27.01.2014 Views

Alfredo Dubra's PhD thesis - Imperial College London

Alfredo Dubra's PhD thesis - Imperial College London

Alfredo Dubra's PhD thesis - Imperial College London

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

A Shearing Interferometer for<br />

the Evaluation of Human<br />

Tear Film Topography<br />

<strong>Alfredo</strong> Dubra Suárez<br />

Submitted in partial fulfilment of the requirements for the degree of Doctor of<br />

Philosophy


Abstract<br />

The evaluation of the optical quality of the human eye beyond spectacle prescription<br />

has recently become a very active field, given the number and importance of<br />

applications that can benefit from it: refractive surgery, retinal imaging, monitoring<br />

of intraocular lens performance, psychophysical experiments and adaptive optics<br />

spectacles for improving visual acuity and increasing accommodation range in presbyopic<br />

subjects. Several ophthalmic wavefront sensors have been recently developed<br />

to measure the imaging properties of the eye, most of them based on the use of a<br />

Shack-Hartmann configuration.<br />

Despite the current widespread use of wavefront sensing based refractive surgery, there<br />

are a number of related issues not yet well understood, one of the most important<br />

being the variability of the wavefront sensor measurements due to eye movements,<br />

heart beat, fluctuations of accommodation and tear film dynamics.<br />

The work in this <strong>thesis</strong> focuses on the study of tear film dynamics and its effects on<br />

the optical quality of the eye. In order to do so, a lateral shearing interferometer to<br />

measure the pre-corneal tear film topography was designed, built and used to study<br />

20 subjects. The collected data was processed with a novel reconstruction technique,<br />

which is applicable to any lateral shearing interferometry experiment, and the results<br />

are discussed in terms of the tear film influence of the optical quality of the eye, which<br />

can be used to understand the effects on vision, wavefront sensing and adaptive optics.<br />

Finally, the feasibility of using interferometry to evaluate the wavefront error of the<br />

eye by modifying the lateral shearing interferometer was also tested.<br />

1


Acknowledgements<br />

Pursuing the <strong>PhD</strong> degree at <strong>Imperial</strong> has been a life changing experience both inside<br />

and outside the lab, and for this, I have to thank a few people.<br />

To begin with, thanks Paula for your support in the mad idea of coming to <strong>London</strong><br />

to start a <strong>PhD</strong> with no funding. I then have to say that I am greatly indebted to<br />

my supervisor Chris Dainty for getting this funding (and Luis Diaz-Santana) and<br />

providing the academic freedom and confidence in the project. I am also grateful<br />

to Chris for the work environment he created both in the Applied Optics groups at<br />

<strong>Imperial</strong> and NUI Galway.<br />

Much of the research undertaken for this <strong>thesis</strong> was the result of several discussions<br />

with Carl Paterson, even before he was my supervisor.<br />

The experiments performed during the <strong>PhD</strong> were only possible thanks to the patience<br />

of and quality of work from, the guys in the mechanical and optics workshop: Paul<br />

Brown, Martin Kehoe, Simon Johnson and Martin Dowman.<br />

Thanks to all the people that volunteered to put their eyes in front of a laser beam and<br />

have their mouths stuck in a wax block for nothing in return. That is true generosity.<br />

It is said that one should not mix business with pleasure, but the most valuable thing<br />

I will take from my <strong>PhD</strong> is the friends I made at work; thanks to Fred Reavell for his<br />

immense patience with my poor engineering skills, the explanation of English culture<br />

and motorcycle tuition. Being in a lab with no windows and having to deal with the<br />

same project for three years (and a series of unsuccessful experiments) could have<br />

driven me insane, had it not been for the company of David Lara, Steve Gruppetta,<br />

Jonathan Brooks, Karen Hampson, Gordon Kennedy, James McIlroy, Ian Munro,<br />

David Catlin, Simon Clay and Allison Craig.<br />

There have been some seriously hard times on a personal level during the <strong>PhD</strong>, which<br />

I only came through with the company and support from Mo and Si. Thanks to them<br />

I came closer to the true English culture (and slang), sometimes difficult to find in<br />

such a cosmopolitan place and more importantly, I found two friends for life.<br />

There was a time when working long hours in front of the computer and a lonely lab<br />

made me lose all perspective of life. Fortunately, Chris Dainty took me to Galway for<br />

2


a few months, where I found it (maybe for the first time). This only happened thanks<br />

to the friendship of David Lara, Andrew Cronin, Larry Murray and David Merino.<br />

Thanks to them for repeatedly telling me what is really important over and over again<br />

(yes, sometimes I can be a bit stubborn).<br />

I also have to thank to all my family for making me feel supported all the way through,<br />

despite the physical distance.<br />

Finally, thanks to Cecilia, for reminding me that there is life outside the lab.<br />

Alf<br />

<strong>London</strong>, June 2004<br />

3


Contents<br />

Contents 4<br />

List of Figures 8<br />

1 Introduction 12<br />

1.1 Wavefront sensing in the human eye . . . . . . . . . . . . . . . . . . . 14<br />

1.1.1 Applications of wavefront sensing in the human eye . . . . . . . 17<br />

1.1.2 Repeatability in wavefront sensing measurements . . . . . . . . 18<br />

1.2 Tear film . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

1.3 Thesis synopsis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

2 Lateral shearing interferometer design 24<br />

2.1 System requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.2 Lateral shearing interferometry . . . . . . . . . . . . . . . . . . . . . . 25<br />

2.3 Interferogram intensity modulation . . . . . . . . . . . . . . . . . . . . 26<br />

2.4 Wedge design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.5 Wedge manufacture . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.6 Illumination branch . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

2.6.1 Source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

2.6.2 Spatial filter and telescope . . . . . . . . . . . . . . . . . . . . . 32<br />

2.6.3 Polarizing beam splitter and quarter waveplate . . . . . . . . . 33<br />

2.6.4 Focusing lens . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

2.7 Imaging branch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

2.7.1 Shear and tilt estimation . . . . . . . . . . . . . . . . . . . . . 39<br />

2.7.2 Polarisation control for optimization of interferogram contrast . 39<br />

3 Data processing 42<br />

3.1 Carrier frequency estimation . . . . . . . . . . . . . . . . . . . . . . . 42<br />

4


CONTENTS<br />

3.1.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

3.1.2 Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

3.1.3 Implementation details . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.1.4 Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.2 Shear estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

3.2.1 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

3.2.2 Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

3.2.3 Implementation details . . . . . . . . . . . . . . . . . . . . . . . 48<br />

3.2.4 Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3.3 Pupil position and size estimation . . . . . . . . . . . . . . . . . . . . 53<br />

3.3.1 Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

3.3.2 Implementation details . . . . . . . . . . . . . . . . . . . . . . . 53<br />

3.3.3 Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

3.4 Phase recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

3.4.1 Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

3.4.2 Implementation details . . . . . . . . . . . . . . . . . . . . . . . 57<br />

3.4.3 Noise introduced by undesired reflections . . . . . . . . . . . . 58<br />

3.4.4 Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

3.5 Phase unwrapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.5.1 Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.5.2 Implementation details . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.5.3 Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

4 Wavefront reconstruction from shear phase maps 67<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.2 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

4.2.1 The transfer function of the reconstruction problem . . . . . . 69<br />

4.2.2 Periodicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

4.2.3 Data extension and extension method . . . . . . . . . . . . . . 70<br />

4.2.4 Filter poles and subdivision method . . . . . . . . . . . . . . . 72<br />

4.3 Computing performance . . . . . . . . . . . . . . . . . . . . . . . . . . 73<br />

4.4 Noise performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

4.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

5 Preliminary experiments 79<br />

5


CONTENTS<br />

5.1 Prydal type experiments . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

5.1.1 Normal illumination experiment . . . . . . . . . . . . . . . . . 80<br />

5.1.2 Focused illumination and estimation of undesired eye reflections 82<br />

5.2 Estimation of errors for tear topography measurements . . . . . . . . . 84<br />

5.2.1 Repeatability test and associated error . . . . . . . . . . . . . . 85<br />

5.2.2 Alignment and shear error propagation . . . . . . . . . . . . . 86<br />

5.2.3 Simulated interferograms and non-linear errors . . . . . . . . . 88<br />

5.3 Validation experiment with a deformable mirror . . . . . . . . . . . . . 88<br />

5.4 Small detail test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

6 Tear topography dynamics experiment 90<br />

6.1 Data collection protocol . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

6.2 Eye movements and effect on tear topography measurements . . . . . 92<br />

6.2.1 Effect of eyeball rotation on tear topography experiment . . . . 92<br />

6.2.2 Effect of head movement on tear topography . . . . . . . . . . 95<br />

6.2.3 Asymmetries in the optical system . . . . . . . . . . . . . . . . 95<br />

6.3 Global error estimation . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

6.4 Wavefront error RMS introduced by the tear topography . . . . . . . . 97<br />

6.4.1 Wavefront error RMS evolution for individual series . . . . . . 98<br />

6.4.2 RMS mean values . . . . . . . . . . . . . . . . . . . . . . . . . 100<br />

6.4.3 RMS mean evolution . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

6.5 Second moment of the AC term . . . . . . . . . . . . . . . . . . . . . . 103<br />

7 Summary and discussion 107<br />

7.1 Future work: Radial shearing interferometry . . . . . . . . . . . . . . . 110<br />

7.2 Publications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

A Preliminary study of feasibility of interferometric wavefront sensing<br />

in the eye 112<br />

A.1 Experimental setups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112<br />

A.1.1 Experimental setup for 632.8 nm . . . . . . . . . . . . . . . . . 114<br />

A.1.2 Experimental setup for 780 nm . . . . . . . . . . . . . . . . . . 115<br />

A.2 Data acquisition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

A.2.1 Data acquired with 632.8 nm source . . . . . . . . . . . . . . . 117<br />

A.2.2 Data acquired with 780 nm source . . . . . . . . . . . . . . . . 118<br />

A.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120<br />

6


CONTENTS<br />

B Laser safety calculations 124<br />

B.1 MPE for He-Ne laser in tear topography experiment . . . . . . . . . . 125<br />

B.2 MPE for He-Ne laser in full ocular wavefront sensor . . . . . . . . . . 125<br />

B.3 MPE for infrared LEDs . . . . . . . . . . . . . . . . . . . . . . . . . . 125<br />

B.4 MPE for infrared laser . . . . . . . . . . . . . . . . . . . . . . . . . . . 126<br />

Bibliography 127<br />

7


List of Figures<br />

1.1 Young’s experiment for demonstrating that the capability of changing<br />

the fixating distance of the human eye is due to the lens. . . . . . . . . 13<br />

1.2 Schematic representation of Scheiner’s principle, used to prescribe spectacles.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

1.3 Gullstrand’s method for estimating the aberrations introduced by the<br />

cornea. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

1.4 Viewing chart proposed by Helmholtz to verify the existence of aberrations<br />

in the human eye. . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2.1 Glass wedge geometry and angle definition for reflected rays. . . . . . 27<br />

2.2 Lateral shearing interferogram spectra produced by a glass wedge. . . 29<br />

2.3 Illumination system of the lateral shearing interferometer. . . . . . . . 32<br />

2.4 Relative intensities of reflections from a 50-50 non- polarising beam<br />

splitter (NPBS) and a polarising beam splitter (PBS). . . . . . . . . . 33<br />

2.5 Interference patterns that result from using different lenses in the illumination<br />

branch of the shear interferometer and an artificial cornea. . 36<br />

2.6 Imaging branches of the double shearing interferometer used for tear<br />

topography estimation. . . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

2.7 Sketch of the 3 dimensional optical setup for tear topography estimation<br />

with double lateral shearing interferometry. . . . . . . . . . . . . . . . 38<br />

2.8 Effect of the orientation of the polarisation state incident on the first<br />

glass wedge on the intensity and contrast of the interferograms produced<br />

by the glass wedges in the optical system illustrated in figure 2.7. . . . 41<br />

3.1 Estimation of the interferogram modulation carrier frequency intensity<br />

modulation from spectra. . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

3.2 As figure 3.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

8


LIST OF FIGURES<br />

3.3 Geometry of autocorrelation of an image consisting of two sheared<br />

copies of an object with finite dimensions. . . . . . . . . . . . . . . . . 48<br />

3.4 Improvement on shear estimation algorithm by filtering in Fourier domain. 49<br />

3.5 Shear estimation algorithm applied to tear film interferograms with<br />

different features. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

3.6 As figure 3.5. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

3.7 Center and radius estimation algorithm applied to tear film interferograms<br />

with different features. . . . . . . . . . . . . . . . . . . . . . . . 55<br />

3.8 As figure 3.7. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

3.9 Outputs of Takeda’s phase recovery method applied to tear interferograms.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

3.10 As figure 3.9. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

3.11 Phase unwrapping applied to different tear topographies. . . . . . . . 65<br />

3.12 As figure 3.11. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

4.1 Illustration of the method for extending the difference data outside the<br />

pupil, by copying the difference values in the boundary of the pupil<br />

outside the pupil. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

4.2 Example of the interlaced subgrids for a grid of dimensions N x = 6 by<br />

N y = 7 and shears s x = 3 and s y = 2. . . . . . . . . . . . . . . . . . . 73<br />

4.3 a) Plot of the number of real multiplications required to perform the<br />

wavefront reconstruction over a square grid with a total of N points and<br />

shear s


LIST OF FIGURES<br />

5.2 Sequence of reflections from the front surface of the tear illustrating<br />

bubble movements on the tear. . . . . . . . . . . . . . . . . . . . . . . 83<br />

5.3 Images obtained by illuminating the eye with a converging beam and<br />

placing the front surface of the eye roughly at the beam focus, and<br />

simulated interferograms. . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

5.4 Reconstructed topography of two simulated small bubbles. . . . . . . . 89<br />

6.1 Light scattered, as opposed to the specular reflections, by the cornea<br />

and the crystalline lens when the tear topography experiment was being<br />

performed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

6.2 Eyeball modelled as two spherical surfaces. . . . . . . . . . . . . . . . 93<br />

6.3 Estimated movement of the center of the illuminated area of the tear. 94<br />

6.4 Correlation between the relative change in the radius of the interferograms<br />

and the estimated change in defocus. . . . . . . . . . . . . . . . 96<br />

6.5 Typical wavefront RMS, defocus and astigmatism evolution with respect<br />

to the diffraction limit. . . . . . . . . . . . . . . . . . . . . . . . 99<br />

6.6 Mean RMS and residual RMS over the data series with respect to the<br />

diffraction limit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

6.7 Temporal evolution of the estimated wavefront error RMS and residual<br />

wavefront error RMS for 30 data series corresponding to 19 different<br />

subjects. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

6.8 Normalized second order moment of the AC term of tear topography<br />

lateral shearing interferograms (M N ) illustrating some of the different<br />

tear topography features. . . . . . . . . . . . . . . . . . . . . . . . . . 106<br />

7.1 Modified imaging branch to convert the lateral shearing interferometer<br />

into a radial shearing interferometer. . . . . . . . . . . . . . . . . . . . 110<br />

7.2 Typical radial shearing interferograms from tear film topography with<br />

no tilt between the interfering wavefronts. Recorded with the experimental<br />

setup described in figure 7.1. . . . . . . . . . . . . . . . . . . . 111<br />

A.1 Illumination branch of the wavefront sensing experiment using a 632.8 nm<br />

He-Ne laser. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114<br />

A.2 Imaging branch of the wavefront sensing experiment using a 632.8 nm<br />

He-Ne laser. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115<br />

10


LIST OF FIGURES<br />

A.3 Illumination branch of the wavefront sensing experiment using a 780 nm<br />

diode laser. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

A.4 Imaging branch of the wavefront sensing experiment using a 780 nm<br />

diode laser. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

A.5 Interferograms and SH spot patterns that resulted from using 3.5, 1.0<br />

and 0.35 µW at λ = 632.8 nm. . . . . . . . . . . . . . . . . . . . . . . . 119<br />

A.6 Interferograms and SH patterns that resulted from the 4 µW at 780 nm<br />

setup. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121<br />

A.7 Interferograms and SH patterns that resulted from the 4 µW at 780 nm<br />

setup with accommodation paralyzed with cyclopentolate hydrochloride<br />

1%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

B.1 Definition of subtended angle α by an apparent source at a distance r. 124<br />

11


Chapter 1<br />

Introduction<br />

Assessing and improving the optical quality of the human eye can be traced back as<br />

far as the 13 th century, when defocus was initially corrected by the use of spectacles.<br />

However, it was not until the 18 th century that monochromatic aberrations in the<br />

optics of the human eye other than defocus were studied by Thomas Young [1]. In his<br />

work, Young describes a series of fundamental experiments related to the imaging and<br />

focusing properties of the eye, like the one shown in figure 1.1, where it is demonstrated<br />

that the focusing capability of the eye is purely due to the change of shape of the<br />

crystalline lens. Young also reported noticing that his own eyes were astigmatic and<br />

consulting with a colleague to find this was not unusual and that some people used<br />

to hold a positive lens obliquely to compensate for it. He investigated this topic<br />

further by modifying a method proposed by Scheiner in 1619 to prescribe spectacles<br />

[2]. The principle of this method is illustrated in figure 1.2: in an refractive error-free<br />

(emmetropic) eye an incoming parallel pencil of rays of monochromatic light will be<br />

focused at the retina as shown in figure 1.2 (a) and therefore if a mask with two<br />

small holes is placed in front of the eye as in (b) the subject will see one spot. If<br />

on the other hand, the eye is ametropic, then the light from the incoming beam will<br />

be focused either in front (myopia) or behind the retina (hyperopia), producing two<br />

spots on the retina instead of one as shown in (c) and (d). Young’s modification to<br />

this experiment consisted of using 4 slits instead of 2 holes, and found that when the<br />

eye of the subject had aberrations other than defocus or astigmatism, there was no<br />

lens sphero-cylindrical correction that could make the image of the four slits coincide<br />

onto the retina, showing the presence of other aberrations in the eye.<br />

12


1. Introduction<br />

Figure 1.1: Young’s experiment for demonstrating that the capability of changing the<br />

fixating distance of the human eye is due to the lens. With the face looking down,<br />

the eye is placed on top of a a small cuvette filled with water so that the cornea<br />

is submerged. At the bottom of the cuvette there is a lens with the same refractive<br />

power as the cornea. Then by looking through this lens, focusing at different distances,<br />

Young observed that the accommodation range was the same as without the cuvette,<br />

confirming then that the accommodation of the eye is due entirely to the lens and<br />

not to the cornea. This image was reproduced from Young’s Bakerian lecture on the<br />

mechanism of the eye [1].<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

Figure 1.2: Schematic representation of Scheiner’s principle, used to prescribe spectacles.<br />

In an emmetropic eye, an incoming parallel pencil of rays will be focused at<br />

the retina as shown in (a). Therefore, by placing a mask with two small holes the<br />

subject will see one spot when the eye is emmetropic (b) and two when the eye is<br />

either myopic (c) or hyperopic (d).<br />

13


1. Introduction<br />

(a)<br />

(b)<br />

Figure 1.3: Gullstrand’s method for estimating the aberrations introduced by the<br />

cornea. The method is a rudimentary corneal topography, consisting of imaging a set<br />

of squares like the one in (a) after reflection from the first corneal surface and measuring<br />

its deformation as shown in (b). This images were reproduced from Helmholtz’s<br />

Handbuch der Physiologischen Optik [3].<br />

One century later, Gullstrand [3] studied extensively the monochromatic aberrations<br />

of the eye, the influence of the concavity of the tear in vision quality, the scattering<br />

introduced by the fibres in the lens and the cornea and the lack of rotational symmetry<br />

shown by the aberrations in the eye. In his work, Gullstrand tried to characterize the<br />

aberrations of the eye in terms of the caustic, and studied the aberrations introduced<br />

by the cornea by performing a very elementary corneal topography, consisting of<br />

imaging a set of distorted squares on the front corneal surface and measuring their<br />

deformation, as shown in figure 1.3.<br />

1.1 Wavefront sensing in the human eye<br />

We can think of wavefront sensing in the eye as the evaluation of the monochromatic<br />

aberrations of the eye other than piston, tip and tilt, as these do not degrade the<br />

imaging properties of the eye. We could also say that Young is the pioneer of pupilplane<br />

wavefront sensing in the eye, while Gullstrand gave the first steps in image-plane<br />

wavefront sensing in the human eye. However, we can consider that the modern era<br />

of wavefront sensing in the eye did not start until 1961, with the work by Smirnov [4]<br />

where a subjective experiment was set up to sample the wavefront slope across the<br />

pupil and then to estimate the point spread function (PSF) of the eye. Smirnov also<br />

suggested the tear film as a possible source for the noticeable inter-subject variability<br />

14


1. Introduction<br />

in the measurements and noticed that the aberrations change with the accommodating<br />

distance. In the same work, Smirnov suggested the use of contact lenses for correcting<br />

high order aberrations. Many of Smirnov’s ideas and observations are the basis of<br />

modern wavefront sensing techniques, both subjective and objective and some of his<br />

ideas, such as the compensation of the aberrations have been put into practice.<br />

Following Smirnov’s work, at least two more subjective or psychophysical wavefront<br />

sensing methods (i.e., those in which the aberrations of the eye are estimated from<br />

responses provided by the subject being examined) have been developed, based on<br />

wavefront slope evaluation in the pupil plane. Firstly, the cross-cylinder aberroscope<br />

proposed by H. and B. Howland [5], in which a grid is sandwiched between two crossed<br />

cylindrical lenses and projected onto the retina. The retinal image of the grid is<br />

distorted by the aberrations of the eye. The subject is asked to draw this grid so<br />

that the aberrations can be estimated the drawings. This technique did not become<br />

popular given the strong reliance on the subject’s ability to replicate the image seen.<br />

Later, in 1992 Webb et al.[6] and then He et al.[7] built more automated versions<br />

of Smirnov’s experiment, where the wavefront slope at different positions within the<br />

pupil is measured by feedback provided by the vernier acuity of the subject.<br />

In 1969 Berny [8] used a completely different approach by performing a Foucault<br />

knife-edge test experiment to measure the wavefront aberration itself, rather than the<br />

wavefront slope.<br />

With the development of more sensitive films and cameras, a number of objective<br />

techniques for evaluating the aberrations of the eye were developed in recent years.<br />

In 1984 Walsh et al. [9, 10, 11] transformed the subjective Howland and Howland<br />

cross-cylinder aberroscope into an objective method by recording the image of a grid<br />

projected onto the retina with a camera.<br />

In 1997, Navarro et al.[12, 13, 14] developed an instrument to evaluate the slope of the<br />

wavefront, sampling the pupil at different locations with a narrow collimated incoming<br />

beam and recording its displacement on the retinal surface with respect to a reference<br />

spot.<br />

Recently, in 1996 Ragazzoni [15] proposed a new slope wavefront sensor, in which the<br />

beam to analyze is moved in a circular pattern over a glass pyramid producing four<br />

images that are combined to estimate local wavefront slopes. This wavefront sensor<br />

15


1. Introduction<br />

was then modified and tested in the eye by Iglesias et al. [16].<br />

Iglesias et al. also proposed an image-plane wavefront sensing method based on applying<br />

phase retrieval algorithms to an experimentally estimated point spread function<br />

(PSF) for the eye [17, 18, 19].<br />

In 1994 Liang et al. [20, 21] introduced the use of the Shack-Hartmann sensor in the<br />

eye, which was then to become the preferred instrument for wavefront sensing in the<br />

eye. In this wavefront sensor, an array of lenslets placed in a plane conjugated to the<br />

pupil of the eye forms an array of spots on a camera. The displacement of these spots<br />

with respect to their reference position is proportional to the average wavefront slope<br />

at the plane of the lenslet. Since the work by Liang et al., an increasing amount of<br />

research has been done in the field of wavefront sensing in the eye, leading to a better<br />

understanding of the field. Artal et al. [22] recognized the double pass problem as an<br />

autocorrelation. Therefore, if equal entrance and exit pupils are used in a wavefront<br />

sensing instrument, the odd aberrations of the eye would be underestimated in the<br />

measurements. Artal then proposed using unequal entrance and exit pupil sizes to<br />

minimize this problem. To tackle the same problem, Díaz and Dainty [23] tested an<br />

original approach, setting up a single pass experiment by using the autofluorescence<br />

of the retinal lipofuscin as a light source for wavefront sensing. Salmon et al. [24]<br />

explored the problem further by performing a comparison between a Smirnov-type<br />

single-pass experiment and a Shack-Hartmann sensor (double-pass) with unequal entrance<br />

pupils, concluding that the results were similar to within experimental errors.<br />

López and Artal [25] and then Llorente et al. [26] explored the difference in the<br />

estimated aberrations for different wavelengths, which is a vital point for clinical applications.<br />

Ideally, one would like to use visible wavelengths, but this causes the iris of<br />

the eye to contract, and therefore pupil dilation drugs would be needed to measure the<br />

aberrations over large pupils. On the other hand, if infrared wavelengths are used, a<br />

large pupil can be used without the need for drugs, but then the presence of chromatic<br />

aberration in the eye and different scattering characteristics need to be accounted for.<br />

Thibos et al. [27] noticed that the Shack-Hartmann measurements can be affected by<br />

tear film break-up and cataracts, which suggests that the measurements from wavefront<br />

sensors should not be used alone without other studies such as retro-illumination<br />

images. More recently, some experimental studies have been performed to characterize<br />

the dynamics of the aberrations [28] and their statistics [29], but theoretical modelling<br />

is still pending.<br />

16


1. Introduction<br />

1.1.1 Applications of wavefront sensing in the human eye<br />

There are a number of applications for wavefront sensing in the eye of which the<br />

most widespread in clinical applications is refractive surgery, in its different techniques:<br />

photorefractive keratectomy (PRK) and laser-assisted in situ keratomileusis<br />

(LASIK), where only sphere and cylinder are corrected, and the more recent wavefrontguided<br />

LASIK, also referred to as custom ablation, where higher order aberrations<br />

are accounted for [30, 31]. These operations are currently being performed on more<br />

than a million people per year. The importance of good understanding of the wavefront<br />

sensing measurements cannot be overemphasized, given the potential permanent<br />

degradation to vision in such large number of individuals. There are currently several<br />

patents related to ophthalmic wavefront sensors [32, 33, 34, 35, 36] and also commercial<br />

instruments available in the market, like the BriteEye Aberrometer from Suzhou<br />

Medical General Factory, the Zywave wavefront aberrometer from Bausch and Lomb,<br />

the WASCA Analyser from Carl Zeiss Meditec, the LADARWave from Alcon, the KR-<br />

9000PW Wavefront Analyzer from Topcon and the COAS from Wavefront Sciences.<br />

Wavefront sensing in the eye in combination with corneal topography, allows the<br />

study of the aberrations of the crystalline lens, which could lead to better design<br />

and evaluation of intraocular lenses. These lenses are currently being implanted in<br />

numbers comparable to or larger than those involved in refractive surgery [37].<br />

Adaptive optics has proven to be the most challenging application of wavefront sensing<br />

in the eye to implement in a reliable manner. Adaptive optics was first proposed by<br />

Babcock in 1953, for compensating the aberrations introduced by the atmosphere in<br />

telescope images [38], and based on the same principles and technology it has recently<br />

been applied to the correction of the aberrations of the human eye [21, 39, 40, 41, 42].<br />

In principle, all ophthalmic imaging instruments such as fundus cameras [43, 44, 45,<br />

46, 47], confocal laser scanning ophthalmoscopes (CLSOs) [48, 49, 50, 41] and optical<br />

coherence tomography systems (OCTs) [51, 52] can benefit by improving the optical<br />

quality of the eye with an adaptive optics system to a resolution in which individual<br />

photoreceptors can be resolved within small retinal patches. If high-resolution imaging<br />

is required over a retinal patch larger than the isoplanatic angle [53, 54, 55], then a<br />

more complex multi-conjugate adaptive optics system would be needed.<br />

Improving the resolution of retinal images down to almost the diffraction limit has also<br />

17


1. Introduction<br />

been achieved by using deconvolution techniques on series of retinal images using the<br />

information on the aberrations of the eye obtained with wavefront sensors [56, 57, 58].<br />

These deconvolution techniques might be cheaper and easier to implement alternatives<br />

to adaptive optics.<br />

A less viable application for wavefront sensing in the eye is the design of phase plates<br />

that can be used as spectacles for superior vision improvement in comparison with<br />

the usual sphero-cylindrical correction [59, 60]. The drawback of this is that the best<br />

optical quality would only be achieved when looking in a particular direction. When<br />

the eye turns to look in a slightly different direction, vision will almost certainly be<br />

worse than with a normal pair of spectacles.<br />

One of the applications that attracts the most attention to the field of wavefront<br />

sensing in the eye is the possibility of achieving what has been called supernormal<br />

vision, through the use of adaptive optics. This has so far only been achieved in<br />

laboratory experimental setups [43, 61, 62], with instruments of sizes and weights<br />

far from the size and weight of a normal pair of spectacles (one and three order of<br />

magnitudes larger respectively). A similarly challenging goal is the use intraocular<br />

adaptive optics either to compensate presbyopia or to improve visual acuity. This<br />

possibility has started to be explored by Vdovin et al. [63] who built and tested a<br />

small liquid crystal correction element of size comparable to that of a crystalline lens.<br />

Adaptive optics in the eye might also allow, in the near future, psychophysical experiments<br />

in which light is focused onto single photoreceptors, to gain knowledge of their<br />

response to different wavelengths. So far this has not been achieved due to technical<br />

difficulties and problems not well understood yet. In the same way, the theoretical<br />

imaging diffraction limit has not yet been achieved experimentally.<br />

1.1.2 Repeatability in wavefront sensing measurements<br />

The variability of wavefront sensing measurements in the eye was identified in the very<br />

early stages of the field by Smirnov, and a number of sources for this variability have<br />

been suggested: fluctuations of accommodation, change in axial length and corneal<br />

shape due to the heart beat, eye and head movement and tear film dynamics. We<br />

believe that the understanding of the effect of these sources in wavefront sensing<br />

measurements could lead to better use of the measurements, for example in the context<br />

18


1. Introduction<br />

Figure 1.4: Viewing chart proposed by Helmholtz to verify the existence of aberrations<br />

in the human eye. When this target is viewed with one eye and a steady fixation,<br />

some rotating sectors can be seen. The time course of the movements correspond to<br />

that of the fluctuations of the accommodation.<br />

of refractive surgery. Reported values for the variability of the root-mean-squared<br />

(RMS) of the wavefront error in the eye are typically of the order of 0.1 µm [4, 24,<br />

64, 17, 21]. These values are not an issue for spectacle prescription. 0.1 µm of sphere<br />

(defocus) for an 8 mm pupil corresponds only to a twentieth of a diopter (D), which is<br />

less than half the minimum step between consecutive ophthalmic lenses. On the other<br />

hand, the variability observed is of concern when considering a diffraction limited<br />

performance imaging system for the eye (e.g. fundus camera), as the wavefront RMS<br />

should be kept below one fourteenth of a wavelength according to Marechal’s criterion,<br />

that is 0.03 − 0.05 µm for visible wavelengths [10].<br />

Is is very likely that the largest source of variability of the optical quality of the eye<br />

is the fluctuations of accommodation, whose amplitude and frequency were studied<br />

by Campbell et al. [65] and Gray et al. [66]. The measured fluctuations depend<br />

on several factors like target luminance, pupil diameter, target form, vergence and<br />

contrast, with amplitudes as large as 0.5 D. If we consider that a typical fluctuation<br />

amplitude is around 0.2 D, which seems reasonable from Charman’s review [67], this<br />

is equivalent to a wavefront error RMS of approximately 0.5 µm for an 8 mm diameter<br />

pupil. This is one order of magnitude larger than the wavefront aberration that is<br />

needed to take a perfect optical system out of the diffraction limit regime. For this<br />

reason, it is common practice when using both wavefront sensors and adaptive optics<br />

in research environments, to paralyze or reduce these fluctuations by the use of drugs<br />

(usually cyclopentolate) [37, 28, 21, 17, 25, 14, 68]. A simple experiment to verify the<br />

existence of these microfluctuations can be performed with the aid of figure 1.4.<br />

19


1. Introduction<br />

A less significant but not negligible source of variability seems to be the axial length<br />

change of the eye when the heart beats. Hofer [28] quotes that typical changes in the<br />

axial length of 4 µm would produce defocus changes of 0.01 D which for an 8 mm pupil<br />

corresponds to a wavefront RMS of 0.02 µm.<br />

The role of eye movement in wavefront sensing measurements has, to our knowledge,<br />

not yet been studied on its own, although the movements of the eye themselves are<br />

well studied and their characteristics appear in current ophthalmology textbooks such<br />

as Adler’s [69]. For example, it has been proved that when the eye is fixating it has<br />

three characteristic movements, namely microsaccades of ∼ 5 arcmin of amplitude at<br />

a frequency of 2 − 3 Hz; drift ∼ 1 arcmin every 2 − 5 min and high frequency tremor<br />

∼ 40 arcsec at 90 Hz.<br />

There seems to be agreement in the ophthalmic wavefront sensing community about<br />

the tear film playing a role in the variability of measurements [4, 70, 39, 46, 71, 28,<br />

58, 72, 21, 24, 68, 60], although to our knowledge before the research described in<br />

this <strong>thesis</strong> there was no quantitative study on the tear topography dynamics and its<br />

effects on wavefront sensing measurements. Most of the research related to this topic<br />

so far had been done for the extreme situations of tear break-up, like that by Tutt et<br />

al. [68] showing that the tear break-up affects the contrast of retinal images. Some<br />

efforts by Himebaugh et al. [71] to provide a technique to study the changes on the<br />

tear topography made in parallel to and independently of our work, led to a retroillumination<br />

instrument, which is yet to produce definite results. A more suitable<br />

technique to study the tear topography, was proposed by Licznerski et al. [73, 74,<br />

75, 76], where information about the tear topography can be obtained using shearing<br />

interferometry, and some semi-quantitative measurements were performed. However,<br />

this experimental setup had an important limitation in that the tear topography<br />

estimation was made using a gradient projection along only one direction.<br />

There are other sources of variability in wavefront sensing measurements that might<br />

not be noticeable if measurements are taken within a few seconds, but are nevertheless<br />

non-negligible. They can significantly bias the wavefront aberration estimation,leading<br />

to, for example, an incorrect corneal ablation profile. They include instrument myopia<br />

[77] which can be as large as 5 D, the change in refraction due to changes in ambient<br />

light [78] and the corneal topography changes that can follow reading due to eyelid<br />

pressure, around 0.25 D sphere and 0.25 D cylinder.<br />

20


1. Introduction<br />

1.2 Tear film<br />

It is accepted that the tear film consists of three layers [76, 79], the outermost being the<br />

lipid layer of around 50 nm thickness, followed by the aqueous layer with a thickness<br />

of around 4 − 8 µm and finally a mucous layer of around 20 − 50 nm. Each of the<br />

layers has a different function, but from the optical point of view we should only be<br />

concerned with their thickness and refractive indices. Actually, from the refractive<br />

index point of view, we will assume, to a first approximation, a mean refractive index<br />

n tear = 1.337 [80], so that we can concentrate on the tear thickness only. What is<br />

more, in the rest of this <strong>thesis</strong>, we will estimate the topography of the front surface<br />

of the film, assuming an unchanged back surface, which is reasonable given the time<br />

scales involved in our experiments (a few seconds).<br />

The first question we should ask ourselves is whether the tear film is thick enough<br />

to produce optical path length differences that are significant in terms of the other<br />

sources of variability of wavefront sensing in the eye measurements. Let us begin by<br />

assuming that the difference between two tear topography maps at two given instants<br />

is a parabolic surface that can be described by the defocus Zernike polynomial. We can<br />

then relate the sag s of this parabolic surface to the wavefront error RMS amplitude<br />

(a 2,0 ) it would produce in the eye through the formula<br />

(<br />

2 √ )<br />

3<br />

s =<br />

a 2,0 (1.1)<br />

n tear − 1<br />

≈ 10 a 2,0 (1.2)<br />

Therefore, if we assume even the most conservative reported values of tear thickness<br />

[81, 82, 83, 84, 85, 86, 87, 88, 89], that is around 3 µm, then a sag of only 15 % of the<br />

tear thickness (which seems plausible) would be enough to take a perfect eye out of the<br />

diffraction limit. To see, under the same assumptions, the sag required to introduce<br />

a defocus amplitude comparable to the intervals in spectacle prescription, we can use<br />

the following<br />

s =<br />

d 2 Φ s<br />

8(n tear − 1)<br />

(1.3)<br />

where d is the pupil diameter and Φ is the prescription in diopters. Therefore, for the<br />

tear to play a noticeable role in vision (let us say a quarter of a diopter) in daylight<br />

conditions with a typical value for d of 3 mm, the sag would have to be 0.8 µm –a<br />

value physically plausible.<br />

21


1. Introduction<br />

It should also be mentioned that decrease in spatial-contrast sensitivity has been<br />

reported by Rolando et al. to be associated with tear disfunction [90], although no<br />

quantitative study of the tear topography was performed for this study.<br />

1.3 Thesis synopsis<br />

Chapter 2 describes the design of and theory behind the double-shearing interferometer<br />

built for the study of the dynamics of human tear topography. The specifications<br />

of the interferometer and role of all the optical elements are explained in sufficient<br />

detail so that anyone trying to replicate or improve the experiment can do so without<br />

great difficulty.<br />

In Chapter 3 the theory, implementation of and elementary testing of the algorithms<br />

used in the data processing from the raw shearing interferograms to the unwrapped<br />

topography difference maps are discussed.<br />

Chapter 4 is devoted to the discrete fourier transform (DFT) based integration methods<br />

used to estimate the tear topography maps from the difference maps. We present<br />

the theory of both methods in detail, and evaluate the computing and noise performance<br />

in comparison with the conventional least-squares pseudoinverse reconstruction<br />

methods.<br />

Chapter 5 presents some preliminary experiments, where the feasibility of using<br />

the interference of the reflections on the front and back surface of the tear to estimate<br />

the tear topography was tested. The repeatability of and some errors in our<br />

double-shearing interferometer are estimated. Finally a simple validation experiment<br />

was performed and the ability of the experiment to detect small details on the tear<br />

topography was tested.<br />

Chapter 6 begins by describing the data collection protocol, follows with a discussion<br />

of the error that can be expected in our data from eye movements and then with<br />

an estimation of the global errors on the data. After this, the estimated wavefront<br />

RMS from the processed tear topography data is analyzed with regards to the optical<br />

quality of the eye. Finally, it is shown based on experimental data, that the method<br />

proposed by Licznerski et al. [74] to evaluate the tear break-up is not adequate either<br />

as originally proposed or after some suggested improvements.<br />

22


1. Introduction<br />

Chapter 7 presents a short summary and suggests the use of radial shearing interferometry<br />

for the study of tear topography instead or double lateral shearing interferometry.<br />

Appendix A explores experimentally the possibility of using lateral shearing interferometry<br />

for wavefront sensing in the eye.<br />

Appendix B shows the safety calculations performed to comply with the British and<br />

European standard BS EN 60825-1:1994 with amendments 1, 2 and 3.<br />

The DVDs attached to this <strong>thesis</strong> contain the following:<br />

• Documents in HTML format that link to MPG movies showing the sequences<br />

of raw tear lateral shearing interferograms and estimated tear topography maps<br />

when these could be obtained.<br />

• Compressed files in ZIP format with the same movies in AVI format (higher<br />

resolution).<br />

• Compressed files in ZIP format with the raw images recorded (TIF format).<br />

• Matlab scripts (M) files containing the experimental configuration for each individual<br />

data series for the tear topography experiments and some notes.<br />

• Matlab binary .MAT files containing the input parameters for the tear topography<br />

data processing and the outcomes (e.g. interferograms center and diameter).<br />

• Raw images for all the other experiments described in this <strong>thesis</strong> in TIF format.<br />

23


Chapter 2<br />

Lateral shearing interferometer<br />

design<br />

Designing and building an instrument to measure the dynamics of the topography of<br />

the pre-corneal tear film has proven a task far more difficult than initially thought. In<br />

what follows we describe the specifications, theory and technical details regarding the<br />

design of the double-shearing interferometer that was built and tested on 20 subjects.<br />

2.1 System requirements<br />

We aimed at producing an instrument able to measure changes in the topography of<br />

the tear film with sub-micron depth resolution and lateral resolution higher than 20<br />

to 40 samples across a pupil diameter of about 7 mm diameter, at a minimum rate of<br />

2 Hz. In this way, we would be able to resolve temporally and spatially tear topography<br />

features such as the rough surface, bubbles, lines and break-ups, not visually obvious<br />

in current SH wavefront sensor images, and study their effect both in the optical<br />

quality of the eye and SH wavefront measurements. The only specification we could<br />

not meet is the area under study in the pupil, given the large numerical aperture<br />

required and the fact that for comfort the instrument was to be placed at least 15 mm<br />

away from the front surface of the cornea. The diameter of the illuminated area over<br />

the tear film was around 3 mm.<br />

The range of tear depth changes to estimate seem appropriate for an interferometric<br />

technique, although not all interferometric methods would be suitable due to the<br />

24


2. Lateral shearing interferometer design<br />

movements of the eye. Self referencing interferometers are able to cope better with<br />

eye movement. If we now look at the techniques to extract the phase from an interferogram,<br />

we find ourselves again restricted by eye movement to single frame techniques<br />

(or synchronized multi-frame acquisition). Based on this and some previous semiquantitative<br />

experiments by Licznerski et al. [73] it was decided to implement a<br />

lateral shearing interferometer as described below.<br />

2.2 Lateral shearing interferometry<br />

In a lateral shearing interferometer (LSI), the beam containing the phase information<br />

to be measured is amplitude divided into two beams, laterally shifted and then recombined.<br />

The lack of a static coherent reference beam makes the resulting interferogram<br />

less sensitive with respect to eye movement. If we assume that the amplitude division<br />

produces two beams with equal intensity and phase profiles, then the equation that<br />

describes the intensity pattern resulting from the interference of the sheared copies of<br />

the electric field E 0 (⃗r) and E 0 (⃗r + ⃗s) is<br />

I(⃗r) = I 0 (⃗r) + I 0 (⃗r + ⃗s) + 2 √ I 0 (⃗r) I 0 (⃗r + ⃗s) cos [φ(⃗r + ⃗s) − φ(⃗r)] , (2.1)<br />

where ⃗r is the position vector in the interferogram plane, ⃗s is the shear between<br />

the two beams, I 0 (⃗r) = |E(⃗r)| 2 , φ is the phase of the complex electric field E(⃗r),<br />

the function of interest. Here we have assumed scalar electric fields, but provided<br />

that when performing the amplitude division the polarisation state of the two sheared<br />

beams remains unchanged with respect to each other, then equation 2.1 remains valid.<br />

It is also implicitly assumed that the spatial and temporal coherence of the beams is<br />

such that the modulus of the complex degree of coherence [91] is virtually one for the<br />

shear and wavefront under consideration.<br />

In order to reconstruct the two-dimensional phase map φ, at least 2 interferograms<br />

with different shear directions (not necessarily orthogonal) are needed, because the<br />

phase difference in equation 2.1 is insensitive to phase maps constant along the direction<br />

of shear.<br />

25


2. Lateral shearing interferometer design<br />

2.3 Interferogram intensity modulation<br />

The fundamental problem in interferometry is to extract the phase from interferograms,<br />

despite fluctuations in the intensity profile across the interferogram.<br />

non-uniform intensity in real interferograms can be due to a number of factors, such<br />

as intensity profile of the source, non-uniform absorption/reflection of the phase object<br />

to be analyzed or diffraction. We will avoid phase-shifting techniques [92] because<br />

they imply the acquisition of multiple interferograms for every phase map to be reconstructed<br />

[93, 94], requiring either non-simultaneous acquisition, which is not desirable<br />

as we do not know how fast the tear topography changes, or a complex optical setup<br />

for the simultaneous acquisition of the interferograms needed. Instead, we modulate<br />

the intensity of the interferogram spatially, by introducing a controlled amount of tilt<br />

between the two interfering beams, and then use the phase recovery algorithm proposed<br />

by Takeda [95]. The spatial frequency of the intensity modulation ⃗ f c acts as a<br />

phase carrier frequency and is related to the tip (ε x ) and tilt (ε y ) angles between the<br />

interfering wavefronts by<br />

The<br />

( tan εx<br />

⃗f c =<br />

λ<br />

, tan ε )<br />

y<br />

. (2.2)<br />

λ<br />

The intensity of the interferogram once the tilt associated to ⃗ f c is introduced is given<br />

by<br />

I(⃗r) = I 0 (⃗r) + I 0 (⃗r + ⃗s) + 2 √ [<br />

I 0 (⃗r) I 0 (⃗r + ⃗s) cos φ(⃗r + ⃗s) − φ(⃗r) + 2πf ⃗ ]<br />

c · ⃗r (2.3) .<br />

The first two terms on the right, are referred to as the DC term and the third one as<br />

the AC term.<br />

2.4 Wedge design<br />

The amplitude division for the shear interferometer can be achieved in different ways,<br />

typically by the use of beam-splitters [96], diffraction gratings [97, 98, 99] and wedged<br />

plates [73]. The use of wedged plates was preferred for their robustness and ease<br />

of production in our optical workshop. Even though a wedge plate can be defined<br />

by only two parameters, namely the angle between the surfaces and the thickness at<br />

one given point, the selection of these two is not trivial in our case. The choice of<br />

tilt to introduce between the interfering wavefronts depends on the optical system,<br />

26


2. Lateral shearing interferometer design<br />

the density of elements in the detector that records the interferograms, the expected<br />

spatial spectra of the tear and source profile and the signal-to-noise ratio.<br />

In the rest of this section the geometry and angle nomenclature is as shown in figure<br />

2.1, and only the reflections on the front surface and first reflection on the back<br />

surface of the wedge will be considered. The consequences of this consideration will<br />

be discussed later in section 3.4.3.<br />

a 1<br />

a 2<br />

a 3<br />

O d<br />

A<br />

a 1<br />

a 5<br />

C<br />

a 4<br />

a 3<br />

B<br />

n w n m<br />

Figure 2.1: Glass wedge geometry and angle definition for reflected rays.<br />

From figure 2.1 and the use of Snell’s law it can be shown that for a ray incident on<br />

the wedge at the point A, the tilt t = |α 5 − α 1 | between the reflection on the front<br />

and back surface is given by<br />

[ ( )]} ∣ t =<br />

1<br />

∣∣∣<br />

∣<br />

{n arcsin w sin 2δ + arcsin sin α 1 − α 1 . (2.4)<br />

n w<br />

where n w is the refractive index of the wedge and we assume that it is in air. This<br />

equation shows that the tilt between each pair of rays resulting from the same incident<br />

ray, depends on the angle of incidence α 1 . Therefore, unless all the rays of the incident<br />

beam at the wedge are parallel, the reflected wavefronts will be distorted. Hence, the<br />

wedges must be placed in a region of the optical system where the incident beam is<br />

parallel.<br />

If we now calculate the shear s = AC introduced by the wedge using basic trigonometry,<br />

we have,<br />

s =<br />

[<br />

sin δ sin<br />

[ ( )]<br />

cos 2δ + arcsin sin α1<br />

n w<br />

( )]<br />

2δ + 2 arcsin sin α1<br />

n w<br />

cos<br />

[<br />

δ + arcsin<br />

(<br />

sin α1<br />

n w<br />

)]AO (2.5)<br />

27


2. Lateral shearing interferometer design<br />

which for δ ≪ 1 and α 1 ≫ δ can be approximated by<br />

[ ( )]<br />

sin 2 arcsin sin α1<br />

n w<br />

s ≈<br />

( )] δ AO (2.6)<br />

cos<br />

[arcsin 2 sin α1<br />

n w<br />

From either of these two equations we can see that the shear is non-uniform across<br />

the wedge, and therefore, the wavefront reflected on the back surface of the wedge will<br />

be non-uniformly stretched in the direction of the shear. The relative change in shear<br />

D s across the beam width w at the wedge, can be expressed as<br />

D s = 2 s max − s min<br />

s max + s min<br />

(2.7)<br />

where s min and s max are the shear of the rays incident on the wedge closest and furthest<br />

from the wedge apex respectively. By using equation 2.5 assuming a collimated<br />

incident beam, the relative change in shear across the beam can be rewritten as<br />

D s =<br />

w<br />

AO mean<br />

(2.8)<br />

where w is the beam width at the wedge. This suggests that the shear non-uniformity<br />

can be kept below a given tolerance T by making the distance AO mean large in comparison<br />

to w or equivalently, making the wedged plate thicker,<br />

w<br />

AO mean<br />

< T (2.9)<br />

2.5 Wedge manufacture<br />

For manufacturing purposes it will be more convenient to invert and combine formulas<br />

2.4, 2.6 and the condition 2.9, so that we have<br />

{<br />

arcsin<br />

AO mean<br />

δ = 1 2<br />

≈<br />

[ ]<br />

sin(α1 ± t)<br />

− arcsin<br />

n w<br />

[ sin(α1 )<br />

n w<br />

]}<br />

( )]<br />

2 s mean cos<br />

[arcsin 2 sin α1<br />

n w<br />

[ ( )]<br />

sin 2 arcsin sin α1<br />

n w<br />

{arcsin<br />

[<br />

sin(α1 )±t<br />

n w<br />

]<br />

− arcsin<br />

(2.10)<br />

[<br />

sin(α1 )<br />

n w<br />

]}(2.11)<br />

AO mean > w T . (2.12)<br />

It should be noticed that if the two equations for AO mean can not be simultaneously<br />

satisfied, then the experiment will need to be modified to change the input parameters<br />

until the three equations are satisfied to within the accepted tolerances. In practice,<br />

28


2. Lateral shearing interferometer design<br />

f AC<br />

max<br />

f DC<br />

max<br />

f c<br />

f sampling<br />

Figure 2.2: Lateral shearing interferogram spectra produced by a glass wedge. The<br />

red circle in the center contains the DC term, while the other two circles contain the<br />

AC terms that carry the phase information. The indicated spatial frequencies are: the<br />

sampling frequency f sampling , the carrier frequency f c and the maximum DC and AC<br />

term frequencies fmax DC and fmax AC respectively.<br />

this is not an important limitation because we have an extra degree of freedom in the<br />

shear, that can be tuned later by sliding the wedge position along the optical axis of<br />

the system, as will be explained in section 2.7.1.<br />

Two conditions must be met to avoid aliasing when spatially sampling the interferogram.<br />

Firstly, the spectra of the AC and DC terms of the interferogram intensity<br />

must not overlap, as it is the case in the example shown in figure 2.2. This is ensured<br />

by choosing a carrier frequency f c (i.e. the wedges angle) larger than the sum of the<br />

highest spatial frequencies of the DC and AC terms, that is<br />

f DC<br />

max + f AC<br />

max < f c . (2.13)<br />

Secondly, according to the sampling theorem, the spatial sampling frequency must be<br />

equal to or larger than twice highest spatial frequency of the interferogram, that is<br />

f c + f AC<br />

max < f sampling /2. (2.14)<br />

If information on the spatial content of the tear topography were available, one could<br />

put figures into these conditions and estimate an adequate angle for the wedges and the<br />

29


2. Lateral shearing interferometer design<br />

number of camera pixels. However, we do not have an established model to describe<br />

the dynamics of tear topography. Therefore, making no assumptions on the spatial<br />

content of the tear topography, the angles of the glass wedges for the interferometer<br />

were chosen so that the carrier frequencies of the interferograms were sampled at<br />

around 8 camera pixels per fringe, with around 450 samples across each interferogram.<br />

These values correspond to a tilt between the reflected wavefronts on the front and<br />

back surfaces of the wedge of 12 mrad, and are a compromise between having high<br />

spatial frequency modulation in the interferograms and reasonable sampling.<br />

We now consider the shear introduced by the wedges, for which we recall equation 2.3<br />

I(⃗r) = I 0 (⃗r) + I 0 (⃗r + ⃗s) + 2 √ [ ]<br />

φ(⃗r + ⃗s) − φ(⃗r)<br />

I 0 (⃗r) I 0 (⃗r + ⃗s) cos<br />

s + 2πf s<br />

⃗ c · ⃗r ,<br />

where for convenience the phase difference has been multiplied and divided by the<br />

shear amplitude s. We can now see that if the shear is small in comparison with the<br />

spatial changes in the phase, the phase difference can be approximated by the phase<br />

slope in the shear direction multiplied by the shear amplitude. We can further assume<br />

that in terms of measured phase, the signal-to-noise ratio (SNR) is proportional to<br />

the shear amplitude. On the other hand, the larger the shear the smaller the area<br />

over which the pupils overlap, and so the poorer the phase estimation. Again, as with<br />

the tilt, an optimum estimation of the shear needed for the experiment would need<br />

some knowledge of the wavefronts to be measured, and therefore the choice of shear<br />

is arbitrary, and in our case is approximately 5 % of the beam diameter for a beam<br />

diameter of 3 mm. Finally, we will consider 2 % as an acceptable tolerance to the shear<br />

variation across the beam.<br />

Taking all the above in to consideration for a beam diameter of 3 mm, we reach the<br />

following values for the wedge angle δ and thickness:<br />

δ ≈ 10 arcmins, (2.15)<br />

thickness ≈ 0.4 mm, (2.16)<br />

The two glass wedges manufactured have angles 11 and 13 arcmins, and mean thicknesses<br />

0.35 and 0.70 mm respectively. The extra thickness of the second wedge will<br />

only affect the shear amplitude, but as will be shown later (see section 2.7.1), this<br />

can be compensated for by displacing the wedge along the optical axis. Both glass<br />

wedges were produced to be used at angle of incidence α 1 of 45 o and made of BK7,<br />

30


2. Lateral shearing interferometer design<br />

He-Ne<br />

laser<br />

633nm<br />

ND filter<br />

microscope objective<br />

spatial filter<br />

f = 250<br />

1<br />

stop aperture<br />

s-polarisation<br />

l/4 f 2 = 40<br />

PBS<br />

eye<br />

p-polarisation<br />

circular-polarisation<br />

Figure 2.3: Illumination system of the lateral shearing interferometer.<br />

which at the wavelength of the He-Ne laser used (632.8 nm) has a refractive index<br />

n w = 1.51509.<br />

2.6 Illumination branch<br />

The illumination system of the shearing interferometer, shown in figure 2.3, consists<br />

of a light source, a neutral density absorption filter, a microscope objective, a pinhole,<br />

a collimating lens, a polarising beam-splitter, a quarter waveplate and a lens. The<br />

purpose of this part of the optical system is to produce a smooth (ideally perfectly uniform)<br />

illumination over a circular region of the front surface of a subject’s precorneal<br />

tear and a fixation target.<br />

2.6.1 Source<br />

The source to be used needs to be quasi-monochromatic because multiple wavelengths<br />

would produce different phase differences and, therefore, multiple overlapping interferograms.<br />

31


2. Lateral shearing interferometer design<br />

A polarized source is desirable to keep the number of optical elements to a minimum<br />

thus reducing undesired reflections, which as we will show in section 3.4.3 can be a<br />

non-negligible source of error in the data processing.<br />

Finally, it would also be desirable to use a source with an intensity profile as uniform<br />

and symmetrical as possible to produce uniform illumination over the cornea and thus<br />

a uniform signal-to-noise ratio across the interferograms. With this in mind we used<br />

a single-mode linearly polarized He-Ne laser with an output of 2.5 mW at 632.8 nm<br />

(Spectra Physics S/N 47075/1919), with a Gaussian intensity profile.<br />

The intensity of the source output is adjusted by using absorption neutral density<br />

filters to keep the power reaching the eye within the safety limits described in Appendix<br />

B.<br />

2.6.2 Spatial filter and telescope<br />

The microscope objective, pinhole, collimating lens and stop aperture shown in figure<br />

2.3 form a telescope to magnify the collimated output of the source and a spatial<br />

filter to produce a smooth intensity profile. The microscope objective is ×40 with a<br />

numerical aperture (NA) 0.65 and a focal length of approximately 3 mm at 632.8 nm,<br />

the pinhole diameter is 10 µm and the collimating lens has a focal length of 250 mm.<br />

In a normal telescope, one would like to match the NAs to make use of all the incoming<br />

light, but in our case, we have chosen the combination of microscope objective and<br />

collimating lens so that the NA of the first one is about 10 times larger than the<br />

second, so that the beam at output of the telescope only takes the central part of the<br />

intensity profile of the laser, to produce a flatter intensity profile. The stop aperture<br />

narrows the beam so that less than the whole diameter of the rest of the optics is<br />

used, providing some tolerance for eye movement.<br />

2.6.3 Polarizing beam splitter and quarter waveplate<br />

The block formed by the polarising beam-splitter (PBS) and the quarter waveplate<br />

serves two purposes, first, to maximize the light reflected back from the eye towards<br />

the imaging branch of the optical setup and second, to minimize the reflections from<br />

the beam-splitter itself.<br />

32


2. Lateral shearing interferometer design<br />

NPBS 50-50<br />

PBS<br />

1<br />

1<br />

2<br />

0.5 x ARC<br />

0.5<br />

1<br />

2<br />

0.5 x ARC<br />

2<br />

0.5 x TR<br />

0.5 x ARC<br />

0.5 x TR<br />

p-polarisation<br />

s-polarisation<br />

ER x ARC<br />

TR<br />

ARC<br />

TR<br />

0.5 x ARC<br />

0.5<br />

ER<br />

Figure 2.4: Relative intensities of reflections from a 50-50 non- polarising beam splitter<br />

(NPBS) and a polarising beam splitter (PBS). ARC is the intensity reflection coefficient<br />

for the antireflection coating, TR is the reflection coefficient of the tear film, and<br />

ER is related to the extinction ratio of the PBS and the polarisation of the incident<br />

beam.<br />

The laser tube was rotated (and thus the polarization) to maximize the intensity<br />

reflected from the PBS towards the eye (s-polarisation). Then, the quarter waveplate<br />

was oriented with its axis at 45 o with respect to the s-polarisation, changing the<br />

polarisation state of the light transmitted towards the eye to circular, so that the<br />

light reflected at the front surface of the tear is fully transmitted through the PBS.<br />

As the reflection on the tear surface occurs at normal incidence, the polarization state<br />

is not altered except for the change in rotation direction.<br />

The minimization of undesired reflections from the beam splitter surfaces is a crucial<br />

point in the experimental setup. The intensity of the light reflected from the tear film<br />

is about 2 % of the incident light (assuming a tear mean refractive index of 1.337 [81].<br />

With reference to figure 2.4 we can see that if we used a non-polarising beam splitter<br />

(NPBS) to combine the illumination and the imaging branches, the relative intensity of<br />

the undesired reflections on the NPBS surfaces would be approximately 2×0.5 2 ×ARC<br />

where ARC is the intensity reflection coefficient of the anti-reflection coating on the<br />

surfaces, which in the best case scenario (laser-line matched) is 0.001. The ratio of<br />

the intensities reflected from the eye to the undesired reflections from the NPBS is<br />

about 20, which is very poor. If on the other hand we use the combination of a PBS<br />

and a quarter waveplate the intensity of the reflections is a bit more complicated to<br />

calculate because it depends on both the extinction ratio of the beam-splitting surface<br />

and the polarisation of the incident beam, which in figure 2.4 we have represented as<br />

33


2. Lateral shearing interferometer design<br />

ER. The experimentally measured value for the ratio of the light reflected from the<br />

tear to that reflected from the PBS is about 2400 that represents an improvement<br />

of two orders of magnitude in intensity and one in SNR. This ratio of intensities is<br />

a good value from the point of view of the detection, because with the camera that<br />

we used the contrast of the undesired interference pattern that would result from the<br />

light reflected from the eye and the undesired reflections from the PBS, is comparable<br />

(a factor of 2 larger) to the readout noise of the camera. Ideally one would make the<br />

contrast between the beam of interest and the undesired reflections equal to or smaller<br />

than the noise of the camera to record the interferogram.<br />

2.6.4 Focusing lens<br />

The final part of the illumination branch of the experimental setup to consider is the<br />

positive lens that makes the beam directed towards the eye converging, focusing at<br />

the center of curvature of the cornea. The two main issues to consider here are the<br />

eye clearance for comfort and the size of the illuminated area.<br />

We considered that the minimum clearance between the closest surface of the lens to<br />

the cornea should be no less than 15 mm; that is the accepted distance for spectacles.<br />

If we bear in mind that the curvature of the wavefront incident on the tear film should<br />

match the curvature of the cornea and the beam before the lens is collimated, then<br />

we have a condition to be met by the focal length f of the lens,<br />

f > 15 mm + R tc (2.17)<br />

where R tc = 8 mm is the radius of a typical cornea, and therefore, f > 23 mm. The<br />

diameter of the illuminated area D illum for a typical cornea is<br />

( ) D<br />

D illum = 2R tc arctan<br />

(2.18)<br />

2f<br />

where D is the minimum of the lens and incoming beam diameters, and f is the<br />

focal length of the lens. This tell us that D illum depends on the F-number of the<br />

lens, assuming the diameter of the incident beam is adjusted accordingly. In the end<br />

we compromised between illuminating the largest possible area over the pupil and<br />

relatively small optics (17 mm clear diameter) in the rest of the experimental setup to<br />

keep costs down, choosing to look at the focal length range 30 − 40 mm. Given the<br />

low F-number, a doublet is preferable to a singlet to keep aberrations low, although<br />

34


2. Lateral shearing interferometer design<br />

this is not crucial because the aberrations of the optical system and corneal surface<br />

are static and will be removed in the data processing. We tried two doublets: an<br />

air-spaced one formed by a 50 mm meniscus and an 80 mm plano-convex lens of the<br />

same material, and an cemented achromat. Also, for the sake of comparison, we<br />

tested a 30 mm biconvex and a 40 mm plano-convex. The shear interferograms they<br />

produce are shown in figure 2.5 and clearly indicate that the preferable lens (in terms<br />

of straighter fringes and hence a smaller spread of the AC and DC spectra)is the 40 mm<br />

achromat, as one would expect. The diameter of the illuminated area for a typical<br />

cornea would then be 3.4 mm if we used the whole clear aperture of the achromat<br />

(17 mm). However, when using the full aperture of the lens, there is no tolerance<br />

of eye misalignment or movement and therefore we traded off some tolerance to eye<br />

misalignment by using only a fraction of the aperture. We have arbitrarily chosen<br />

15 mm of usable aperture which has proven to be reasonable for the tested subjects,<br />

providing ±0.1 mm sideways and ±0.3 mm depth tolerance, without vignetting.<br />

2.7 Imaging branch<br />

The imaging branch consists of: a focusing lens, shared with the illumination branch,<br />

a telescope to demagnify the image of the tear film to a size suitable for the camera;<br />

the purpose-built glass wedges; and two 4f systems to allow the placement of the<br />

wedges. Given a corneal radius of curvature R c , we can calculate the position d of<br />

the planes T ′ and T ′′ conjugated to the front surface of the tear, assuming that to<br />

first order the image is flat (the sag is only 2 % for a typical corneal radius and the<br />

illuminated area in our experiment), by using the formulas<br />

(<br />

d T ′ = f 4 1 − f 2 2 f )<br />

4<br />

f3 2 R<br />

(<br />

d T ′′ = f 6 1 − f 2 2 f 4 2 f )<br />

6<br />

f3 2 f 5 2 R<br />

(2.19)<br />

(2.20)<br />

where the notation is that indicated in figure 2.6. The magnifications M at those<br />

planes are given by<br />

M T ′ = f 2 f 4<br />

f 3 R<br />

(2.21)<br />

M T ′′ = f 2 f 4 f 6<br />

f 3 f 5 R . (2.22)<br />

35


2. Lateral shearing interferometer design<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

Figure 2.5: Interference patterns that result from using different lenses in the illumination<br />

branch of the shear interferometer and an artificial cornea: (a) is from a 30 mm<br />

biconvex lens, (b) from a 40 mm plano-convex lens, (c) from a 30 mm air-spaced doublet<br />

and(d) from a 40 mm achromat.<br />

In theory, we should adjust the camera position along the optical axis for it to be<br />

conjugated to the tear, according to the radius of curvature of the cornea of every<br />

subject, but in practice, we set the experiment assuming a typical corneal radius of<br />

curvature of 8 mm. Under this assumption we choose the focal lengths of the lenses<br />

for the magnification of the imaging system to be unity.<br />

The glass wedges are introduced in the imaging system to produce sheared and tilted<br />

copies of the incident beams by being used in reflection at 45 o . This produces reflections<br />

from the front and back surface of the wedges with almost equal intensity and<br />

thus results in very high contrast interferograms. The first of the wedges is also used<br />

36


2. Lateral shearing interferometer design<br />

T''<br />

CCD<br />

camera<br />

f 6 = 150<br />

CCD<br />

camera<br />

T''<br />

f 6 = 150<br />

f 5 = 150<br />

T'<br />

f 4 = 30 f 3 = 150<br />

f = 40 2<br />

T<br />

wedge 2<br />

wedge 1<br />

300 180<br />

180 190<br />

eye<br />

Figure 2.6: Imaging branches of the double shearing interferometer used for tear<br />

topography estimation. For ease of understanding we have drawn both branches with<br />

their optical axis in the paper plane, but in reality the beam after wedge number 2<br />

and the CCD camera are in the perpendicular plane.<br />

in transmission and we will neglect the undesired reflections in both transmission and<br />

reflection modes, which are around 0.5 % for linear polarisation incident at 45 o with<br />

respect to the reflection plane.<br />

From the imaging point of view, the glass wedges can be thought of as mirrors, and<br />

being so thin (less than a millimeter) we will neglect the optical path difference between<br />

the reflections in their front and back surface.<br />

As explained previously, we need to produce two lateral shearing interferograms with<br />

different shear directions. Given the symmetry of the camera pixels and for ease of<br />

data processing we have chosen orthogonal directions. In the diagram 2.6 we show<br />

both of the branches after the wedges in the same plane for ease of visualization, but<br />

in the experimental setup one of the two is in the plane perpendicular to the paper,<br />

so that the shears introduced are orthogonal. If the branches were left as shown,<br />

two synchronized cameras would be needed. Instead both branches were folded using<br />

mirrors at 45 o making the optical axis go along the edges of a rectangular prism, thus<br />

superimposing the image planes T ′′ (figure 2.7).<br />

37


P<br />

P<br />

2. Lateral shearing interferometer design<br />

CCD camera<br />

wedge<br />

P<br />

P<br />

P<br />

PP<br />

P<br />

He-Ne (632.8nm)<br />

ND filter<br />

microscope objective<br />

spatial filter<br />

P<br />

l/2 plate<br />

wedge<br />

aperture stop<br />

PBS<br />

l/4 plate<br />

eye<br />

Figure 2.7: Sketch of the 3 dimensional optical setup for tear topography estimation<br />

with double lateral shearing interferometry. After the light is reflected back from the<br />

front surface of the tear, the first glass wedge produces two horizontally sheared and<br />

tilted copies of the incident beam by reflection and a third copy by transmission. The<br />

second wedge allows the first pair of copies of the beam to go through unchanged,<br />

while on reflection produces a second pair of copies of the beam carrying information<br />

from the eye, sheared and tilted in the perpendicular direction.<br />

Even though so far the imaging system has been considered in a conventional way, it<br />

is very important to notice that there is only one ray to be traced from each point at<br />

the tear surface because the front surface of the tear behaves like a mirror and the<br />

illumination is such that there is only one ray reaching each point of it. Geometrically<br />

this would give an infinite depth of focus, however, diffraction makes the depth of<br />

focus finite. Without entering into the problem of how to define depth of focus in this<br />

experiment, we will considered that in our experimental setup it is large enough to<br />

allow for all the radii of the corneas imaged to be in focus.<br />

2.7.1 Shear and tilt estimation<br />

Using the paraxial equations, we will now estimate the tilts t 1f and t 2f and shears<br />

s 1f and s 2f introduced by the wedges in the camera plane from the tilts t 1 and t 2 and<br />

shears s 1 and s 2 introduced by the wedges. It can be shown that the tilts between the<br />

38


2. Lateral shearing interferometer design<br />

front and back reflections on the wedge surfaces at the camera plane are given by,<br />

( )<br />

f5<br />

t 1f = t 1 (2.23)<br />

f 6<br />

and that the shears are<br />

s 1f = f 5 t 1 + f 6 t 1 +<br />

t 2f = t 2 (2.24)<br />

( ) ( ) ( )<br />

f5 t 1<br />

f6 f6 t<br />

d T ′′ − √ 1<br />

s 1 − d 5 (2.25)<br />

f 6 f 5 2 f 5<br />

s 2f = s 2<br />

√<br />

2<br />

+ t 2 d 6 (2.26)<br />

where d 5 is the distance from the first wedge to the lens labelled f 5 and d 6 is the<br />

distance from the second wedge to the camera. These equations show that by sliding<br />

the optical wedges along the optical axis, the shear amplitude can be tuned, and this<br />

was used to achieve the desired value of 5 % of the beam diameter.<br />

2.7.2 Polarisation control for optimization of interferogram contrast<br />

When observing the interferograms produced by the experimental setup described<br />

above, one can observe a clear difference in their intensities.<br />

This is due to the<br />

dependence of the reflection coefficients of the glass wedges on the incident state of<br />

polarisation. It is therefore necessary to calculate the interferogram intensities and<br />

contrasts as a function of the incident state of polarisation. The Fresnel reflection r<br />

and transmission t coefficients related to air-wedge transitions will be denoted with<br />

the subindex a while the inverse transitions will be denoted with the subindex w. We<br />

can calculate the relative amplitudes of the reflected beams from the front and back<br />

surface of the first glass wedge E 1F and E 1B respectively. Assuming an incident beam<br />

with unit amplitude and elliptical polarization described by the angles θ (azimuth)<br />

and β (ellipticity) and ignoring multiple reflections, the coefficients are given by<br />

⃗ E 1F = (t ‖a t ‖w r ⊥a cos θ, t ⊥a t ⊥w r ‖a e iβ sin θ) (2.27)<br />

⃗ E 1B = (t ‖a t ‖w t ⊥a r ⊥w t ⊥w cos θ, t ⊥a t ⊥w t ‖a r ‖w t ‖w e iβ sin θ) (2.28)<br />

Similarly for E 2F and E 2B ,<br />

⃗ E 2F = (t ‖a t ‖w r ⊥a e iβ sin θ, t ⊥a t ⊥w r ‖a cos θ) (2.29)<br />

⃗ E 2B = (t ‖a t ‖w t ⊥a r ⊥w t ⊥w e iβ sin θ, t ⊥a t ⊥w t ‖a r ‖w t ‖w cos θ ) (2.30)<br />

39


2. Lateral shearing interferometer design<br />

1.001<br />

C 12<br />

0.16<br />

I 12<br />

1<br />

C 34<br />

0.14<br />

I 34<br />

inteferogram contrast<br />

0.999<br />

0.998<br />

0.997<br />

inteferogram intensity<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.996<br />

0.02<br />

0.995<br />

0 50 100 150<br />

polarization orientation (deg)<br />

0<br />

0 50 100 150<br />

polarization orientation (deg)<br />

(a)<br />

(b)<br />

Figure 2.8: Effect of the orientation of the polarisation state incident on the first glass<br />

wedge on the intensity and contrast of the interferograms produced by the glass wedges<br />

in the optical system illustrated in figure 2.7. Notice there is virtually no dependence<br />

for the visibility but there is a very significant dependence for the intensity.<br />

where it should be noted that due to the way the interferometer is set, the axes<br />

of the system are swapped between the wedges by the reflections from two mirrors<br />

between the first and second wedge. The first thing to notice from examination of<br />

equations 2.27 to 2.30 is that neither the contrast nor the intensities will depend on the<br />

ellipticity β. Thus, the only effect of polarisation on both the contrast and intensity<br />

of the interferograms will arise from the azimuth of the polarisation state, suggesting<br />

that we only need to rotate the polarisation state incident on the first wedge with, for<br />

example, a linear retarder of half a wavelength to tune these parameters. Figure 2.8<br />

(a) shows that the contrast is almost maximum for any polarisation state, while the<br />

situation is very different for the interferogram intensities shown in (b), where there<br />

are two orthogonal directions in which the intensities will be equalized, although not<br />

maximized. A half-wave plate was placed in the optical system just before the first<br />

glass wedge as shown in figure 2.7, so that by rotating this plate the intensity of both<br />

lateral shearing interferograms were equalized.<br />

40


Chapter 3<br />

Data processing<br />

The methods used to process the recorded data, from the raw interferograms to the<br />

topography slope maps are now presented and discussed. The algorithms described<br />

here might not be optimum either mathematically or computationally, but they keep<br />

the complexity to a minimum and, within reasonable levels, minimize the user input<br />

(automation). The quantitative use of an interferometric technique to study the tear<br />

film topography is already a novel field, and we think that by keeping the data processing<br />

simple, not only will this help in its understanding, but it will also keep artifacts<br />

to a minimum. We have sometimes, therefore, traded performance and accuracy for<br />

simplicity.<br />

Before processing, a visual inspection of all the raw interferograms was made, and<br />

those with vignetting, a blink, or where the eye was too far (or too close) to the<br />

experimental setup were tagged as non-usable. From now on we will refer to all the<br />

other interferograms as usable, and these are the only ones to be processed as described<br />

below.<br />

3.1 Carrier frequency estimation<br />

3.1.1 Theory<br />

The first step of the data processing is the estimation of the carrier frequency ⃗ f c of<br />

the interferograms described by equation 2.3 that we recall here once more<br />

I(⃗r) = I o (⃗r) + I o (⃗r + ⃗s) + 2 √ [<br />

I o (⃗r) I o (⃗r + ⃗s) cos φ(⃗r + ⃗s) − φ(⃗r) + 2πf ⃗ ]<br />

c · ⃗r .<br />

41


3. Data processing<br />

If we look at the square modulus of the Fourier transform of the intensity of the<br />

interferograms (spectra), under certain conditions (smooth tear topography) we should<br />

be able to identify three peaks. The major central peak corresponds to the DC term,<br />

and the two other peaks, distributed symmetrically with respect to the zero frequency<br />

correspond to the AC term.<br />

Ideally, we would like to estimate the frequency of the intensity modulation ⃗ f c , produced<br />

by the tilt between the two interfering wavefronts. However, tilt between the<br />

two interfering wavefronts is not the only source of linear phase terms in lateral shearing<br />

interferograms. Both defocus and astigmatism would produce linear terms as well<br />

with an associated frequency ⃗ f lin . This can be understood by looking back at equation<br />

2.15 where the phase difference term is to first order a derivative, hence, as defocus<br />

and astigmatism are second order polynomials, their derivative will produce linear<br />

terms. Actually, second order polynomials will not be the only ones leading to linear<br />

terms in the phase difference term, but we assume these are the dominant ones,<br />

as is corroborated later with the experimental data. To estimate absolute values of<br />

defocus and astigmatism in the tear topography would require calibrating the carrier<br />

frequency. In our case this is not important, because we are not interested in measuring<br />

the absolute topography of the tear film, only its changes with respect to an<br />

initial topography map. We will therefore ignore the offset in the carrier frequency estimation<br />

due to the presence of defocus and astigmatism in our optical setup, corneal<br />

topography and eye misalignment.<br />

3.1.2 Algorithm<br />

The algorithm for the estimation of the carrier frequency has two steps: calculating<br />

the modulus of the spectrum of the raw interferogram and then looking for the spectra<br />

maximum in a region of the Fourier space that excludes the DC term. The coordinates<br />

of this maximum with respect to the origin of coordinates in the Fourier space<br />

is what from now on we will refer to as the estimated carrier frequency. As we have<br />

just explained, we should be aware that the algorithm will not retrieve the frequency<br />

associated with the tilt between the wavefronts, but the vectorial sum of the carrier<br />

frequency and the linear term in the difference of the sheared wavefronts. This algorithm<br />

relies on the implicit assumption that there is no other spatial frequency in the<br />

AC term that has more energy than that corresponding to the sum f ⃗ c + f ⃗ lin .<br />

42


3. Data processing<br />

3.1.3 Implementation details<br />

In practice the algorithm calculates the square modulus of the discrete Fourier transform<br />

(DFT) of the raw interferograms, and then performs the search of the maximum<br />

value over a semi-annular region centered at the zero spatial frequency, with the inner<br />

and outer radii corresponding to the limits of the range we expect the carrier<br />

frequency to have. This search range was set to spatial frequencies with wavelengths<br />

between 5 and 17 image pixels. The maximum of the AC term is only found within<br />

pixel accuracy, sub-pixel accuracy by spectra interpolation was not pursued. The only<br />

consequence of this coarse estimation is to leave a residual tip/tilt on the retrieved<br />

phase maps that is removed later.<br />

3.1.4 Tests<br />

To illustrate the performance and limitations of the algorithms described in this chapter,<br />

we will show the results of their application to 8 interferograms that display the<br />

different features we have found in the almost 5000 usable interferograms. These results<br />

are by no means an exhaustive test, although they provide a good sample of all<br />

the possible scenarios that could be found.<br />

Figure 3.1 shows a series of raw interferograms and the corresponding frontal view<br />

of the 3 dimensional plot of the calculated spectra. The region of the spectra within<br />

the red vertical lines was excluded from the search for the maxima (inner diameter<br />

of the maximum search area), and the DC term can be clearly recognized in it for<br />

all figures. However, the AC terms in the far left and right of the spectra plots<br />

can be difficult to identify in some of the interferograms, because when the phase or<br />

intensity of the interferogram contains rapid variations, the spectra of the DC and AC<br />

term broaden and overlap, as can be seen in figures 3.2 b), f) and h). These figures<br />

illustrate two very important problems that are critical for automated processing of<br />

the data. First, the algorithm will always find a maximum which might not be related<br />

to the modulation frequency, but to other features on the interferograms, such as<br />

the speckle grain size in figures 3.2 f) and h). Second and more important is the<br />

issue of the spectra overlapping, since the method we use to retrieve the phase of<br />

the interferogram is based on the assumption that the DC and AC spectral terms do<br />

not overlap. When this is not the case there will be a non-linear influence of the DC<br />

43


3. Data processing<br />

term on the retrieved phase. We should therefore either visually inspect the spectra<br />

of all the interferograms to be processed, or find an automated method to decide on<br />

whether or not we consider the DC and AC spectra do overlap, to discard them as<br />

not usable 1 .<br />

3.2 Shear estimation<br />

3.2.1 Theory<br />

Let us assume that we have an image that consists of the sum of the intensities of<br />

two copies of the same object A (represented by a triangle in figure 3.3) shifted with<br />

respect to each other by a vector ⃗s. The normalised autocorrelation C AA of this image<br />

can be defined as<br />

∫ ∫ [ A( r ⃗′ ) + A( ⃗ [<br />

r ′ + ⃗s)]<br />

A( r ⃗′ − ⃗r) + A( ⃗ ]<br />

r ′ + ⃗s − ⃗r) d 2 r ′<br />

C AA (⃗r; ⃗s) =<br />

∫ ∫ [ A( r ⃗′ ) + A( r ⃗ ] 2<br />

(3.1)<br />

′ + ⃗s) d 2 r ′<br />

where we assume that A is a real function, and in practice will only take non-zero<br />

values over a finite range. For an autocorrelation, we expect to get a maximum value<br />

when ⃗r is zero, but in this particular case, we will also find that C AA has two other<br />

peaks at ⃗r = −⃗s and ⃗r = ⃗s. These peaks correspond to the situation illustrated in<br />

figure 3.3 b) and c) respectively. Thus, the position of the side peaks are at ± the<br />

shear. This is the idea behind the algorithm used to estimate the shear between the<br />

pupils in the shearing interferograms.<br />

3.2.2 Algorithm<br />

The shearing interferograms are more complex than the illustration of figure 3.3, they<br />

also contain the interference (AC) term. The autocorrelation of such interferograms,<br />

has multiple side fringes with the same spacing as the fringes in the interferogram as<br />

illustrated by figure 3.4 (a) and (b). If we low-pass filter the raw interferogram (c) to<br />

remove the AC term, one could resolve central peak and two side peaks (f). Although<br />

1 If we were to automate the evaluation of the spectra overlapping, we could use as a first step<br />

the second order moment of the area around the estimated maximum, as defined in section 6.5 and<br />

then setting a threshold value to determine when the AC term corresponds to a relatively smooth<br />

topography.<br />

44


3. Data processing<br />

mk1/46<br />

(a)<br />

(b)<br />

st3/15<br />

(c)<br />

(d)<br />

im4/17<br />

(e)<br />

(f)<br />

pb2/92<br />

(g)<br />

(h)<br />

Figure 3.1: Estimation of the interferogram modulation carrier frequency intensity<br />

modulation (left) from spectra (right). The frequencies between the red lines indicate<br />

the area where we expect to have only the DC term. In the top row, we see a normal<br />

smooth front tear surface, the second row shows a rough surface typical after blink, in<br />

the third small bubbles on the tear surface can be noticed and at the bottom a clear<br />

bump in the tear formed by the one of the eyelids is shown.<br />

45


3. Data processing<br />

mk1/33<br />

(a)<br />

(b)<br />

kh3/47<br />

(c)<br />

(d)<br />

te2/49<br />

(e)<br />

(f)<br />

te2/13<br />

(g)<br />

(h)<br />

Figure 3.2: As figure 3.1. The top row shows a very undulated tear surface due<br />

to blink prevention, the following row corresponds to a tear topography with a hole<br />

(break-up) and the two bottom ones are interferograms produced by extremely rough<br />

tear surfaces in front of contact lenses.<br />

46


3. Data processing<br />

y<br />

y<br />

y<br />

x<br />

x<br />

s<br />

x<br />

-s<br />

(a)<br />

(b)<br />

(c)<br />

Figure 3.3: Geometry of autocorrelation of an image consisting of two sheared copies<br />

of an object with finite dimensions: a) the image to autocorrelate, b) and c) the<br />

situations in which the autocorrelation has local maxima other than at the origin.<br />

the peaks are visible, they are not strong compared with the background, and in order<br />

to improve this and therefore the sensitivity of the algorithm, we also remove the very<br />

low spatial frequencies of the object. We therefore virtually retain only the edges of<br />

the figure as illustrated in figure 3.4 (e). This stage requires the non-overlapping AC<br />

and DC spectra.<br />

3.2.3 Implementation details<br />

The interferogram band-pass filtering and the autocorrelation calculation were performed<br />

using the DFT and the autocorrelation theorem. There are two advantages<br />

of using the DFT rather than the straightforward calculation of the autocorrelation<br />

here: firstly a significant speed increase by making use of the fast implementations of<br />

the DFT and secondly, the bandpass filtering in the Fourier domain becomes simply<br />

a multiplication by a binary mask.<br />

Before using the DFT we pad the data with zeros so that the area of the padded data<br />

is no less than twice the size of the interferogram plus the shear to avoid overlapping<br />

of the periodized autocorrelation (resulting from using the DFT).<br />

The mask that performs the band-pass filtering is simply a binary annular mask. The<br />

inner radius was chosen so that the spatial frequencies within the shear search range<br />

[s min , s max ] would be preserved and all the lower frequencies removed. The outer<br />

radius was chosen to keep all the frequencies within the carrier frequency search range<br />

and remove those above it.<br />

Finally, once the correlation of the filtered interferograms has been evaluated, the<br />

47


3. Data processing<br />

mk1/46<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

(e)<br />

(f)<br />

Figure 3.4: Improvement on shear estimation algorithm by filtering in Fourier domain:<br />

(a) raw shearing interferogram, (c) and (e) the same interferogram after low- and bandpass<br />

filtering respectively, (b), (d) and (f) show frontal views of the corresponding<br />

autocorrelation plots. The position of the two side peaks with respect to the central<br />

peak in the autocorrelation plot in (d) is the shear.<br />

48


3. Data processing<br />

maximum value within the annular section s min < r < s max is found. The position<br />

of the maximum was only determined with 1 pixel accuracy. Sub-pixel search of<br />

the maximum by interpolation would not provide any advantage in our case because<br />

the algorithm that reconstructs the wavefront from the slope of the phase maps only<br />

uses integer values of shear amplitude. If we assume that the error in the maximum<br />

position determination is half a pixel, then the relative error for a typical shear value<br />

(30 pixels) is about 2%. In practice the error of the algorithm is higher, due to noise<br />

in the image acquisition and pixelation.<br />

3.2.4 Tests<br />

The value of the autocorrelation at the side peaks can not be determined theoretically<br />

unless one knows the image function A and the shear ⃗s. However, we can estimate it<br />

roughly if we use the fact that the actual overlapping of the images over areas with<br />

non-zero values is small in comparison with the non-overlapping areas. Evaluating<br />

formula 3.1 for ⃗r = −⃗s (and expanding the denominator) gives,<br />

∫ ∫ [<br />

C AA (−⃗s; ⃗s) = 1 A( r ⃗′ )A( r ⃗′ + ⃗s) + A( r ⃗′ )A( ⃗ ]<br />

r ′ + 2⃗s) d 2 r ′<br />

2 + 2 ∫ ∫ [ A 2 ( r ⃗′ ) + A( r ⃗′ )A( r ⃗ ] , (3.2)<br />

′ + ⃗s) d 2 r ′<br />

which shows that if the overlap between the two sheared images is small, then the value<br />

of the normalized autocorrelation at the side peaks is around 0.5. This is the case<br />

for our filtered interferograms that only have non-negligible values at the edges of the<br />

pupil and tear features. This can be clearly seen in figures 3.5 and 3.6, where the first<br />

columns show raw interferograms, the images in the second columns are the band-pass<br />

filtered interferograms and the three dimensional plots are the central region of the<br />

autocorrelation. By comparing the values of the autocorrelation side peaks indicated<br />

in the figures, and the filtered interferograms, it can be seen that other than when the<br />

tear surface is so rough that the fringes can not be recognized, the values at the side<br />

peaks are close to 0.5. More interesting is that even in the situation when the fringes<br />

can not be identified at all, the side peaks seem clearly identifiable, illustrating the<br />

robustness of the algorithm.<br />

49


3. Data processing<br />

0.43<br />

mk1/46<br />

(a)<br />

(b)<br />

(c)<br />

0.39<br />

st3/15<br />

(d)<br />

(e)<br />

(f)<br />

0.42<br />

im4/17<br />

(g)<br />

(h)<br />

(i)<br />

0.46<br />

pb2/92<br />

(j)<br />

(k)<br />

(l)<br />

Figure 3.5: Shear estimation algorithm applied to tear film interferograms with different<br />

features. The first column contains the input images (raw interferograms), the<br />

second shows the band-pass filtered images and the third one is the central region of<br />

the autocorrelation of the images on the second column.<br />

50


3. Data processing<br />

0.44<br />

mk1/33<br />

(a)<br />

(b)<br />

(c)<br />

0.43<br />

kh3/47<br />

(d)<br />

(e)<br />

(f)<br />

0.32<br />

te2/49<br />

(g)<br />

(h)<br />

(i)<br />

0.23<br />

te2/13<br />

(j)<br />

(k)<br />

(l)<br />

Figure 3.6: As figure 3.5.<br />

51


3. Data processing<br />

3.3 Pupil position and size estimation<br />

There are three parameters to estimate from the interferograms that are crucial for<br />

the phase map estimation and wavefront integration algorithms: the position, size and<br />

shape of our region of interest (pupil) on the raw interferograms. Ideally, the pupils at<br />

the detector plane would be circular. However, in practice the recorded interferograms<br />

have shapes which are not exactly circular, sometimes clearly elliptical, due to corneal<br />

astigmatism, and some times very irregular due to very rough tear surface and tear<br />

break-ups. Despite this asymmetry, to first approximation, the pupils can in general<br />

be modelled as circular and the algorithm proposed below is based on this assumption.<br />

3.3.1 Algorithm<br />

To estimate the pupil position and radius, a normalised correlation between a binarised<br />

low pass-filtered version of the raw interferograms and a model of a pupil was<br />

maximized, by varying the parameters of the model.<br />

The raw interferograms are low-pass filtered to remove the AC term and then binarised<br />

to eliminate the non-uniformities in illumination, then they are correlated with a series<br />

of models consisting of binarised sums of pairs of circles shifted with respect to each<br />

other by the estimated shear, with different radii (r model ). Finally, we take the value<br />

of r model that maximizes the normalised correlation as the pupil size.<br />

3.3.2 Implementation details<br />

Again, as with the shear estimation, it is convenient for speed and ease of filtering to<br />

use the DFT and the correlation theorem, again padding as before. The low-pass filter<br />

is a circular binary mask in the Fourier space, with a radius equal to the minimum<br />

modulation frequency we considered for all the algorithms.<br />

The intensity profiles of the interferograms change from one series of data to the next,<br />

due to fluctuations of intensity of the source, eye position, etc., requiring different<br />

threshold values to binarize the low-pass filtered images. Despite several attempts<br />

to find a satisfactory automated method for determining the threshold value (e.g.<br />

histogram based) that was able to cope with diffraction rings, tear film features,<br />

52


3. Data processing<br />

etc., we ended up having to set this value manually for each individual series of<br />

interferograms.<br />

The algorithm calculates the peak of the correlation between the binarised low-pass<br />

filtered interferogram and the pupil model by changing the radius of the model r model<br />

in steps of one pixel. In order to reduce the number of computations performed,<br />

the algorithm does not actually calculate the correlation for each value in the range<br />

passed to the algorithm. Instead, two series of calculations are performed, first a coarse<br />

estimation in which the step size between consecutive values of r model is set to √ N,<br />

where N is the number of integers in the radii search range, and a second, in which<br />

the step is set to one, and a smaller range of width 2 √ N centered in the value of the<br />

coarse estimation that yielded the maximum normalised correlation. This works on<br />

the assumption of the correlation being a slow-varying function of the value r model , and<br />

reduces the number of correlation calculations from N to 2 √ N. In a typical situation,<br />

the size of the search range needed to be around 100 pixels, and thus the reduction<br />

on the number of calculations achieved by this two-step calculation was 5. As in the<br />

shear estimation algorithm, we make no attempt to achieve sub-pixel position or size<br />

resolution because the integration algorithm is only applicable to integer values.<br />

3.3.3 Tests<br />

In order to illustrate the performance of the method, we show in figures 3.7 and<br />

3.8 raw interferograms with the edge of the pupils that give the highest normalised<br />

correlation indicated by two red circles, the corresponding binarized low-pass filtered<br />

interferograms used for the correlation calculation and the calculated correlation peak<br />

values for the two-step algorithm.<br />

It can be seen from the figures, that the model seems adequate, both visually on the<br />

left column and numerically on the correlation peak value on the right column, but<br />

it is also clear that there is room for improvement both in the pupil model and the<br />

thresholds election. A further step to improve the algorithm would be to consider<br />

fitting elliptical pupils, though the dimension of the search parameter space would<br />

then increase from one to three (ellipse major and minor radii and orientation).<br />

53


3. Data processing<br />

(a)<br />

(d)<br />

(g)<br />

(j)<br />

mk1/46<br />

st3/15<br />

im4/17<br />

pb2/92<br />

(b)<br />

(e)<br />

(h)<br />

(k)<br />

normalised correlation<br />

normalised correlation<br />

normalised correlation<br />

normalised correlation<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

1 1.2 1.4 1.6<br />

pupil radius (m m)<br />

(c)<br />

1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9<br />

pupil radius (m m)<br />

(f)<br />

1.2 1.3 1.4 1.5 1.6 1.7 1.8<br />

pupil radius (m m)<br />

(i)<br />

1.2 1.3 1.4 1.5 1.6 1.7 1.8<br />

pupil radius (m m)<br />

(l)<br />

Figure 3.7: Center and radius estimation algorithm applied to tear film interferograms<br />

with different features. The first and second columns show the estimated pupils (red<br />

circles) superimposed on the raw interferograms and corresponding binarised low-pass<br />

filtered interferograms respectively. On the right we can see the calculated peak values<br />

of the normalised correlation within the r model search range. The values from the first<br />

step of the search algorithm are plotted in blue and those on the second step in red.<br />

54


3. Data processing<br />

1<br />

0.9<br />

0.8<br />

normalised correlation<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

(a)<br />

mk1/33<br />

(b)<br />

0.1<br />

1<br />

1 1.2 1.4 1.6<br />

pupil radius (m m)<br />

(c)<br />

0.9<br />

0.8<br />

normalised correlation<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

(d)<br />

kh3/47<br />

(e)<br />

0.1<br />

1<br />

1.2 1.3 1.4 1.5 1.6 1.7 1.8<br />

pupil radius (m m)<br />

(f)<br />

0.9<br />

0.8<br />

normalised correlation<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

(g)<br />

te2/49<br />

(h)<br />

0.1<br />

1<br />

1.2 1.3 1.4 1.5 1.6 1.7 1.8<br />

pupil radius (m m)<br />

(i)<br />

0.9<br />

0.8<br />

normalised correlation<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

(j)<br />

te2/13<br />

(k)<br />

0.1<br />

1.2 1.3 1.4 1.5 1.6 1.7 1.8<br />

pupil radius (m m)<br />

(l)<br />

Figure 3.8: As figure 3.7.<br />

55


3. Data processing<br />

3.4 Phase recovery<br />

The phase recovery algorithm we now describe is that proposed by Takeda [95], to<br />

extract the phase map of an interferogram with tilt fringes.<br />

3.4.1 Algorithm<br />

Let us begin by rewriting the equation describing the intensity of the interferograms<br />

(Eq. 2.3) in terms of complex exponentials for convenience,<br />

I(⃗r) = I o (⃗r) + I o (⃗r + ⃗s)<br />

+ e +2πi f ⃗ √ c·⃗r +i [φ(⃗r+⃗s)−φ(⃗r)]<br />

I o (⃗r) I o (⃗r + ⃗s) e<br />

+ e −2πi ⃗ f c·⃗r √ I o (⃗r) I o (⃗r + ⃗s) e −i [φ(⃗r+⃗s)−φ(⃗r)] (3.3)<br />

Recalling the Fourier shift theorem and assuming that the carrier frequency is such<br />

that it separates the DC term from the AC term in the Fourier domain, we filter one<br />

of the AC terms in the Fourier domain simply by masking out the other terms, and<br />

then remove the tilt associated with the carrier frequency (using the shift theorem<br />

again) by shifting the AC term to the origin of coordinates of the Fourier domain.<br />

The resulting intensity I Takeda is now the complex magnitude<br />

I Takeda (⃗r) = √ I o (⃗r) I o (⃗r + ⃗s) e ± i [φ(⃗r+⃗s)−φ(⃗r)] (3.4)<br />

were the plus or minus sign depends on which of the AC terms is masked out. If we<br />

now take the complex logarithm of both sides of the equation, and then retain the<br />

imaginary part, we can separate the phase information,<br />

W {± [φ(⃗r + ⃗s) − φ(⃗r)]} = I [ln I Takeda (⃗r)] (3.5)<br />

from the magnitude information, obtained by taking the modulus of equation 3.4<br />

√<br />

Io (⃗r) I o (⃗r + ⃗s) = |I Takeda (⃗r)| (3.6)<br />

Note that the recovered estimated phase is wrapped (W) in the interval [−π, π].<br />

3.4.2 Implementation details<br />

The algorithm calculates the DFT of the raw interferograms, padded as before, then<br />

everything except for one of the AC terms is masked out with a circular binary mask,<br />

56


3. Data processing<br />

centered at the carrier frequency (or its negative) and with radius half of the carrier<br />

frequency. The remaining AC term is translated to the center of the Fourier domain<br />

and then the inverse DFT is computed to obtain I Takeda .<br />

Finally the phase and<br />

magnitude maps are obtained by taking the imaginary part of the complex logarithm<br />

and the modulus respectively.<br />

Note that is has been assumed that the DC and AC terms do not overlap. Overlapping<br />

would give rise to aliasing both in the phase and intensity maps, thus producing<br />

misleading results. The algorithm should not be used blindly, but accompanied with<br />

some estimation of the overlapping of the DC and AC terms in the Fourier domain.<br />

3.4.3 Noise introduced by undesired reflections<br />

Two sources of noise in the experiment are the unwanted reflections from the PBS and<br />

the glass wedges. We now estimate the magnitude of undesired reflections in Takeda’s<br />

phase recovery method.<br />

We can describe the electric field at the output of the tear film experiment as the<br />

superposition of two sheared and tilted copies of the electric field to be studied (E T )<br />

and the undesired reflections (E R ). The observed intensity at the output of the interferometer<br />

will then be<br />

I T+R (⃗r) =<br />

∣ E T (⃗r) + E T (⃗r + ⃗s) e 2πi ⃗ f·⃗s + E R (⃗r) + E R (⃗r + ⃗s) e 2πi ⃗ f·⃗s<br />

∣ 2 (3.7)<br />

where we assume that the source has a coherence length greater than the length of<br />

the optical system. If we now apply Takeda’s algorithm, we get the recovered phase<br />

ψ rec<br />

ψ rec (⃗r) = arctan<br />

{ }<br />

IT,T sin ψ T,T + I R,R sin ψ R,R + I R,T sin ψ R,T + I R,T sin ψ T,R<br />

I T,T cos ψ T,T + I R,R cos ψ R,R + I R,T cos ψ R,T + I R,T cos ψ T,R<br />

(3.8)<br />

where I a,b (⃗r) = √ I a (⃗r)I b (⃗r + ⃗s) and φ a,b = φ a (⃗r) − φ b (⃗r + ⃗s). If the amplitude of<br />

the undesired reflections is much smaller than the amplitude of the light reflected<br />

from the tear film, i.e. I R /I T ≪ 1 and it is assumed that the intensity profiles are<br />

approximately constant over the pupil, then, to first order, we can neglect the terms<br />

57


3. Data processing<br />

in I R,R and approximate the recovered phase by<br />

ψ rec (⃗r) ≈ ψ T,T (3.9)<br />

( )<br />

IR,T<br />

+ [(sin ψ R,T + sin ψ T,R ) cos ψ T,T − (cos ψ R,T + cos ψ T,R ) sin ψ T,T ]<br />

I T,T<br />

If we now define the phase error ψ error as the difference between the recovered phase<br />

and ψ T,T , it is possible to show that this error is bounded by<br />

( )<br />

IR,T<br />

max {|ψ error |} ≈ 2 . (3.10)<br />

I T,T<br />

If we now assume that the scale of the spatial intensity fluctuations is smaller than<br />

the shear amplitude (i.e. I a (⃗r) ≈ I a (⃗r + ⃗s)), then to first order<br />

√<br />

I R,R<br />

max {|ψ error |} ≈ 2<br />

(3.11)<br />

I T,T<br />

We can now use this approximate formula and the measured value for (I R,R /I T,T ) of<br />

about 1/2000 to see that the reflections introduced by the polarizing beam-splitter in<br />

our experimental setup would introduce a maximum error in the reconstructed phase<br />

of around 0.05 radians. Of more concern though are the unwanted reflections from the<br />

glass wedges which have values I R /I T ≈ 1/210, producing a maximum phase error<br />

of about 0.14 radians, equivalent to a wavefront error of λ/50.<br />

It is important to<br />

remember that this error is not on the tear topography but on the phase difference<br />

maps, that to first order are gradients of the tear topography.<br />

3.4.4 Tests<br />

Takeda’s algorithm produces two outputs from each interferogram, a phase map<br />

wrapped between −π and π and a map containing information on the amplitude<br />

of the electric fields. In principle, one would think that the information of interest<br />

is contained purely within the phase map, but the information in the amplitude map<br />

should not be overlooked, as is now explained. Figures 3.9 and 3.10 show a series<br />

of raw interferograms, and the corresponding normalised amplitude maps (values between<br />

0 and 1), and wrapped phase maps retrieved by the algorithm. Even though<br />

we designed the optical system to achieve almost uniform illumination across the interferograms,<br />

it is clear that the amplitude maps are not so. There are a number of<br />

reasons for this:<br />

58


3. Data processing<br />

• Diffraction rings at the pupil edge (seen on the right side of figure 3.9 b). This<br />

could be reduced by conjugating the aperture stop to the tear film for each<br />

subject, but we believe this is not an important issue. In practice, to reduce the<br />

influence of these rings, the pupil radius used for data processing was 95 % of<br />

the estimated pupil radius.<br />

• Dependence of the tear front surface reflection coefficient on the angle of incidence,<br />

that results in changes in the amplitude when the topography departs<br />

from a smooth spherical surface (3.9 e). This has no implications in the phase<br />

map estimation.<br />

• The steepness of the tear film with respect to a spherical shell can reflect some<br />

of the incident rays out of the optical system (vignetting) producing completely<br />

dark areas in the amplitude map, such as the ones in figures 3.9 h and k. There<br />

is little we can about this, because we can only reduce it by increasing the<br />

numerical aperture of the lens closest to the eye. It is a problem that will always<br />

be present, and will lead to areas with undefined phase.<br />

• When the tear breaks up, the optically rough mucus layer of the tear film is<br />

exposed, and most of the light is scattered out of the optical system, again<br />

producing dark areas (3.10 e).<br />

• Finally, when the optical surface is rough the AC and DC terms overlap considerably,<br />

giving non-linear aliasing of phase and amplitude (3.10 b, h and k). This<br />

effect can only be minimized by the use of a higher intensity modulation frequency,<br />

accompanied by a larger sampling density, but we are currently limited<br />

by the number of pixels in our camera.<br />

If we now look at wrapped phase maps in figures 3.9 and 3.10, we can see that there<br />

are two types of discontinuities, the ones that are a closed line or begin and end at<br />

the pupil edge, and the ones that are lines that begin and/or end within the pupil.<br />

The latter correspond to discontinuities in the phase, and to see how they are related<br />

to vignetting and areas with almost zero or zero intensity, we placed a red spot on<br />

each pixel of the intensity map where the amplitude is less than one per cent of the<br />

maximum value. As we can see in figures 3.9 and 3.10, whenever we have a red spot<br />

indicating low amplitude, we have a matching phase discontinuity in the wrapped<br />

phase map.<br />

59


3. Data processing<br />

Phase discontinuities are physically produced in some situations like atmospheric propagation<br />

or with diffractive elements, but because of surface tension true phase discontinuities<br />

cannot be produced at the tear film surface.<br />

3.5 Phase unwrapping<br />

Phase unwrapping is a subject in itself, not only in optical interferometry, but also<br />

synthetic aperture radar, tomography and nuclear magnetic resonance among other<br />

fields [100]. The basic idea behind many algorithms consists of calculating the wrapped<br />

gradients from the wrapped phase, and then integrating these. The main difficulty in<br />

practise is performing the integration when there are certain types of discontinuities<br />

in the initial data that result in residues in the calculated gradient maps.<br />

3.5.1 Principle<br />

For the sake of simplicity, we have chosen one of the simplest to implement and, we<br />

believe, most intuitive unwrapping algorithm, that is a least-square estimation. The<br />

algorithm finds the phase values ψ unwrapped that minimize the sum of the squared<br />

differences between the data gradients and the gradients of the estimated phase map.<br />

In practise, we do not calculate the exact gradients of the data ψ rec or the estimated<br />

wavefront ψ estimated , instead we calculate the differences between neighboring points,<br />

and minimize the error ɛ 2 defined as<br />

ɛ 2 = ∑ i,j<br />

{[∆ x ψ estimated (i, j) − ∆ x ψ rec (i, j)] 2 + [∆ y ψ estimated (i, j) − ∆ y ψ rec (i, j)] 2}<br />

where ∆ x ψ(i, j) = ψ(i, j) − ψ(i − 1, j) and ∆ y ψ(i, j) = ψ(i, j) − ψ(i, j − 1).<br />

(3.12)<br />

3.5.2 Implementation details<br />

The algorithm performs three steps. First, it calculates the differences, then wraps<br />

the differences in the interval [−π, π], and finally performs the integration using an<br />

optimized least-square integrator described in the next chapter. All that should be<br />

pointed out here is that not all the gradients are given the same weight, the implications<br />

of which is that the solution will deviate slightly from the least square solution<br />

60


3. Data processing<br />

0 0.2 0.4 0.6 0.8 1 -3 -2 -1 0 1 2 3<br />

mk1/46<br />

(a)<br />

(b)<br />

(c)<br />

st3/15<br />

(d)<br />

(e)<br />

(f)<br />

(g)<br />

im4/17<br />

(h)<br />

(i)<br />

(j)<br />

pb2/92<br />

(k)<br />

(l)<br />

Figure 3.9: Outputs of Takeda’s phase recovery method applied to tear interferograms.<br />

The first column is a series of raw interferograms, the central column shows<br />

the normalised amplitude maps, and on the right we show the estimated phase maps<br />

wrapped in the interval [−π, π]. The red spots on the amplitude maps indicate the areas<br />

where the amplitude value is less than one percent of the maximum value, and can<br />

be compared with the phase discontinuities in the same locations in the corresponding<br />

wrapped phase maps.<br />

61


3. Data processing<br />

0 0.2 0.4 0.6 0.8 1 -3 -2 -1 0 1 2 3<br />

mk1/33<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

kh3/47<br />

(e)<br />

(f)<br />

(g)<br />

te2/49<br />

(h)<br />

(i)<br />

(j)<br />

te2/13<br />

(k)<br />

(l)<br />

Figure 3.10: As figure 3.9.<br />

62


3. Data processing<br />

in the presence of noise.<br />

3.5.3 Tests<br />

The algorithm was tested on the wrapped phase maps from figures 3.9 and 3.10, and<br />

the results are shown in figures 3.11 and 3.12. Comparing with figures 3.9 and 3.10, we<br />

see that the phase discontinuities have been smoothed out in the unwrapping process.<br />

63


3. Data processing<br />

8<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

mk1/46<br />

-8<br />

(a)<br />

(c)<br />

st3/15<br />

(b)<br />

(d)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

(e)<br />

im4/17<br />

(f)<br />

8<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

(g)<br />

pb2/92<br />

(h)<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

Figure 3.11: Phase unwrapping applied to different tear topographies. The central<br />

column contains the corresponding unwrapped phase maps corresponding to the interferograms<br />

on the left. Note the difference in the colorbar scales.<br />

64


3. Data processing<br />

8<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

mk1/33<br />

-6<br />

(a)<br />

(b)<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

(c)<br />

kh3/47<br />

(d)<br />

-8<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

(e)<br />

te2/49<br />

(f)<br />

-8<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

(g)<br />

te2/13<br />

(h)<br />

-4<br />

Figure 3.12: As figure 3.11.<br />

65


Chapter 4<br />

Wavefront reconstruction from<br />

shear phase maps<br />

In this chapter we present the algorithm used to reconstruct the tear topography<br />

maps from the estimated finite difference maps extracted from the interferograms as<br />

described in the previous chapter. The need for this algorithm arose due to the impossibility<br />

of using conventional pseudoinverse reconstruction techniques and currently<br />

available computing capabilities, given the large number of samples over which the<br />

tear topography was estimated. The work described in this chapter has been published<br />

in Applied Optics [101]<br />

4.1 Introduction<br />

Discrete zonal wavefront reconstruction from shear phase maps is, in principle, a simple<br />

and well understood linear algebra problem [102, 103, 104]. The forward problem<br />

is modelled as a set of linear equations and arranged in matrix form, where the forward<br />

matrix contains the coefficients of the difference equations. Then, the pseudoinverse of<br />

this matrix is calculated using iterative algorithms [105] and finally the reconstructed<br />

wavefronts are obtained by multiplying the discrete difference data by the calculated<br />

pseudoinverse. This method retrieves the best fit wavefronts in terms of the least<br />

square solution with respect to the data.<br />

In theory, there is no restriction to the problem size, that is, the number of samples over<br />

which the wavefront is to be reconstructed. However, in practice, for large number<br />

66


4. Wavefront reconstruction from shear phase maps<br />

of samples the computing and storing of the pseudoinverse, as well as the matrix<br />

multiplication can be extremely computationally demanding. The need for faster and<br />

less computer-demanding methods for when fast reconstruction is required or when<br />

there are a large number of wavefront samples, led to the formulation of the problem as<br />

a linear filter using fast implementations of the discrete Fourier transform (DFT) [106,<br />

107, 108, 109, 110, 111], grouped under the name of FFTs [112]. The use of the DFT<br />

for wavefront reconstruction from difference maps was first proposed by Freischlad and<br />

Koliopoulos [106]. In their work, they calculate the transfer function corresponding to<br />

problems with particular geometries related to Shack-Hartmann sensors (e.g. Hudgin<br />

[102] and Fried [103]) for a rectangular pupil and a shear amplitude of one sample.<br />

Elster and Weingärtner [108] treated the one dimensional (1D) case in which the<br />

amplitude of the shear s is a divisor of the total number of samples N along the<br />

direction of the shear. They also suggested a 1D data extension method for cases in<br />

which N is not a multiple of s, but inconsistencies between the extended data and the<br />

forward model introduce errors even in the absence of noise. The 2D data extension<br />

for non-rectangular pupils has been approached by Roddier and Roddier [107], who<br />

proposed an iterative solution that can in principle cope with any pupil shape, and<br />

recently by Poyneer et al. [111], who suggested the extension of the data outside a<br />

pupil so that it remains consistent with the data inside the pupil. In both works the<br />

shear amplitude is one. However, in shearing interferometry, it is often desirable to<br />

be able to adjust the shear to trade resolution for signal-to-noise ratio, and thus a<br />

general reconstruction algorithm is desirable.<br />

In this work, we treat the more general case of 2D wavefront reconstruction from<br />

difference maps originated from sheared phase maps, for any integer shear amplitudes<br />

and convex pupils using two different methods. In the first method we enlarge the<br />

problem dimensions by extending the data to a larger domain while keeping it consistent<br />

with the forward model of the problem. In the second, the original problem<br />

is subdivided into a set of smaller problems with shear amplitude one. The solutions<br />

retrieved by either method will have minimum norm, in the sense that all the modes<br />

in the null space of the problem will have zero amplitude. The performance of the<br />

proposed methods is evaluated by calculating their associated noise coefficient.<br />

67


4. Wavefront reconstruction from shear phase maps<br />

4.2 Theory<br />

4.2.1 The transfer function of the reconstruction problem<br />

We will first generalize the linear filter proposed in reference [106] to any integer shear<br />

and to allow for different weights for x- and y-difference data. We want to estimate<br />

a wavefront φ, sampled over a regular grid of points labelled (n, m), inside a pupil<br />

P , from an experimentally measured set of differences S x and S y defined over the<br />

sub-pupils P x and P y contained within the pupil P . Under the forward model, the<br />

phase differences are given by<br />

D x (n, m) = φ(n + s x , m) − φ(n, m), (4.1)<br />

D y (n, m) = φ(n, m + s y ) − φ(n, m). (4.2)<br />

We can now try to estimate the wavefront by least-square fitting these phase differences<br />

to the data over a rectangular grid of dimensions N x × N y , by minimizing the error<br />

E 2 = ∑ n,m<br />

{<br />

|S x (n, m) − D x (n, m)| 2 + |S y (n, m) − D y (n, m)| 2} . (4.3)<br />

If we now assume that the wavefront is periodic (with period N x × N y samples), and<br />

band-limited, it can be proven [106] that minimizing E 2 leads to<br />

[ ( ) [ ( )<br />

exp −2πi k sx<br />

N x<br />

− 1] ˜Sx (k, l) + exp −2πi l sy<br />

N y<br />

− 1] ˜Sy (k, l)<br />

˜φ(k, l) =<br />

( ) ( )] , (4.4)<br />

4<br />

[sin 2 π k sx<br />

N x<br />

+ sin 2 π l sy<br />

N y<br />

where ˜g represents the DFT of g. There are three problems to address in applying<br />

equation 4.4 to estimate φ from the experimental data S x , S y : the assumption of<br />

periodicity on φ; data extension from the domains P x and P y to rectangular grids<br />

that contain them; and the removal of the poles of the filter defined by equation 4.4.<br />

4.2.2 Periodicity<br />

It can be seen from equations 4.1 and 4.2 that for each value of m there are s x less<br />

experimental differences S x than there are values of φ, and similarly for each value of<br />

n there are s y less experimental differences S y than there are values of φ. Therefore,<br />

for a rectangular pupil P with dimensions N x × N y , the dimensions of the pupils P x<br />

and P y will be (N x − s x ) × N y and N x × (N y − s y ) respectively. Thus, we need to<br />

68


4. Wavefront reconstruction from shear phase maps<br />

add s x columns to the array of differences D x and s y rows to the array of differences<br />

D y , while simultaneously being consistent with the forward model and meeting the<br />

periodicity condition. Elster and Weingärtner [108] noticed that when N x is divisible<br />

by s x and N y by s y (which we will henceforth refer to as the divisibility condition)<br />

suitable values for these extra rows and columns can be calculated in terms of the<br />

available data, thus,<br />

S x (N x − s x + k x , m) = −<br />

S y (n, N y − s y + k y ) = −<br />

Nx<br />

sx −1<br />

∑<br />

j=1<br />

Ny<br />

sy −1<br />

∑<br />

j=1<br />

S x (s x (j − 1) + k y , m), (4.5)<br />

S y (n, s y (j − 1) + k y ), (4.6)<br />

for k x = 1, 2, . . . , s x and k y = 1, 2, . . . , s y . It is difficult to extend the wavefront<br />

periodization by Elster and Weingärtner to the cases in which the divisibility condition<br />

is not met, since this would require knowledge of differences not described by equations<br />

4.1 and 4.2, and therefore not available in the data. It would also be undesirable to<br />

use the periodization method described by Ghiglia and Pritt [100], based on mirror<br />

reflections of the initial grid of data, as it quadruples the problem dimension and thus<br />

requires significantly more operations.<br />

4.2.3 Data extension and extension method<br />

We now describe a data extension method for the more general case of a convex pupil<br />

contained within a rectangular pupil, which also solves the cases where the divisibility<br />

condition is not met. Our extension method, is a generalization of that described by<br />

Poyneer et al.[111] for shears of amplitude one. It can be seen that for equations 4.1<br />

and 4.2 to be valid, the following loop continuity condition has to be met<br />

D x (n, m) + D y (n + s x , m) − D x (n, m + s y ) − D y (n, m) = 0. (4.7)<br />

If we consider the difference data as a discrete gradient of the scalar field φ, we can<br />

think of the loop continuity condition as the discrete equivalent of requiring the curl<br />

of the gradient to be zero.<br />

We can now use the loop continuity condition to extend the data outside the pupil, by<br />

copying the difference values in the boundary of the pupil outside the pupil at intervals<br />

69


4. Wavefront reconstruction from shear phase maps<br />

d<br />

e<br />

d<br />

e<br />

b<br />

d<br />

b<br />

e<br />

b<br />

b<br />

d<br />

c<br />

e<br />

b<br />

c<br />

a<br />

d<br />

a<br />

P<br />

a<br />

d<br />

a<br />

P<br />

(a)<br />

(b)<br />

Figure 4.1: Illustration of the method for extending the difference data outside the<br />

pupil, by copying the difference values in the boundary of the pupil outside the pupil.<br />

The black spots are the points over which the wavefront φ is to be evaluated, the gray<br />

area indicates the pupil P and the curved arrows indicate the two values of the grid<br />

that define the difference with coordinates at the origin of the arrow.<br />

given by s x and s y . The method is illustrated with two examples in figures 4.1 (a) and<br />

(b). The black spots are the points at which the wavefront φ is to be reconstructed,<br />

the gray area indicates the pupil P and the curved arrows indicate phase differences<br />

(solid lines measured data, dashed lines extended data). In figure 4.1 (a), the shears<br />

are s x = s y = 1. The available difference data at the boundary of the pupil (indicated<br />

a, b, c, d and e) are repeatedly copied outside the pupil boundary as indicated by the<br />

dashed lines in the figure to the edge of the grid. It can be seen that for any of the<br />

cells outside the pupil the loop continuity is satisfied. In figure 4.1 (b), the shears are<br />

s x = 2 and s y = 1. In this case the x-differences are copied at intervals of s y = 1 as<br />

before, but the y-differences are copied at intervals s x = 2. It can be seen that the<br />

loop continuity condition is satisfied around loops of size s x × s y .<br />

As there are no limits to extending the data using the loop continuity, we can choose<br />

the dimensions of the rectangular grid to satisfy the divisibility condition. Then, the<br />

data can be made consistent with a periodic wavefront using equations 4.5 and 4.6<br />

and the reconstruction problem solved using equation 4.4.<br />

It should be noted that whenever one extends the difference data, other than to meet<br />

periodicity, one is no longer estimating the wavefront by the conventional least square<br />

fit derived from equation 4.3. Instead, one has a weighted least square fit, in which the<br />

data inside the pupil has weight one and the data in the boundary has a weight given<br />

70


4. Wavefront reconstruction from shear phase maps<br />

by the number of copies required for the extension. This suggests that unnecessary<br />

data extension should be avoided.<br />

4.2.4 Filter poles and subdivision method<br />

In this section, addressing of filter poles problem in equation 4.4 will lead us to an<br />

alternative method for approaching the reconstruction problem when the dimensions<br />

of the original grid do not meet the divisibility condition.<br />

From equations 4.1 and 4.2, it can be seen that there are some wavefronts, with period<br />

s x × s y , which give rise to zero for all the differences. These define the null space and<br />

correspond to the poles of the reconstruction filter (equation 4.4).<br />

By setting the<br />

values of the filter at the poles (which occur when ks x /N x and ls y /N y are integers) to<br />

zero, one eliminates only these modes unsensed by the equations defining the forward<br />

problem, thus, producing a solution with minimum norm. Therefore, we replace the<br />

reconstruction filter of equation 4.4 by<br />

⎧<br />

⎪⎨<br />

0 when k sx<br />

[ ( )<br />

[ ( )<br />

N x<br />

˜φ(k, l) = exp −2πi −1]<br />

⎪⎩<br />

k sx ˜Dx (k,l)+ exp −2πi −1] l sy ˜Dy (k,l)<br />

Nx [<br />

)<br />

)]<br />

Ny<br />

elsewhere.<br />

4<br />

sin 2( π k sx<br />

Nx<br />

+sin 2( π l sy<br />

Ny<br />

and l sy<br />

N y<br />

are integers<br />

(4.8)<br />

Equivalently, we can see that sample points in the original problem are only linked<br />

by difference data if they are separated by multiples of the shear distances s x , s y .<br />

The original sample grid can by split into s x × s y interlaced subgrids with sample<br />

spacing s x , s y which can be treated independently from each other, since points on<br />

different subgrids are not linked by difference data (figure 4.2). Thus, the original<br />

reconstruction problem can be subdivided into s x × s y independent problems each of<br />

which can be treated as having shears s x = s y = 1, and which can therefore be solved<br />

by direct application of equation 4.8. The resulting phase maps are re-interlaced onto<br />

the original grid. Since there is no information about the relative piston term of these<br />

interlaced grids in the original measurements, the piston term (i.e. the mean phase<br />

over the pupil) for each of the reconstructed subgrids is set to zero. It is equivalent<br />

to setting the null mode terms to zero, resulting in the minimum norm solution. We<br />

shall refer to this method as subdivision.<br />

71


4. Wavefront reconstruction from shear phase maps<br />

x<br />

y<br />

subgrid (1,1) subgrid (2,1) subgrid (3,1)<br />

subgrid (1,2) subgrid (2,2) subgrid (3,2)<br />

Figure 4.2: Example of the interlaced subgrids for a grid of dimensions N x = 6 by<br />

N y = 7 and shears s x = 3 and s y = 2. The relative phases between points in different<br />

subgrids is undetermined by the data.<br />

4.3 Computing performance<br />

An important consideration when comparing alternative reconstruction methods is<br />

their computational requirements. Figure 4.3 a) plots the number of real multiplications<br />

required to perform the wavefront reconstruction over a square grid with a<br />

total of N points and shear s


4. Wavefront reconstruction from shear phase maps<br />

and subdivision methods (using Matlab 6.1 on a computer with a Pentium 4 processor<br />

at 2.2 GHz and 500 MB RAM memory). Figure 4.3 b) shows the measured<br />

time for an individual reconstruction against the number of reconstruction points N.<br />

The trends of the plot are in agreement with those of the figure 4.3 a) for large N.<br />

Differences between the methods are attributable to the exact details of the computational<br />

implementations. If we compare the performance of the DFT methods<br />

against reconstruction using the pseudoinverse for the case N = 3 × 10 3 , the speed<br />

improvement factor is around 50 (which is comparable to that predicted by the ratio<br />

of the number of multiplications, 2 N 2 /3 N log 2 N ≈ 170). If we try to put this in<br />

the context of our tear topography estimation problem, we have that each slope map<br />

is around 400 samples across, that is N ≈ 1 × 10 5 . This would imply that if we<br />

were using a pseudoinverse reconstruction method, each topography map calculation<br />

would take 3 orders of magnitude longer than with either of the DFT algorithms (i.e.<br />

≈ 200 s/reconstruction). However, this was not the main reason for not using a pseudoinverse<br />

method. The actual calculation of the pseudoinverse itself was not possible<br />

at all with current technology in reasonable times (≈ 100 years with the available<br />

computing power).<br />

4.4 Noise performance<br />

We now compare the DFT methods against the least square pseudoinverse in terms<br />

of their associated noise coefficient[104], which quantifies the effect of white noise in<br />

the data on the reconstructed phase,<br />

C = 1 ∑<br />

R ω R ω , (4.9)<br />

N pupil<br />

where N pupil is the number of wavefront samples inside the pupil of interest P , ω is an<br />

index that runs through all the elements of the reconstruction matrix R. Figures 4.4<br />

(a) and (b) show the noise coefficients for different pupil sizes and unit shear, without<br />

any data extension. For unit shear no subdivision or data extension is necessary<br />

for the DFT reconstruction. Note that the pseudoinverse has a lower (better) noise<br />

coefficient for both pupil geometries.<br />

We now investigate the effect of extending the data by enlarging the original grid<br />

as previously explained. Figures 4.5 a) and b) show the noise coefficients versus<br />

the number of extra rows/columns used for circular and square pupils respectively,<br />

ω<br />

73


4. Wavefront reconstruction from shear phase maps<br />

Number of multiplications required for reconstruction<br />

10 9<br />

10 8<br />

10 7<br />

10 6<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

10 1 10 2 10 3 10 4 10 5<br />

10 10 Number of samples<br />

Reconstruction time (s)<br />

10 2<br />

10 1<br />

10 0<br />

10 1<br />

10 2<br />

10 3<br />

10 4<br />

10 5<br />

10 1 10 2 10 3 10 4 10 5<br />

10 3 Number of samples<br />

(a)<br />

(b)<br />

Figure 4.3: a) Plot of the number of real multiplications required to perform the<br />

wavefront reconstruction over a square grid with a total of N points and shear s


4. Wavefront reconstruction from shear phase maps<br />

each of width 32 samples and with unit shear. There are two points to note from<br />

these plots. Firstly, the noise coefficient actually decreases with small data extension.<br />

This suggests that all the existing DFT methods could be improved by applying data<br />

extension. Secondly, we can see that extending the pupil around all sides leads to<br />

a smaller noise coefficient than keeping the pupil in a corner of the grid. This can<br />

be understood by considering the noise variance of the various difference data points.<br />

Recall that in order to use the DFT reconstruction methods, the gradient data need<br />

to be made to meet the periodicity condition, using equations 4.5 and 4.6. That is, for<br />

row (or column) of data in a subgrid, an extra data point given by the negative sum<br />

of the data in the row is added to ensure that the gradients along the row sum to zero<br />

(corresponding to a periodic phase function). The noise in the extra data is given by<br />

the sum of the noise in the row of data. If we consider Gaussian noise of variance σ 2<br />

in the original data, then these extra data points have noise variance N x /s x σ 2 (N x /s x<br />

being the number of data points in the row of the subgrid). This suggests that for<br />

large grids, there will be a considerable noise contribution around the edge due to<br />

the extra periodization data. For the 2D reconstructor this will have greatest impact<br />

on the reconstructed phase near the edges. Extending the data pushes these edges<br />

further from the original pupil, reducing its effect in the reconstructed phase. This is<br />

illustrated in figures 4.6 (a) and (b), which plot the spatial distribution of the noise<br />

in the reconstructed phase maps (i.e. the noise variance) for square 32 × 32 sample<br />

pupils with no data extension and with 4 rows and columns (two around each side)<br />

data extension respectively. The plots correspond to the first and third circles of plot<br />

4.5 (a).<br />

The plots of spatial distribution of noise variance in figures 4.6 a) and b) correspond<br />

to square pupils 32 samples across, unit shear, without data extension (first circle in<br />

figure 4.5 a) and with 2 columns/rows at the pupil sides (third circle in figure 4.5 a).<br />

4.5 Conclusions<br />

We have presented two alternative methods to extend DFT-based wavefront reconstruction<br />

in a convex pupil from difference maps to cope with arbitrary problem dimensions<br />

and arbitrary integer shears, with performance comparable to that of the<br />

pseudoinverse in terms of the noise propagation coefficient. Non-optimized algorithm<br />

implementations showed significant speed increase with respect to the least squares<br />

75


4. Wavefront reconstruction from shear phase maps<br />

N ext/2 N ext/2 Next<br />

N ext/2 N ext/2 Next<br />

1.5<br />

1.5<br />

1.4<br />

1.4<br />

1.3<br />

1.3<br />

Noise coefficient<br />

1.2<br />

1.1<br />

Noise coefficient<br />

1.2<br />

1.1<br />

1<br />

1<br />

0.9<br />

0.9<br />

0.8<br />

0 5 10 15<br />

N ext<br />

0.8<br />

0 5 10 15<br />

N ext<br />

(a)<br />

(b)<br />

Figure 4.5: Effect on noise coefficient of extending the data by enlarging a square (a)<br />

and a circular (b) pupil of 32 samples across and unit shear against N ext , the number<br />

of columns/rows of padding for extension from two sides of the grid only (dots) and for<br />

symmetric extension from all the four sides of the grid (circles). The noise coefficient<br />

of the pseudoinverse is shown as a solid line for reference.<br />

6<br />

6<br />

5<br />

5<br />

4<br />

4<br />

3<br />

3<br />

2<br />

2<br />

1<br />

1<br />

0<br />

30<br />

0<br />

25<br />

20<br />

15<br />

10<br />

5<br />

5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

(a)<br />

(b)<br />

Figure 4.6: Spatial distribution of noise variance for DFT methods for a cornered (a)<br />

and centered (b) square pupil of 32 samples across.<br />

76


4. Wavefront reconstruction from shear phase maps<br />

pseudoinverse reconstruction method. Our analysis shows that the noise performance<br />

of the DFT reconstruction can actually be improved by judicious application of the<br />

extension method.<br />

77


Chapter 5<br />

Preliminary experiments<br />

5.1 Prydal type experiments<br />

Prydal et al. [83] tried to measure the tear thickness by using the interference between<br />

the front and back surfaces of the tear film using a coherent source, reproducing an experiment<br />

by Green et al. [82] for the estimation of the corneal thickness. According to<br />

the mean refractive index values for the tear (1.337) [81] and epithelium (1.401) [113],<br />

if the boundary surfaces were optically smooth, then a large fringe visibility, around<br />

0.44, was to be expected. However, this was not the case, and the fringes present<br />

led to a tear thickness value an order of magnitude larger than previous estimations,<br />

in clear disagreement with other mechanical and interferometric techniques. Prydal’s<br />

model of the tear proved to be an oversimplified one, because the tear has a complex<br />

three layer structure, each with a different refractive index. More importantly, the<br />

deepest of these, the mucus layer, is very rough on the optical scale. It is believed<br />

that what Prydal actually measured was the thickness of the tear and epithelium.<br />

We want to study the tear topography and not its absolute thickness, and therefore<br />

we want to obtain an interference pattern resulting from back reflections from the<br />

front surface of the tear and some other surface within the eye. Provided that the<br />

interferogram contrast is high enough, a modified version of Prydal’s experiment could<br />

be used to serve this purpose. Two experiments performed to test the viability of this<br />

approach are now described and their results discussed.<br />

Both experiments use the same experimental setup: that described in sections 2.6 and<br />

2.7 with the glass wedges removed.<br />

78


5. Preliminary experiments<br />

5.1.1 Normal illumination experiment<br />

For this particular experiment, the eye was placed so that the spherically convergent<br />

illumination was everywhere normal to the surface of the tear film. Two sequences<br />

of 100 images at a 2.5 Hz rate were recorded from each of the two subjects (series<br />

coded bb1, bb2, dm1 no wedges and dm2 no wedges). Figures 5.1 and 5.2 show two<br />

sequences of 12 images each, where the illuminated area over the tear was around<br />

3.1 mm in diameter. Note the diffraction rings at the pupil edge, which decrease in<br />

thickness towards the center of the pupil, the bubbles and other tear features, and<br />

finally the small sets of Newton rings with very low contrast (in comparison to the<br />

diffraction rings) that are due to interference within optical elements and/or their<br />

coatings.<br />

These particular sets of images were selected because they illustrate some of the most<br />

interesting things that can be seen using this technique. The first sequence (figure<br />

5.1) starts with what is probably the most representative situation, a smooth tear<br />

surface with some small bubbles distributed pretty much randomly over the pupil.<br />

Before the next frame the top eyelid moved down and then up a fraction of the pupil<br />

diameter, leaving a clearly delimited region full of bubbles which smooths out slowly in<br />

the following 5 frames (approximately 2 seconds). The subject felt the need to blink<br />

and tried to prevent it in the eighth frame of the sequence, where the tear surface<br />

becomes very rough. Finally the subject blinked between the eighth and ninth frame,<br />

producing again a smooth tear surface that remains so for the rest of the sequence<br />

apart from the displacement of a small line of bubbles.<br />

The second sequence (figure 5.2) shows a phenomenon that could be seen every time<br />

this subject blinked. After the blink shown in the first image, a set of relatively<br />

large bubbles are left near the top of the image pupil, and as time passes these set<br />

of bubbles move downwards while their size is reduced until they almost disappear<br />

before the next blink.<br />

We failed to identify clearly an interference pattern on any of the 400 images recorded<br />

with this experimental configuration that we could attribute with confidence to interference<br />

between the reflections from the front surface and another surface within the<br />

eye.<br />

79


5. Preliminary experiments<br />

bb2 (t = 0.000s)<br />

bb2 (t = 0.390s)<br />

bb2 (t = 0.781s)<br />

bb2 (t = 1.172s)<br />

bb2 (t = 1.562s)<br />

bb2 (t = 1.953s)<br />

bb2 (t = 2.344s)<br />

bb2 (t = 2.734s)<br />

bb2 (t = 3.125s)<br />

bb2 (t = 3.515s)<br />

bb2 (t = 3.906s)<br />

bb2 (t = 4.297s)<br />

Figure 5.1: Sequence of reflections from the front surface of the tear film (series coded<br />

bb2).<br />

80


5. Preliminary experiments<br />

The recorded patterns look very similar, if not identical, to what the subject sees when<br />

the experiment is being performed. We believe the image projected on the retina not<br />

only shows information about the tear topography, but also about floaters within the<br />

eye, and although this was not explored further, it should also show opacity in the<br />

eye due to for example cataracts. Typically when the subject participates for the first<br />

time in the experiment, asked to concentrate on one particular spot of these images<br />

and to remain looking always in the same direction, they would choose a floater or<br />

a bubble on the tear. Then because this feature is rigidly attached to the eye, the<br />

subject would start moving the eye trying (unsuccessfully) to bring that object to the<br />

fovea, resulting in a very frantic eye movement (one could push the analogy with a<br />

dog chasing its own tail to its very limit). The subject would then be asked to try to<br />

fixate on one of the set of small Newton rings that are due to the optics (and therefore<br />

static with respect to the optical setup). In any case, it is an interesting experiment<br />

to participate in as a subject, not only because it allows one to see what happens on<br />

ones own tear film but also to see other structures (e.g. floaters) within ones own eye.<br />

5.1.2 Focused illumination and estimation of undesired eye reflections<br />

Despite the failure of the previous experiment to provide a quantitative tool for studying<br />

the tear topography, a similar experiment with the same optical setup was performed,<br />

but with the eye further away from the experiment, so that the focus of the<br />

beam incident on the eye coincided with the front surface of the tear. In this way, if<br />

there was a reflection from a surface within the eye other than that from the tear surface,<br />

it would show up by producing an interferogram with concentric circular fringes.<br />

These fringes would carry information on the optical path between the front surface<br />

of the tear and the second reflecting surface. This is a similar experiment to that performed<br />

by Prydal, the only difference being that in this case the axis of the incident<br />

beam is perpendicular to the surface of the tear rather than at 45 o .<br />

For this experiment, a series of 100 images on two subjects (series coded bb focused<br />

and lm focused) were recorded. The two images shown at the top in figure 5.3<br />

are representative of the recorded series of images, and show three types of ringing:<br />

those at the edges, caused by diffraction, small sets due to interference from reflections<br />

within optical elements in the setup and the very dim set almost centered covering the<br />

81


5. Preliminary experiments<br />

dm1_no_wedges (t = 0.000s)<br />

dm1_no_wedges (t = 0.188s)<br />

dm1_no_wedges (t = 0.391s)<br />

dm1_no_wedges (t = 0.594s)<br />

dm1_no_wedges (t = 0.781s)<br />

dm1_no_wedges (t = 0.969s)<br />

dm1_no_wedges (t = 1.172s)<br />

dm1_no_wedges (t = 1.360s)<br />

dm1_no_wedges (t = 1.563s)<br />

dm1_no_wedges (t = 1.750s)<br />

dm1_no_wedges (t = 1.9380s)<br />

dm1_no_wedges (t = 2.156s)<br />

Figure 5.2: Sequence of reflections from the front surface of the tear illustrating bubble<br />

movements on the tear. (series coded dm1).<br />

82


5. Preliminary experiments<br />

whole pupil, which we believe results from the interference between the front surface<br />

of the tear and some surface near the back of the cornea. This last set of fringes has<br />

a contrast around 0.16 (estimated in the white rectangle on the left image away from<br />

the edge to avoid confusion with the greater contrast diffraction rings. This means<br />

that the undesired reflection from the surface near the back of the cornea is about 150<br />

times less intense than the reflection from the front surface of the tear, assuming both<br />

reflections are completely coherent. This is equivalent to a maximum error of λ/40 in<br />

the phase recovery algorithm (see section 3.4.3).<br />

Even though this matter will not be pursued in depth, it should be noticed that from<br />

counting the fringes within the image (or at least in the central part, away from the<br />

diffraction rings) one should be able to estimate the distance between the two reflecting<br />

surfaces. To illustrate this point, the interferograms that would result if the reflecting<br />

surface was 55 µ and 510 µm away from the front surface of the tear, typical for the<br />

front and back surface of the cornea respectively, were simulated (bottom two images<br />

of figure 5.3).<br />

The simulation accounted for the focal length of the lens used, the refractive index of<br />

the cornea (taken as 1.376), the double pass, the beam diameter, the wavelength and<br />

the estimated contrast. Comparing the simulations with the data, it seems that one<br />

of the reflections seem to come from a surface in between the epithelium and the back<br />

of the stroma, or otherwise indicates thin corneas in both subject, although the fringe<br />

visibility is quite poor for a good estimation of the number of fringes, not to mention<br />

the difficulty in distinguishing the interference rings from the diffraction rings.<br />

5.2 Estimation of errors for tear topography measurements<br />

If the topography of the tear film was static multiple measurements could be used to<br />

estimate a statistical error on the topography measurements. However, in principle<br />

this cannot be done because the characteristic times involved in the tear topography<br />

dynamics are not known.<br />

83


5. Preliminary experiments<br />

bb_focused<br />

lm_focused<br />

simulated thicknes = 55 mm<br />

simulated thickness = 510 mm<br />

Figure 5.3: The two top images were obtained by illuminating the eye with a converging<br />

beam and placing the front surface of the eye roughly at the beam focus.<br />

5.2.1 Repeatability test and associated error<br />

In an attempt to estimate the errors introduced in the topography estimation by<br />

the optical setup itself, the following experiment was performed. Three glass spheres<br />

of different radii were placed one at a time in the position where the cornea of the<br />

subject would be in a normal tear topography estimation experiment (see figure 2.7),<br />

and a sequence of 50 interferograms was recorded for each of them. The spheres<br />

were calibration accessories from a Humphrey-Zeiss Atlas Corneal topographer, having<br />

radii 9.65, 8.00 and 6.15 mm. The data was then processed (series coded ae 965mm,<br />

84


5. Preliminary experiments<br />

ae 8mm and ae 615mm) and the corresponding topography maps reconstructed. The<br />

resulting mean variation in the RMS of the topography is of the order of 10 nm. It<br />

can therefore be considered that the error associated to the experimental setup, that<br />

is, all the sources of noise related to the acquisition with a CCD camera, vibrations,<br />

etc. in the estimated tear topography will be of the order 10 nm.<br />

5.2.2 Alignment and shear error propagation<br />

Although it is difficult to justify mathematically, experience would indicate that the<br />

major sources of errors when the topography of the tear is relatively smooth (e.g. no<br />

vignetting due to bubbles or tear film break-up) are the estimation of the position<br />

of the interferograms and the shears not being perfectly horizontal or vertical. The<br />

former problem is mainly due to the non-sphericity of the cornea of some subjects,<br />

the slight asymmetries that exist between the two branches of the imaging system and<br />

the non-uniform illumination of the cornea, while the latter is mostly related to eye<br />

movement.<br />

Estimation of the center of interferograms is not crucial for our experiment. The<br />

crucial issue is the relative alignment of the two interferograms (horizontal and vertical<br />

shears) corresponding to each tear topography map. Actually a systematic error in<br />

estimating the position of the interferograms can be afforded, provided this error was<br />

the same for each pair of interferograms and it was the same for the each series of<br />

topography maps to be estimated.<br />

A good estimation of the error introduce by this misalignment is difficult if not impossible,<br />

because some idea of the statistics of the misalignment itself would be needed,<br />

and this is not trivial to obtain. To estimate the distribution of misalignment errors six<br />

series of interferograms were taken (series coded ad1, cp1, cp2, cp3, cw1 and cw3) with<br />

a total of 550 pairs of interferograms. The centers estimated by the correlation-based<br />

algorithm (see section 3.3) were compared with a visual estimation. These particular<br />

series were chosen by alphabetical order and are not representative in the sense that<br />

the subjects ad and cp have significant corneal astigmatism (about 2.5D) compared<br />

to the other subjects, resulting in a poorer performance of the center estimation algorithm,<br />

so one could argue that the typical misalignment will now be over-estimated.<br />

We then assumed that the manually estimated position of the interferograms were<br />

85


5. Preliminary experiments<br />

the correct ones, and calculated the standard deviation of the misalignment for the<br />

550 pairs of interferograms, and found values around 4 camera pixels. Some of the<br />

histograms of these misalignments were also plotted to study their distributions and<br />

different distribution shapes were found, with their mean typically non-zero, indicating<br />

a systematic error within each of the series associated with the particular corneal<br />

astigmatism of each subject.<br />

The estimated shears were not perfectly horizontal or vertical. Typically when the<br />

shear in the desired direction was about 35 camera pixels, the shear in the perpendicular<br />

direction was 1 camera pixel (standard deviation). This is strongly related<br />

to eye movement, because it changes from subject to subject and even within each<br />

series of interferograms. Had this error in the orientation of the shear been due to a<br />

misalignment of the wedges, the error would have been similar (i.e. same direction)<br />

for all interferograms.<br />

It is clear that the effect of misalignment and shear errors in the reconstructed topography<br />

depends on the particular topography to be estimated. However, to get<br />

an idea of how significant the resulting errors are, 7 topography maps (second and<br />

third order Zernike polynomials) and their sheared phase maps were simulated, using<br />

typical shear values (30 camera pixels) and pupil sizes (450 camera pixels). Then the<br />

topography was estimated using the DFT wavefront reconstructor after introducing<br />

all the possible combinations of misalignment error within the range −4 to 4 camera<br />

pixels and the shear error along the perpendicular direction to the main shear in the<br />

range −1 to 1 pixels. The relative RMS error for each of these were calculated and in<br />

a somewhat conservative approach the maximum, 13 % was taken. Notice that this<br />

error can be expressed as a percentage because the reconstructor is linear. It should<br />

be taken into account that the topographies used for the simulation are relatively<br />

smooth and therefore one should expect them to be less sensitive to small misalignments,<br />

while in a real tear topography, small and abrupt features such as tear bubbles<br />

and break-ups will be present. This features will be smoothed out in the wavefront<br />

reconstruction if there is misalignment between the horizontal and vertical shearing<br />

interferograms.<br />

86


5. Preliminary experiments<br />

5.2.3 Simulated interferograms and non-linear errors<br />

We now describe a test of all the data processing software as a whole. Three series of<br />

12 interferograms with similar characteristics to those in our data were simulated, i.e.<br />

440 samples across the pupil, 20-sample shear amplitudes and similar carrier frequencies,<br />

with the second, third and fourth radial order Zernike aberrations. The RMS<br />

amplitude of the polynomials in each series was λ/2, λ and 5λ/2. From processing<br />

the simulated interferograms topography maps were obtained, and two numbers were<br />

extracted: the amplitude error, ɛ amplitude , that is the error in the amplitude of that<br />

particular polynomial, and the residual error ɛ residual that is the RMS of the topography<br />

with that particular polynomial subtracted. The mean values obtained for each<br />

of the series are shown in the table 5.2.3.<br />

Series RMS ɛ amplitude (%) ɛ residual (%)<br />

λ/2 6 22<br />

λ 4 7<br />

5λ/2 18 80<br />

5.3 Validation experiment with a deformable mirror<br />

As a final check on the software and experimental setup as a whole, a simple experiment<br />

with a 37-channel 15 mm diameter deformable membrane mirror (OKO technologies)<br />

was performed to verify that the units and scaling in the data processing are<br />

correct. For this experiment the experimental setup described in sections 2.6 and 2.7<br />

was used after replacing the focusing lens that is closest to the eye by the surface of<br />

the deformable mirror.<br />

Five different voltage configurations were applied to the 37 hexagonal actuators ,<br />

where the voltages were either the maximum (on) or minimum (off). The estimated<br />

topographies from the different voltage configurations were: 1) all actuators off, 2)<br />

all but one actuator off, 3) all but a horizontal line of actuators off, 4) one actuator<br />

and its immediate neighbors on, while the rest remain off and 5) all the actuators<br />

on. The mirror was tested using these 5 configurations with a commercial interferometer,<br />

µPhase2 from Fisba Optik, and the peak-to-valley values of the estimated<br />

mirror surface measured. Then the experiment was repeated with the lateral shearing<br />

interferometer, and the peak-to-valley values obtained were compared with those from<br />

87


5. Preliminary experiments<br />

1<br />

10<br />

1<br />

20<br />

15<br />

0<br />

5<br />

0<br />

10<br />

5<br />

-1<br />

0<br />

-1<br />

0<br />

-1 0 1<br />

-1 0 1<br />

Figure 5.4: Reconstructed topography of two simulated small bubbles, of around 0.13<br />

and 0.26 mm in diameter, with heights (depths) of 11 and 22 nm respectively. The axis<br />

units are millimeters and the units on the color scale are nanometers. The error in<br />

the peak-to-valley value in each of the reconstructed topographies versus the original<br />

simulated phase is 15 % and 10 % respectively.<br />

the µPhase2, and it was found that the values differ by around 6%.<br />

5.4 Small detail test<br />

When developing the software the question of whether any small detail could be seen<br />

in the final tear topography arose, because of the non-linear filtering in the Fourier<br />

domain and the smoothing that occurs in the unwrapping and integration processes.<br />

Therefore, some interferograms that would result from a flat tear topography and a<br />

series of bubbles with increasing size were simulated. Figure 5.4 shows the reconstructed<br />

topography of two small bubbles, of around 0.13 and 0.26 mm in diameter,<br />

with heights (depths) of 11 and 22 nm respectively. The errors in the peak-to-valley<br />

in each of the reconstructed topographies with respect to the original simulated topography<br />

are 15% and 10% respectively, suggesting that the presence of small details<br />

can in principle be detected.<br />

88


Chapter 6<br />

Tear topography dynamics<br />

experiment<br />

6.1 Data collection protocol<br />

The subjects for all the experiments described in this <strong>thesis</strong> were volunteers. They were<br />

informed of the experimental protocol, approved by the St. Mary’s Local Research<br />

Ethics Committee of the Kengsinton, Chelsea and Westminster Health Authority. All<br />

the information was coded and strictly confidential. The laser radiation levels were for<br />

all cases well below 2% of the maximum permissible exposures (MPEs) as indicated<br />

on the British and European standard Safety of laser products (BS EN 60825-1:1994<br />

with amendments 1, 2 and 3). The calculations of the MPEs are detailed in Appendix<br />

B.<br />

The measurements for each subject were taken in one session that lasted around one<br />

hour, starting with a description of the experiment, followed by a verification of laser<br />

radiation levels with a calibrated power meter and a demonstration of what the subject<br />

was expected to do during the experiment. A bite registration with soft dental wax<br />

was made and mounted on a x-y-z translation stage (from now on we will refer to<br />

this block as the bite bar) for precise positioning of the subject with respect to the<br />

experiment. The subject was then positioned in front of the experiment while using<br />

the bite bar to align the subject’s dominant eye with the optical setup. This was<br />

achieved by asking the subject to concentrate on one of the small features within the<br />

big red circle projected by the laser onto the retina (only around 2 % of the light<br />

89


6. Tear topography dynamics experiment<br />

Figure 6.1: On the left we can see the light scattered, as opposed to the specular<br />

reflections, by the cornea (large disc) and the crystalline lens (cone) when the tear<br />

topography experiment was being performed (the exposure was 15 seconds). The<br />

picture on the right is a superposition of the picture on the left on top of a picture of<br />

the same subject with the room lights on. It should be noticed that when performing<br />

the experiment, the intensity of the scattered light was extremely dim, and it was<br />

enhanced in this picture for illustrative purposes only.<br />

reaching the eye is reflected back by the front surface of the tear, most of the rest goes<br />

through up to the retina producing a circular spot in which several structures can be<br />

seen as described in section 5.1.1). Finally, when the eye was aligned with respect<br />

to the optical system, the subject was asked to remain as steady as possible and to<br />

blink normally, while sequences of 100 pairs of shearing interferograms (horizontal<br />

and vertical shear) were recorded at a frame rate of 5 Hz. The radii of curvature of<br />

the cornea was also measured with the Atlas corneal topographer (manufactured by<br />

Humphrey-Zeiss).<br />

The data analyzed in this chapter corresponds to 2450 pairs of usable interferograms,<br />

each pair leading to one topography map. No attempt was made to process the data<br />

in which there was vignetting (due to eye movement), the eyelids or eyelashes blocked<br />

part of the illuminated area on the tear or the eye was either too close or too far<br />

from the experiment so that the density of fringes in one of the interferograms was<br />

too high to be sampled properly. The tolerance to alignment in depth is determined<br />

by the fringes introduced by defocus. This can be understood by noticing that to<br />

first order a shear interferometer is a slope sensor, so defocus will produce fringes that<br />

either increase or decrease the number of fringes produced by the tilt introduced by<br />

the wedges, depending on the orientation of the wedges. In our experiment the wedges<br />

were oriented so that when the eye was too close, the number of fringes in one of the<br />

90


6. Tear topography dynamics experiment<br />

z<br />

cornea<br />

rotated cornea<br />

Rc<br />

h(x,y)<br />

q<br />

x<br />

y<br />

Rs<br />

sclera<br />

Figure 6.2: Eyeball modelled as two spherical surfaces.<br />

interferograms increased while decreased in the other one and vice-versa when the eye<br />

was too far, providing then a visual aid for alignment. The area illuminated on the<br />

tear film was typically around 3 mm in diameter.<br />

6.2 Eye movements and effect on tear topography measurements<br />

Before presenting the estimated topography maps the effect that eye movements would<br />

have on the tear topography estimation shall be discussed. For the sake of simplicity,<br />

the rotation of the eyeball with respect to the center of the sclera and the movement<br />

of the head will be considered independent.<br />

6.2.1 Effect of eyeball rotation on tear topography experiment<br />

Let us model the eyeball as a rigid body formed by two spherical surfaces, the sclera<br />

with a typical radius (R s ) of approximately 11 mm and the cornea with a radius (R c )<br />

of around 8 mm (typical values range from 7 to 9 mm). It will be also assumed that<br />

the eyeball rotates around the center of the sclera which is a distance d ≈ 6 mm away<br />

from the center of the corneal surface. Now, for the sake of simplicity but without<br />

loss of generality only rotations in the y − z plane around the center of the sclera will<br />

be considered, that as shown in figure 6.2 coincides with the origin of the system of<br />

coordinates. Then, the change in corneal height ∆h with respect to the x − y plane<br />

due to an eyeball rotation of angle θ is given by<br />

91


6. Tear topography dynamics experiment<br />

∆h(θ, x, y) = d (cos θ − 1) + √ R 2 c − x 2 − (y − d sin θ) 2 − √ R 2 c − x 2 − y 2<br />

∀ (x, y) | x 2 + y 2 ≤ R 2 pupil < R2 c. (6.1)<br />

In a real situation, when the eye is fixating the amplitude of the eyeball rotation is of<br />

the order of 10 −3 rad [69], and thus we can approximate to first order in θ<br />

∆h(|θ| ≪ 1, x, y)<br />

≈<br />

y d<br />

√<br />

R 2 c − x 2 − y 2 θ<br />

∀ (x, y) | x2 + y 2 ≤ R 2 pupil < R2 c.(6.2)<br />

The RMS of the change in topography ∆h for small angles over a pupil with diameter<br />

R pupil < R c is,<br />

RMS(θ) =<br />

[∫∫<br />

≈ d √<br />

2<br />

[<br />

] 1/2<br />

pupil ∆h2 (θ, x, y)dxdy<br />

∫∫<br />

pupil dxdy ( )<br />

−<br />

R2 c<br />

R 2 pupil<br />

ln<br />

1 − R2 pupil<br />

R 2 c<br />

− 1] 1/2<br />

θ. (6.3)<br />

We now need to remove the RMS that is due to piston, tip and tilt, both because<br />

these modes do not affect the optical quality of the eye, and because they cannot be<br />

sensed by the experiment. Due to the angular symmetry, the piston and tip terms<br />

cancel for the case where the rotation is in the y − z plane, leaving only the tilt which<br />

is calculated as<br />

RMS tilt (θ) =<br />

≈<br />

2<br />

∣πRpupil<br />

3 0<br />

[<br />

2 Rc<br />

2 2 d<br />

3 Rpupil<br />

2 +<br />

∫ 2π ∫ Rpupil<br />

0<br />

(<br />

1<br />

3 + 2 3<br />

∆h(ρ, φ, θ) ρ sin φ ρ dρ dφ<br />

∣<br />

) √ ]<br />

R 2 c<br />

R 2 pupil<br />

R 2 c<br />

R 2 pupil<br />

− 1<br />

θ (6.4)<br />

We can now subtract the tilt component from the overall change in topography, where<br />

it should be noticed that even for large eye movement of one degree, around 99.5% of<br />

the overall RMS is due to tilt when the typical corneal radius of curvature are used<br />

and for a pupil radius (that is, the area imaged on our experiment as opposed to the<br />

radius of the pupil of the eye) of 1.5 mm. Thus, the RMS of the change in topography<br />

expected to be measured in our experiment purely due to eye movement, with the<br />

assumptions made in our model of eyeball rotation can be estimated as<br />

RMS(θ) ≈ (2.4 µm) × θ. (6.5)<br />

To convert this topography RMS value to wavefront error RMS in full wavefront<br />

sensing measurements in the eye, one only needs to multiply the topography RMS by<br />

the difference in refractive indices between air (n air ≈ 1) and cornea (n cornea ≈ 1.37).<br />

92


6. Tear topography dynamics experiment<br />

0.2<br />

0.1<br />

0.1<br />

vertical displacement (mm)<br />

0.1<br />

0<br />

-0.1<br />

-0.2<br />

-0.2 0 0.2<br />

horizontal displacement (mm)<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

-0.5 0 0.5<br />

horizontal displacement (mm)<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

-0.5 0 0.5<br />

vertical displacement(mm)<br />

Figure 6.3: Estimated movement of the center of the illuminated area of the tear: on<br />

the left a typical movement pattern over 30 seconds is shown (the radius of the red<br />

circle corresponds to one standard deviation), and in the center and right we can see<br />

the histograms of the x− and y−displacement of the interferograms center for all the<br />

2450 usable frames. Both distributions have zero mean and 0.16 and 0.12 mm standard<br />

deviations respectively, which would be equivalent to around 9 nm of wavefront error<br />

RMS.<br />

The movement of the eye was estimated from the recorded interferograms by estimating<br />

the evolution of the center of each interferogram with respect to the first one of<br />

the series. Figure 6.3 a) shows a typical eye movement pattern from the series coded<br />

cp1 over a period of 30 seconds. From the standard deviation of the eye movement,<br />

one can estimate the RMS of the wavefront error that it would be produced by eye<br />

movement of those amplitudes. These can then be taken as an estimation of the<br />

uncertainty in RMS due to eye movement, that is ≈ λ/70 for λ = 632.8 nm.<br />

A completely different approach to estimate the effect of eye movement in the estimated<br />

tear topography maps was also used, based on the correlation of the dominant<br />

low-order modes (defocus and astigmatism) with the pupil displacement for each series<br />

of data. In this way a set of correlation coefficients were generated, providing some<br />

measure of how the changes in these mode, are linked to the movement of the eye.<br />

These correlation coefficients provide a measure of how confidently one can say that<br />

the estimated tear topography dynamics are due to eye movement or true change in<br />

the tear topography. High correlation values would mean that the estimated changes<br />

in the tear topography might be due more to eye movement than to true tear topography<br />

changes.<br />

When the correlation coefficients between eye movement and defocus and astigmatism<br />

were calculated for all the series of data, it was found that in about half the series the<br />

93


6. Tear topography dynamics experiment<br />

values where higher than 0.7 and some were as high as 0.99.<br />

6.2.2 Effect of head movement on tear topography<br />

It was noticed in the course of the experiments that despite the use of the bite bar,<br />

the head of the subject pivots a little bit around the biting point during the 20 to 30<br />

seconds of the data acquisition in a plane vertical plane. This shows very clearly in<br />

the raw data as noticeable fluctuations in the interferogram diameter of up to 10 %,<br />

and on the number of fringes across the interferograms due to the changes in defocus<br />

when the front tear surface changes its distance to the experimental setup along the<br />

optical axis.<br />

For this type of eyeball displacement no modelling was made, only the correlation<br />

coefficient calculation used for eyeball rotation was performed. In this case, we look<br />

for correlation between the defocus component of the measured topography maps and<br />

the relative change in pupil diameter. Figure 6.4 shows a correlation value of 0.67<br />

when all the 2450 usable pairs of interferograms are considered.<br />

Again, as with the eye movement one could use the correlation coefficients as a measure<br />

of how much of the changes in our measurements is due to head movement and how<br />

much is due to true change in the tear topography.<br />

6.2.3 Asymmetries in the optical system<br />

The third and last issue with the experiment that was noticed through the movement<br />

of the eye is the presence of small asymmetries between the two interferometer<br />

branches, which should in principle be identical. However, after studying the data<br />

small asymmetries of two types were noticed: a difference in magnification, which is<br />

not constant throughout the different series and a difference in the movement of the<br />

pair of interferograms on the interferometer output plane. If the pupil magnifications<br />

are different, this will affect the wavefront integration, unless some pupil scaling (and<br />

thus data interpolation) is done. In order to keep the data processing to a minimum,<br />

such scaling was avoided, although it is shown later that the consequences of this do<br />

not seem to have a large effect in the RMS estimation.<br />

We should mention that if we were to repeat the experiments, the alignment of the<br />

94


6. Tear topography dynamics experiment<br />

1.2<br />

1<br />

0.8<br />

(m m)<br />

0.6<br />

RMS measured<br />

0.4<br />

0.2<br />

0<br />

0.2<br />

0.4<br />

0.6<br />

0.85 0.9 0.95 1 1.05 1.1 1.15 1.2<br />

relative pupil radii<br />

Figure 6.4: Correlation between the relative change in the radius of the interferograms<br />

and the estimated change in defocus. The correlation coefficient was calculated considering<br />

all the interferograms from all the series together, and the obtained value was<br />

0.67.<br />

system should be improved until these asymmetries are negligible. The main practical<br />

difficulty when trying to do this is not in focusing the optics, but in positioning the<br />

mirrors that fold the optical path of the two branches so that the two interferograms<br />

can be recorded by the same camera.<br />

6.3 Global error estimation<br />

Based on the experiments and simulations in section 5.2, and the data analysis in<br />

section 6.2, a very rough estimation of the uncertainty in the wavefront error RMS<br />

can be made.<br />

To recap: the repeatability of the experimental setup plus the software was tested<br />

to be around (n tear − n air )λ/60 (see section 5.2.1); it was shown that the effect of<br />

reflections was lower than λ/50 in the slopes estimation, while the typical slope RMS<br />

is of the order of a wavelength, so we can neglect this source of error with respect<br />

to the others (see section 3.4.3); from the eyeball rotation estimation and model, the<br />

estimated RMS error was around λ/70 (see section 6.2.1); from the pupil alignment<br />

and shear estimation errors we estimated a 13 % (see section 5.2.2); and finally, the<br />

95


6. Tear topography dynamics experiment<br />

errors caused by the software and the hypo<strong>thesis</strong> on which it is based (like correctly<br />

sampled band-limited spatial spectra) which for the amplitudes measured (comparable<br />

to or smaller λ/2) we estimate around 6 %. If we take a very conservative approach,<br />

the total uncertainty in our tear topography RMS (σ RMS ) can be coarsely estimated<br />

by combining the errors assuming their sources are independent, getting σ RMS ≈<br />

[<br />

(λ/50) 2 + (14 %) 2] 1/2 . This estimation does not consider the errors due to head<br />

movement because we could not find a simple way to get an estimate.<br />

6.4 Wavefront error RMS introduced by the tear topography<br />

Once the tear topography maps have been estimated as described in chapters 3 and<br />

4, they can be converted to wavefront error maps by a simple multiplication by the<br />

difference in refractive index between the tear film and air (which is approximately<br />

0.134). In this way, the effect that the changes in the tear topography have on the<br />

optical quality of the eye, which is the main goal of the experiment, can be studied.<br />

The results will be now be presented in various formats, that will help to address<br />

different issues: data confidence, relevance of low order aberrations, dynamics of the<br />

optical quality and inter-subject variability.<br />

All the parameters calculated below to describe the optical quality of the eye, are<br />

calculated under the hypo<strong>thesis</strong> of the rest of the optics of the eye being perfect. This<br />

is clearly an unreal situation, because we are already aware of a number of dynamic<br />

factors in the eye, fluctuations of accommodation being dominant from the optical<br />

quality point of view. However, it will serve for our purpose of studying the effects of<br />

the tear film independently from the other changing features (though not completely<br />

de-coupled from eye movement).<br />

6.4.1 Wavefront error RMS evolution for individual series<br />

A simple way to evaluate the imaging performance of a given optical system when<br />

the amplitude of its aberrations are small is by looking at the wavefront error RMS.<br />

If the RMS is smaller than λ/14, then the system is usually referred to as diffraction<br />

limited, which means that it can be considered that the dominant effect in degrading<br />

96


6. Tear topography dynamics experiment<br />

the image quality is diffraction. This is a criterion proposed by Maréchal, and is the<br />

one to be used as a reference for evaluating the optical quality of the tear. Let us now<br />

look at the RMS evolution of a particular data series (code named cp3) to illustrate<br />

some of the findings from the data analysis.<br />

Figure 6.5 (a) shows the diffraction limit as a red continuous line and three different<br />

plots. The blue continuous line is the RMS of the difference between each wavefront<br />

and the mean wavefront of the series, thus providing information about how the optical<br />

quality of the tear does evolve. From now on, whenever we refer to RMS without any<br />

other reference, we will be referring to the RMS of the wavefront calculated in this<br />

way. It was found that the fluctuations of the RMS are comparable to the diffraction<br />

limit (i.e. 0.045µm for λ = 632.8nm), but this will be discussed further in the next<br />

section. The dotted green and dashed black lines show the RMS of the wavefronts<br />

reconstructed using only the horizontal and vertical shear interferograms respectively<br />

with respect to the mean wavefront of the series (as in the blue continuous plot), after<br />

a straightforward modification of the wavefront integration algorithm. These plots<br />

are to be compared with the continuous blue plot for two reasons, first as a peace<br />

of mind check that all three are of the same order of magnitude, to verify that the<br />

pupil misalignments and difference in pupil magnification do not lead to catastrophic<br />

results in the integration algorithm, and second, to have an idea of how different<br />

would the results be if instead of performing a full wavefront estimation with two<br />

shear interferograms, a partial estimation was performed by using only one shear and<br />

thus using a simpler experimental setup and software. It was found that even though<br />

the plots may look very dissimilar in some series, all three have comparable amplitudes,<br />

and when defocus and astigmatism are removed, they are virtually identical.<br />

Figure 6.5 (b) shows the same plots as (a) but now with the defocus and astigmatism<br />

components of the wavefronts removed, and we will refer to these RMS values as<br />

residual RMS. In this way, if we were not to have confidence in the defocus and<br />

astigmatism estimation because of high correlation with eye movement, it can be seen<br />

that the amplitude of the remaining aberrations would be in most cases below the<br />

diffraction limit.<br />

Finally, figure 6.5 (c) shows the evolution of the defocus and astigmatism components<br />

of the wavefront difference. When studying these plots for all the series, it was found<br />

that in most cases defocus is the dominant of the three aberrations, the two astigma-<br />

97


6. Tear topography dynamics experiment<br />

RMS evolution<br />

Residual RMS<br />

second order zernike evolution<br />

0.08<br />

diff limit ( l /14)<br />

RMS<br />

RMS hs<br />

0.08<br />

diff limit ( l /14)<br />

diff limit ( l /14)<br />

RMS vs astig 2<br />

RMS<br />

0.08<br />

astig<br />

1<br />

RMS hs<br />

defocus<br />

amplitude (m m)<br />

0.06<br />

0.04<br />

RMS vs<br />

0 10 20<br />

amplitude (m m)<br />

0.06<br />

0.04<br />

amplitude (m m)<br />

0.06<br />

0.04<br />

0.02<br />

0.02<br />

0.02<br />

0<br />

0<br />

0<br />

0 10 20<br />

0 10 20<br />

time (s)<br />

time (s)<br />

time (s)<br />

(a) (b) (c)<br />

Figure 6.5: Typical wavefront RMS evolution. The red horizontal line in the plots<br />

indicates the diffraction limit for the wavelength used in the experiment. Plot (a)<br />

shows the evolution of the RMS with respect to the initial wavefront estimated using<br />

both interferograms (blue), and only one interferogram (green and black). Plot (b) is<br />

similar to (a) but with defocus and astigmatism removed from the wavefronts, (c) plots<br />

the evolution of the defocus and astigmatism components of the estimated wavefront.<br />

tisms either being comparable to or below the diffraction limit. In order to analyze<br />

these plots further, the correlation coefficients between these aberrations and pupil<br />

position (associated to eyeball rotation) and pupil radius change (associated to head<br />

movement) were used. For this particular case, it was found that the correlation of<br />

defocus with head movement was 0.9 which is quite high, indicating that the large<br />

fluctuations of the defocus component might be due to head movement rather than<br />

true tear topography dynamics. Similarly, one of the astigmatisms correlates well<br />

(again 0.9) with eyeball rotation, and again one should ask if the measured changes<br />

are mostly due to eye movement rather than tear topography dynamics.<br />

6.4.2 RMS mean values<br />

Let us now look at the wavefront error results in a more global way, by plotting<br />

the mean RMS and residual RMS over the data series with more than 75 usable<br />

topography maps, corresponding to time intervals of around 30 seconds (see figure<br />

6.6). Again, the red horizontal line shows the diffraction limit for the wavelength used<br />

in the experiment for reference. The data is separated in three groups: on the left,<br />

over the area with light gray background, we can see those series in which there were<br />

no blinks during the data recording; on the center we have those series in which there<br />

was at least one blink; and on the right over the dark gray area we show those series<br />

98


6. Tear topography dynamics experiment<br />

in which the subject was wearing contact lenses at the time of the experiment. The<br />

numbers on top of the error bars are the maximum correlation coefficient between the<br />

corresponding RMS and the eye and head movement. One can think of these numbers<br />

as indicators of the likelihood of the RMS being an artifact due to the undesired eye<br />

movements as opposed to true topography estimation.<br />

The first thing to notice from the figure, is that in most cases the residual error<br />

remains below the diffraction limit, while the RMS mean values including the defocus<br />

and astigmatism components are comparable to the diffraction limit.<br />

It is important to be aware that the results shown in figure 6.6 correspond only to a<br />

sample of 14 subjects and within those, to the data series with no tear breakup, and<br />

that all the high spatial frequencies higher than half that of the interferogram fringes<br />

are lost, i.e. the smoothest tear surfaces. Thus, we are somehow underestimating<br />

the RMS values, although not significantly, as the processed corresponds to the most<br />

representative situation within the recorded data.<br />

It seems from the data shown in the figure that the tear in front of the contact lenses is<br />

no different from the wavefront RMS point of view to that in front of the cornea when<br />

no contact lens is being worn. However, the processed data in this case represents a<br />

small portion of the recorded data from contact lens users (less than 33 %). In the rest<br />

of the data, the roughness of the tear surface is such that no fringes can be identified,<br />

and therefore no quantitative analysis could be performed.<br />

The conclusions from this section can therefore be summarized as follows:<br />

1) The processed data for non contact lens users represents the typical tear topography<br />

scenario within our set of recorded data, with a smooth tear topography that does<br />

not degrade the eye beyond the diffraction limit.<br />

2) The evolution of the residual wavefront error RMS in a normal situation seems<br />

comparable to the diffraction limit, so we can say that the optical quality of the eye<br />

is not significantly affected by the tear topography dynamics under normal conditions<br />

(i.e. no tear break up or contact lenses).<br />

3) In some cases the correlation coefficients between eye and/or head movement suggest<br />

that the estimated values of defocus and astigmatism estimated from our experiment<br />

might actually be partly due to eye and/or head movement and not only to tear<br />

99


6. Tear topography dynamics experiment<br />

topography changes. It is important to notice that even when these movements might<br />

lead to an overestimation of the RMS, the estimated optical quality of the eye does<br />

not seem to degrade far beyond the diffraction limit.<br />

4) The estimated RMS mean amplitudes are consistent with the RMS variability<br />

measurements reported for full wavefront sensors in the eye, ranging from 0.02 to<br />

0.10 µm [4, 14, 28, 21], with and without paralyzed accommodation.<br />

RMS residual<br />

RMS<br />

no blink blink contact lens on<br />

0.1<br />

0.1<br />

0.2<br />

amplitude ( m m)<br />

0.05<br />

0.5<br />

0.2<br />

0.2<br />

0.2<br />

0.2<br />

0.1<br />

0.4<br />

0.2<br />

0.3<br />

0.3<br />

0.4<br />

0.5<br />

0.5<br />

0.2<br />

0.1 0.7<br />

0.6<br />

0.3<br />

0.3<br />

0.2<br />

0.2<br />

0.6<br />

0.1<br />

0.3<br />

0.3<br />

0.1<br />

0.4<br />

0.3<br />

0.1<br />

0.2<br />

0.3<br />

0.2<br />

0.1<br />

0.3<br />

0.4<br />

0.4<br />

0.3<br />

0.2<br />

0.2<br />

0.4<br />

0.5<br />

0.2<br />

0<br />

cp1 cp2 cp3 cw1 jm1 jm2 jn1 kg2 kh4 pb3 sc3 sj1 cw2 cw3 jp2 sc2 st3 te4 im1 fr1 fr2 fr3<br />

Figure 6.6: Mean RMS and residual RMS over the data series with more than 75<br />

usable topography maps, corresponding to time intervals of 30 seconds. The red<br />

horizontal line shows the diffraction limit (λ = 632.8 nm). The data is separated in<br />

three groups: on the left, over the area with light gray background, we can see the<br />

series in which there were no blinks during the data recording; in the center we have<br />

the series in which there was at least one blink and on the right over the dark gray area<br />

we show the series in which the subject was wearing contact lenses at the time of the<br />

experiment. The numbers on top of the error bars are the the maximum correlation<br />

coefficient between the corresponding RMS and the eye and head movement.<br />

6.4.3 RMS mean evolution<br />

We can now consider the temporal behavior of the RMS by looking at figure 6.7<br />

where the evolution of estimated RMS (b) and residual RMS (a) (this time with<br />

respect to the first topography map instead of the mean topography map as in the<br />

100


6. Tear topography dynamics experiment<br />

previous sections) for 30 data series are plotted, using around 2300 topography maps<br />

corresponding to 19 different subjects. The black lines are the mean RMS at the<br />

corresponding time and the gray areas delimited by the blue lines are the areas within<br />

one standard deviation from the mean. Again, the red horizontal line indicates the<br />

diffraction limit at 632.8 nm.<br />

The average blink rate is approximately 1 every 5 seconds [68], with considerable<br />

variation between individuals and visual tasks. Having this in mind, it can be noticed<br />

that the residual RMS stays below the diffraction limit even for periods up to 25<br />

seconds. If we now look at the RMS including defocus and astigmatism, we find that<br />

(ignoring the possibility of the measurements being biased to larger values due to eye<br />

movement), it remains around twice the diffraction limit. This is interesting in the<br />

context of wavefront sensing for refractive surgery, because it shows that the effect<br />

of the tear topography dynamics is negligible. If the whole of the RMS was due to<br />

defocus, for an 8 mm diameter pupil (and assuming that the RMS values measured<br />

for a 3 mm pupil are similar to what they would be over a larger pupil), then an RMS<br />

comparable to diffraction limit corresponds to 1/55 D and twice that to 1/25 D. These<br />

values, from the refraction point of view are negligible if we bear in mind that the<br />

graduation of spectacles go in steps of 1/8 D and in practice nobody prescribes with<br />

more than 1/4 D accuracy.<br />

If we now look at the same numbers in the context of adaptive optics for high resolution<br />

retinal imaging or psychophysical experiments on individual photoreceptors,<br />

even though the degradation of the optical quality of the eye is not very significant,<br />

is by no means negligible if structures of the size of photoreceptors are to be resolved,<br />

and some adaptive correction is needed, probably with a bandwidth of the order of<br />

1 Hz. Again, the results obtained here are consistent with Roorda’s experience [114]<br />

with static wavefront correction on a scanning laser ophthalmoscope (SLO), where<br />

he reported some image resolution degradation when trying to resolve photoreceptors<br />

after preventing the blink for a few seconds while having accommodation paralyzed.<br />

6.5 Second moment of the AC term<br />

Licznerski et al. [74] proposed to use the second moment of the AC term of lateral<br />

shearing tear film interferograms as a tool to study the tear film break-up time. How-<br />

101


6. Tear topography dynamics experiment<br />

0.25<br />

0.25<br />

0.2<br />

0.2<br />

RMS residual<br />

(mm)<br />

0.15<br />

0.1<br />

RMS (mm)<br />

0.15<br />

0.1<br />

0.05<br />

0.05<br />

0<br />

0 5 10 15 20 25<br />

time (s)<br />

0<br />

0 5 10 15 20 25<br />

time (s)<br />

Figure 6.7: Temporal evolution of the estimated wavefront error RMS (b) and residual<br />

wavefront error RMS (a) for 30 data series (around 2300 topography maps) corresponding<br />

to 19 different subjects. The blue dots correspond to the estimated RMS<br />

(b) and residual RMS (a) wavefront aberration introduced by the tear for all the 30<br />

data series. The black lines are the mean RMS at the corresponding time and the gray<br />

areas delimited by the blue lines are the areas within one standard deviation from the<br />

mean. Again, the red horizontal line indicates the diffraction limit at 632.8 nm.<br />

ever, the second moment as defined by Licznerski depends not only on the surface of<br />

the tear, but also on the intensity profile of the image, on the total intensity, on the<br />

shear amplitude, spatial coherence of the source and pupil shape. Thus, monitoring<br />

of the second order moment of the AC term (in the Fourier space) can only be used to<br />

measure relative changes within a series of interferograms from a given subject with<br />

unchanged conditions. The effect of shear amplitude, total intensity and coherence of<br />

the source could be eliminated by defining a normalised second moment,<br />

∫ ∫<br />

f 2 |AC( f)|<br />

M N =<br />

⃗ 2 d 2 f<br />

fR 2 ∫ ∫<br />

|AC( f)| ⃗ 2 d 2 f<br />

(6.6)<br />

where the integral is performed over a circle with radius f R and AC is the part of the<br />

spectrum of the interferogram centered on the carrier frequency. The moment defined<br />

in this way ranges from 0 (when the AC term is a delta function) to 1 (when the AC<br />

term is a ring of radius f R ). In practice, one has a discretised (pixelated) image, and<br />

a discrete Fourier transform, and thus the need for the discrete version of equation<br />

6.6, that is<br />

M N =<br />

∑ m 2 +n 2 ≤R 2<br />

m,n<br />

(<br />

m 2 + n 2) |AC(m, n)| 2<br />

R 2 ∑ m 2 +n 2 ≤R 2<br />

m,n<br />

|AC(m, n)| 2 (6.7)<br />

102


6. Tear topography dynamics experiment<br />

where (m, n) are the frequency coordinates in pixels or sample units with the carrier<br />

frequency as origin of coordinates. Before performing any analysis of real data, one<br />

can already see that the magnitude just defined will depend on the intensity profile<br />

of the illumination, the pupil shape (this depends on the non-spherical corneal topography<br />

of the subject under study) and the radius of the integration area. Thus, the<br />

numbers obtained can only be compared with measurements from instruments with<br />

equal illumination profiles and subjects with similar corneal topography. One also<br />

requires that the spectra of the DC and AC terms do not overlap.<br />

Figure 6.8 shows results for the data of figures 3.1 and 3.2 with the radius of the<br />

integration area equal to half the estimated carrier frequency. Looking at the numbers<br />

obtained and the corresponding figures, we notice firstly, that M N is not a good tool<br />

to study tear break-up, as (f) shows a clear break-up without significant modification<br />

of the topography other than in the surroundings of the break-up area, and N M is<br />

comparable to that for (a) where the tear shows no break-up. Secondly, we can see<br />

that N M seems to give an estimation of the roughness of the topography, roughness<br />

being any departure of the tear topography from a sphere that cannot be described<br />

as a second order polynomial. If the surface of the tear was so rough that the Fourier<br />

transform could be thought of as white noise (equal amplitude for all frequencies), then<br />

the corresponding value for N M (normalised moment of inertia of a disc) would be 0.5.<br />

Figures 6.8 (e), (g) and (h) show values very close to this theoretical case, although<br />

for different reasons, the first two because the spectrum of the DC term overlaps with<br />

the spectrum of the AC term, and the second shows a more white noise appearance<br />

(see figures 3.1 and 3.2). Although these few examples are not an exhaustive test, we<br />

believe that the above defined normalised second moment is not a good number to<br />

describe tear film break-up.<br />

We can also test whether the normalised second moment of the shear interferograms<br />

is related to the tear wavefront error RMS or RMS 2 (that if the RMS ≪ λ as it is<br />

the case, is related to the Strehl ratio). This was put to the test by calculating the<br />

correlation coefficients between M N and the total RMS (0.23), residual RMS (0.36),<br />

total RMS squared (0.23) and residual RMS squared (0.32) for 2150 interferograms.<br />

The low correlation values suggest that the normalised second order moment of the<br />

interferograms on its own does not provide an idea of the optical quality of the tear<br />

film. This was to be expected because to assess the optical quality of the tear we<br />

should look at the tear topography and not at its derivatives, that to first order is<br />

103


6. Tear topography dynamics experiment<br />

what the shear interferograms provide.<br />

104


6. Tear topography dynamics experiment<br />

M = 0.134<br />

N<br />

M = 0.193<br />

N<br />

M = 0.254<br />

N<br />

mk1/46<br />

st3/15<br />

(a) (b) (c)<br />

im4/17<br />

M = 0.240<br />

N<br />

M = 0.484<br />

N<br />

M = 0.154<br />

N<br />

pb2/92<br />

mk1/33<br />

kh3/47<br />

(d) (e) (f)<br />

M = 0.412<br />

N<br />

M = 0.489<br />

N<br />

(g)<br />

te2/49<br />

(h)<br />

te2/13<br />

Figure 6.8: Normalized second order moment of the AC term of tear topography<br />

lateral shearing interferograms (M N ) illustrating some of the different tear topography<br />

features.<br />

105


Chapter 7<br />

Summary and discussion<br />

The design of a double shearing interferometer for evaluating the dynamics of the<br />

front surface of the pre-corneal tear film has been described. The instrument was<br />

built and tested on 20 subjects, and the data was analyzed with purpose-developed<br />

software. In order to achieve the topography integration from finite differences, we<br />

generalized an existing reconstruction method based on the DFT, which allows the<br />

processing of data volumes that can not be achieved with current computing powers<br />

and conventional pseudoinverse reconstruction techniques.<br />

The data processing software was tested in a set of recorded interferograms that<br />

illustrate the different features found on the tear topography, to become familiar with<br />

the limits of both the software algorithms and the theory on which these algorithms<br />

are based.<br />

We found that observing the back-reflection of the tear surface with normal coherent<br />

illumination might be a qualitative tool for studying the tear topography evolution,<br />

without the need for data processing and with relatively high resolution. The only<br />

problem that might prevent this technique from being used clinically is the small<br />

tolerance to eye misalignment.<br />

A modified version of Prydal’s experiment was performed to estimate the tear film<br />

thickness [83], with the purpose of estimating the amplitude of back-reflections from<br />

surfaces within the eye other than the front surface of the tear. It was found, based<br />

on fringe contrast estimation that the intensity of these reflections are around 150<br />

times lower than those of the front surface of the tear, thus introducing very little<br />

106


7. Summary and discussion<br />

error in Takeda’s phase recovery algorithm [95]. It was also noticed on the concentric<br />

ring pattern observed that counting the number of fringes within a certain radius<br />

allows one to estimate the distance between the two back-reflecting surfaces. After<br />

performing some computer simulations based on a simple model, we believe the deeper<br />

reflecting surface is not always the same for all subjects. In the first subject, the<br />

inner back-reflecting surface was likely to be the front surface of the cornea (around<br />

50 − 60 µm deep) while on the second subject it was more likely to be the back surface<br />

of the cornea (around 500 µm deep). This results may explain why Prydal estimated<br />

a tear film thickness of around 40 µm, which is one order of magnitude larger than<br />

the accepted values.<br />

We tested the instrument repeatability and simulated the effect of pupil error alignments<br />

and general data processing error for the current experimental setup, to get a<br />

feeling of the amplitude of the errors or uncertainties to be expected in our measurements.<br />

We also performed a simple validation experiment on the experiment against<br />

a commercial interferometer.<br />

Ethical approval and insurance cover was obtained for acquiring data from a number<br />

of subjects, and 20 volunteers were recruited for the experiments. From the analyzed<br />

data, the amplitude of the theoretical errors due to eye movement in the tear<br />

topography maps was estimated. Then with this estimation and the results from previous<br />

experiments and some simulations, a global error figure for the estimated tear<br />

topography RMS was obtained.<br />

We found that the contribution of the tear film to the degradation of the optical<br />

quality of the eye even when the defocus and astigmatism are considered (not reliably<br />

estimated from our measurements due to eye movement) is not significant, that is,<br />

assuming the rest of the optics of the eye to be perfect (Strehl ratio = 1). It can be<br />

concluded that the changes in the tear film topography are not large enough to bring<br />

the optical quality of the eye below the diffraction limit. It should be kept in mind<br />

that this conclusion has been drawn from the data that could be processed, which<br />

corresponds to the smoothest tear surface, therefore leading to an underestimation of<br />

the degradation of the optical quality of the eye by the tear film. Rough surfaces, can<br />

degrade the optical quality of the eye more significantly, but these rough surfaces are<br />

less common than the smoother surfaces found in the processed data.<br />

107


7. Summary and discussion<br />

The values found for RMS variability are consistent with the reported variability in<br />

full wavefront sensing in the eye measurements, with paralyzed accommodation.<br />

It was noticed that the front surface of the tear in front of contact lenses is very rough<br />

and although a quantitative analysis of the topography was not possible when the<br />

typical roughness was present, we believe the subject should be studied further given<br />

the high number of contact lens users.<br />

From the data it can also be infered that the characteristic times associated with<br />

change in the tear topography are of the order of seconds, so if one was thinking of<br />

using an AO system in the eye just to correct the tear fluctuations, a bandwidth of<br />

the order of 1 Hz or faster should be considered.<br />

We put to the test the use of the amplitude of the second moment of the intensity<br />

of a lateral shearing interferogram with tilt fringes as a tool to study tear break up<br />

times. We first suggested a normalization of this moment to make it more robust<br />

and independent of certain experimental parameters, and then showed, with a few<br />

experimental examples, that this number is no good on its own for the estimation of<br />

the break up time. It was also found that there is very little correlation of the second<br />

moment with the actual optical quality of the tear, either in terms of wavefront error<br />

RMS or Strehl ratio.<br />

Even though the data processing could not be completely automated and despite<br />

the inherent problems of using back-reflections from a moving eye, we believe that<br />

interferometry (lateral shearing, radial shearing or else) could be useful as a tool<br />

to study the tear film of subjects pre- and post-refractive surgery, and with tear<br />

conditions, but only if the issue of eye movement is addressed in far more depth than<br />

it was here.<br />

After the tear topography experiments, the experimental setup was modified to be<br />

used as a full wavefront sensor in the eye and a similar one was built for a different<br />

wavelength. Each instrument recorded simultaneous lateral shearing interferograms<br />

and SH spot patterns, for a qualitative study on the feasibility of using conventional<br />

interferometry for wavefront sensing in the eye. The experiments performed suggest<br />

that conventional interferometry is not as straightforward to use as other types of<br />

wavefront sensing due to speckle. When recording interferograms with short exposure<br />

times, the interferograms will be severely affected by speckle, while if one tries to<br />

108


7. Summary and discussion<br />

f6<br />

f6<br />

CCD<br />

camera<br />

T''<br />

BS<br />

f5<br />

f5<br />

BS<br />

T'<br />

f 4= 30 f 3= 150<br />

f 2= 40<br />

T<br />

eye<br />

180 190<br />

Figure 7.1: Modified imaging branch to convert the lateral shearing interferometer<br />

into a radial shearing interferometer.<br />

increase the exposure times to reduce the speckle, the interferogram contrast will<br />

decrease.<br />

7.1 Future work: Radial shearing interferometry<br />

Radial shearing interferometry is a self-referencing interferometric technique that can<br />

allow the estimation of a phase map with only one interferogram, as opposed to lateral<br />

shearing interferometry, that requires two interferograms [92]. In this technique, two<br />

copies of the beam to study are made to interfere, with one of the beams expanded<br />

with respect to the other. The intensity changes due to the beam profile spatial<br />

fluctuations could be separated from the phase changes in the same way as it is done<br />

in the lateral shearing interferometer described in Chapter 2 by introducing tilt fringes<br />

and applying a similar phase recovery technique.<br />

The imaging branch of the experimental setup of the lateral shearing interferometer<br />

was modified in the way described by figure 7.1. The two copies of the beam are<br />

created by the first beam splitter, and recombined through the second one. Each<br />

branch of the interferometer would be a 4f system with different focal lengths, the<br />

ratio of the focal lengths being the relative magnification of the pupils. This type of<br />

configuration, with 50:50 beam splitters, would make better use of the light reflected<br />

from the tear, because while in our lateral shearing interferometer only around 12 % of<br />

the light is used, in this setup we make use of 50 % of it, a gain factor of more than 4.<br />

More important would be the simplification of the data processing and elimination of<br />

sources of error like the pupil alignment and asymmetries between the optical branches<br />

109


7. Summary and discussion<br />

ao2_radial_shear (t = 3.625s)<br />

ao2_radial_shear (t = 4.203s)<br />

Figure 7.2: Typical radial shearing interferograms from tear film topography with<br />

no tilt between the interfering wavefronts. Recorded with the experimental setup<br />

described in figure 7.1.<br />

of the interferometer.<br />

7.2 Publications<br />

The following contributions related with this work have been published:<br />

A. Dubra, C. Paterson, and J.C. Dainty. Wave-front reconstruction from shear phase<br />

maps by use of the discrete Fourier transform. App. Opt. 43 (5) 1108-1113.<br />

A. Dubra, J.C. Dainty, and C. Paterson. Measuring the Effect of the Tear Film on<br />

the Optical Quality of the. Eye Invest. Ophthalmol. Vis. Sci. 2002 43: E-Abstract<br />

2045.<br />

A.Dubra, C.Paterson and J.C. Dainty. Tear film topography dynamics measurement<br />

with a shear interferometer. Eye Invest. Ophthalmol. Vis. Sci. 2004 45: E-Abstract<br />

2777.<br />

A. Dubra, J.C. Dainty, and C. Paterson. Double-lateral shearing interferometer for<br />

the quantitative measurement of the tear film topography (submitted to App. Opt.).<br />

A. Dubra, J.C. Dainty, and C. Paterson. Study of the tear topography dynamics using<br />

a lateral shearing interferometer (in preparation, to be submitted to Optics Express).<br />

110


Appendix A<br />

Preliminary study of feasibility<br />

of interferometric wavefront<br />

sensing in the eye<br />

As mentioned in the introduction, a number of wavefront sensors have been used to<br />

assess the optical quality of the human eye. In this chapter the feasibility of using interferometric<br />

techniques for evaluating the optical quality of the human eye is explored<br />

through an experimental qualitative comparison of a lateral shearing interferometer<br />

and a Shack-Hartmann wavefront sensor under similar conditions.<br />

Because the retinal reflectance and scattering dependence on wavelength is known to<br />

be very complex and subject-dependent [115], it is likely that different wavelengths<br />

perform differently both for a SH wavefront sensor and an interferometer. It was<br />

therefore decided to set up two similar experiments for two different wavelengths,<br />

632.8 nm and 780 nm. These wavelengths were chosen because of their widespread<br />

use in research laboratories [20, 24, 14, 28, 39] and commercially available wavefront<br />

sensors (e.g. Zywave by Bausch & Lomb).<br />

A.1 Experimental setups<br />

There are a number of practical issues related to setting up a wavefront sensor for<br />

the eye [116, 56], like the low light levels that can be used for safety reasons and the<br />

double pass problem [22]. We do not intend to deal with these issues in depth here,<br />

111


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

but only to briefly describe how the experiments were setup for the data acquisition.<br />

Most Shack-Hartmann wavefront sensors and laser ray tracers in the eye work by<br />

delivering a narrow collimated beam (1 − 2 mm in diameter) into the eye, parallel to<br />

the optical axis of the instrument and the eye, the latter assumed to be the line of<br />

sight [117], and slightly off the pupil center to eliminate the corneal reflection. The<br />

purpose of the narrow incoming beam is to try to reduce the loss of information on odd<br />

aberrations due to symmetric double pass, by using an entrance pupil (beam diameter)<br />

over which the optics of the eye is assumed to be almost diffraction limited. In this<br />

way, the wavefront error of the outgoing beam will be mainly that of the second pass<br />

through the optics of the eye after the reflection on the retina, when using a larger exit<br />

pupil. It has also been shown that the phase information on the first pass through the<br />

optics of the eye is almost completely eliminated after the non-specular reflection on<br />

the retina, due to its roughness and movement. By delivering the incoming beam off<br />

the optical axis but parallel to it, the corneal reflections are eliminated by vignetting.<br />

After the incoming beam is reflected back from the retina, the wavefront emerging<br />

from the eye is imaged by one or more pairs of lenses forming 4f systems, that produce<br />

a copy (ignoring diffraction and lens aberrations) of both phase and intensity of the<br />

beam exiting the eye on a plane we will denote output plane. In SH wavefront sensors<br />

a lenslet array in the output plane focuses the light onto the surface of a detector<br />

(typically a CCD camera) forming an array of spots. The displacement of these spots<br />

from a reference position is proportional to the average slope (tilt) of the wavefront<br />

over the corresponding lenslet, assuming the intensity over the lenslet is uniform.<br />

This last hypo<strong>thesis</strong> is not always true due to speckle, as we will see later, but which<br />

can be reduced by time averaging using the random eye movements, or equivalently,<br />

using a high frequency (a few KHz) scanning mirror [28]. In the lateral shearing<br />

interferometers described below, either a glass wedge or plate with parallel sides is<br />

placed after the last lens before the output plane (where the beam is collimated) at<br />

45 o to the optical axis to produce the lateral shearing interferogram from the reflections<br />

from the front and back surface of the wedge.<br />

112


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

A.1.1<br />

Experimental setup for 632.8 nm<br />

This instrument is a modification of the interferometer described in sections 2.6 and<br />

2.7, used for the tear topography estimation. The illumination branch shown in figure<br />

A.1 uses a spatially filtered He-Ne single mode laser (632.8 nm) to deliver a 1 mm<br />

diameter collimated beam to the eye, off-center by about two millimeters.<br />

The polarising beam-splitter and the λ/4 waveplate are left as they were in the tear<br />

film experiment to reduce the reflections from the beam-splitter, although this might<br />

not be the best configuration to divert most of the light reflected back from the retina<br />

down to the imaging branch due to the birefringent and depolarising properties of the<br />

eye [118] which have great inter-subject variability.<br />

He-Ne<br />

laser<br />

633nm<br />

ND filter<br />

microscope objective<br />

spatial filter<br />

f 1= 250<br />

stop aperture<br />

(off axis)<br />

s-polarisation<br />

PBS<br />

l/4<br />

eye<br />

p-polarisation<br />

circular-polarisation<br />

Figure A.1: Illumination branch of the wavefront sensing experiment using a 632.8 nm<br />

He-Ne laser, which delivers an off-center collimated 1 mm beam to the eye.<br />

The imaging branch remains virtually unchanged from that described in 2.7 apart<br />

from minor modifications that can be seen in figure A.2. The focusing lens that was<br />

shared with the illumination branch, the first wedge and λ/2 waveplate were removed<br />

113


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

T''<br />

T''<br />

CCD<br />

camera<br />

f 6 = 150<br />

f 5 = 150<br />

T'<br />

f 4= 30 f 3= 150<br />

T<br />

CCD<br />

camera<br />

eye<br />

lenslet<br />

array<br />

wedge 2<br />

300 180<br />

180 170<br />

Figure A.2: Imaging branch of the wavefront sensing experiment using a 632.8 nm<br />

He-Ne laser. The exit pupil of the eye (T) is conjugated to a SH lenslet array (T”)<br />

and a CCD camera that will record the laterally sheared interferograms (T”’).<br />

and a second camera is placed after the second (and now only) glass wedge. The<br />

resulting optical system produces a replica of the electric field at the pupil plane of<br />

the eye demagnified by a factor of 5.<br />

A.1.2<br />

Experimental setup for 780 nm<br />

The optical setup built for the wavefront sensing experiments at 780 nm (see figures A.3<br />

and A.4) is quite similar to the one just described, although the illumination branch<br />

is slightly simpler because the source, which in this case is a diode laser emitting at a<br />

peak wavelength between 780 and 800 nm (depending on current and temperature) is<br />

coupled to a single mode optical fiber, and thus the spatial filtering is produced within<br />

the optical fiber by attenuation of all but the fundamental modes (gaussian intensity<br />

profile and orthogonal polarisation states).<br />

After the source there is a collimating lens, which we can now afford to choose with a<br />

short focal length for compactness because the condition on uniform intensity over the<br />

beam entering the eye is now not as demanding as it was for the tear topography experiment,<br />

the aperture stop now being only 0.6 mm in diameter (as opposed to 15 mm<br />

in the tear related experiments). Then the non-polarising beam splitter (NPBS) reflects<br />

50 % of the light towards the 4f system in front of the eye with a magnification<br />

of 2. This pair of lenses was not strictly needed but it was added thinking of using it<br />

to compensate the sphere of the refractive error of the subjects. This is achieved by<br />

displacing the subject and the lens together along the optical axis of the system, in<br />

an arrangement that is known as the Badal optometer [77].<br />

In this setup polarisation control was deliberately avoided, thus having the advantage<br />

of collecting not only the light with the same polarisation state as the input beam, as<br />

114


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

a<br />

diode laser<br />

@ 790nm<br />

single mode optical fiber<br />

f 1= 60<br />

stop aperture (off axis)<br />

NPBS<br />

f 2= 40<br />

f 3= 80<br />

eye<br />

120 100<br />

a/2<br />

Figure A.3: Illumination branch of the wavefront sensing experiment using a 780 nm<br />

diode laser, which delivers an off-center collimated 1.2 mm beam to the eye.<br />

in the previous experiments, but also the light with the orthogonal polarisation state.<br />

The price to pay for this gain is the tilting of the beam splitter and illumination<br />

branch to eliminate the reflections from the beam splitter, which in this case are even<br />

stronger in comparison with the signal in the tear interferometry experiment, because<br />

the total power reflected from the retina is lower than that reflected by the tear. It<br />

can be demonstrated by using the reflection law that if the optical axis of the part of<br />

the illumination branch before the beam splitter is rotated an angle α and the beam<br />

splitter is rotated half that angle, then the collimated beam after the beam splitter<br />

will be parallel to the optical axis of the Badal optometer and that the undesired<br />

back reflections from the beam splitter (leaving the beam splitter through the left<br />

surface on the figure) will go off-axis at an angle α, allowing us to spatially filter these<br />

reflections.<br />

The imaging branch of the experiment (figure A.4) shares the Badal optometer and<br />

beam-splitter with the illumination branch. The only purpose of the 4f system to the<br />

left of the beam-splitter is to focus the light coming back from the eye so that a stop<br />

aperture can be placed in the focal plane of this telescope to filter the off-axis beam<br />

splitter reflections. Again, the final optical element before the planes conjugated to<br />

the exit pupil of the eye is the wedge at 45 degrees, that produces the two reflections<br />

on the camera that will register the interferograms and lets through the light that<br />

goes to the SH sensor.<br />

115


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

a/2<br />

CCD<br />

camera<br />

CCD<br />

camera<br />

T'' T''<br />

wedge<br />

f 5= 40 f 4= 40<br />

NPBS<br />

f 2= 40<br />

f 3= 80<br />

T<br />

eye<br />

lenslet<br />

array<br />

40<br />

80<br />

80<br />

stop aperture (BS backreflection filter)<br />

120 100<br />

Figure A.4: Imaging branch of the wavefront sensing experiment using a 780 nm diode<br />

laser. The exit pupil of the eye (T) is conjugated to a SH lenslet array (T’) and a<br />

CCD camera that will record the laterally sheared interferograms (T”).<br />

A.2 Data acquisition<br />

The interferograms and SH spot patterns were recorded with two CCD cameras, Retiga<br />

1300 and Retiga EX (manufactured by QImaging), synchronized by an external TTL<br />

signal. The exposures of the cameras were equal at all times, and the gain of the<br />

electronics of each camera was adjusted to try to optimise the use of the dynamic<br />

range and take into account the fact that the SH camera receives approximately 11<br />

times more energy than the one recording the interferograms, and also the light is<br />

focused into an array of spots as opposed to the interferogram recording camera in<br />

which the energy is more evenly distributed over the pupil. The lenslet array used<br />

to produce the SH spot patterns was a Fresnel microlens array on polycarbonate,<br />

produced by WelchAllyn, with a focal lens of 7.5 mm and 200 µm pitch.<br />

For the data acquisition, the subject was set in front of the experiment and the head<br />

kept in position with the aid of a bite bar. In all the wavefront sensing experiments the<br />

optical power entering the eye was kept below 1 % of the MPEs for 10 s, as calculated<br />

in Appendix B.<br />

A.2.1<br />

Data acquired with 632.8 nm source<br />

A subject (coded name ad) was tested with three different optical power levels entering<br />

the eye and exposure times chosen so that the energy recorded by each camera on each<br />

frame was always the same. The optical power entering the eye was adjusted by the<br />

use of neutral density absorption filters placed between the laser and the spatial filter<br />

to reduce the effect of undesired reflections. The experiment was repeated twice, with<br />

116


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

the wedge in opposite orientations (180 degrees rotation), producing in one a large<br />

shear comparable to the pupil diameter, and in the other an almost negligible shear.<br />

This could be used to get an idea of the spatial coherence of the light emerging from<br />

the eye.<br />

Figure A.5 shows the interferograms and SH patterns that resulted from using 3.5, 1.0<br />

and 0.35 µW with camera exposures 0.2, 0.7 and 4.0 s respectively on each row. The<br />

first and second columns display the lateral shearing interferograms with the wedge in<br />

the orientations that produces large and small shear respectively. The third column<br />

shows the associated SH spot pattern. Apart from the pupil size change in response<br />

to the intensity of the light entering the eye, it can be clearly noticed that the speckle<br />

both in the interferograms and the SH spots is reduced with increasing exposures. It<br />

can also be seen that for the interferograms corresponding to a small lateral shear,<br />

vertical fringes with very poor contrast seem to appear for the 4 second exposure but<br />

these fringes are not of the spatial frequency produced by the wedge. We also looked<br />

at the spectra of the interferograms looking for a peak at the frequency of the fringes<br />

that the wedge would produce, and we found that there is indeed a peak, although<br />

based on the experience gained from processing the tear film data for the previous<br />

chapter, we believe no reliable quantitative data can be extracted.<br />

A.2.2<br />

Data acquired with 780 nm source<br />

The same subject was tested for different exposures and this time with equal optical<br />

power of 4 µW. The rows in figure A.6 show the interferograms and SH patterns<br />

that resulted from the exposures of 0.5, 1.0 and 3.0 s respectively. Notice that the<br />

optical magnification of the system is different from the previous experiment, hence<br />

the increase in number of lenslets over the pupil in the SH spot patterns.<br />

Again, as in the previous experiment, reduction of speckle can be noticed for longer<br />

exposures. Fringes with poor contrast (≈ 0.1) can be observed for the interferograms<br />

with small shear and no fringes at all can be noticed for the larger shearing interferograms.<br />

If one could vary the shear in a continuous way like in the interferometer<br />

proposed by Lohmann [97] and the spatial coherence characteristics of the source were<br />

known, then the fringe contrast could be used to study the spatial coherence of the<br />

retinal reflection. When looking at the amplitude of the spectra at the spatial fre-<br />

117


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

4<br />

4<br />

4<br />

2<br />

2<br />

2<br />

mm<br />

0<br />

mm<br />

0<br />

mm<br />

0<br />

-2<br />

-2<br />

-2<br />

mm<br />

mm<br />

-4<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

ad / 3.20 mW / 0.2s -4<br />

-4<br />

-5 0 5<br />

mm<br />

(a) (b) (c)<br />

-5 0 5<br />

mm<br />

(d) (e) (f)<br />

-2<br />

ad / 0.35 mW / 4.0s -4<br />

-5 0 5<br />

-4 -2 0 2 4<br />

mm<br />

mm<br />

mm<br />

mm<br />

4<br />

4<br />

2<br />

2<br />

0<br />

-2<br />

-2<br />

ad / 1.00 mW / 0.7s -4<br />

-4<br />

4<br />

4<br />

2<br />

2<br />

0<br />

-4 -2 0 2 4<br />

mm<br />

-4 -2 0 2 4<br />

mm<br />

-4 -2 0 2 4<br />

mm<br />

(g) (h) (i)<br />

mm<br />

mm<br />

0<br />

0<br />

-2<br />

-4<br />

-4 -2 0 2 4<br />

mm<br />

-4 -2 0 2 4<br />

mm<br />

Figure A.5: Interferograms and SH spot patterns that resulted from using 3.5, 1.0<br />

and 0.35 µW at λ = 632.8 nm with camera exposures 0.2, 0.7 and 4.0 s respectively<br />

on each row (subject ad). The first and second columns display the lateral shearing<br />

interferograms with the wedge in the orientations that produces a large and small<br />

shear respectively. The third column shows the associated SH spot pattern.<br />

118


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

quency of the fringes, we found a peak. Also a small decrease in amplitude, with<br />

respect to the DC term, with increasing exposure was observed, we believe it was not<br />

significant enough to assert that the fringe contrast increases with exposure times.<br />

When comparing the SH spot patterns it can be noticed that at 780 nm there is a<br />

uniform halo (not speckled) surrounding every spot, not present at 632.8 nm. The<br />

presence of this scattering phenomenon at near infrared wavelengths is well known<br />

[26, 25].<br />

Finally, to rule out the possibility of the blur in the fringes being caused by fluctuations<br />

of accommodation, the measurements were repeated under the same experimental<br />

conditions, after paralyzing the accommodation (and pupil dilation as a side effect)<br />

with cyclopentolate hydrochloride 1 % (see figure A.7). It was found that again as<br />

the exposure increases the speckle is reduced, but no improvement in fringe contrast<br />

was observed with respect to the measurements with the same light source and nonparalyzed<br />

accommodation.<br />

A.3 Conclusions<br />

We have tested, we believe for the first time the use of a lateral shear interferometer for<br />

wavefront sensing in the eye, using two different wavelengths and two shear amplitudes.<br />

For the experiment at 632.8 nm we could not observe interference fringes for either<br />

of the shear amplitudes (0.75 mm and 0.17 mm). In the experiment using 780 nm we<br />

could not observe interference fringes for the larger shear amplitude (2.4 mm), but<br />

fringes of contrast 0.1 could be seen for the smaller shear (0.17 mm) and all three<br />

exposure times (0.5 s, 1 s and 3 s).<br />

The fringe visibility of the interferograms depends on the complex degree of coherence<br />

of the electric field at the pupil plane over the integration time [91]. The experiments<br />

that were performed do not provide enough information to plot the fringe visibility<br />

as a function of the shear amplitude. However, it can be said that the width of the<br />

visibility function (in terms of the shear) is smaller than 0.17 mm at 632.8 nm and<br />

of the order of 0.1 mm at 780 nm. This width tells us which is the maximum shear<br />

amplitude for which interference fringes can be observed.<br />

119


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

4<br />

4<br />

4<br />

2<br />

1<br />

2<br />

2<br />

mm<br />

0<br />

mm<br />

0<br />

mm<br />

0<br />

-2<br />

-2<br />

-2<br />

-4<br />

4<br />

ad / 4.00 mW / 0.5s<br />

-4 -2 0 2 4<br />

mm<br />

-4<br />

-4 -2 0 2 4<br />

mm<br />

4<br />

4<br />

-4<br />

-4 -2 0 2 4<br />

mm<br />

(a) (b) (c)<br />

2<br />

2<br />

2<br />

mm<br />

0<br />

mm<br />

0<br />

mm<br />

0<br />

-2<br />

-2<br />

-2<br />

-4<br />

4<br />

ad / 4.00 mW / 1.0s<br />

-4 -2 0 2 4<br />

mm<br />

-4<br />

-4 -2 0<br />

mm<br />

2 4<br />

4<br />

4<br />

-4<br />

-4 -2 0<br />

mm<br />

2 4<br />

(d) (e) (f)<br />

2<br />

2<br />

2<br />

mm<br />

0<br />

mm<br />

0<br />

mm<br />

0<br />

-2<br />

-2<br />

-2<br />

-4<br />

ad / 4.00 mW / 3.0s<br />

-4 -2 0 2 4<br />

mm<br />

-4<br />

`-4 -2 0<br />

mm<br />

2 4<br />

-4<br />

-4 -2 0<br />

mm<br />

2 4<br />

(g) (h) (i)<br />

Figure A.6: Interferograms and SH patterns that resulted from the 4 µW at 780 nm<br />

setup with exposures of 0.5, 1.0 and 3.0 s. The first and second columns display the<br />

lateral shearing interferograms with the wedge in the orientations that produces a<br />

large and small shear respectively. The third column shows the associated SH spot<br />

pattern.<br />

120


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

From the experiments performed, it seems that conventional interferometry (e.g.<br />

shearing interferometers or Twyman-Green) is not as straightforward to use as other<br />

types of wavefront sensing (such as a Shack-Hartmann sensors), due to speckle. When<br />

recording interferograms with short exposure times, the interferograms will be severely<br />

affected by speckle, while if one tries to increase the exposure times (or scan the incoming<br />

beam over the retina) to reduce the speckle, the interferogram contrast (and<br />

thus the SNR) will decrease.<br />

121


A. Preliminary study of feasibility of interferometric wavefront sensing in the<br />

eye<br />

5<br />

5<br />

5<br />

3<br />

3<br />

3<br />

1<br />

1<br />

1<br />

mm<br />

-1<br />

mm<br />

-1<br />

mm<br />

-1<br />

-3<br />

-3<br />

-3<br />

-5<br />

ad / 4.00 mW / 0.5s<br />

-6 -4 -2 0 2 4 6<br />

mm<br />

-5<br />

-5 0 5<br />

mm<br />

5<br />

5<br />

5<br />

-5<br />

-5 0 5<br />

mm<br />

(a) (b) (c)<br />

3<br />

3<br />

3<br />

1<br />

1<br />

1<br />

mm<br />

-1<br />

mm<br />

-1<br />

mm<br />

-1<br />

-3<br />

-3<br />

-3<br />

-5<br />

ad / 4.00 mW / 1.0s -5<br />

-6 -4 -2 0 2 4 6 -5 0 5<br />

mm<br />

mm<br />

5<br />

5<br />

5<br />

-5<br />

-5 0 5<br />

mm<br />

(d) (e) (f)<br />

3<br />

3<br />

3<br />

1<br />

1<br />

1<br />

mm<br />

-1<br />

mm<br />

-1<br />

mm<br />

-1<br />

-3<br />

-3<br />

-3<br />

-5<br />

ad / 4.00 mW / 3.0s -5<br />

-6 -4 -2 0 2 4 6 -5 0 5<br />

mm<br />

mm<br />

-5<br />

-5 0 5<br />

mm<br />

(g) (h) (i)<br />

Figure A.7: Interferograms and SH patterns that resulted from the 4 µW at 780 nm<br />

setup with exposures of 0.5, 1.0 and 3.0 s, with accommodation paralyzed with cyclopentolate<br />

hydrochloride 1%.<br />

122


Appendix B<br />

Laser safety calculations<br />

In all the experiments conducted for this <strong>thesis</strong>, the maximum permissible exposures<br />

(MPEs) of the human eye to electromagnetic radiation were calculated based on the<br />

British and European standard Safety of laser products (BS EN 60825-1:1994 with<br />

amendments 1, 2 and 3). The MPE calculations below will only consider retinal<br />

thermal damage as the cornea, aqueous humor, lens and vitreous humor are considered<br />

transparent for all the wavelengths of the sources used in the experiments (i.e. within<br />

the range 600 − 1400nm).<br />

apparent<br />

source<br />

a<br />

eye<br />

r<br />

Figure B.1: Definition of subtended angle α by an apparent source at a distance r.<br />

The diameter of the pupil of the eye applicable to measuring laser irradiance exposure<br />

for a wavelength between 400 and 1400 nm is 7 mm for t < 3 × 10 4 s, thus, the area of<br />

the pupil to be considered is<br />

A pupil ≈ 4 × 10 −5 m 2 .<br />

(B.1)<br />

123


B. Laser safety calculations<br />

B.1 MPE for He-Ne laser in tear topography experiment<br />

The wavelength λ of the He-Ne laser used for illuminating the precorneal tear film<br />

in the shear interferometer is 632.8 nm, and the angular subtense α of the apparent<br />

source for the tear topography experimental setup was greater than 300 mrad. Thus,<br />

the MPE as a function of time is given by<br />

{ 100<br />

MP E(t; λ = 632.8 nm, α > 100 mrad) = 18 C 6 A pupil ×<br />

−0.25 Wm −2 for t > 100 s<br />

t 0.75 Jm −2 for t ≤ 100 s<br />

(B.2)<br />

where the factor C 6 depends only on the angular subtense of the apparent source. In<br />

this case, α > α max , being α max = 100 mrad, therefore making C 6 take its maximum<br />

value, i.e. α max /α min = 67 and the MPE becomes<br />

{ 14.7 mW for t > 100 s<br />

MP E(t; λ = 632.8 nm, α > 100 mrad) =<br />

46 t 0.75 mJ for t ≤ 100 s<br />

(B.3)<br />

The exposure times for the experiments performed varied between 20 and 200 seconds<br />

maximum, thus the most restrictive MPE is<br />

MP E(t > 100 s; λ = 632.8 nm, α > 100 mrad) ≈ 15 mW<br />

(B.4)<br />

where this exposure is valid for up to 8.3 hours.<br />

B.2 MPE for He-Ne laser in full ocular wavefront sensor<br />

When the shear interferometer is modified to be a full ocular wavefront sensor, the<br />

beam illuminating the eye is collimated, making the angular subtense to be used in<br />

the calculations α min = 1.5 mrad. If under these conditions we consider the most<br />

restrictive exposures, i.e. 10s < t < 8.3 hs, the MPE should be calculated as<br />

MP E(t > 10 s; λ = 632.8 nm, α = 1.5 mrad) = A pupil × 10 Wm −2 (B.5)<br />

which is<br />

MP E(t > 10 s; λ = 632.8 nm, α = 1.5 mrad) = 0.4 mW<br />

(B.6)<br />

B.3 MPE for infrared LEDs<br />

The sources used for pupil (iris) illumination were 4 LEDs with a central wavelength<br />

(i.e. emission peak) λ c ≃ 880 nm. The following safety calculations are performed as<br />

124


B. Laser safety calculations<br />

if the LEDs were lasers, as indicated by the standard. The LEDs have a built-in lens<br />

after the diode, making the apparent angular size of each LED the diameter of this<br />

lens, that is 5mm, which then divided by the distance from the eye (50 mm), gives an<br />

apparent subtended angle α ≈ 100 mrad. Then, the MPE for one LED is given by<br />

MP E 1 (t > 100 s; λ = 880 nm, α > 100 mrad) = 18 C 4 C 6 C 7 A pupil × 100 −0.25 Wm −2<br />

(B.7)<br />

where C 4 and C 7 are wavelength dependent factors which for 880nm take the values<br />

2.29 and 1, respectively. Again, we will consider the most restrictive case, i.e. exposures<br />

greater than 100s and a value of 67 for C 6 , which gives the MPE for a single<br />

LED<br />

MP E 1 (t > 100 s; λ = 880 nm, α > 100 mrad) ≈ 33 mW (B.8)<br />

If we have a very restrictive approach, we can assume that the power of the 4 LEDs<br />

is coming from a subtended angle equal to that subtended by only one of them,<br />

simplifying the calculation. Thus, we can safely say that the MPE for the combination<br />

of the 4 LEDs at 50mm from the eye is<br />

MP E 1 (t > 100 s; λ = 880 nm, α > 100 mrad) ≥ 33 mW<br />

(B.9)<br />

B.4 MPE for infrared laser<br />

The source used for the Shack-Hartmann sensor was a laser diode emitting at λ =<br />

780 nm. The apparent angular size of the source to be considered is 1.5 mrad as the<br />

beam entering the eye is collimated, making the viewing distance infinite. For these<br />

conditions and exposures times larger than 10s, the MPE is given by<br />

MP E(t > 10 s; λ = 780 nm, α = 1.5 mrad) = 10 C 4 C 7 A pupil Wm −2<br />

(B.10)<br />

where C 4 and C 7 are wavelength dependent factors which for 780 nm take the values<br />

1.45 and 1, respectively, yielding<br />

MP E(t > 100 s; λ = 880 nm, α > 100 mrad) ≈ 0.6 mW<br />

(B.11)<br />

125


Bibliography<br />

[1] T. Young. The Bakerian lecture. on the mechanism of the eye. Phil. Trans.,<br />

91:23–88, 1801.<br />

[2] Franz Daxecker. Christoph scheiner’s eye studies. Documenta Ophthalmologica,<br />

81:27–35, 1992.<br />

[3] H. Helmholtz. Physiological optics, volume I. Optical Society of America, 1924.<br />

pages 160–203 and 416–443. Translated version of Handbuch der Physiologischen<br />

Optik, by J.P.C. Southall 1924.<br />

[4] H.S. Smirnov. Measurement of wave aberration in the human eye. Biophysics,<br />

6:52–66, 1961.<br />

[5] H.C. Howland and B. Howland. A subjective method for the measurement of<br />

monochromatic aberrations of the eye. J. Opt. Soc. Am., 67(11):1508–1518,<br />

1977.<br />

[6] Robert H. Webb, C. Murray Penney, and Keith P. Thompson. Measurement of<br />

ocular local wavefront distortion with a spatially resolved refractometer. App.<br />

Opt., 31(19):3678–3686, 1992.<br />

[7] J. C. He, S. Marcos, R. H. Webb, and S. A. Burns. Measurement of the wavefront<br />

aberration of the eye by a fast psychophysical procedure. J. Opt. Soc. Am.<br />

A, 15(9):2449–2456, 1998.<br />

[8] Francçoise Berny. Etude de la formation des images retiniennes et determination<br />

de l’aberration de sphericite de l’oeil humain. Vis. Res., 9.<br />

[9] G. Walsh, W.N. Charman, and H.C. Howland. Objective technique for the<br />

determination of monochromatic aberrations of the human eye. J. Opt. Soc.<br />

Am. A, 1(9):987–992, 1984.<br />

126


BIBLIOGRAPHY<br />

[10] G. Smith, R.A. Applegate, and D.A. Atchinson. Assesment of the accuracy of<br />

the crossed-cylinder aberroscope technique. J. Opt. Soc. Am. A, 15(9):2477–<br />

2487, 1998.<br />

[11] M.J. Cox and R. Calver. Reassesing the theoretical accuracy of the crossedcylinder<br />

aberroscope technique. J. Opt. Soc. Am. A, 16(10):2343–2351, 1999.<br />

[12] R. Navarro and M.A. Losada. Aberrations and relative efficiency of light pencils<br />

in the living human eye. Optom. Vis. Sci., 74(7):540–547, 1997.<br />

[13] R. Navarro and E. Moreno-Barriuso. Laser ray-tracing method for optical testing.<br />

Opt. Lett., 24(14):951–953, 1998.<br />

[14] E. Moreno-Barriuso and R. Navarro. Laser ray tracing versus Hartmann-Shack<br />

sensor for measuring optical aberrations in the human eye. J. Opt. Soc. Am. A,<br />

17(6):974–985, 2000.<br />

[15] Roberto Ragazzoni. Pupil plane wavefront sensing with an oscillating prism.<br />

[16] Ignacio Iglesias, Roberto Ragazzoni, Yves Julien, and Pablo Artal. Extended<br />

source pyramid wave-front sensor for the human eye. Opt. Exp., 10(9):419–428,<br />

2002.<br />

[17] I. Iglesias, E. Berrio, and P. Artal. Estimates of the ocular wave aberration from<br />

pairs of double pass retinal images. J. Opt. Soc. Am. A, 15(9):2466–2476, 1998.<br />

[18] J. Santamaría, P. Artal, and J. Bescós. Determination of the point-spread function<br />

of human eyes using a hybrid optical-digital method. J. Opt. Soc. Am. A,<br />

4(6):1109–1114, 1987.<br />

[19] I. Iglesias, N. López-Gil, and P. Artal. Reconstruction of the point-spread function<br />

of the human eye from two double-pass retinal images by phase-retrieval<br />

algorithms. J. Opt. Soc. Am. A, 15(2):326–339, 1998.<br />

[20] J. Liang, B. Grimm, S. Goelz, and J.F. Bille. Objective measurement of wave<br />

aberrations of the human eye with the use of a Hartmann-Shack wave-front<br />

sensor. J. Opt. Soc. Am. A, 11(7):1949–1957, 1994.<br />

[21] J. Liang and D.R. Williams. Aberrations and retinal image quality of the normal<br />

human eye. J. Opt. Soc. Am. A, 14(11):2873–2883, 1997.<br />

127


BIBLIOGRAPHY<br />

[22] P. Artal, S. Marcos, R. Navarro, and D.R. Williams. Odd aberrations and<br />

double-pass measurements of retinal image quality. J. Opt. Soc. Am. A,<br />

12(2):195–201, 1995.<br />

[23] Luis Diaz-Santana Haro and J. Chris Dainty. Single-pass measurements of the<br />

wave-front aberrations of the human eye by use of retinal lipofuscin autofluorescence.<br />

Opt. Lett., 24(1):61–63, 1999.<br />

[24] T.O. Salmon, L.N. Thibos, and A. Bradley. Comparison of the eye’s wave-front<br />

aberration measured psycophysically and with the Shack-Hartmann wave-front<br />

sensor. J. Opt. Soc. Am. A, 15(9):2457–2465, 1998.<br />

[25] Norberto Lopez-Gil and Pablo Artal. Comparison of double-pass estimates of<br />

the retinal-image quality obtained with green and near-infrared light. J. Opt.<br />

Soc. Am. A, 14(5):961–971, 1997.<br />

[26] Lourdes Llorente, Luis Diaz-Santana, David Lara-Saucedo, and Susana Marcos.<br />

Aberrations of the human eye in visible and near infrared illumination. Optom.<br />

Vis. Sci., 80(1):26–35, 2003.<br />

[27] L.N. Thibos and Xin Hong. Clinical applications of the shack-hartmann aberrometer.<br />

Optom. Vis. Sci., 76(12):817–825, 1999.<br />

[28] H. Hofer, P. Artal, B. Singer, J.L. Aragón, and D.R. Williams. Dynamics of the<br />

eye’s aberration. J. Opt. Soc. Am. A, 18(3):497–506, 2001.<br />

[29] Manuel P. Cagigal, Vidal F. Canales, José F. Castejón-Mochón, Pedro M. Prieto,<br />

Norberto López-Gil, and Pablo Artal. Statistical description of wave-front<br />

aberration in the human eye. Opt. Lett., 27(1):37–39, 2002.<br />

[30] Laura Oliveira-Soto and W. Neil Charman. Some possible longer-term ocular<br />

changes following excimer laser refractive surgery. Ophthal. Physiol. Opt., 22.<br />

[31] J. Schwiegerling. Wavefront guided lasik. Optics & Photonics News, 15(2):26–29,<br />

2000.<br />

[32] David R. Williams and Junzhong Liang (University of Rochester). Method and<br />

apparatus for improving vision and the resolution of retinal images. United<br />

States Patent (US 5,777,719), 1998.<br />

128


BIBLIOGRAPHY<br />

[33] Josef Bille. Method and apparatus for measurement of the refractive properties<br />

of the human eye. European Patent Office (EP 1 059 061 A2), 2000.<br />

[34] Rudolph W. Frey, James H. Burkhalter, Neil Zepkin, Edward Poppeliers, and<br />

John Alfred Campin (Alcon Universal Ltd.). Apparatus and method for objective<br />

measurements of optical systems using wavefront analysis. United State Patent<br />

(US 6,460,997 B1), 2002.<br />

[35] Daniel R. Neal, Darrell J. Armstrong, Daniel M. Topa, and Inc.) Richard J.<br />

Copland (Wavefront Sciences. Dynamic range extension techniques for a wavefront<br />

sensor including use in ophthalmic measurement. United States Patent<br />

(US 6,550,917 B1), 2003.<br />

[36] Gerhard Youssefi (Bausch and Lomb incorporated). Objective manifest refraction.<br />

World Intellectual Property Organization (WO 02/098290 A2), 2002.<br />

[37] Pablo Artal and Antonio Guirao. Contributions of the cornea and the lens to<br />

the aberrations of the human eye. Opt. Lett., 23(21):1713–1715, 1998.<br />

[38] H. W. Babcock. The possibility of compensating astronomical seeing.<br />

[39] Luis Diaz-Santana, Cristiano Torti, Ian Munro, Paul Gasson, and Chris Dainty.<br />

Benefit of higher closedloop bandwidths in ocular adaptive optics. Opt. Exp.,<br />

11(20):2597–2605, 2003.<br />

[40] E.J. Fernández, I. Iglesias, and P. Artal. Closed-loop adaptive optics in the<br />

human eye. Opt. Lett., 26(10):746–748, 2001.<br />

[41] Austin Roorda, Fernando Romero-Borja, William J. Donnelly III, Hope<br />

Queener, Thomas J. Hebert, and Melanie C. W. Campbell. Adaptive optics<br />

scanning laser ophthalmoscopy. Opt. Exp., 10(9):405–412, 2002.<br />

[42] Nathan Doble, Geunyoung Yoon, Li Chen, Paul Bierden, Ben Singer, Scott<br />

Olivier, and David R. Williams. Use of microelectromechanical mirror for adaptive<br />

optics in the human eye. Opt. Lett., 27(17):1537–1539, 2002.<br />

[43] Junzhong Liang, David R. Williams, and Donald T. Miller. Supernormal vision<br />

and high-resolution retinal imaging through adaptive optics. J. Opt. Soc. Am.<br />

A, 14(11):2884–2892, 1997.<br />

129


BIBLIOGRAPHY<br />

[44] Austin Roorda and David R. Williams. The arrangement of the three cone<br />

classes in the living human eye. Nature, 397(11):520–522, 1999.<br />

[45] Jean-François Le Gargasson, Marie Glanc, and Pierre Léna. Retinal imaging<br />

with adaptive optics. C. R. Acad. Sci. Paris, t. 2, Série IV, pages 1131–1138,<br />

2001.<br />

[46] M. Glanc, E. Gendron, F. Lacombe, D. Lafaille, J.-F. Le Gargasson, and P. Léna.<br />

Towards wide-field retinal imaging with adaptive optics. Opt. Comm., 230:225–<br />

238, 2004.<br />

[47] H. Hofer, L. Chen, G.Y. Yoon, B. Singer, Y. Yamamuchi, and D.R. Williams.<br />

Improvement in retinal image quality with dynamic correction of the eye’s aberrations.<br />

Opt. Exp., 8(11):631–643, 2001.<br />

[48] Robert H. Webb, George W. Hughes, and Francois C. Delori. Confocal scanning<br />

laser ophthalmoscope. App. Opt., 26(8):1492–1499, 1987.<br />

[49] Daniel X. Hammer, R. Daniel Ferguson, John C. Magill, Michael A. White,<br />

Ann E. Elsner, and Robert H. Webb. Compact scanning laser ophthalmoscope<br />

with high-speed retinal tracker. App. Opt., 42(22):4621–4632, 2003.<br />

[50] A. R. Wade and F. W. Fitzke. In vivo imaging of the human cone-photoreceptor<br />

mosaic using a confocal lso. Laser and Light in Ophthalmology, 8.<br />

[51] Adrian Gh. Podoleanu, George M. Dobre, David J. Webb, and David A. Jackson.<br />

Simultaneous en-face imaging of two layers in the human retina by low-coherence<br />

reflectometry. Opt. Lett., 22(13):1039–1041, 1997.<br />

[52] Adrian Gh. Podoleanu, John A. Rogers, and David A. Jackson. Three dimensional<br />

oct images from retina and skin. Opt. Exp., 7(9):292–298, 2000.<br />

[53] R. Navarro, P. Artal, and D.R. Williams. Modulation transfer of the human eye<br />

as a function of retinal eccentricity. J. Opt. Soc. Am. A, 10(2):201–212, 1993.<br />

[54] R. Navarro, E. Moreno-Barriuso, and C. Dorronsoro. Monochromatic aberrations<br />

and point-spread functions of the human eye across the visual field. J.<br />

Opt. Soc. Am., 15(9):2522–2529, 1998.<br />

130


BIBLIOGRAPHY<br />

[55] Salvador Bará and Rafael Navarro. Wide-field compensation of monochromatic<br />

eye aberrations: expected performance and design trade-offs. J. Opt. Soc. Am.<br />

A, 20(1):1–10, 2003.<br />

[56] David Peter Catlin. High resolution imaging of the human retina. <strong>PhD</strong> <strong>thesis</strong>,<br />

<strong>Imperial</strong> <strong>College</strong> of Science, Technology and Medicine, University of <strong>London</strong>,<br />

2001.<br />

[57] David Catlin and Christopher Dainty. High-resolution imaging of the human<br />

retina with a Fourier deconvolution technique. J. Opt. Soc. Am. A, 19(8):1515–<br />

1523, 2002.<br />

[58] Ignacio Iglesias and Pablo Artal. High-resolution retinal images obtained by<br />

deconvolution from wave-front sensing. Opt. Lett., 25(24):1804–1806, 2000.<br />

[59] R. Navarro and E. Moreno-Barriuso. Phase plates for wave-aberration compensation<br />

in the human eye. Opt. Lett., 25(4):1–3, 2000.<br />

[60] F. Vargas-Martin, P.M. Prieto, and P. Artal. Correction of the aberrations<br />

in the human eye with a liquid-crystal spatial light modulator: limits to its<br />

performance. J. Opt. Soc. Am. A, 15(9):2552–2562, 1998.<br />

[61] Geun-Young Yoon and David R. Williams. Visual performance after correcting<br />

the monochromatic and chromatic aberrations of the eye. J. Opt. Soc. Am. A,<br />

19(2):266–275, 2002.<br />

[62] Karen Hampson. Personal communication. 2003.<br />

[63] Gleb Vdovin. On the possibility of intraocular adapative optics. Opt. Exp.,<br />

11(7):810–817, 2003.<br />

[64] Luis Diaz-Santana. Lecturer, Applied Vision Research Centre, Department of<br />

Optometry an Visual Science, City University, <strong>London</strong>, UK. Personal communication.<br />

2004.<br />

[65] F.W. Campbell, J.G. Robson, and G. Westheimer. Fluctuations of accommodation<br />

under steady viewing conditions. J. Phys., 145.<br />

[66] L.S. Gray, B. Winn, and B. Gilmartin. Effect of target luminance on microfluctuations<br />

of accommodation. Ophthal. Physiol. Opt., 13.<br />

131


BIBLIOGRAPHY<br />

[67] W.N. Charman and G. Heron. Fluctuations in accommodation: a review. Ophthal.<br />

Physiol. Opt., 8.<br />

[68] R. Tutt, A Bradley, C Begley, and L.N. Thibos. Optical and visual impact of<br />

tear break-up in human eyes. Invest. Ophth. Vis. Sci., 41(13):4117–4123, 2000.<br />

[69] W. H. Hart Jr. Adler’s physiology of the eye: clinical application. Mosby-<br />

Year Book Inc., 11830 Westline Industrial Drive, St. Louis, Missouri 63146, 9th<br />

edition, 1992.<br />

[70] Xu Cheng, Nikole L. Himebaugh, Pete S. Kollbaum, Larry N. Thibos, and<br />

Arthur Bradley. Testretest reliability of clinical shack-hartmann measurements.<br />

Invest. Ophth. Vis. Sci., 45(1):351–360, 2004.<br />

[71] Nikole L. Himebaugh, Annette R. Wright, Arthur Bradley, Carlolyn G. Begley,<br />

and Larry Thibos. Use of retroillumination to visualize optical aberrations<br />

caused by tear film break-up. Optom. Vis. Sci., 80(1):69–78, 2003.<br />

[72] Shizuka Koh, Naoyuki Maeda, Teruhito Kuroda, Yuichi Hori, Hitoshi Watanabe,<br />

Takashi Fujikado, Yasuo Tano, Yoko Hirohara, and Toshifumi Mihashi. Effect of<br />

rear film break-up on higher-order aberrations measured with wavefront sensor.<br />

Am. J. of Ophthalmol., 134(1):115–117, 2002.<br />

[73] T.J. Licznerski, H.T. Kasprzak, and W. Kowalik. Two interference techniques<br />

for in vivo assesment of the tear film stability on a cornea and contact lens.<br />

Proc. SPIE, 3320:183–186, 1998.<br />

[74] T.J. Licznerski, H.T. Kasprzak, and W. Kowalik. Analysis of shearing interferograms<br />

of tear film using fast Fourier transform. J. Biomed. Opt., 3(1):32–37,<br />

1998.<br />

[75] Monika I. Lechna-Marczynska, Tomasz J. Licznerski, and Henryk T. Kasprzak.<br />

Interferometry for in-vivo testing of artificial tears on the surface of the cornea.<br />

Proc. SPIE, 3820:386–389, 1999.<br />

[76] Henryk Kasprzak. The role of tear film in image quality. Agean 2 nd Summer<br />

School in Visual Optics, July 2003.<br />

[77] G. Smith and D.A. Atchison. The eye and visual optical instrumentation. Cambridge<br />

University Press, Cambridge, U.K., 1st edition, 1997.<br />

132


BIBLIOGRAPHY<br />

[78] G. Wald and D. R. Griffin. The change in refractive power of the human eye in<br />

dim and bright light. J. Opt. Soc. Am., 37(5):321–336, 1947.<br />

[79] Maurizio Rolando and Manfred Zierhut. The ocular surface and tear film and<br />

their dysfunction in dry eye disease. Surv. Ophthalmol., 45(2):S203–S210, 2001.<br />

[80] J.P. Craig, P.A. Simmons, S. Patel, and A. Tomlinson. Refractive index and<br />

osmolality of human tears. Optom. Vis. Sci., 72(10):718–724, 1995.<br />

[81] J.L. Prydal and F.W. Campbell. Study of precorneal tear film thickness and<br />

structure by interferometry and confocal microscopy. Invest. Ophth. Vis. Sci.,<br />

33(6):1996–2005, 1992.<br />

[82] D.G. Green, B.R. Frueh, and J.M. Saphiro. Corneal thickness measured by<br />

interferometry. J. Opt. Soc. Am., 65(2):119–123, 1975.<br />

[83] J.L. Prydal, P. Artal, H. Woon, and F.W. Campbell. Study of precorneal tear<br />

film thickness and structure using laser interferometry. Invest. Ophth. Vis. Sci.,<br />

33(6):2006–2011, 1992.<br />

[84] Y. Danjo, M. Nakamura, and T. Hamano. Measurement of the precorneal tear<br />

film thickness with a non–contact optical interferometry film thickness measurement<br />

system. Jpn. J. Ophthalmol., 38:260–266, 1994.<br />

[85] P.E. King-Smith, B.A. Fink, N. Fogt, K.A. Kinney, and R.M. Hill. Is the thickness<br />

of the tear film about 40µm or about 3µm? Invest. Ophth. Vis. Sci.,<br />

40(4):2876B751, 1999.<br />

[86] J.H. Wang, D. Fonn, T. L. Simpson, and L. Jones. Pre-corneal and pre- and postlens<br />

tear film thickness measured with optical coherence tomography. Invest.<br />

Ophth. Vis. Sci., 43:3078 Suppl., 2002.<br />

[87] N. Fogt and P.E. King-Smith. Interferometric measurement of the tear film<br />

thickness by use of spectral oscillations. J. Opt. Soc. Am. A, 15(1):268–275,<br />

1998.<br />

[88] P. Ewen King-Smith, Barbara A. Fink, Nick Fogt, Kelly K. Nichols, Richard M.<br />

Hill, and Graeme S. Wilson. The thickness of the human precorneal tear film:<br />

evidence from reflection spectra. Invest. Ophth. Vis. Sci., 41(11):3348–3359,<br />

2000.<br />

133


BIBLIOGRAPHY<br />

[89] J.P. Craig, P.A. Simmons, S. Patel, and A. Tomlinson. In vivo tear-film thickness<br />

determination and implications for tear-film stability. Optom. Vis. Sci.,<br />

72(10):718–724, 1995.<br />

[90] M. Rolando, M. Iester, A. Macri, and G. Calabria. Low spatial-contrast sensitivity<br />

in dry eyes. 17(4):376–379, 1998.<br />

[91] M. Born and E. Wolf. Principle of optics. Pergamon Press, Headlington Hill<br />

Hall, Oxford OX3 0BW, England, sixth (corrected) edition, 1980.<br />

[92] D. Malacara. Optical shop testing. John Wiley and Sons, 1 edition, 1978.<br />

[93] José A. Ferrari, Erna M. Frins, Daniel Perciante, and <strong>Alfredo</strong> Dubra. Robust<br />

one-beam interferometer with phase-delay control. Opt. Lett., 24(18):1272–1274,<br />

1999.<br />

[94] D. Malacara. Optical shop testing. John Wiley and Sons, 2 edition, 1992.<br />

[95] M. Takeda, H. Ina, and S. Kobayashi. Fourier-transform method of fringepattern<br />

analysis for computer based topography and interferometry. J. Opt.<br />

Soc. Am., 72(1):156–160, 1982.<br />

[96] Chung-Pin Cherng, Theodore C. Salvi, Marek Osinski, and John G. McInerney.<br />

Near field wavefront measurements of semiconductor laser arrays by shearing<br />

interferometry. App. Opt., 29(18):2701–2706, 1990.<br />

[97] Adolf Lohmann and Olof Bryngdahl. A lateral wavefront shearing interferometer<br />

with variable shear. App. Opt., 6(11):1934–1937, 1967.<br />

[98] Chris L. Koliopoulos. Radial grating lateral shear heterodyne interferometer.<br />

App. Opt., 19(9):1523–1528, 1980.<br />

[99] J. C. Wyant. Double frequency grating lateral shear interferometer. App. Opt.,<br />

12(9):2057–2060, 1973.<br />

[100] Dennis C. Ghiglia and Mark D. Pritt. Two-dimensional Phase Unwrapping:<br />

Theory, Algorithms and Software. John Wiley & Sons, Inc., 1998.<br />

[101] <strong>Alfredo</strong> Dubra, Carl Paterson, and J. Christopher Dainty. Wave-front reconstruction<br />

from shear phase maps by use of the discrete Fourier transform. App.<br />

Opt., 43(5):1108–1113, 2004.<br />

134


BIBLIOGRAPHY<br />

[102] R.H. Hudgin. Wave-front reconstruction for compensated imaging. J. Opt. Soc.<br />

Am., 67(3):375–378, 1977.<br />

[103] D.L. Fried. Least-square fitting a wave-front distortion estimate to an array of<br />

phase-difference measurements. J. Opt. Soc. Am., 67(3):370–375, 1977.<br />

[104] W.H. Southwell. Wave-front estimation from wave-front slope measurements.<br />

J. Opt. Soc. Am., 1980.<br />

[105] E. Anderson, Z. Bai, C. Bischof, S. Blackford, J. Demmel, J. Dongarra, J. Du<br />

Croz, A. Greenbaum, S. Hammarling, A. McKenney, and D. Sorensen. LAPACK<br />

Users’ Guide. Society for Industrial and Applied Mathematics, Philadelphia,<br />

PA, third edition, 1999.<br />

[106] Klaus R. Freischlad and Chris L. Koliopoulos. Modal estimation of a wave front<br />

from difference measurements using the discrete Fourier transform. J. Opt. Soc.<br />

Am. A, 3(11):1852–1861, 1986.<br />

[107] François Roddier and Claude Roddier. Wavefront reconstruction using iterative<br />

Fourier transforms. App. Opt., 30(11):1325–1327, 1991.<br />

[108] Clemens Elster and I. Weingärtner. Exact wave-front reconstruction from two<br />

lateral shearing interferograms. J. Opt. Soc. Am. A, 16(9):2281–2285, 1999.<br />

[109] Clemens Elster and Ingolf Weingärtner. Solution to the shearing problem. App.<br />

Opt., 38(23):5024–5031, 1999.<br />

[110] Clemens Elster. Exact two-dimensional wave-front reconstruction from lateral<br />

shearing interferograms with large shears. App. Opt., 39(29):5353–5359, 2000.<br />

[111] Lisa A. Poyneer, Donald T. Gavel, and James M. Brase. Fast wave-front reconstruction<br />

in large adaptive optics systems with use of the Fourier transform. J.<br />

Opt. Soc. Am. A, 19(10):2100–2111, 2002.<br />

[112] Alan V. Oppenheim and Ronald W. Schafer. Digital Signal Processing. Prentice-<br />

Hall, Inc., Englewood Cliffs, New Jersey 07632, 1975.<br />

[113] Dan Z. Reinstein. Consultation Section. J. Ref. Surg., 27(9), 2001.<br />

[114] Austin Roorda. Associate professor, university of houston, college of optometry,<br />

houston tx 77204-2020, US. Personal communication. 2002.<br />

135


BIBLIOGRAPHY<br />

[115] Frangois C. Delori and Kent P. Pflibsen. Spectral reflectance of the human<br />

ocular fundus. App. Opt., 28(6):1061–1077, 1989.<br />

[116] Luis Diaz-Santana. Wavefront Sensing in the Human Eye with a ShackHartmann<br />

Sensor. <strong>PhD</strong> <strong>thesis</strong>, <strong>Imperial</strong> <strong>College</strong> of Science, Technology and Medicine,<br />

University of <strong>London</strong>, 2000.<br />

[117] L.N. Thibos, R.A. Applegate, J.T. Schwiegerling, R. Webb, and VSIA Standards<br />

Taskforce Members. Standards for reporting the optical aberrations of<br />

eyes. 2000.<br />

[118] Gijsbertus Johannes Van Blokland. The optics of the human eye studied with<br />

respect to polarized light. <strong>PhD</strong> <strong>thesis</strong>, Rijksuniversiteit Utrecht, 1986.<br />

136

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!