordffect. Based on this observation, we recently introduced aonfocal scanning laser ophthalmoscope (c-SLO) designedo make use of the directionality of reflected light in high-esolution retinal imaging.7 This technique relies on anccurate alignment of the incident beam (chosen suffi-iently narrow to reduce the influence of aberrations) athe pupil point where its visibility is maximized, i.e., athe peak of the Stiles–Crawford effect. In this way theoupling of light at the retina to the underlying photore-eptors is maximized (allowing for the largest signal).10,14

t the confocal detection a sufficiently small pinhole ishosen to highlight the contribution of directional backre-ected light produced by the waveguiding of thehotoreceptors.7 The photoreceptors permit waveguidingith negligible absorption if bleached or if near-IR illumi-ation is used, and a major fraction of the returning lightan therefore be directional.11,15

Generally, this light has suffered less from aberrationshan broadly scattered light since it exits the eye mostlyhrough a fraction of the entire pupil area.11 A valid alter-ative would be a reduced exit pupil at the eye althoughhis would presumably require additional transverselignment to be properly centered at the peak of the (op-

1084-7529/05/122606-7/$15.00 © 2

ical) Stiles–Crawford effect (which is often displacedrom the apparent pupil center).8,12,13 The use of a smallonfocal pinhole has the additional advantage of high-ighting the contribution of scattering sources (bothuided and nonguided light) located in the plane of theetina conjugate to the pinhole while discriminatinggainst scattered light from elsewhere that could deterio-ate image quality.

In this paper we make a numerical analysis of the im-ging capabilities of the technique in order to gain furthernsight and, in particular, to examine under what circum-tances the c-SLO may be used to directly access param-ters of the photoreceptor mosaic even in the absence ofavefront correction. The model will be used on light thatriginates in photoreceptor waveguiding (as fromleached or nonabsorbing cones), but it should be stressedhat the same approach can easily be generalized toandle imaging of arbitrary scattering sources (not onlyhotoreceptor waveguides) by a simple modification ofheir pupil field. The paper is organized as follows. In Sec-ion 2 a model of the directional light c-SLO is described.ection 3 contains a numerical analysis in which the im-ortance of various physical parameters (pinhole size,hotoreceptor diameter, etc.) is studied in order to eluci-ate their influence on high-resolution retinal images,nd the outcome is compared with experimental results.n Section 4 we present our conclusions.

. MODEL OF THE DIRECTIONAL LIGHTCANNING LASER OPHTHALMOSCOPE

n this section the model constructed to analyze our re-ently proposed directional light c-SLO is discussed.7 Theescription includes only directional light produced byone photoreceptor waveguiding; consequently scatteringrom all other retinal structures has been omitted. Thisssumption is justified by the often large contribution ofirectional light compared with broadly scatteredight11,15 and by the selectivity of the confocal pinhole.7

On the basis of a simulation of the optical Stiles–rawford effect, and in agreement with the findings ofthers,16 we have recently found that the light reflected

005 Optical Society of America

ffTttwieosdtudtt

ifsttbcctti

fboesciit

swa

Npfi

wtwmtwmdjattiespdtrtoac

itrtatccppfifpitsatpia

pp

Fwtdf2a4

Vohnsen et al. Vol. 22, No. 12 /December 2005 /J. Opt. Soc. Am. A 2607

rom each cone photoreceptor may be dominated by theundamental waveguide mode of each inner segment.14

his mode can be well approximated by a Gaussian func-ion of the type exp�−r2 /wm

2 �, where r is the radial dis-ance from the photoreceptor axis and wm is the half-idth of the mode (which is related to the diameter of the

nner segment). Thus, in the following we assume thatach illuminated photoreceptor radiates a Gaussian beamf light directed toward the center of the pupil17 as shownchematically in Fig. 1. It may be noted that the intensityistribution produced at the pupil for a single photorecep-or is Gaussian, too, in agreement with the commonlysed model for the (optical) Stiles–Crawford effect when airectionality parameter is sought.14 A shift away fromhe pupil center can easily be incorporated by translatinghe Gaussian distribution in the pupil plane.

For simplicity, the incident beam scanning of the c-SLOs omitted at first. Instead, an image is considered to beormed by scanning with the confocal pinhole across ateady image plane (where all contributing photorecep-ors are brought to a focus) while recording the transmit-ed light power versus position. Leaving out the incidenteam scanning simplifies the analysis of the imaging pro-ess and corresponds to the realistic case of a poorly fo-used beam that simultaneously illuminates several pho-oreceptors. Nevertheless, it should be kept in mind thathe dual scanning of an actual confocal system can lead tomproved image resolution and contrast.18

Wavefront aberrations too will have a detrimental ef-ect on the quality of recorded images but may be reducedy either customized phase plates19 or adaptiveptics.1,2,4,20 The wavefront aberrations �WA of the realye and system, as measured with a Hartmann–Shackensor in a plane conjugate to the eye pupil,21 can be in-luded in the model by introducing them on each Gauss-an beam. It may be assumed that each beam suffersdentical aberrations (apart from the tilt that separateshem) when scanning only a small retinal area. A similar

ig. 1. Schematic of the model used to describe c-SLO imagingith directional light detection including Gaussian beam radia-

ion from each cone photoreceptor (C), confocal pinhole (P), andetector (D). Also shown is (a) point-spread function calculatedrom the (b) measured wavefront aberration of an eye shown on a� phase map (6.0 mm pupil at a wavelength of 0.785 �m). Theberration, although dominated by astigmatism, includes up toth-order Zernike polynomials with a total RMS value of0.67 �m.

ituation arises in experiments with wavefront correctionhere an average is used during the rapid imagecquisition.1,2,4,20

Thus, the field in the plane of the confocal pinhole fromsimultaneously illuminated photoreceptors can be ex-

ressed via the Fourier transform of the (scaled) pupileld as

F��pupil� = �j=1

N

F�A exp�i�WA�x,y� −x2 + y2

wp2 � � ��r − rj�,

�1�

here � denotes a convolution. In this expression wp ishe spot size at the pupil of the reflected Gaussian beamhich at wavelength � is related to the photoreceptorode through wp=�feye/�neyewm [i.e., the standard rela-

ion between pupil field width and focal spot size (here theaveguide mode) for a Gaussian beam in a homogenousedium]. Here feye=22.2 mm and neye=1.33 in the re-

uced eye model. In Eq. (1) the unaberrated image of theth photoreceptor is centered at rj in the pinhole plane,nd the Fourier transform is carried out at frequencieshat take into account the system magnification as well ashe focal length of the lens nearest the pinhole. Althought has been written explicitly for Gaussian beams, thequation may also be used to simulate imaging of retinaltructures other than photoreceptor waveguides by sim-ly replacing the Gaussian amplitude weighting with aifferent pupil field (a uniform pupil would be represen-ative, e.g., of a pointlike scattering object located in theetinal plane, and out-of-focus features can be consideredhrough their phase variation at the pupil). Thus, withnly small modifications the model would be suitable tonalyze imaging of retinal structures other than photore-eptor waveguides.

In the present situation, the best possible cone mosaicmage will show an extended bright spot (correspondingo the mode width) at the location of each imaged photo-eceptor. This differs somewhat from the common situa-ion in confocal microscopy where a point object is imageds a spot because of the finite size of the numerical aper-ure (here the waveguide objects have a finite width). Foronvenience of modeling, it has been assumed that allontributing photoreceptors radiate with equal power andhase as contained in the amplitude factor A. This sim-lification can easily be undone by inclusion of a complexeld scaling factor at each rj to take into account, e.g., dif-erences in the phase of beams radiated by neighboringhotoreceptors.14,22 It should be mentioned, though, thatf the exposed retinal area is very small, as in a c-SLO,he simultaneously illuminated photoreceptors are highlyimilar,23 and differences in phase of the reradiated lightre therefore expected to be small. In contrast, for an ex-ended retinal area, with numerous photoreceptors ex-osed to light in unison (as in fundus photography), thiss presumably no longer the case and the phase may play

more important role.14,22

The image I�r� obtained by scanning with the confocalinhole P�r� can (apart from a scaling constant) be ex-ressed as

AainntflptVpvpbtcfr

sttpiptatpeseacre

wdttwesimtitctp

qopf

ewphscteisd

3CTcr2t�mst2=bcmprmh

Ffr1hew

2608 J. Opt. Soc. Am. A/Vol. 22, No. 12 /December 2005 Vohnsen et al.

I�r� = �F��pupil��2 � P�r�. �2�

lthough we have assumed that all photoreceptors returnn identical amount of light this can be expected to varyn a real experiment. This may result in an inhomoge-eous distribution of brightness that in the presence ofoise can make the photoreceptor density seem lowerhan it actually is. This can occur if the photoreceptor re-ectivity changes (spatially or temporally), or if the cou-ling efficiency varies as a result of nonuniform illumina-ion, photoreceptor tilt, or the scanning in itself.ariations should also be expected when different sizedhotoreceptors are scanned. Actually, time-dependentariations in the amount of reflected light from individualhotoreceptors have been reported on a long time scale,15

ut this factor has hardly a measurable influence duringhe rapid acquisition of a single c-SLO image. Reflectivityhanges, however, may influence subsequently recordedrames (in our present experimental system the frameate is 15 Hz).

The incident beam scanning can be modeled by inclu-ion of a position-dependent parameter T�u�, where u ishe displacement at the retina of the incident beam fromhe photoreceptor axis, and in a first approximation by re-lacing the pinhole in Eq. (2) with T�u�P�r�. This takesnto account changes in the photoreceptor-to-light cou-ling as the focused beam sweeps across each photorecep-or aperture during the raster scan [ideally, T�u� shouldct on the object before the convolution with the pinholehereby increasing the complexity of the model]. This ap-roximation is most valid when the beam illuminates sev-ral receptors simultaneously. If focused tightly, to a spotize comparable with the photoreceptor spacing, a differ-nt modeling would be required as the light may couple tosingle photoreceptor at a time. If aberrations of the in-

ident light can be ignored, as may be the case for a nar-ow beam, a position-dependent coupling efficiency can bestimated as14

T�u� = � 2wrwm

wr2 + wm

2 �2

exp� − 2u2

wr2 + wm

2 � , �3�

here wr is the focused spot size at the retina of an inci-ent Gaussian beam. Equation (3) results from a calcula-ion of the overlap integral between the incident ampli-ude field and the fundamental mode of the photoreceptoraveguide. Inclusion of this position-dependent param-ter can be expected to reduce the blur of each cone inimulated images, although a better estimate should alsonclude the aberrations of the incident beam. It should be

entioned that the retinal object may be extracted fromhe resulting image by inverting Eq. (2) and removing thenfluence of measured aberrations. However, because ofhe low-pass filtering of images by the finite-sized confo-al pinhole, noise in the recorded data, and the uncer-ainty in measurements of �WA, this is not easily accom-lished.In summary, to resolve the photoreceptor mosaic re-

uires obviously either a pinhole smaller than the spacingf the imaged receptors, or alternatively that only a singlehotoreceptor is exposed to light at a time as by a tightlyocused (aberration-corrected) incident beam.4 The influ-

nce of T�u� on the recorded image quality may, togetherith a faster acquisition, plausibly explain why the ap-arently best high-resolution cone mosaic images to dateave been obtained with flood illumination systems,1,2

ince in this case T�u� may be considered approximatelyonstant. In a scanning system, its influence may changehe amount of coupled light (and thereby also the signal)ach time a given photoreceptor is scanned. In the follow-ng, we examine numerically the possibilities of cone mo-aic imaging with a tiny pinhole in the c-SLO tailored toirectional-light detection.

. NUMERICAL RESULTS ANDOMPARISON WITH EXPERIMENTShe actual c-SLO has been calibrated by scanning aopper-on-glass mesh of 63.5 �m period through an ach-omatic lens (25 mm focal length). A conversion factor of80 �m/degree has been used for the eye, and based onhis calibration all of the following images cover 140140 �m at the retina (with 256�256 pixels). Experi-entally obtained results for different pinhole sizes are

hown in Fig. 2. To reduce the influence of aberrations onhe incident beam its diameter has been restricted to.0 mm at the eye pupil. For the near-IR wavelength �0.785 �m used in the experiments, this corresponds atest to an �8 �m bright spot on the retina. With a suffi-iently small pinhole the presence of the photoreceptorosaic becomes apparent in the recorded images as re-

orted previously.7 To simulate the situation, cone photo-eceptors have been distributed in a hexagonal arrange-ent with added random jitter (each one bounded in aexagonal area to avoid overlapping cones) thereby re-

ig. 2. Experimental images (correlation of 6–8 consecutiverames) obtained with a 0.785 �m wavelength laser diode in theight eye at �2° nasal with different-sized pinholes: (a) 200, (b)00, (c) 50, (d) 30 �m. In (d) a small signal-to-noise ratio hindersigh image quality. Ocular motion prevents repeated recording ofxactly the same retinal location and images are therefore some-hat shifted with respect to each other.

sb

(=iaccFfGeswts

csd

astntlpcieiooaTotsi

tdFrbrs

F11tt(tao

F2

Vohnsen et al. Vol. 22, No. 12 /December 2005 /J. Opt. Soc. Am. A 2609

embling the in vitro photoreceptor distributions studiedy Curcio et al.23

Figures 3 and 4 show simulated distributions of conesGaussian beam intensity images) for densities �10,000/mm2 and 20,000/mm2, respectively (correspond-

ng to typical cone spacings, lc=1.075/�1/2, of �10.75 �mnd �7.60 �m), and simulated c-SLO images for differentonfocal pinholes. The wavefront aberration used in thealculation is the experimentally recorded one shown inig. 1. To match the situation at the periphery of the

ovea we have chosen wm=1.50 �m for the quasi-aussian mode (corresponding to an inner segment diam-ter of approximately 3.6 �m14). This mode results in apot size of wp=2.78 mm at the pupil for the near-IRavelength used. A smaller width at the pupil can be ob-

ained with a larger wm (i.e., for wider cones) or for ahorter wavelength of illumination.

From the results obtained it is apparent that a suffi-iently small pinhole (comparable to the photoreceptorpacing when imaged onto the retina) can reveal the un-erlying photoreceptor mosaic as suggested by Eq. (2) and

ig. 3. Simulated (a) retinal section for a cone density of0,000/mm2 and corresponding c-SLO images for pinholes of (b)00, (c) 50, (d) 30 �m. The imaged retina is 6-times magnified inhe plane of the pinhole. Corresponding two-dimensional ampli-ude spectra of (a) and (d) are shown in (e) and (f), respectively,to enhance contrast of the hexagonal frequency pattern the cen-ral part of each spectrum has been suppressed). The wavefrontberration used in the simulation is the experimentally obtainedne shown in Fig. 1(b).

s verified also by the spectral images. The larger pinholemears out any detailed structure with only some con-rast produced by the variation in the number of simulta-eously scanned cones (as contained within the area ofhe pinhole) and by interference among the contributingight beams. In reality, the simulated case of the largeinhole is perchance somewhat exaggerated since the in-ident beam hardly couples equally well to all contribut-ng photoreceptors, thereby violating the assumption ofqual radiation from all (see Fig. 6 below). As the pinholes reduced the image resembles more closely the structuref the original object, but the relationship is not a simplene because of the overlap of light from neighboring conesnd the loss in resolution due to the finite pinhole size.hus, only in some cases can bright spots be unambigu-usly identified with the presence of a corresponding pho-oreceptor. Moreover, the greater the photoreceptor den-ity the harder it becomes to identify contributions fromndividual receptors.

To examine the image-deteriorating effect of aberra-ions more closely, the process has been simulated withifferent scalings of �WA producing the results shown inig. 5. It can be seen that even with all aberrations cor-ected, the image differs from the object because of thelurring by the pinhole [cf. Fig. 5(d) and Fig. 3(a)]. Theesulting resolution loss prevents the distinction of closelypaced photoreceptors although an increased brightness

ig. 4. Simulated (a) retinal section for a cone density of0,000/mm2; otherwise, details are the same as in Fig. 3.

(Ibcbcs

t2

(G(tiwcip

lv

F1wFt

Fs

F35aaw3e

2610 J. Opt. Soc. Am. A/Vol. 22, No. 12 /December 2005 Vohnsen et al.

when in phase) indicates the presence of more than one.f the phase of each Gaussian beam is allowed to vary ar-itrarily, which may be the case if many photoreceptorsontribute at a time with a poorly focused incidenteam,14 the image appearance does not change signifi-antly apart from some redistribution of brightness (forimplicity not shown here).

We have also examined the consequences of includinghe incident beam scanning in the model by taking a.0 mm wide (untruncated) incident beam at the eye pupil

ig. 5. Simulated c-SLO images for a cone density of0,000/mm2 and a 30 �m pinhole for different amounts of theavefront aberration �WA shown in Fig. 1(b): (a) �WA same as inig. 3(d), (b) �WA/2, (c) �WA/4, (d) no aberrations (correspondingo a completely corrected wavefront).

ig. 6. Same as Fig. 3 but with the inclusion of incident beamcanning. For simplicity the spectral images are not shown.

corresponding to wr�4.2 �m), and letting it act as aaussian weighting of the pinhole in accordance with Eq.

3). The results obtained (Fig. 6) show a sharpening of de-ails, but perhaps most apparent when compared to Fig. 3s the absence of circular structures (pinhole images)hen larger pinholes are used. Thus, the situation comes

loser to the experimentally realized one, although propernclusion of the incident beam aberrations in T�u� is ex-ected to produce an even greater resemblance.Finally, the influence of the chosen illumination wave-

ength has been considered. Near-IR has some clear ad-antages over the visible in the context of imaging apart

ig. 7. Simulated c-SLO images (b), (c), (d) and (f), (g), (h) with0 �m pinhole for objects (a) 10,000/mm2 with wm=1.50 �m, (e)000/mm2 with wm=2.00 �m, respectively. Images of the objectsre shown at wavelengths of (b), (f) 0.785 �m; (c), (g) 0.543 �m,ll with the aberration map shown in Fig. 3(b); (d), (h) 0.543 �mithout aberrations. Note that (a) and (b) are identical to Fig.(a) and Fig. 3(d), respectively, but have been reproduced here forase of comparison.

fwllpatlrola7cbqoHili

4Iccfbbdmfvsfsiwettftaamet

ATfBIR

i

R

1

1

1

1

1

1

1

1

1

1

2

2

2

2

2

Vohnsen et al. Vol. 22, No. 12 /December 2005 /J. Opt. Soc. Am. A 2611

rom the comfort of the subject: More light can be usedithout compromising safety, and a larger fraction of

ight will travel the full photoreceptor length as a result ofow absorption in the outer segment thereby not requiringrebleaching. However, the width of the directional lightt the pupil scales with the wavelength squared24 and isherefore more prone to suffer aberrations at longer wave-engths. Moreover, the overall amount of reflected (nondi-ectional) light increases with wavelength25,26 making themission of a uniform background at the pupil more prob-ematic in the present model. The outcome of simulationst two different illumination wavelengths is shown in Fig.(again leaving out the incident beam scanning). For

omparison also a low-density photoreceptor section haseen simulated (with slightly wider cones). The imageuality is slightly better at the lower wavelength becausef the smaller influence of aberrations as discussed above.owever, the improvement is not dramatic and again the

maged cones cannot be clearly distinguished unless theirocal density is low or, alternatively, wavefront corrections used together with a tiny pinhole.

. CONCLUSIONSn this paper, our recently proposed directional light-SLO7 has been examined in further detail with numeri-al simulations of cone-mosaic images obtained under dif-erent conditions. Since the photoreceptors are known toe wave-guiding structures, the description is formulatedy considering each photoreceptor to radiate indepen-ently a Gaussian beam (resembling the fundamentalode of the inner segment) when exposed to light. We

ound that the results agree qualitatively well with obser-ations made in the real eye confirming that although amall pinhole can reveal the photoreceptor mosaic off theovea it requires wavefront correction to be reliably re-olved when the density is large. For simplicity of model-ng the focused beam on the retina has been assumedider than the intercone distance, thereby exposing sev-ral photoreceptors at a time. In experiments, the situa-ion can be expected to improve since the small width ofhe incident beam, together with any fine structure in theocus, may act as a resolution-enhancing filter by favoringhe coupling of light to a single photoreceptor at a timend thereby highlighting its contribution. Although suchsituation is not controllable for accurate photoreceptorosaic imaging, it may suffice to reveal areas of misori-

nted photoreceptors27 as well as other retinal abnormali-ies.

CKNOWLEDGMENTShe authors gratefully acknowledge financial support

rom the Ministerio de Educación y Ciencia (grantFM2001-0391 and FIS2004-02153), Spanish Fondo de

nvestigación Sanitaria (grant red IM3 G03/185), andamón y Cajal research contract RYC2002-006337.

Corresponding author Brian Vohnsen’s e-mail addresss [email protected].

EFERENCES1. J. Liang, D. R. Williams, and D. T. Miller, “Supernormal

vision and high-resolution retinal imaging throughadaptive optics,” J. Opt. Soc. Am. A 14, 2884–2892(1997).

2. A. Roorda and D. R. Williams, “The arrangement of thethree cone classes in the living human eye,” Nature(London) 397, 520–522 (1999).

3. E. J. Fernández, I. Iglesias, and P. Artal, “Closed-loopadaptive optics in the human eye,” Opt. Lett. 26, 746–748(2001).

4. A. Roorda, F. Romero-Borja, W. J. Donnelly III, H. Queener,T. J. Hebert, and M. C. W. Campbell, “Adaptive opticsscanning laser ophthalmoscopy,” Opt. Express 10, 405–412(2002).

5. A. R. Wade and F. W. Fitzke, “In vivo imaging of the humancone-photoreceptor mosaic using a confocal laser scanningophthalmoscope,” Lasers Light 8, 129–136 (1998).

6. A. R. Wade and F. W. Fitzke, “A fast, robust patternrecognition system for low light level image registrationand its application to retinal imaging,” Opt. Express 3,190–197 (1998).

7. B. Vohnsen, I. Iglesias, and P. Artal, “Directional imagingof the retinal cone mosaic,” Opt. Lett. 29, 968–970 (2004).

8. W. S. Stiles and B. H. Crawford, “The luminous efficiency ofrays entering the eye pupil at different points,” Proc. R.Soc., London, Ser. B 112, 428–450 (1933).

9. A. W. Snyder and C. Pask, “The Stiles–Crawford effect-explanation and consequences,” Vision Res. 13, 1115–1137(1973).

0. P. J. Delint, T. T. J. M. Berendschot, and D. van Norren,“Local photoreceptor alignment measured with a scanninglaser ophthalmoscope,” Vision Res. 37, 243–248 (1996).

1. S. A. Burns, S. Wu, F. Delori, and A. E. Elsner, “Directmeasurement of human-cone-photoreceptor alignment,” J.Opt. Soc. Am. A 12, 2329–2338 (1995).

2. S. A. Burns, S. Wu, J. C. He, and A. E. Elsner, “Variationsin photoreceptor directionality across the central retina,” J.Opt. Soc. Am. A 14, 2033–2040 (1997).

3. J. C. He, S. Marcos, and S. A. Burns, “Comparison of conedirectionality determined by psychophysical andreflectometric techniques,” J. Opt. Soc. Am. A 16,2363–2369 (1999).

4. B. Vohnsen, I. Iglesias, and P. Artal, “Guided light anddiffraction model of human-eye photoreceptors,” J. Opt.Soc. Am. A 22, 2318–2328 (2005).

5. A. Pallikaris, D. R. Williams, and H. Hofer, “Thereflectance of single cones in the living human eye,” Invest.Ophthalmol. Visual Sci. 44, 4580–4592 (2003).

6. J.-M. Gorrand and F. C. Delori, “A model for assessment ofcone directionality,” J. Mod. Opt. 44, 473–491 (1997).

7. A. M. Laties and J. M. Enoch, “An analysis of retinalreceptor orientation,” Invest. Ophthalmol. Visual Sci. 10,69–77 (1971).

8. R. H. Webb, G. W. Hughes, and F. C. Delori, “Confocalscanning laser ophthalmoscope,” Appl. Opt. 26, 1492–1499(1987).

9. S. A. Burns, S. Marcos, A. E. Elsner, and S. Bara, “Contrastimprovement of confocal retinal imaging by use of phase-correcting plates,” Opt. Lett. 27, 400–402 (2002).

0. B. Vohnsen, I. Iglesias, and P. Artal, “Confocal scanninglaser ophthalmoscope with adaptive optical wavefrontcorrection,” in Proc. SPIE 4964, 24–32 (2003).

1. P. M. Prieto, F. Vargas-Martín, S. Goelz, and P. Artal,“Analysis of the performance of the Hartmann–Shacksensor in the human eye,” J. Opt. Soc. Am. A 17, 1388–1398(2000).

2. S. Marcos, S. A. Burns, and J. C. He, “Model for conedirectionality reflectometric measurements based onscattering,” J. Opt. Soc. Am. A 15, 2012–2022 (1998).

3. C. A. Curcio, K. R. Sloan, R. E. Kalina, and A. E.Hendrickson, “Human photoreceptor topography,” J. Comp.Neurol. 292, 497–523 (1990).

4. N. P. A. Zagers, T. T. J. M. Berendschot, and D. van Norren,“Wavelength dependence of reflectometric cone

2

2

2

2612 J. Opt. Soc. Am. A/Vol. 22, No. 12 /December 2005 Vohnsen et al.

photoreceptor directionality,” J. Opt. Soc. Am. A 20, 18–23(2003).

5. N. P. A. Zagers, J. van de Kraats, T. T. J. M. Berendschot,and D. van Norren, “Simultaneous measurement of fovealspectral reflectance and cone-photoreceptor directionality,”Appl. Opt. 41, 4686–4696 (2002).

6. J. J. Vos, A. A. Munnik, and J. Boogaard, “Absolute

spectral reflectance of fundus oculi,” J. Opt. Soc. Am. 55,573–574 (1965).

7. M. Kono, J. M. Enoch, E. Strada, P. Shih, R. Srinivasan, V.Lakshminarayanan, W. Susilasate, and A. Graham,“Stiles–Crawford effect of the first kind: assessment ofphotoreceptor alignments following dark patching,” Vision

Res. 41, 103–118 (2001).