143
Aberration correction with sensorless adaptive optics for imaging the mouse retina by Daniel John Wahl B.S., University of Northern British Columbia, 2014 Thesis Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in the School of Engineering Science Faculty of Applied Sciences © Daniel John Wahl 2019 SIMON FRASER UNIVERSITY Summer 2019 Copyright in this work rests with the author. Please ensure that any reproduction or re-use is done in accordance with the relevant national copyright legislation.

Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

Aberration correction with sensorless adaptive

optics for imaging the mouse retina

by

Daniel John Wahl

B.S., University of Northern British Columbia, 2014

Thesis Submitted in Partial Fulfillment of the

Requirements for the Degree of

Doctor of Philosophy

in the

School of Engineering Science

Faculty of Applied Sciences

© Daniel John Wahl 2019

SIMON FRASER UNIVERSITY

Summer 2019

Copyright in this work rests with the author. Please ensure that any reproduction or re-use is done in accordance with the relevant national copyright legislation.

Page 2: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

ii

Approval

Name:

Degree:

Title:

Examining Committee:

Date Defended/Approved:

Daniel John Wahl

Doctor of Philosophy

Aberration correction with sensorless adaptive optics for imaging the mouse retina

Chair: Bonnie Gray Professor

Marinko V. Sarunic Senior Supervisor Professor

Yifan Jian Supervisor Assistant Professor

Mirza Faisal Beg Supervisor Professor

Robert J. Zawadzki Supervisor Associate Research Professor

Pierre Lane Internal Examiner Associate Professor of Professional Practice

Jennifer Hunter External Examiner Associate Professor Flaum Eye Institute Department of Ophthalmology Department of Biomedical Engineering Center for Visual Science The Institute for Optics University of Rochester

May 2, 2019

Page 3: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

iii

Ethics Statement

Page 4: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

iv

Abstract

Small animals, such as mice, are commonly used in biomedical research as models for

studying human diseases. Imaging the retina in a living animal can provide valuable

insights into the causes and mechanisms of vision loss. However, often imaging in vivo

results in low resolution due to optical aberrations that can be caused by the biological

tissue in front of the retina. Imaging systems that could non-invasively image the mouse

retina with cellular-level resolution would be beneficial to many vision scientists.

Adaptive optics (AO) is a technology that was originally developed for astronomers to

image through the turbulent atmosphere. AO technology has been extended for

microscopy and ophthalmoscopy to restore imaging performance lost due to optical

aberrations from biological samples. Often, AO systems employ a wavefront sensor for

direct measurement of the aberrations. Alternatively, Sensorless AO (SAO) has been

implemented for imaging into tissue with multiple scattering layers, which can confound

the optical wavefront measurements from a single imaging plane.

In this thesis, I present several imaging systems for imaging the mouse retina with cellular-

level resolution by using custom and novel SAO methods. The imaging modalities include

Scanning Laser Ophthalmoscopy with fluorescence detection, Optical Coherence

Tomography, and Two-Photon Excited Fluorescence imaging. The simple and robust

optical designs in this thesis feature wide imaging field of views for navigation and a

compactable system layout. Using SAO enables depth-resolved aberration correction in

the different layers of the mouse retina. My results demonstrate detailed non-invasive

cellular imaging capabilities in the living mouse eye of GFP labelled cells, nerve fibers

bundles, volumetric imaging of vasculature, as well as the RPE mosaic of the outer retina.

Keywords: Sensorless Adaptive Optics; Scanning Laser Ophthalmoscopy; Optical

Coherence Tomography; Two-Photon Excited Fluorescence; Mouse

Retina

Page 5: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

v

Dedication

To my grandfather, Edward Hark.

Page 6: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

vi

Acknowledgements

I would like to acknowledge the contributions of my supervisor, Dr. Marinko

Sarunic. His dedication to research, teaching, and his students is above and beyond,

which has enabled the work presented in this thesis. I would like to thank all of the

supervisors including, Drs. Marinko Sarunic, Yifan Jian, Mirza Faisal Beg, and Robert

Zawadzki for sharing their knowledge through mentorship and guidance during my

program at Simon Fraser University. It has been a privilege to work with this group of

advisors who have inspired me to keep striving.

I am grateful to all of the past and present members of the Biomedical Optics

Research Group for making this a great place to be. I am thankful for the collaborative and

friendly environment created by the people in our group.

There have been many people who have helped me along the way. However, in

particular I would like to thank Dr. Pengfei Zhang for his contributions to the work

presented in Chapter 5. Also, I would like to thank Ms. Christine Huang for her

contributions in Chapter 4 during her undergraduate thesis. And, a special thanks to Mr.

Ringo Ng for supporting everyone in the lab with great designs and fabrication work,

presented in Chapter 6.

Finally, I would like to thank my all of my family and friends for supporting me

during many years of post-secondary school and into the future.

Page 7: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

vii

Table of Contents

Approval .......................................................................................................................... ii

Ethics Statement ............................................................................................................ iii

Abstract .......................................................................................................................... iv

Dedication ....................................................................................................................... v

Acknowledgements ........................................................................................................ vi

Table of Contents .......................................................................................................... vii

List of Tables ................................................................................................................... x

List of Figures................................................................................................................. xi

List of Acronyms .......................................................................................................... xviii

Chapter 1. Introduction .............................................................................................. 1

1.1. Overview ............................................................................................................... 1

1.2. The mouse eye ...................................................................................................... 2

1.3. Imaging the mouse retina ...................................................................................... 6

1.4. Outline ................................................................................................................... 8

1.5. Contributions ......................................................................................................... 9

Chapter 2. Background on retinal imaging systems and adaptive optics ............ 10

2.1. Scanning Laser Ophthalmoscopy ........................................................................ 10

2.2. Optical Coherence Tomography .......................................................................... 11

2.3. Fluorescence imaging ......................................................................................... 13

2.4. Adaptive optics for ophthalmic imaging ................................................................ 15

2.5. Summary ............................................................................................................. 19

Chapter 3. Wavefront sensorless adaptive optics fluorescence biomicroscope for in vivo retinal imaging in mice .................................................................... 20

3.1. Introduction .......................................................................................................... 20

3.2. Methods .............................................................................................................. 22

3.2.1. Mouse handling ........................................................................................... 23

3.2.2. Biomicroscope optical setup ........................................................................ 23

3.2.3. Image acquisition and optimization .............................................................. 25

3.3. Results ................................................................................................................ 27

3.3.1. WSAO f/c biomicroscope resolution ............................................................. 27

3.3.2. In vivo WSAO confocal fluorescence imaging of retinal ganglion cells ......... 28

3.3.3. In vivo WSAO confocal fluorescence imaging of retinal microglia cells ........ 30

3.4. Discussion ........................................................................................................... 31

3.5. Summary ............................................................................................................. 34

Chapter 4. Pupil segmentation adaptive optics for in vivo mouse retinal fluorescence imaging ........................................................................................ 35

4.1. Introduction .......................................................................................................... 35

4.2. Methods .............................................................................................................. 37

4.3. Discussion ........................................................................................................... 42

Page 8: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

viii

4.4. Summary ............................................................................................................. 44

Chapter 5. Adaptive optics in the mouse eye: Wavefront sensing based vs. image-guided aberration correction ................................................................. 45

5.1. Introduction .......................................................................................................... 45

5.2. Methods .............................................................................................................. 47

5.2.1. AO SLO system description ......................................................................... 47

5.2.2. WFS AO description .................................................................................... 49

5.2.3. WFS-less AO algorithm. .............................................................................. 50

5.2.4. WFS and WFS-less AO system calibration .................................................. 51

5.2.5. Animal handling and image processing ....................................................... 53

5.3. Results ................................................................................................................ 54

5.3.1. WFS and WFS-less AO for phantom imaging, comparison of performance . 54

5.3.2. WFS and WFS-less AO comparison on mouse photoreceptor mosaic ........ 57

5.3.3. AO SLO reflectance imaging of an albino mouse strain ............................... 60

5.3.4. AO SLO fluorescence imaging of EGFP microglia cells ............................... 61

5.4. Discussion ........................................................................................................... 64

5.5. Summary ............................................................................................................. 68

Chapter 6. Multi-modal imaging .............................................................................. 69

6.1. Introduction .......................................................................................................... 69

6.2. Methods .............................................................................................................. 70

6.2.1. Optical design .............................................................................................. 70

6.2.2. Sensorless adaptive optics .......................................................................... 75

6.2.3. Animal handling ........................................................................................... 76

6.2.4. Image processing ........................................................................................ 76

6.3. Results ................................................................................................................ 77

6.3.1. Imaging without adaptive optics ................................................................... 77

6.3.2. Structural imaging with sensorless adaptive optics OCT and SLO ............... 79

6.3.3. Fluorescence imaging with sensorless adaptive optics ................................ 81

6.4. Discussion ........................................................................................................... 86

6.5. Summary ............................................................................................................. 88

Chapter 7. Non-invasive cellular-resolution imaging of the retina with two-photon excited fluorescence ............................................................................ 89

7.1. Introduction .......................................................................................................... 89

7.2. Methods .............................................................................................................. 90

7.2.1. System setup ............................................................................................... 90

7.2.2. Animal handling and image processing ....................................................... 93

7.3. Results ................................................................................................................ 95

7.3.1. Fluorescein angiography ............................................................................. 95

7.3.2. GFP and YFP labelled cells ......................................................................... 98

7.3.3. RPE imaging ............................................................................................. 101

7.4. Discussion ......................................................................................................... 105

7.5. Summary ........................................................................................................... 107

Page 9: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

ix

Chapter 8. Future work and conclusion................................................................ 108

8.1. Technology refinement ...................................................................................... 108

8.2. Non-confocal Scanning Laser Ophthalmoscopy ................................................ 109

8.3. Extensions of two-photon excited florescence technology ................................. 110

8.4. Conclusion......................................................................................................... 111

References ................................................................................................................. 113

Page 10: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

x

List of Tables

Table 2.1. Zernike polynomials, names and index up to the 5th radial order. ........... 16

Table 5.1. Key optical parameters of the AO-SLO system components .................. 48

Table 7.1. Laser specifications used for each fluorescent sample and the calculated resolution. .............................................................................................. 92

Table 7.2. Summary of mice that were used in this report. ...................................... 93

Page 11: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

xi

List of Figures

Figure 1.1. Simplified schematic of the mouse eye compared to a human eye [6]. ..... 3

Figure 1.2. Organization of the retinal layers. The image is from Webvision [8] and used under the Creative Commons Licenses. .......................................... 4

Figure 1.3. The visual cycle and the location of each step. Image is from Wikimedia.org and used here under the Creative Commons License. ...... 5

Figure 1.4. The options for focusing light onto the mouse retina: collimated light or focused light from an objective lens.......................................................... 7

Figure 2.1. Simplified Jablonski diagram of single photon excited fluorescence and two-photon excited fluorescence. The excitation light, λex, and emission light, λem. ................................................................................................ 14

Figure 3.1. Schematic of the WSAO f/c biomicroscope using 488 nm excitation from an Ar/Kr laser. Relay lenses are achromatic doublets. Other optical elements: 80/20 beam splitter (BS), dichroic mirror (DC), deformable mirror (DM), zero-order quarter wave plate (QWP), objective lens (OBJ), linear polarizer (LP), pinhole (PH), variable lens (VL), galvanometer scanning mirrors (GM). Electronic elements: avalanche photo diode (APD), photo multiplier tube (PMT). The images on the computer icon are representative images of the structural and fluorescence imaging channels................................................................................................. 25

Figure 3.2. WSAO modal hill-climbing algorithm flowchart for the fluorescence image optimization process; deformable mirror (DM), variable lens (VL)........... 27

Figure 3.3. US Air Force resolution target with line width 2.19 µm highlighted by the red rectangle to demonstrate the reflectance resolution. Scale bar: 50 µm. ............................................................................................................... 28

Figure 3.4. Images of 2.1 µm diameter fluorescent beads acquired (a) before WSAO optimization and (b) after optimization. (c) The line plots for a bead before and after optimization. Scale bars: 10 µm. ............................................. 28

Figure 3.5. (a,b) Ganglion cells labelled by EGFP comparing the images acquired before and after the WSAO optimization. These images are an average of 50 frames of an off-axis ganglion cell. Scale bars: 20 µm. ...................... 29

Figure 3.6. (a) The Zernike coefficients applied to the DM (deformable mirror) after the optimization. (b) The impact of the optimization on the intensity-based merit function are plotted for each mode. The intensity is normalized from zero when the DM is flat. The Zernike coefficients are reported by the OSA standard for optical aberrations of eyes [68]. (c) The intensity plot of a dendrite on the EGFP-labelled ganglion cell at the location and in the direction indicated by the arrows. ........................................................... 30

Figure 3.7. Images of EGFP-labelled retinal microglia cells acquired in vivo before and after WSAO correction with different field of views: (a), (b), and (c). Images (b) and (c) were taken at the same location with different field of views as indicated by the red dashed box. Each image is an average of 50 frames. Scale bars: 10 µm. ............................................................... 31

Figure 4.1. Schematic of the Scanning Laser Ophthalmoscope: 488 nm laser; dichroic mirror (DC); deformable mirror (DM); variable lens (VL);

Page 12: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

xii

galvanometers (GM); {L1,L2, L3,L4,L5,L6} = {200,200,150,100,50,19} mm. (a),(b) PS-AO on 6 µm fluorescent beads with aberration correction (AO On) and without (AO Off). These images are an average of 30 frames. Scale bar: 8 µm. (c) Wavefront aberration map. (d) Normalized intensity plotted at the location indicated by the dashed lines with a ~30% increase in the peak intensity after correction. (e) The Zernike coefficients for the corrected wavefront. ............................................................................... 38

Figure 4.2. Aberration correction performed with both hill-climbing and PS-AO. (a) Image without aberration correction. (b) Correction performed with hill-climbing. (c) Correction performed with PS-AO. (d) The Zernike coefficients for the corrected wavefronts. ............................................... 40

Figure 4.3. PS-AO aberration correction on (a) static and (b, c) moving samples. The aberration correction was performed with (b) using the multiple intra-frame reference images and (c) a single reference image. ..................... 41

Figure 4.4. (a), (b) PS-AO for retinal fluorescein angiography with aberration correction (AO On) and without (AO Off) for two mice. In each panel, the top row of images (angular FOV 5.2°) is an optically zoomed in section of the bottom row of images (angular FOV 10.4°). Scale bars: 20 µm. (c) Zernike coefficients for the corrected wavefront. (d) On the top panel, the normalized intensity plot at the location indicated by the dashed lines had a ~30% increase in the peak intensity after correction, and (d) on the bottom panel, the wavefront aberration map. ......................................... 42

Figure 5.1. Adaptive Optics Scanning Laser Ophthalmoscopy (AO-SLO) system schematic. The layout is presented in a scale drawing. Abbreviations: L#, lens; F#, filter; BS#, beamsplitter; M, mirror; SM, spherical mirror; DM, deformable mirror; D#, dichroic mirror; Hsc, horizontal resonant scanner; Vsc, vertical scanner; PMT, photomultiplier tube; P (circled in blue) optical planes conjugate with the pupil; SLD, superluminescent diode. Collimated beams are marked as dashed lines and focusing beams are marked as solid lines. The on-axis beams are represented by red lines and scanned beams by green and blue. Image credit: Pengfei Zhang. ........................ 48

Figure 5.2. Phantom imaging of fluorescent beads and wavefront measurements during Wavefront Sensor Adaptive Optics (WFS AO) and Wavefront Sensorless Adaptive Optics (WFS-less AO). (a) Fluorescence images of 30 µm beads on white paper with a 100 mm focal length model eye before AO, after WFS AO, and after WFS-less AO. For the inset image before AO, the pixel intensity values were multiplied by 8, so the beads could be visualized. (b) The increase in the fluorescence image quality during the WFS-less AO optimization. (c) The wavefront RMS excluding defocus, tip and tilt during WFS AO correction. (d) The wavefront RMS excluding defocus, tip and tilt during WFS-less AO optimization. (e) The Zernike decomposition of the wavefront measured before and after each method of AO correction. ....................................................................... 56

Figure 5.3. Imaging the mouse photoreceptor mosaic with Wavefront Sensor based Adaptive Optics (WFS AO) and Wavefront Sensorless Adaptive Optics (WFS-less AO). (a,b) Images after WFS AO and WFS-less AO. Scale bar: 10 µm. (c) The image quality improvement during WFS-less AO optimization. (d) The wavefront RMS during WFS-less AO optimization.

Page 13: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

xiii

(e) The Zernike decomposition of the wavefront measured before and after each method of AO. ....................................................................... 58

Figure 5.4. (a, b) Further mouse photoreceptor imaging with Wavefront Sensor Adaptive Optics (WFS AO) and Wavefront Sensorless Adaptive Optics (WFS-less AO). Images mouse photoreceptor mosaic after WFS AO and WFS-less AO. Scale bar 10 µm. The Zernike decomposition of the wavefront measured before and after each method of AO. The wavefront RMS during WFS-less AO optimization. The image quality improvement during WFS-less AO optimization. .......................................................... 59

Figure 5.5. SH-WFS measurements from an Albino mouse strain (BALB/cJ) retina. (a) The SH-WFS centroids of an albino mouse compared to a pigmented mouse. (b) The RMS of the wavefront measurement without defocus. (c) The image quality metric during WFS-less AO optimization. .................. 60

Figure 5.6. Imaging the inner retinal of an Albino mouse (BALB/cJ) retina with Wavefront Sensorless Adaptive Optics (WFS-less AO). Images of the retina vasculature before and after WFS-less AO in the Nerve Fiber Layer (NFL), and after WFS-less AO in the Plexiform Layer (IPL), and Outer Plexiform Layer (OPL). Scale bar: 10 µm. .............................................. 61

Figure 5.7. Imaging EGFP labelled microglia with Wavefront Sensor Adaptive Optics (WFS AO). (a) Reflectance imaging in the inner retinal blood vessels. (b) Fluorescence imaging of EGFP labelled microglia. (c) The fluorescence image superimposed in green on the reflectance image in magenta. Scale bar: 20 µm. (d) The measured wavefront RMS during WFS AO without defocus. (e) The wavefront measurements in Zernike decomposition before and after the WFS AO aberration correction. ............................... 62

Figure 5.8. (a) Imaging EGFP labeled microglia within the inner retina of a mouse with Wavefront Sensor based Adaptive Optics (WFS AO) and Wavefront Sensorless Adaptive Optics (WFS-less AO). Fluorescence image with WFS AO aberration correction (left). Fluorescence image with WFS-less AO aberration correction (middle). Fluorescence images before and after WFS-less AO with a ~40 µm FOV (right). Scale bar: 20 µm. (b) The intensity line plot between the red arrows on the WFS AO image and between the blue arrows on the WFS-less AO image............................. 63

Figure 5.9. (a) Imaging EGFP labeled microglia within the inner retina of a mouse with Wavefront Sensor based Adaptive Optics (WFS AO) and Wavefront Sensorless Adaptive Optics (WFS-less AO). Fluorescence image after WFS AO (left). Fluorescence image after WFS AO and WFS-less AO aberration correction of residual aberration (middle). Fluorescence image with a smaller FOV of microglia dendrites superimposed in green on the reflectance image of the retinal blood vessels in magenta (right). (b) The Zernike decomposition of the wavefront measured before WFS AO and after both methods of AO. Scale bar: 20 µm........................................... 64

Figure 6.1. (a) Schematic of Optical Coherence Tomography (OCT) and confocal Scanning Laser Ophthalmoscopy (SLO) system. The cyan represents the beam path of only 488 nm light, the green represents the beam path of only the fluorescence emission and the red represents the beam path of only the SLD light. The pink represents the co-aligned beam path of the 488 nm light, fluorescence emission, and SLD light. System components: Superluminescent diode (SLD), fiber coupler (FC), polarization controller

Page 14: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

xiv

(PC), polarization beam splitter (PBS), dichroic mirror (DC), emission filter (EF), cold mirror (CM), variable focus lens (VFL), deformable mirror (DM), galvanometer-scanning mirrors (GM), quarter wave plate (QWP), photomultiplier tube (PMT), dispersion compensation block (DCB), mirror (M). Achromatic doublet lenses: L1=50mm, L2=150mm, L3=300mm, L4=75mm, L5=2x125mm, L6=2x50mm. (b) Computer simulation of optical layout on custom optical mounts using OpticStudio and SolidWorks. ..... 72

Figure 6.2. (a) Spot diagrams of the OCT light at 820 nm (red), 840 nm (pink) and 860 nm (purple) across a 15-degree FOV, where the black circle represents the Airy disk with a 2.1 µm radius. Spot diagrams of the 488 nm (blue) SLO light scanned across a 15-degree FOV with 0 D of vergence at the sample pupil plane and 7-degress with 20 D of vergence at the sample pupil plane where the black circle represents the Airy disk with a 1.2 µm radius. (b) The boundary of the imaging beam at the final pupil plane of the system. The black circle represents a 2 mm aperture. Each color represents a different scan position across a 15-degree and 7-degree FOV to simulate the pupil wander due to the space between the scanning mirrors in the optical design. ................................................... 74

Figure 6.3. (a) OCT B-scan across 50 degrees in the mouse retina and en face projection of the outer plexiform layer (OPL) across 44 degrees. The B-scan is an average of 200 consecutively acquired cross-sectional frames and the en face OCT image is an average of 5 frames. (b,c) Average of 5 adjacent OCT B-scans and an average of 5 en face OCT frames of the OPL. The B-scans are located at the position of the red dashed lines. Vertical scale bar: 50 µm. Horizontal scale bars: 100 µm. ...................... 78

Figure 6.4. Confocal SLO images of a mouse retina with 488 nm light. (a) Structural image of the nerve fiber layer from back-scattering. (b) Fluorescein angiography composited with a MIP from images of three different vascular layers. Scale bar: 100 µm. ....................................................... 79

Figure 6.5. (a) En face OCT-A images of the OPL in a mouse retina. (b) En face OCT intensity image from the same image data. (c) En face OCT-A images that were generated from the OPL (red), IPL (green), and NFL (blue). Scale bar: 50 µm. ................................................................................... 79

Figure 6.6. (a) En face images of the outer plexiform layer (OPL, top row, ~250 µm FOV) and nerve fiber layer (NFL, bottom row, ~280 µm FOV) retinal layers before and after Sensorless Adaptive Optics (SAO). SAO-OCT B-scans with the imaging focal plane on the OPL (red arrows) and NFL (blue arrows). (b) The normalized image quality for each step in the SAO optimization over two iterations and the Zernike coefficients selected for each iteration. Vertical scale bars: 50 µm. Horizontal scale bars: 20 µm.80

Figure 6.7. (a) Confocal SLO images before and after Sensorless Adaptive Optics (SAO) of the nerve fiber layer (NFL) with a FOV ~250 µm. Images of the outer plexiform layer (OPL) after SAO. (b) The normalized image quality metric values for each step used for the SAO optimization for each iteration. The Zernike coefficients selected for each iteration. Scale bar: 20 µm. .................................................................................................... 81

Figure 6.8. Confocal SLO images of a mouse retina with labelled retinal ganglion cells (Tg(Thy1-EGFP)MJrs/J). (a) Fluorescence images before and after Sensorless Adaptive Optics (SAO) and an intensity line plot between the

Page 15: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

xv

blue arrows (before SAO) and red arrows (after SAO). (b) The left column presents structural images focused on the nerve fiber layer at a ~750 µm FOV (top) and ~230 µm FOV (bottom). The right column presents the structural image in magenta overlaid by the fluorescence image in green. The fluorescence image was composited from two different focal planes for the axon and the dendrites of the RGC. Scale bars: 50 µm. .............. 82

Figure 6.9. Confocal SLO fluorescein angiography of a mouse retinal vasculature after Sensorless Adaptive Optics. Images (left to right) of the nerve fiber layer (NFL), inner plexiform layer (IPL), outer plexiform layer (OPL), and the MIP with the NFL in red, IPL in green, and NFL in blue. Scale bar: 50 µm. ......................................................................................................... 83

Figure 6.10. Confocal SLO images with Sensorless Adaptive Optics of EGFP labelled microglia in the mouse retina (B6.129P-Cx3cr1{tm1Litt}/J) acquired at different focal position between the outer plexiform layer (OPL) and the nerve fiber layer (NFL) selected from Visualization 1 of reference [16]. The microglia images were color-coded in depth between the OPL and the NFL of the retina and rendered in 3D for Visualization 2 reference [16]. Scale bar: 20 µm. ........................................................................... 84

Figure 6.11. (a) Confocal SLO fluorescence images with Sensorless Adaptive Optics of EGFP labelled microglia in the mouse retina (B6.129P-Cx3cr1{tm1Litt}/J) from three time points in the time-lapse video from Visualization 3 reference [16]. (b) The microglia images color-coded with time. The white arrows 1-4 note areas of significant growth and retraction. Scale bar: 20 µm. ................................................................................... 85

Figure 6.12. (a) Confocal SLO fluorescence images with Sensorless Adaptive Optics of EGFP labelled microglia in the mouse retina (B6.129P-Cx3cr1{tm1Litt}/J) from three time points in the time-lapse video from Visualization 4 of reference [16] with an increase in laser power at 39 minutes. (b) The microglia images color-coded with time. The white arrows 1-2 note areas of significant growth and retraction. Scale bar: 20 µm. ......................................................................................................... 86

Figure 7.1. Schematic of the Sensorless Adaptive Optics (SAO) Optical Coherence Tomography (OCT) and Two-Photon Excitation Fluorescence (TPEF) imaging system. The imaging system was constructed with a pellicle beam splitter (PeBS), a variable focus lens (VFL), a deformable mirror (DM), a dichroic mirror (DcM), galvanometer-scanning mirrors (GM), emission filters (EF), a photo-multiplier tube (PMT), dispersion compensation (DC), and the following lenses: L1=100 mm, L2=300 mm, L3=400 mm, L4=100 mm, L5=2×125 mm, L6=2×50 mm. The reference arm denoted as a dashed line. ............................................................... 91

Figure 7.2. Optical Coherence Tomography (OCT) and Two-Photon Excited Fluorescence (TPEF) images of the mouse retina before (top row) and after (bottom row) OCT-guided Sensorless Adaptive Optics (SAO). The improvement in the OCT B-scan is shown in the left column, the improvement in the en face OCT is shown in the middle column, and the improvement in the TPEF is shown in the right column. The yellow arrows represent the imaging focal position and the line between the blue arrows represents the cross-sectional location of the OCT B-scans. Scale bars: 50 µm. .................................................................................................... 96

Page 16: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

xvi

Figure 7.3. (a) OCT B-scans (top row), OCTA en face (middle row), and TPEF (bottom row) with the focal plane at the Outer Plexiform Layer (OPL), Inner Plexiform Layer (IPL), and Nerve Fiber Layer (NFL). In the right column, the images of the vascular layers were composited with a MIP. The red arrows point out connecting vessels in the TPEF. (b) Cross-sectional TPEF images (left) of the inner retinal vasculature before and after Adaptive Optics (SAO) acquired with a 25-step z-stack that was interpolated to 75 image pixels. The axial intensity profile plot between the red and blue arrows of the TPEF cross-sectional images. Scale bars: 50 µm. .................................................................................................... 98

Figure 7.4. TPEF imaging of GFP labelled microglia (B6.129P2(Cg)-Cx3cr1{tm1Litt}/J) in the mouse retina. (a) Single TPEF frame (left) and an average of 100 frames (right) at a ~0.8 mm FOV. The red square represents a 100 µm FOV to represent the scale of the microglia. Scale bar: 100 µm. (b) TPEF images of a GFP labelled microglia cells before (left) and after (right) Sensorless Adaptive Optics (SAO). (c) TPEF image after SAO. Scale bars: 20 µm. ................................................................ 99

Figure 7.5. Comparison of a GFP labelled retinal ganglion cell that was imaged using SAO TPEF (left) and using SAO SPEF with the same 200 µm FOV (middle). A SPEF image is also shown at a ~1.3 mm FOV (right), where the red square represents the 200 µm FOV that was used for the other images. Left scale bar: 20 µm. Right scale bar: 100 µm. ...................... 100

Figure 7.6. OCT B-scans (top row) and TPEF (middle row) imaging with the focal plane at the Nerve Fiber Layer (NFL), Inner Plexiform Layer (IPL), and Outer Plexiform Layer (OPL) of a Thy-1 YFP-16 Line (B6.Cg-Tg(Thy1-YFP)16Jrs/J) transgenic mouse. The blue arrow and yellow arrow point at fluorescently labelled cell bodies. The red arrow points at fluorescently labelled axons. In the bottom row, the OCTA en face image (magenta) was composited with the TPEF image (green). Vertical scale bar: 50 µm. Horizontal scale bars: 20 µm. ............................................................... 101

Figure 7.7. (a) The SAO-OCT B-scans in linear scale (top row) and the en face OCT (bottom row) with the focal plane at the Nerve Fiber Layer (NFL), Outer Plexiform Layer (OPL), and Retinal Pigment Epithelium (RPE) in the mouse retina. The en face OCT images were extracted between the cyan arrows (NFL), yellow arrows (OPL), and green arrows (RPE). The OCT B-scans were located between the red arrows on the en face OCT image. (b) TPEF images of the RPE of the mouse retina before and after SAO. (c) An intensity line plot between the blue arrows and the red arrows on the TPEF images of the RPE mosaic. Scale bars 50 µm. ..................... 102

Figure 7.8. (a) TPEF images of the RPE (left), en face OCT (middle), and OCT B-scans (right). (b) TPEF images of the RPE (left), en face OCT (middle), and OCT B-scans (right) from the same mouse four days later. (c) The digital enlargement of the TPEF images on day 1 (green) and day 4 (magenta), which were combined with a MIP. Scale bars 50 µm. ......... 103

Figure 7.9. TPEF from the RPE layer of the mouse retina in three different mouse strains, including a pigmented B6 mouse (C57BL/6J), an albino B6 mouse (B6(Cg)-Tyr{c-2J}/J), and a pigmented rpe65 mouse (B6(A)-Rpe65{rd12}/J). Scale bar 100 µm. ...................................................... 104

Page 17: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

xvii

Figure 7.10. TPEF image of a pigmented rpe65 mouse (B6(A)-Rpe65{rd12}/J) with different central wavelengths, including 760 nm, 780 nm, 800 nm, and 820 nm. The red arrow highlights an RPE cell where the fluorescence near the cell membrane is reduced with longer wavelengths. Scale bar 50 µm. ....................................................................................................... 105

Figure 8.1. Volumetric averaging of 150 OCT volumes. Scale bar: 50 µm. ............ 109

Page 18: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

xviii

List of Acronyms

ADD Airy Disk Diameters

AMD Age-Related Macular Degeneration

ANSI American National Standards Institute

AO Adaptive Optics

APD Avalanch Photodiode

Ar/Kr Argon/Krypton

BS Beam Splitter

CMOS Complementary Metal-Oxide-Semiconductor

CNIB Canadian National Institute for the Blind

CS Coordinate Search

DC Dichroic Mirror

DM Deformable Mirror

DMD Digital-Micro-Device

DONE Data-based Online Nonlinear Extremum-seeker

EGFP Enhanced Green Fluorescent Protein

f/c Fluorescence/Confocal

FA Fluorescein Angiography

FAD Flavin Adenine Dinucleotide

FLIM Fluorescence Lifetime Imaging Microscopy

FOV Field of View

FWHM Full Width at Half Maximum

GCL Ganglion Cell Layer

GFP Green Fluorescent Protein

GM Galvanometer Mirror(s)

GPU Graphics Processing Unit

IPL Inner Plexiform Layer

IS Inner Segment

LP Linear polarizer

MEMS Micro-Electro-Mechanical Systems

MIP Maximum Intensity Projection

MIRT Medical Image Registration Toolbox

MPE Maximum Permissible Exposure

Page 19: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

xix

MSPS Mega-Samples Per Second

NA Numerical Aperture

NADH Nicotinamide Adenine Dinucleotide

NFL Nerve Fiber Layer

NI National Instruments

NIR Near Infrared

OBJ Objective Lens

OCT Optical Coherence Tomography

OCT-A Optical Coherence Tomography Angiography

ONH Optic Nerve Head

OPL Outer Plexiform Layer

OS Outer Segment

OSA Optical Society of America

PBS Polarization Beam Splitter

PH Pinhole

PMT Photomulitpler Tube

PS-AO Pupil Segmentation Adaptive Optics

PSF Point Spread Function

QWP Quarter Wave Plate

RGC Retinal Ganglion Cell

RMS Root-Mean-Squared

ROI Region of Interest

RPE Retina Pigment Epithelium

SAO Sensorless Adaptive Optics

SD Spectral Domain

SH Shack-Hartmann

SH-WFS Shack-Hartmann Wavefront Sensor

SLD Superluminescent Diode

SLM Spatial Light Modulator

SLO Scanning Laser/Light Ophthalmoscopy

SNR Signal-to-Noise Ratio

SPEF Single Photon Excited Fluorescence

SS Swept Source

SVD Singular-Value Decomposition

Page 20: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

xx

TPEF Two-Photon Excited Fluorescence

TWCoG Thresholded Weighted Center of Gravity

USAF United States Air Force

UV Ultra Violet

VFL Variable Focus Lens

VIS Visible

VL Variable Lens

WFS Wavefront Sensor

WFS-AO Wavefront Sensor Adaptive Optics

WFS-less Wavefront Sensorless

WSAO Wavefront Sensorless Adaptive Optics

YFP Yellow Fluorescent Protein

Page 21: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

1

Chapter 1. Introduction

1.1. Overview

Vision is an invaluable sense and losing visual function is life changing for many

people. Currently, it is estimated by the Canadian Nation Institute for the Blind (CNIB)

that ~0.5 million Canadians are blind or partially sighted and ~5.6 million Canadians have

an eye disease that could lead to irreversible vision loss [1]. Many of us are or will be

directly affected by the current state of diagnostics and therapies available. Retinal

diseases that disrupt the ability of the eye detect light, including age-related macular

degeneration and diabetic retinopathy, are among the leading causes of vision loss.

Along with most technology, the ability to non-invasively image the retina has

experienced rapid development in the past couple decades. Several imaging modalities

are commercially available and they are used routinely by clinicians for disease diagnosis.

Fundus photography and confocal scanning laser ophthalmoscopy (SLO) provide en face

imaging of the large features of the retina, such as the blood vessels, the optic nerve head,

and the fovea. SLO systems can also include fluorescence detection for alternative

sources of contrast including, autofluorescence imaging from fluorophores that are

intrinsic to the retina, or fluorescein angiography (FA) through the intravenous injection of

fluorescent dye. Optical Coherence Tomography (OCT) provides volumetric and cross-

sectional imaging to visualize the different layers of the retina and OCT-Angiography

provides a non-invasive method for visualizing blood flow. The inclusion of adaptive optics

(AO) with these modalities has provided unprecedented spatial resolution of cellular

structures of the human retina [2]. These technologies provided ophthalmologist and

vision scientists with the ability to track degeneration and evaluate treatments. However,

future imaging tools will allow for non-invasive functional imaging of cellular processes in

the human retina, which can be used to develop novel therapies and to discover earlier

indicators of vision loss [3].

The study of animal models of human diseases is used extensively throughout the

therapy development process. In particular, the mouse has become a very important

Page 22: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

2

animal model for many of reasons, including the availability, easy handling, institutional

management, and short reproductive cycles. Furthermore, one of the primary benefits to

using mice is the unparalleled availability of genetically engineered strains for a wide range

of purposes including disease modelling and fluorescence labelling. Conventional ex vivo

immunohistochemistry is often used to study the eye and it provides exquisite cellular

contrast and high cellular resolution of the retina, but only at a single point in time. This

results in studies with large cohorts of animals multiplied by the number of time points that

are needed.

Non-invasive imaging is highly desirable for longitudinal studies for many reasons

including, reducing the effects of inter-animal variation, reducing the number of animals

required for a study, and thereby reducing the development time and the cost of new

therapies [4]. Furthermore, in vivo imaging allows for the study of physiological processes

that are not possible to study ex vivo. With technological advancements, high-resolution

systems have been able to study anatomy and physiology in vivo at the cellular level [2,3].

There would be profound benefits and advancements if more researchers had access to

high-resolution in vivo imaging systems with the functional and structural detection

capabilities that previously were only attainable through histology.

The topic of this thesis is focused on technological development, in order to expand

the ability of in vivo imaging techniques for the mouse retina. The small size of the mouse

eye has consequences for optical imaging. However, it is useful to study because the

mouse retina has similarities to the human retina.

1.2. The mouse eye

The eye is the organ responsible for gathering light and image formation. Much

like a camera, most vertebrate eyes have focusing elements in the front and detecting

elements in the back, as shown in Figure 1.1. The amount of light that enters the eye is

determined by the size of the pupil, which is modulated by the contraction and dilation of

the iris. The light entering the eye is refracted by the cornea and lens to be focused onto

the retina. The human eye has an axial length of 22-25 mm from cornea to retina and a

maximum pupil size of ~9 mm. The mouse eye is ~8 times smaller with an axial length of

~3.3 mm and a maximum pupil size of ~2 mm, which also creates a larger Numerical

Aperture (NA) for light entering the eye. However, the structure and function of the mouse

Page 23: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

3

retina has many similarities to the human retina. The thickness of the mouse retina is ~220

µm on average, which is comparable to the human retina at ~250 µm on average [5].

Figure 1.1. Simplified schematic of the mouse eye compared to a human eye [6].

The retina is the tissue at the back of the eye that is responsible for detecting the

light and forwarding the stimuli to the brain for visual processing into images for perception.

The mouse retina is organized in the same way as the human retina with various

differences, for example the absence of a macula for high-acuity vision and the greater

ratio of rod photoreceptors to cones photoreceptors for low light vision [6].

The retina consists of several layers with different types of cells, as shown in Figure

1.2. Starting with the outermost layer in the retina and moving inwards, first there is a

single layer of cells containing melanin called the retina pigment epithelium (RPE), which

is responsible for supporting the light sensing cells. The next layer inwards contain the

photoreceptors, which include rods for high sensitivity to visible light, and cones for colour

vision. The photoreceptors extend inwards with the outer segment (OS), the inner

segment (IS), cell body, and then the synaptic terminals in the outer plexiform layer (OPL).

Bipolar cells extend axially and they can transfer signal directly to the ganglion cell

dendrites in the inner plexiform layer (IPL). Within the inner retina there are also lateral

connections to horizontal cells and amacrine cells, which regulate signals and combine

signals from multiple photoreceptors before transferring signals onwards. The ganglion

cell bodies and the axons form the innermost layers of the retina, which includes ganglion

cell layer (GCL), and then the nerve fiber layer (NFL). The optic nerve, as shown in Figure

Page 24: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

4

1.1, exits the retina at the optic nerve head (ONH), which is where the ganglion cell axons

leave of the eye carrying the neural signals to the brain. Vasculature also enters the eye

from the ONH to supply blood to the inner retina and distinct vascular layers are found in

the NFL, IPL, and OPL.

Other types of cells in the retina support the retina, including Müller cells, and

microglia. Müller cells structurally and functionally support the other cells in the retina.

They stretch axial from the outer retina to the inner retina. Microglia can be found in many

of the retinal layers and they perform surveillance tasks, help to maintain homeostasis,

and perform phagocytosis of degenerating retinal neurons. They are highly sensitive to

changes to their micro-environment and can be stimulated into activation by disturbances

such as optic nerve damage, light injury, and disease. Activated microglia change their

morphology from having extended branches to a rounder amoeboid shape, and attempt

to repair the damage [7].

Figure 1.2. Organization of the retinal layers. The image is from Webvision [8] and used under the Creative Commons Licenses.

Light must travel through many retinal layers before detection by the

photoreceptors. The photoreceptors enable the conversion of light into electrochemical

signals through a process called phototransduction. In phototransduction, a photon is

Page 25: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

5

absorbed by a visual pigment molecule (photopigment) in the outer segment of the

photoreceptor, which causes a cascade of chemical reactions that results in an

electrochemical potential that can be transmitted through the neural retina to the brain.

After phototransduction, the photopigment must be regenerated before it can absorb

another photon. The regeneration process is carried out by the visual cycle within the outer

segment of the photoreceptor and RPE. In the photoreceptor, the absorption of a photon

causes 11-cis-retinal to be converted to all-trans-retinal and dissociate from the

photopigment. A series of reactions convert all-trans-retinal back to 11-cis-retinal, which

occurs primary in the RPE [9]. The location of each step is summarized in Figure 1.3.

Figure 1.3. The visual cycle and the location of each step. Image is from Wikimedia.org and used here under the Creative Commons License.

Page 26: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

6

In the outer retina, the visual cycle and normal cellular metabolism generate

compounds that are also fluorescent. This provides an opportunity to detect these

compounds that are critical to visual function, which enables measurements that could be

used to asses visual function [10–12].

1.3. Imaging the mouse retina

The efficient transmission of light through the eye enables non-invasive optical

imaging of the retina unlike any other internal tissue, but the refractive properties of the

eye must be considered by the imaging system. The anatomical differences in the mouse

eye to the human eye are significant enough that specialized equipment should be

developed for imaging the mouse retina with optimal performance.

Similar to human eye imaging systems, the mouse retina can be imaged with

modalities such as fundus photography, scanning laser ophthalmoscopy (SLO), and

optical coherence tomography (OCT). Fundus photography detects scattered light from

the retina on a two-dimensional pixel array. Alternatively, techniques such as SLO and

OCT, require light to be scanned across the retina and interrogate each point individually

for each image pixel. The diffraction-limited lateral resolution in the sample is related to

the smallest spot size of the focused light on the sample, which is dependent on the

numerical aperture (NA) into the eye. The NA is the half angle of the cone of light being

focused to a point, given by the Equation 1.1:

𝑁𝐴 = 𝑛 sin 𝜃, (1.1)

where 𝑛 is the index of refraction and 𝜃 is the angle from the optical axis. An advantage

of the point scanning system is that a confocal pinhole can be placed in the detection path

to remove light scattered from outside the focal plane, providing an axial sectioning ability

within biological specimen. For OCT, the lateral resolution is determined in a similar way

as SLO, but the axial resolution is not dependent on the focal spot size. Instead, the axial

resolution for OCT is determined by the spectral bandwidth. Therefore, OCT typically

provides superior axial resolution for cross-sectional imaging. These modalities are further

explained in the next chapter.

Non-invasive optical imaging of the mouse eye is enticing due to the large

Numerical Aperture (NA) available that would permit a small focal spot on the retina.

Page 27: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

7

Theoretically, the geometry of the mouse eye allows for sub-micrometer resolution

imaging. There are two options for focusing light onto the mouse retina, as shown in Figure

1.4. Focusing light onto the retina can be achieved with an objective lens, like in the

traditional confocal microscope. In this configuration, the refraction caused by the cornea

must be canceled out with plano-concave lens or ‘fundus lens’ that matches the curvature

of the cornea. Alternatively, the focusing ability of the eye can be used by the imaging

system if collimated light is directed into the mouse eye. The benefits of each method are

discussed in Chapter 3. For either method, the NA is related to the diameter of the beam

across the eye upon entry. However, for imaging with a large NA, optical aberrations are

introduced by the refractive elements, including the biological tissues in the optical path to

the retina. Optical aberrations from the tear film, cornea, and lens of the mouse enlarge

the size of the spot on the retina, reducing the actual imaging resolution.

Figure 1.4. The options for focusing light onto the mouse retina: collimated light or focused light from an objective lens.

Aberrations from the mouse eye can be corrected with adaptive optics (AO), in

order to obtain diffraction-limited imaging. The goal of the AO within an imaging system is

to restore the smallest possible spot on the sample, which has been enlarged by

aberrations. Since every mouse will have different errors in the focusing system, this will

result a variety of aberrations for a given area on the sample, known as the isoplanatic

patch. Therefore, AO systems need a method to correct the wavefront, as well as a

method of determining the aberrations to be corrected. To enable cellular resolved

imaging in the mouse retina, there are high-orders of aberrations that must be corrected

for optimal performance.

Page 28: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

8

Conventional AO employs a method to directly measure the aberrations from the

sample, and a corrective element to restore optical performance. Alternatively, the images

from the system can be used to indirectly determine the optimal aberration correction

required. These image-based methods are usually referred to as Sensorless Adaptive

Optics (SAO).

Performing accurate wavefront measurements for AO imaging in a small animal

retina typically requires a high level of system complexity due to the short length of the

eye creating an optically thick sample with multiple scattering surfaces that can confound

the wavefront measurement [5]. Although despite the difficulty in performing good

measurements in the mouse eye, retinal imaging system that use SH-WFS AO have been

reported in the Literature with high quality state-of-the-art performance, such as in

references [4,13,14]. Sensorless AO (SAO) imaging methods developed in this thesis can

avoid the complexities of WFS measurements at the cost of execution time. SAO allows

for AO retinal imaging with systems that are compact, easily operated, and robust.

1.4. Outline

The remaining chapters of this thesis are organized as follows. Chapter 2 presents

background information on the imaging modalities and methods for adaptive optics used

in this thesis. Chapter 3 details a preliminary optical design for imaging the mouse retina,

where we used focused light into the eye with a fundus lens to cancel the refraction from

the cornea. This chapter demonstrates the ability of image-based adaptive optics for

fluorescence imaging of fluorescently labelled cells. Chapter 4 demonstrates a novel

method for image-based adaptive optics that was developed for imaging the retina, which

was based on a technique called ‘pupil segmentation’. In this chapter, we also transition

the optical design to use collimated light into the retina. In Chapter 5, wavefront

measurements from the mouse eye show that the image-based approaches to AO are

indeed providing aberration correction in the mouse eye. Chapter 6 demonstrates

significant improvements to the image quality with a novel optical design in a compact

form factor suitable for translational research. Chapter 7 shows that the optical techniques

developed in the previous chapters can be applied to a Two-Photon Excited Fluorescence

(TPEF) system for high resolution imaging of both endogenous and exogenous

fluorophores in the retina. Finally, Chapter 8 discusses future research directions to

advance imaging technology for the mouse retina.

Page 29: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

9

1.5. Contributions

The over-arching goal of my work has been to develop advanced imaging

technology that has the potential to be translated to scientists that are not specialized in

Adaptive Optics (AO), which has required that AO imaging systems be easy-to-use,

robust, and compact. My first project demonstrated the performance of a lens-based

optical design with SAO to provide aberration correction for fluorescence imaging, which

is described Chapter 3 and published by Biomedical Optics Express in reference [15]. This

work was extending into a compact design, which is described in Chapter 6 and published

by Biomedical Optics Express in reference [16]. The compact form factor imaging system

also included SAO-OCT for improved multi-modal functionality. The results in this work

demonstrated that SAO could provide imaging resolution that are comparable to traditional

wavefront sensing methods in the mouse eye. Also, my work with vision science

collaborators has resulted in measurements that helped explain underlying mechanisms

of a retinal disease in a mouse model, described in reference [17].

While working on these optical systems, I also improved the image-based SAO

techniques. The ability of SAO to correct for aberrations in the mouse eye was investigated

using a wavefront sensor and the final image quality to evaluate the performance. This

investigation is described in Chapter 5, which was published in Biomedical Optics Express

[18]. I also contributed to further investigations on AO for the mouse eye using a contact

lens, which is described in reference [19]. I demonstrated the proof-of-principle of a novel

image-based AO technique for retinal imaging using pupil segmentation, which is

described in Chapter 4 and published by Optics Letters in reference [20]. My contributions

to the SAO methods were used in human imaging systems as well, which are described

in references [21,22].

Using the methods developed in previous projects, I contributed to the success of

the TPEF and the visible-light OCT imaging systems with my optical designs and imaging

experiments, which are described in references [23,24]. Finally, I also incorporated these

methods into an improved TPEF imaging system capable of imaging the RPE mosaic of

the mouse retina. The TPEF imaging system is described in Chapter 7 and published in

Biomedical Optics Express [25].

Page 30: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

10

Chapter 2. Background on retinal imaging systems and adaptive optics

2.1. Scanning Laser Ophthalmoscopy

Scanning Laser Ophthalmoscopy (SLO) is an imaging tool for an en face view of

a living eye, which is based on confocal scanning laser microscopy. However, the eye is

used as the objective lens to focus the light onto the retina. For SLO, the illumination light

is focused onto a single point on the sample and the back-scattered light returns to a

detector to be measured. The focused point is scanned across the retina to perform a

measurement for each pixel in the image. Light can be scanned across the sample with

different types of scanners, including galvanometer mirrors, resonant scanners, and

MEMS-based scanners, and usually video frame rates are achievable.

An advantage of sampling each point in time is that a pinhole can be optically

conjugated to the sample plane, which will remove light that is scattering from out of the

focal plane and improve the axial sectioning ability of the imaging system. A pinhole in this

configuration can be called a confocal aperture. Also sampling each pixel in time allows

for SLO systems to use highly sensitive detectors such as photo-multiplier tubes and

avalanche photodiodes that have a single detector. Similar to conventional microscopy,

optical filters can be used to isolate the back-scattered light from fluorescence emission

from the sample, further described in Section 2.3.

In order to define the diffraction-limited resolution of SLO, a point illumination

imaging system can be characterized by the point spread function (PSF). Consider an

ideal system where the objective lens is illuminated with a circular aperture of uniform

intensity. The light will focus into the Airy pattern, which will be the PSF or impulse

response (h) of the system. At a given NA and wavelength (λ), the radius to the first

minimum of the Airy pattern in the lateral direction at the focal plane is given by Equation

2.1:

𝑟𝐴𝑖𝑟𝑦 =

0.61 𝜆

𝑁𝐴 . (2.1)

Page 31: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

11

In the axial direction, the distance from the center of the diffraction pattern will have a first

minimum (𝑧𝑚𝑖𝑛) at a distance given by Equation 2.2:

𝑧𝑚𝑖𝑛 =

2 𝑛 𝜆

𝑁𝐴2 , (2.2)

where 𝑛 is the index of refraction of the medium [26,27] . The FWHM of the axial PSF can

be calculated by a multiplying the half width (𝑧𝑚𝑖𝑛 ) by a factor of ~.84 [26]. Using a

confocal aperture that is close to the size of the Airy radius can provide an additional

improvement to the lateral imaging resolution. However, in this thesis, a confocal aperture

several times the size of the Airy disk was used in order to balance signal with depth

sectioning. Therefore, the non-confocal PSF calculations were used to approximate the

theoretical diffraction-limited resolutions.

The final intensity distribution, 𝐼, that is measured by the detector of the SLO

system will be a convolution of the PSF of the system, ℎ, with the pinhole (circle function)

[26,28], described by Equation 2.3:

𝐼 = |ℎ|2 (|ℎ|2 ⊗ 𝐶𝑖𝑟𝑐) (2.3)

In a real imaging situation, the sample or system will introduce aberrations that

must be corrected to restore diffraction-limited resolution, which is further discussed in

Section 2.4.

2.2. Optical Coherence Tomography

Optical Coherence Tomography (OCT) provides high-resolution cross-sectional

and volumetric imaging of the retina. OCT was initially developed as a tool for

ophthalmology to better visualize the layers of the retina but also has been adopted by

other areas of biomedical imaging. Other advancements in OCT systems have enabled

polarization sensitivity [29], as well as functionally imaging of blood flow often called OCT-

Angiography (OCT-A). OCT-A is generated by analysing changes to the cross-sectional

images that are caused by the blood moving through the vessels [30–33].

OCT volumes are generated by acquiring adjacent depth intensity profiles (A-

scans) to generate a cross-section (B-Scan) into the sample. Then, B-scans can be

acquired adjacently in the remaining dimension for volumetric imaging. The OCT volumes

Page 32: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

12

typically have comparable resolution in both the axial and lateral directions, therefore en

face views can be generated at any depth within the sample.

Fourier domain OCT can be separated into two types: Spectral Domain (SD) OCT

and Swept Source (SS) OCT. Both types are used in modern systems and both types

share the same principle based on the Michelson interferometer using a low-coherence

light source. The difference between each type is that SD-OCT relies on a spectrometer

to measure the interference pattern on an array of detectors, and SS-OCT uses a single

detector and a laser that sweeps through a spectral bandwidth. The typical bandwidth of

an OCT system for imaging the retina is 50 nm to 100 nm, commonly with a near infrared

center wavelength.

For an OCT system, the imaging light is separated into a reference path and the

sample path, and then recombined at a fiber coupler. The interference pattern modulations

correspond to the path mismatch between the reference arm and the sample arm where,

higher frequency fringes correspond to a larger mismatch (∆𝑧). Therefore, by calculating

the Fourier transform of the fringes, the axial location of light that is scattered from the

sample can be determined. For example, consider a single reflector at position ∆𝑧. The

measured intensity on the detector (𝐼𝐷) will be a function of wavenumber (𝑘) from the

interference pattern, as described by Equation 2.5:

𝐼𝐷(𝑘) = 𝑆(𝑘){[𝐼𝑅 + 𝐼𝑆] + 2√𝐼𝑅𝐼 cos(2∆𝑧𝑘)}, (2.4)

where 𝐼𝑅 is the reference light, 𝐼𝑆 is the sample light, and 𝑆(𝑘) is the intensity of the source

spectrum. After the Fourier transform, the location of the reflector, ∆𝑧, is revealed as delta

functions convolved with the Fourier transform of the source spectrum, �̂�(𝑘), in Equation

2.6.

𝐼(𝑧) ∝ �̂�(𝑘) ⊗ {[𝐼𝑅 + 𝐼𝑆]𝛿(𝑧) + 2√𝐼𝑅𝐼𝑠(𝛿(𝑧 − ∆𝑧) + 𝛿(𝑧 + ∆𝑧))}. (2.5)

The axial resolution of an OCT system is determined by the coherence length (𝑙𝑐), given

by Equation 2.7:

𝑙𝑐 =

2 ln 2

𝜋∙

𝜆𝑜2

∆𝜆, (2.6)

Page 33: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

13

which is dependent on the center wavelength of the light source, 𝜆𝑜, and the spectral

bandwidth, ∆𝜆. OCT typically provides a 1 µm - 10 µm axial resolution, with a depth

penetration on the order of millimeters. The lateral resolution is dependent on the focal

spot size similar to traditional microscopy, as described in the SLO section.

OCT systems have been increasing in speed with advancement in the detectors

and lasers available. SD OCT systems are often limited by the speed the camera can

acquire A-scans. Similarly, SS OCT systems are often limited by the speed that the laser

can sweep the spectral bandwidth. For real-time processing and display, OCT images

often require GPU accelerated processing programs to keep up with the A-scan

acquisition rate.

2.3. Fluorescence imaging

Fluorescence imaging is an essential tool that has been used in microscopy to

provide contrast that may not be available by other methods. Fluorophores are often

added to the biological specimen to identify cells, blood flow or specific molecules. Green

Fluorescent Protein (GFP) and its derivatives have been used extensively to label selected

cell classes in a variety of organisms, including mice. Labeling otherwise transparent cells

with fluorescent compounds allows them to be imaged with fluorescence detection. Also,

many biological samples have intrinsic fluorescent properties or autofluorescence that can

be imaged to provide insights into the health of the tissue [34–36].

The process, known as fluorescence, starts when a fluorophore is excited by the

absorption of a photon. Then, some of the energy gained by the initial photon is lost

through non-radiative processes, and finally a photon with a longer wavelength is released

from the fluorophore. This chromatic difference, known as the Stokes shift, can be used

to isolate the fluorescence emission from the fluorescence excitation. Various

fluorophores have different excitation and emission spectra.

Alternatively, fluorescence emission can occur from a process called Two-Photon

Excited Fluorescence (TPEF). The principle of TPEF is that two photons can provide

sufficient energy as a single photon to excite the same fluorophore, as shown in Figure

2.1. This requires that the pair of lower energy photons arrive at the fluorophore at

Page 34: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

14

practically the same time. Therefore, TPEF imaging uses ultrashort pulses of light to

improve the probability that two-photon absorption occurs.

Figure 2.1. Simplified Jablonski diagram of single photon excited fluorescence and two-photon excited fluorescence. The excitation light, λex, and emission light, λem.

Two-Photon Excited Fluorescence (TPEF) can be used as an imaging technique

that enables fluorescence imaging deeper into tissue than the equivalent single photon

excitation [37,38]. Often TPEF uses near infrared light to excited the same fluorophores

that require UV or visible light for excitation, which has advantages for imaging the retina

that are further discussed in Chapter 7. For TPEF imaging, signal can be improved by

optimizing the probability of two-photon absorption. The number of photons absorbed (𝑛)

is related to the intensity of the fluorescence emission, which is described by Equation 2.8

[38]:

𝑛 ≈𝑃2𝜎

𝜏𝑝𝑓𝑝2 (

𝑁𝐴2

ℎ𝑐 𝜆)

2

. (2.7)

Factors that affect the absorption include, the average incident power (P), the cross-

section of the fluorophore (𝜎), the laser pulse width (𝜏𝑝), the laser repetition rate (𝑓𝑝), and

the NA of the incident light (𝜆).

The axial and lateral resolution of fluorescence imaging in point scanning systems

are related to the size of the spot on the sample. For this thesis, we will use the

conventional spot size given by the PSF in order to determine the theoretical resolution of

the systems, as described in Section 2.1.

Page 35: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

15

2.4. Adaptive optics for ophthalmic imaging

AO technology is used to improve the performance of many types of optical

systems. Originally, AO was developed for astronomy to dynamically correct for

aberrations caused by the transmission of light through atmospheric turbulence [39]. Later,

AO techniques started to be used to correct for aberrations caused by imperfections

through biological specimens, such as imaging of the brain. Naturally, AO was also

developed for retinal imaging to correct focusing errors from many different eyes, including

humans, primates, and rodents. AO has been used for imaging the eye with many

modalities, including fundus photography, SLO, and OCT [2]. An AO system consists of a

device to correct the optical performance, and a method for determining the correction

required for the sample.

There are a wide variety of wavefront correctors commercially available, such as

deformable mirrors. The deformable mirror can change shape to modify the incident

wavefront. In this thesis, a segmented DM and a continuous membrane DM are both used.

The segmented DM offers quick MEMS based actuators, and low flatness of the reflection

surface, <20 nm. The continuous membrane DM provides large stroke with the magnetic

actuators, ~80 µm (peak to valley) and low settle time, ~0.5 ms. For the AO systems in

this thesis, it is ideal for the deformable mirrors to operate quickly and accurately with little

hysteresis and drift in the actuators position over time.

AO systems for imaging the retina were developed using a Shack-Hartmann (SH)

wavefront sensor (WFS) to directly measure the aberrations [40–42]. A SH-WFS is

constructed by an array of identical micro lenses, often called lenslets, that are mounted

a focal length away from a 2D detector. For measuring a wavefront, the lenslets should

be positioned at a pupil plane and the wavefront will form an array of spots on the detector.

Each lenslet of the WFS will gather a sample of the wavefront. If a measured wavefront

has no aberrations, the spots on the detector will form an even grid, corresponding to

centers of the lenslets in the array. However, a local gradient in the wavefront will cause

a displacement in the spot on the sensor, where magnitude and direction can be used to

reconstruct the aberrations in the pupil plane.

Page 36: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

16

A pupil plane of a system is related to the PSF by the Fourier transform, which

means that modulations due to aberrations in the pupil plane will enlarge the size of the

focal spot on the sample [43]. However, if there are no aberrations in the pupil plane, then

the system is only limited by diffraction. Wavefront aberrations are commonly described

by a decomposition into the Zernike polynomials [39]. The aberrations in the pupil

plane, 𝑊(𝜌, 𝜃), can be described by Equation 2.9:

𝑊(𝜌, 𝜃) = ∑ 𝑎𝑗𝑍𝑗(𝜌, 𝜃)

𝑗=0

, (2.8)

where 𝑍𝑗 is a Zernike polynomial with a given coefficient, 𝑎𝑗. Table 1 lists the Zernike

polynomials with the corresponding index (j) that is used for this thesis.

Table 2.1. Zernike polynomials, names and index up to the 5th radial order.

Index (j) Radial order (n) Aberration term Zernike Polynomial

𝒁𝒋(𝝆, 𝜽)

0 0 Piston 1 1 1 Tilt 2𝜌 sin 𝜃 2 1 Tip 2𝜌 cos 𝜃 3 2 Oblique astigmatism √6𝜌2 sin 2𝜃 4 2 Defocus √3(2𝜌2 − 1) 5 2 Vertical astigmatism √6𝜌2 cos 2𝜃 6 3 Vertical trefoil √8𝜌3 sin 3𝜃 7 3 Vertical coma √8(3𝜌3 − 2𝜌) sin 𝜃 8 3 Horizontal coma √8(3𝜌3 − 2𝜌) cos 𝜃 9 3 Oblique trefoil √8𝜌3 cos 3𝜃 10 4 Oblique quadrafoil √10𝜌4 sin 4𝜃 11 4 Oblique secondary

astigmatism √10(4𝜌4 − 3𝜌2) sin 2𝜃

12 4 Primary spherical √5(6𝜌4 − 6𝜌2 + 1) 13 4 Vertical secondary

astigmatism √10(4𝜌4 − 3𝜌2) cos 2𝜃

14 4 Vertical quadrafoil √10𝜌4 cos 4𝜃 15 5 Higher orders √12𝜌5 sin 5𝜃 16 5 √12(5𝜌5 − 4𝜌3) sin 3𝜃 17 5 √12(10𝜌5 − 12𝜌3 + 3𝜌) sin 𝜃 18 5 √12(10𝜌5 − 12𝜌3 + 3𝜌) cos 𝜃 19 5 √12(5𝜌5 − 4𝜌3) cos 3𝜃 20 5 √12𝜌5 cos 5𝜃

A basic SH-WFS AO system will illuminate a spot on the retina and measure the

optical wavefront that returns to the SH-WFS. The wavefront corrector will attempt to

Page 37: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

17

remove the aberrations with the complex conjugate of the measured wavefront. This

process will repeat in what is referred to as closed-feedback loop, until the measured

wavefront aberrations are reduced to an acceptable flatness. The WFS AO will require

optical elements to re-direct the scattered light from the sample to the WFS in a different

optical path from the illumination and detection components.

In confocal microscopy, AO can be used for imaging into volumetric biological

samples. The ideal configuration is to measure the wavefront from the focal plane that is

being imaged by the system. Confocal imaging systems rely on a pinhole to reject light

from out of the focal planes in order to image a specific depth in the sample. However,

most wavefront sensors do not have the ability to reject out of focus light. Therefore, back-

scattered light from multiple planes in the sample can corrupt the measurement from the

imaging plane. However, an alternative to using a direct measurement from the sample is

to use the images from the system to indirectly infer the aberrations, thereby using the

confocal pinhole to provide depth discrimination.

As described in Section 1.3, the short focal length of the mouse eye and a large

NA illumination create multiple scattering surfaces from the volumetric layering of retina.

In this thesis, SAO methods were developed to determine optimal aberration correction

for the chosen focal plane in the retina.

A common image-based AO method operates by applying aberrations to the

wavefront corrector and recording the effect on the image, which is then used to determine

the best correction. This is called open-loop control. The image quality due the aberration

on the corrective element can be quantified with a sharpness or brightness metric. Then,

the image metric values can be used as the merit function for an optimization problem that

uses a chosen number of degrees of freedom (Zernike modes) to find the optimal image.

A few versions of the hill-climbing coordinate search optimization algorithm are presented

in each chapter. Section 5.2.3 and Section 5.2.2 provide be best descriptions for each

case.

Another type of image-based AO investigated in this thesis uses a computational

algorithm based on pupil segmentation [44–46]. A pupil plane can be divided into sub-

regions by an active element, such as a segmented deformable mirror or spatial light

modulator. If only one sub-region illuminates the sample, a smaller diameter beam or

Page 38: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

18

‘beamlet’ will illuminate a point in the sample. An image can be acquired using the beamlet

in the center of the pupil, which will be defined as the reference. A sample with aberrations

will cause the beamlets from the other pupil regions will be deflected from the reference

focal point and the images formed will be shifted from the reference image. The amount

of translation between the images from each sub-region of the pupil and the reference can

be related to the aberrations in the entire pupil.

The translated distance, ∆𝑥, ∆𝑦, of the images can be used as wavefront slopes,

which can be reconstructed into Zernike coefficients that approximate the wavefront in a

similar way to how a wavefront sensor would perform the calculation. The local slope of

the wavefront in the x and y direction is related the Zernike polynomials, Z, by Equation

2.10 and 2.11.

∆𝑥

𝑓=

𝜕𝑊

𝜕𝑥= ∑ 𝑎𝑗

𝑗

𝜕𝑍

𝜕𝑥 , (2.9)

∆𝑦

𝑓=

𝜕𝑊

𝜕𝑦= ∑ 𝑎𝑗

𝑗

𝜕𝑍

𝜕𝑦 , (2.10)

where the partial derivatives of the wavefront, W, can be calculated on the Zernike

polynomials, Z, to the jth term. Using this relationship, a conversions matrix, Z, can be

constructed to calculate the slope of the wavefront at n positions of the pupil, 𝐬 =

[∆𝑥1, ∆𝑦1 … , ∆𝑥𝑛, ∆𝑦𝑛 ]𝑇, for a given vector of Zernike coefficients, a = [𝑎0 , … , 𝑎𝑗]𝑇 ,

described by Equation 2.12:

𝐬 = 𝐙𝐚 , (2.11)

Then, any wavefront can be approximated by a vector of Zernike coefficients, 𝐚𝒍𝒔, for the

measured slope values, s, by a least-squared fit in Equation 2.13:

𝐚𝒍𝒔 = 𝐙ϯ𝐬 . (2.12)

The active element that divides the imaging pupil can be a separate device than

the wavefront corrector. As further described in Chapter 4, we use the segmented

deformable mirror for both aberration correction and pupil segmenting.

Page 39: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

19

2.5. Summary

This chapter presented a background to the imaging modalities SLO, OCT, and

TPEF that are used in this thesis. Also, the background information for AO used in

ophthalmoscopy was presented, including an introduction to SAO with optimization

algorithms and with pupil segmentation. The remaining chapters utilize these topics and

provide additional depth.

As a note, in the remainder of this thesis, various acronyms were used for the

abbreviation of “Sensorless Adaptive Optics” and “Wavefront Sensorless Adaptive

Optics”. All of the acronyms are listed in the directory and SAO, WSAO, and WFS-less

can be used interchangeably.

Page 40: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

20

Chapter 3. Wavefront sensorless adaptive optics fluorescence biomicroscope for in vivo retinal imaging in mice

3.1. Introduction

Small animal models of diseases are a vital component in vision research because

they facilitate the understanding of underlying biological processes, the identification of

potential causative genes for human disorders, and the development of therapies against

vision-robbing diseases. Mice are commonly used for preclinical vision research due to

the significant anatomical and functional similarity of their eyes to human eyes and to the

availability of transgenic strains that model human diseases. Non-invasive in vivo retinal

imaging has the potential to reduce the number of animals required for a study, which in

turn reduces the development time and the cost of new therapies [4]. Transgenic mice

expressing endogenous fluorescent markers, such as Enhanced Green Fluorescent

Protein (EGFP), are particularly important for vision research. The ability to image

molecular markers has the potential to accelerate vision research by allowing retinal

function to be observed in vivo and by permitting longitudinal studies of the same animal

[13]. Research animals expressing EGFP in neuronal cells, including retinal ganglion cells

and axons, are useful for studying retinal neurodegenerative diseases such as glaucoma

[47]. Similarly, mice with EGFP-labelled microglia enable the in vivo study of the retinal

response to diabetic retinopathy, glaucoma, and age-related macular degeneration [48–

50]. Additional examples of vision research based on fluorescence imaging are described

in Zhang et al. [49], and Alt et al. [50], just to name a few. Non-invasive fluorescence

imaging of the mouse retina with even higher resolution is desirable, but requires

correction of optical aberrations in the mouse eye [5].

Adaptive optics (AO) for ophthalmoscopy is an important tool for ophthalmologists

and vision scientists, permitting cellular-resolution imaging of the retina. For non-invasive

in vivo imaging in humans, AO has been demonstrated to improve the resolution for fundus

photography, Scanning Laser Ophthalmoscopy (SLO), and Optical Coherence

Tomography (OCT); there are several reviews on this topic, including [2,39,51–53]. The

geometry of the mouse eye allows for focusing light on the retina with a higher numerical

Page 41: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

21

aperture (NA) than in humans, which allows for even higher resolution imaging in vivo.

Due to the higher NA, the mouse eye is even more sensitive to aberrations induced when

imaging with a large diameter beam, with aberrations introduced from the ocular tissues

(i.e. cornea, lens and vitreous humour) [5]. AO facilitates retinal imaging with diffraction-

limited performance at cellular resolution in mice.

Conventional AO compensates for aberrations in the wavefront with an adaptive

element such as a deformable mirror (DM) controlled by a Shack-Hartmann (SH)

wavefront sensor (WFS) in a closed feedback loop. The common approach of wavefront

sensing is to use an extra light source (beacon) with low NA to measure the wavefront

aberrations with the SH-WFS [4,14,53]. However, the beacon adds to the limited light

power that is allowed into the eye and contributes to non-common path errors [54]. The

ability of the WFS-based AO system to correct wavefront aberrations can be limited by the

WFS design (its accuracy and dynamic range) and a geometry mismatch between the

WFS and the adaptable element, leading to wavefront correction errors [55]. The SH-WFS

performance is also susceptible to specular reflection from lenses and optical elements

within the system [56]. Many of these issues have been addressed in different ways with

AO system modifications. For example, efficient spherical mirror-based telescopes can be

implemented to maintain signal and minimize back-reflections into the SH-WFS [57]. More

compact AO systems have been developed with lens-based optics and polarizing

elements to reduce these back-reflections [56]. For applications in vision science and

small animal retinal imaging, a SH-WFS is further hampered by the ‘small eye artifact’, in

which multiple reflecting/backscattering surfaces in the retina affect the wavefront

measurement [5].

Some of the challenges associated with the SH-WFS can be better managed

through Wavefront Sensorless Adaptive Optics (WSAO), which uses information from the

quality of the image to guide aberration correction. Wavefront sensorless techniques have

been applied in microscopy [58] and ophthalmic imaging for both human and mouse eyes

[59–61].

In this chapter, we expand on previous work combining WSAO with Fourier-

domain OCT for mouse retinal imaging, and present fluorescence images of mouse retina.

AO fluorescence retinal imaging in mice has been presented with different modalities and

methods throughout the Literature [4,13,14,62–65], including both with and without a

Page 42: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

22

WFS. Notably, WFS-AO for in vivo subcellular-resolution imaging has been presented

using an annular beam for the beacon, a high-resolution SH-WFS, and a mirror-based

optical system [4,13,14]. We have implemented a lens-based WSAO system with a modal

hill-climbing optimization algorithm using fluorescent image intensity as a metric. We used

the combination of a variable focus lens and a small-stroke MEMS

(microelectromechanical system) deformable mirror to perform defocus and aberration

correction in the mouse eyes. The details of the experimental methods are described in

the next section. The system performance was evaluated on phantoms, and

representative images acquired from mouse retina in vivo are presented. Our results

demonstrate that AO can be simple and compact with cellular resolution.

3.2. Methods

We have implemented a compact and low-cost confocal biomicroscope in order to

acquire both reflectance (structural) and fluorescence (functional) images from mouse

retina simultaneously. The system used off-the-shelf lenses with a smaller footprint and

simpler design compared to the AO configurations that are based on SH-WFS and

constructed from curved mirrors. A benefit of the lens-based system is that it permits a

relatively wide field of view (FOV) to be imaged on the mouse retina [56,66], albeit without

diffraction-limited imaging performance outside of the isoplanatic patch. The wide field

structural image was used in real-time to navigate on the retina using features such as

blood vessels and the optic nerve head. Once centered at the desired location using a

FOV of about 0.8 mm, the FOV was reduced to 0.2 mm or smaller for acquisition of the

high-resolution fluorescence images that are presented below.

During imaging, the mouse was aligned to the imaging system using a plano-

concave ‘fundus lens,’ with a 2 mm contact diameter and no magnification (Volk Optical

Inc, Mentor, OH) that canceled out most of the refractive power from the mouse cornea;

we utilized an external objective to focus the light on the mouse retina. For this reason,

we refer to the imaging system as a fluorescence confocal (f/c) biomicroscope rather than

an ophthalmoscope. The fundus lens approach to retinal imaging has several benefits: it

facilitates alignment of the mouse, provides mechanical stability, and retains the moisture

of the mouse cornea during imaging.

Page 43: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

23

3.2.1. Mouse handling

The two strains of EGFP-labelled mice, B6.Cg-Tg(Thy1-EGFP)MJrs/J (ganglion

cells) and B6.129P-Cx3cr1tm1Litt/J (microglia cells), that were imaged in this report were

obtained from Jackson Laboratories (Bar Harbor, ME). The mouse imaging sessions were

performed under protocols compliant to the Canadian Council on Animal Care, and with

the approval of the University Animal Care Committee at Simon Fraser University. Prior

to the imaging experiment, the mice were anesthetized with a subcutaneous injection of

ketamine (100 mg/kg of body weight) and dexmedetomidine (0.1 mg/kg of body weight).

Next, the eyes were dilated with a drop of topical solution (Tropicamide, 1%) and a couple

minutes later, a drop of topical anesthetic (Alcaine, 0.5%) was applied. Artificial tear gel

(Alcon, Fort Worth, TX) was applied liberally to protect the cornea from dehydration. The

anesthetized mouse was placed on a translation stage and the eye was gently aligned

with direct contact to the fundus lens. The laser power at the fundus lens was ~150 µW.

The mice were recovered after the experiment using atipamezole injected at 1.8 mg/kg of

body weight. The local anesthetic was applied to reduce potential irritation to the mouse

after recovery due to the contact with the fundus lens.

3.2.2. Biomicroscope optical setup

The optical system schematic is presented in Figure 3.1. We used a fiber-coupled

Ar/Kr ion laser and a diffraction grating to select the 488 nm spectral line as the excitation

source. At the fiber output, the beam was collimated to be approximately 3.5 mm in

diameter. The first element was the segmented MEMS deformable mirror (DM) with modal

control and a 5 µm stroke (PTT111, Iris AO, Inc, Berkeley, CA), which also defined the first

pupil plane. Two lenses, f1 = 200 mm and f2 = 200 mm, relayed the conjugate plane to the

tunable liquid lens (Varioptics, Arctic 39N0) for focus control and for optical sectioning of

the different retinal layers. Relay lenses with focal lengths of f3 = 150 mm, and f4 = 100

mm were used to decrease the beam’s diameter and to relay the pupil to the scanning

mirrors. The beam position was scanned over the sample with two 6210H galvanometer

mounted mirrors (Cambridge Technology Inc.) for each direction. Next, two lenses, f5 = 25

mm, and f6 = 75 mm expanded the beam’s diameter and relayed the pupil to the final

objective lens. The beam was focused with a NA = 0.17 to a final spot (Airy disk radius) of

1.8 µm and the FWHM of the axial point spread function was 38 µm with a Mitutoyo infinity-

corrected long working distance 10x objective lens. The backscattered light (reflectance)

Page 44: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

24

was coupled into a 100 µm multimode fiber (3.9 Airy disks in diameter) and detected by

an Avalanche Photo Diode (APD, Hamamatsu S5343). We also used a quarter-wave plate

(QWP) to act on the linearly polarized light from the laser and a linear analyzer at the

detector to reduce the specular reflection from the optical elements [56]. The backward-

directed fluorescence signal was isolated by a dichroic mirror and a long-pass filter

(Semrock 496 nm blocking edge BrightLine, FF01-496/LP-25). The fluorescence light was

focused by an 80 mm lens through a 50 µm confocal pinhole with diameter 1.8 times the

size of the Airy disk, and a Photo-Multiplier Tube (PMT, Hamamatsu H7827-002) was used

to detect the weak fluorescence signal. An analog-to-digital converter digitized the signal

from the APD and PMT simultaneously, and the imaging system speed was limited to 1.00

Mega-Samples Per Second (MSPS) per channel by the NI PCIe-6361. The galvanometers

were driven by a 1 kHz sinusoidal waveform scanning pattern, and acquired data from

both forward and backward sweeps of the scan. Image distortion caused by the non-linear

scanning pattern was corrected with de-warping in real-time using custom software

developed in C/C++ for acquisition and display. This allowed for 400x400 reflectance and

fluorescence samples per frame at 5 fps, which was used for aligning the mouse eye and

for data streaming. The AO optimization was performed at 400x100 samples which

corresponded to an acquisition and display rate of 20 fps. The icon of a computer in Fig.

3.1. has representative images for each channel displayed during imaging.

Page 45: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

25

Figure 3.1. Schematic of the WSAO f/c biomicroscope using 488 nm excitation from an Ar/Kr laser. Relay lenses are achromatic doublets. Other optical elements: 80/20 beam splitter (BS), dichroic mirror (DC), deformable mirror (DM), zero-order quarter wave plate (QWP), objective lens (OBJ), linear polarizer (LP), pinhole (PH), variable lens (VL), galvanometer scanning mirrors (GM). Electronic elements: avalanche photo diode (APD), photo multiplier tube (PMT). The images on the computer icon are representative images of the structural and fluorescence imaging channels.

3.2.3. Image acquisition and optimization

The anesthetized mouse was placed in front of the fundus lens to initiate imaging.

The retinal imaging location was determined based on landmarks such as the vascular

pattern and the optic nerve head. The maximum FOV on the mouse retina was ~0.8 mm

with the mouse cornea approximately perpendicular to optical axis of the system. The

position of the fundus lens was fixed in the center of the optical path, and different

eccentricities on the mouse retina were imaged by rotating the orientation of the mouse

with respect to the fundus lens. The focus was adjusted using the Varioptic lens in order

to get the best qualitative image appearance [67].

The WSAO optimization algorithm that we used is a modified version from our

previous report for WSAO mouse imaging with OCT [61]. During optimization, the frame

rate of the WSAO f/c biomicroscope presented in this report was 20 fps with a frame size

Page 46: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

26

of 400x100 samples. Following optimization, higher quality images were acquired with

400x400 samples per frame at 5 fps. The optimization algorithm used a modal control to

build up an optimal shape of the DM that corrected for the wavefront aberrations. For each

Zernike mode, the algorithm searched through 21 coefficient values; the search range for

a particular Zernike mode was selected from typical aberration amplitudes for mouse eyes.

A larger range was searched for lower order aberrations and a smaller range for higher

order aberrations. We were able to correct for up to the 20th Zernike mode (OSA

convention [68]) in ~30 seconds.

The WSAO used a hill-climbing search algorithm to find the best set of Zernike

coefficients that corrected for aberrations based on the image quality metric. The overall

intensity of an image was calculated as the sum of each pixel, defining the merit function

J(k) as in Equation 3.1:

𝐽(𝐤) = ∑ Iw(𝐤)(𝑥, 𝑦),

𝑥,𝑦

(3.1)

where k is a vector of Zernike coefficients and Iw(k) is the acquired intensity if the pixel at

the image coordinates x, y. The wavefront shape, w(k), applied to the DM is given by

Equation. 3.2:

𝑤(𝐤) = ∑ 𝑘𝑛𝑍𝑛

20

𝑛=3

. (3.2)

The flowchart in Fig. 3.2 is a summary of the optimization process. The algorithm

applied a linearly spaced range of coefficients (kn) for each Zernike mode (Zn) to the DM

and recorded an image for each pre-set coefficient value. The optimal image, and thus the

optimal coefficient value, was determined as the one that corresponded to the highest

value of the merit function. The optimization began with defocus (n = 4) [68] and first three

Zernike modes (piston, tip and tilt) were assigned a coefficient value of zero and not

included as part of the optimization. The Zernike modes were optimized in the order as

presented on the abscissa of Fig. 3.6(b)

Page 47: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

27

Figure 3.2. WSAO modal hill-climbing algorithm flowchart for the fluorescence image optimization process; deformable mirror (DM), variable lens (VL).

3.3. Results

3.3.1. WSAO f/c biomicroscope resolution

The biomicroscope system design was computer simulated with Zemax (ZEMAX

Development Corporation, Bellevue, WA) to model the spot size on the mouse retina off

of the optical axis. The simulation results predicted diffraction-limited performance within

a 0.2 mm FOV. The FOV on the retina was limited to ~0.8 mm by lens L6 in Fig. 3.1, note

that this imaging mode was not diffraction-limited, but useful for navigation to landmarks

on the retina.

The performance of the optical imaging system was evaluated for both reflectance

and fluorescence channels on resolution targets. The performance of the reflectance

imaging system was measured by setting the DM to the flat position and placing a US Air

Force (USAF) resolution target (Fig. 3.3) in the retinal plane. We imaged groups 6 and 7

of the USAF target and we were able to resolve the smallest element of group 7 that has

a line width of 2.19 µm.

Page 48: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

28

Figure 3.3. US Air Force resolution target with line width 2.19 µm highlighted by the red rectangle to demonstrate the reflectance resolution. Scale bar: 50 µm.

The resolution of the fluorescence channel was measured by imaging fluorescent

beads (190508, Polyscience, Inc.) that had a diameter of 2.1 µm with a standard deviation

of 0.018 µm. The WSAO optimization was performed on the fluorescent signal from the

beads and the intensity profile across the diameter of a representative bead was plotted

to demonstrate the system resolution and AO performance (Fig. 3.4). A line plot of a bead

after optimization shows a 5% intensity increase and a narrower full width at half

maximum, suggesting that the optimization corrected small system aberrations. The RMS

of the Zernike polynomial coefficients obtained from this optimization was 0.02 µm, which

is below the Maréchal criterion for diffraction-limited imaging (λ/14).

Figure 3.4. Images of 2.1 µm diameter fluorescent beads acquired (a) before WSAO optimization and (b) after optimization. (c) The line plots for a bead before and after optimization. Scale bars: 10 µm.

3.3.2. In vivo WSAO confocal fluorescence imaging of retinal ganglion cells

We acquired retinal images of anesthetized mice to demonstrate the WSAO f/c

biomicroscope performance in vivo. Images of an EGFP-labelled ganglion cell are

displayed before and after the optimization in Fig. 3.5. A video of the change in the

appearance of the ganglion cell during the optimization process as displayed on the

Page 49: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

29

screen is included in Visualization 1 of [15]. The images in Fig. 3.5 were produced using

identical processing steps, and by averaging 50 registered frames. The Medical Image

Registration Toolbox (MIRT) for Matlab was used to register the frames prior to averaging

[69]. Given the small amplitude and slow speed of intra-frame motion in the anesthetized

mice during the 5 frames per second acquisition, we used a non-rigid cubic B-spline

registration algorithm with a sum of squared differences similarity metric. In the presence

of larger amplitude or faster motions, more advanced registration algorithms, for example

strip-based registration [70–72], may achieve better performance. The intensity-based

similarity measurement performed well for the frames with a strong signal from Fig. 3.5(b),

although the registration was less effective for the dendrites in Fig 3.5(a) due to the lower

signal intensity. The dendrites in Fig. 3.6(b) appear to be less resolved on the right side of

the image; this is due to the dendrites being out of the focal plane.

Figure 3.5. (a,b) Ganglion cells labelled by EGFP comparing the images acquired before and after the WSAO optimization. These images are an average of 50 frames of an off-axis ganglion cell. Scale bars: 20 µm.

The optimization results for these images are plotted in two ways: the optimum

Zernike coefficient value determined by the search algorithm (Fig. 3.6(a)) and the

corresponding merit function for each optimized mode in the order of the optimization (Fig.

3.6(b)). The intensity profile of a dendrite is displayed in Fig. 3.6(c) to demonstrate the AO

performance. Ganglion cells are classified into types based on the cell structure with

properties including soma size and dendrite patterns; a short review of this topic was

presented in Geng et al. [4]. The image of the ganglion cell in Fig. 3.5 appears to have a

round soma that is approximately 20 µm in diameter with thick and straight dendrites.

Based on this description, this ganglion cell may belong to the RGA2 category as described

Page 50: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

30

by Sun et al. [73]. This classification is further supported with the description by Coombs

et al. [74] and images by Geng et al. [4].

Figure 3.6. (a) The Zernike coefficients applied to the DM (deformable mirror) after the optimization. (b) The impact of the optimization on the intensity-based merit function are plotted for each mode. The intensity is normalized from zero when the DM is flat. The Zernike coefficients are reported by the OSA standard for optical aberrations of eyes [68]. (c) The intensity plot of a dendrite on the EGFP-labelled ganglion cell at the location and in the direction indicated by the arrows.

3.3.3. In vivo WSAO confocal fluorescence imaging of retinal microglia cells

We also imaged EGFP-labelled microglia cells with similar results. The images of

microglia presented in Fig. 3.7 before and after the optimization were processed by the

same method as the ganglion cell images. In Fig. 3.7(b) the fluorescence signal on the left

side of the image has lower intensity after the optimization due to the microglia being at a

different depths within the retina. This demonstrates the optical axial sectioning effects of

the confocal pinhole and the ability of WSAO to reject out-of-focus features.

Page 51: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

31

Figure 3.7. Images of EGFP-labelled retinal microglia cells acquired in vivo before and after WSAO correction with different field of views: (a), (b), and (c). Images (b) and (c) were taken at the same location with different field of views as indicated by the red dashed box. Each image is an average of 50 frames. Scale bars: 10 µm.

3.4. Discussion

The mouse eye is commonly used as a model of the human eye for vision

research. High quality images of the retina can be acquired without AO in some mice, in

particular for animals with healthy eyes [49,50]. Time course studies in mice and research

on transgenic mouse models of degenerative retinal diseases stand to benefit from the

incorporation of AO with the imaging system for increased resolution across a wider range

of animals, improved consistency in image resolution between time points, and for

locations that are off the optical axis of the eye. Additionally, AO provides the ability to

control the focus plane of the imaging system, which allows for easy and controlled

transition between layers of interest within the retina. Developing accessible AO that is

low-cost and small in size has the potential to be more widely used across multiple

research specialties.

We demonstrated WSAO for non-invasive in vivo fluorescence imaging of the

mouse retina. Our results showed cellular-resolution images acquired using a lens-based

Page 52: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

32

AO system without the difficulties associated with implementing a WFS for mouse retinal

imaging. Our system included two electronically controllable elements on optically

conjugated pupil planes: a tunable lens controlling defocus for layer selection (which can

loosely be referred to as a ‘woofer’) and a deformable mirror for higher order corrections

(‘tweeter’) [75–77]. We used an intensity-based image quality metric to search for the

Zernike coefficients that would produce the strongest fluorescent signal and hence better

resolution. Recent research results suggest that using an intensity-based metric

exclusively is not sufficient for confocal scanning ophthalmoscopes, due to differences in

the illumination and the collection paths. Sulai et al. have demonstrated an approach using

a sharpness metric to optimize the point spread function of the illumination path in order

to increase the resolution [54]. In future work, we will investigate the use of a sharpness

metric, for example as defined in [54], in place of the intensity metric used in this chapter.

We anticipate that in cases where the fluorescent signal arises from non-planar structures

that a sharpness metric will result in an improved optimization performance.

The WSAO f/cSLO method used the same source of light (the fluorescence)

confined to a single retinal layer for both imaging and for guiding the aberration correction

(since the merit function was derived from the fluorescence images). In common AO

imaging configurations using a SH-WFS, the sensing (‘beacon’) and imaging wavelengths

are different. According to Zhou et al. [78], the chromatic defocus between the sensing

light and the imaging light are significant in the small animal eye, and can introduce higher

order aberrations in addition to defocus. This further accents a strength of the WSAO

technique, in that the aberration correction is performed at the same retinal depth section

that is being imaged, for example with the ganglion cells in the inner retina.

Our optimization algorithm required ~30 seconds to perform an exhaustive search

up to the 20th Zernike mode with 21 steps per mode. In addition to the modal hill-climbing

algorithm demonstrated in this chapter, the performance of other approaches such as the

simulated annealing algorithm or the stochastic parallel gradient descent algorithm could

also be explored [79,80]. For human imaging, the WSAO optimization speed is essential

due to the motion of the patient [61]. However, for an anesthetized mouse the amount of

motion is low and the optimization speed is not as crucial; this is supported by a recent

report by Palczewska et al. [81], where they used WSAO for two-photon microscopy and

required 4-6 minutes for an optimization.

Page 53: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

33

The approach to WSAO retinal imaging in mice described in this chapter utilized a

fixed microscope objective lens and a plano-concave lens with radius of curvature

matched approximately to the curvature of the eye, reducing the refraction at the cornea.

This configuration constrained the FOV by the size of the mouse pupil and the mismatch

between the field curvature at the image plane of the objective lens and the curvature of

the mouse retina. Delivering a collimated beam into the mouse eye and allowing the

cornea and lens to focus it on the retina would provide a larger FOV. However, with a

collimated beam approach, a significantly larger amount of focus correction at the final

pupil plane would be required to optically section through the retina, on the order of ~ 30

– 40 diopters [5]. Another possible solution for our system would be to design a custom

multi-element objective lens with a scan pivot closer to the pupil in order to permit a wider

field of view. An advantage of this approach is in the potential ease of reconfiguring the

system for AO retinal imaging in different animal species, such as rats (see recent work

by Geng et al. [82]). This would require only a change in the final objective lens to

accommodate a different NA and the final concave surface to accommodate a different

corneal curvature.

A significant benefit of the WSAO algorithm is that the aberration correction is less

sensitive to multiple reflections from the sample and optical elements. Even in the

presence of undesired reflections that would affect SH-WFS wavefront measurement,

WSAO could still perform the optimization if these reflections do not change significantly

with the DM shape or if they are removed from the region of interest used for the merit

function with image processing techniques. As reported by Biss et al. [62] and Geng et al.

[5], challenges in wavefront sensing can arise due to reflections from different layers within

the relatively thick mouse retina as well as due to increased scattering. This effect is even

more pronounced in albino animals, in which the choroid and sclera layers also generate

a large backscattered signal. We employed a confocal detection which allowed the WSAO

to optimize the image signal within the depth of focus and minimize the out of focus signal.

The mice used for imaging in this chapter were pigmented; however, since the WSAO

algorithm does not rely on wavefront sensing, similar performance is anticipated in non-

pigmented animals. It has recently demonstrated WSAO OCT in albino mice with no

difference in performance with respect to pigmented mice [83].

Page 54: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

34

3.5. Summary

In this chapter, we demonstrated WSAO for non-invasive in vivo fluorescence

imaging of the mouse retina. We imaged transgenic mice with EGFP-labelled ganglion

and microglia cells and used WSAO to increase the image resolution by correcting for

wavefront aberrations introduced by the eye. The AO system demonstrated cellular-

resolution imaging with a low-cost, simple and robust lens-based system.

In the next chapter, we investigate an alternative method for image-based AO

using pupil segmentation and provide a comparison to the hill-climbing optimization. We

transition to using collimated light into the eye with a 0-diopter contact lens placed on the

eye to prevent dehydration instead of focused light with a fundus lens.

Page 55: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

35

Chapter 4. Pupil segmentation adaptive optics for in vivo mouse retinal fluorescence imaging

4.1. Introduction

Adaptive Optics (AO) is a technology used for human retinal imaging [2] that is

also of high interest for pre-clinical research. Vision research commonly uses small animal

models of vision robbing diseases, particularly mice, because they are inexpensive, and

are versatile to genetic manipulations. Non-invasive optical imaging of the mouse retina

permits retinal diseases to be characterized and the effects of potential therapies to be

studied in vivo and longitudinally. The mouse eye is well suited for high resolution non-

invasive optical imaging due to its large Numerical Aperture (NA). The large NA

exacerbates aberrations and necessitates AO in the case of a non-ideal eye to restore

diffraction-limited performance, and to allow for in vivo high resolution analysis [62].

Conventional AO employs a Shack-Hartmann (SH) Wavefront Sensor (WFS) to

measure the aberrations in combination with a deformable element to perform the

correction [2]. SH-WFS AO has been reported in the Literature for mouse retinal imaging

achieving high quality performance, for example [4,13,14]. While this method has

demonstrated excellent aberration correction ability for rodent imaging, SH-WFS based

approaches can be challenging as they are sensitive to wavefront reconstruction errors

produced by non-common path errors, multiple reflective retinal planes (the ‘small eye

artifact’) [5], and specular reflections [78].

Wavefront sensorless adaptive optics (WSAO) is an alternative method that uses

images acquired with the optical system to determine the optimal shape of a deformable

element to correct the wavefront aberrations. WSAO has demonstrated promising results

in microscopy as well as retinal imaging in humans and mice [15,62,63,80,84,85] A recent

review of WSAO algorithms can be found in [86]. A method that is common to many WSAO

reports in the Literature is iteratively changing the shape of the deformable mirror while

optimizing an image quality metric [86]. The quality of the aberration correction obtained

with merit function based WSAO is often sensitive to the number of images used to

Page 56: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

36

optimize the shape of the deformable element, which comes at the cost of an increased

optimization time. Reducing the time required for aberration correction becomes a critical

goal to translate WSAO for in vivo mouse retinal imaging applications in vision science. In

this chapter, we demonstrate an alternative method of WSAO that uses the acquired

images to indirectly measure the wavefront aberrations in the sample; this approach is

known as Pupil Segmentation Adaptive Optics (PS-AO). This approach to ocular

aberration measurement is also conceptually similar to that reported in [87].

PS-AO indirectly measures a wavefront using images acquired with different

segments of the imaging pupil to determine the gradient of the wavefront at each pupil

region. In the case where no aberrations are present, all of the rays across the pupil of the

imaging system will converge at the sample to a focal spot size limited by diffraction.

However, in the presence of aberrations, the rays at different segments across the pupil

are deflected to different lateral positions at the focal plane due to heterogeneity in the

index of refraction and imperfections in the shape of the ocular structures. PS-AO

measures the deflection of the beam at each pupil segment through comparison of the

image acquired at that position with a reference image, commonly selected as the central

region of the pupil, in order to determine the local wavefront slope at that pupil segment.

A set of images acquired at each segment across the pupil. The wavefront gradient at

each segment is determined by measuring the lateral shift in the corresponding image

with respect to the reference image. These indirect measurements of the wavefront slope

using PS-AO are conceptually similar to the operation of a SH-WFS. The aberrations are

corrected by shaping the deformable mirror into the phase conjugate of the measured

wavefront. The PS-AO method has been demonstrated with great success by Ji et al. for

in vivo mouse brain imaging [44–46,88], which encourages the extension of this AO

method to in vivo retinal imaging modalities.

In this chapter, we present PS-AO for fluorescence Scanning Laser

Ophthalmoscopy (SLO) in mice. PS-AO for in vivo retinal imaging requires considerations

for the motion from the living sample as this has potential to corrupt the image-based

wavefront measurement. Our implementation used a Micro-Electro-Mechanical (MEMS)

segmented deformable mirror to segment the imaging pupil and also to correct the

aberrations. Our results demonstrate the improvement in the resolution of images

acquired of mouse retinal capillaries for fluorescein angiography.

Page 57: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

37

4.2. Methods

We have implemented a compact, low cost, lens-based AO system for

fluorescence SLO to acquire mouse retinal images. The imaging system schematic in Fig.

4.1 represents the optical layout that is based upon the system described in Chapter 3

[15], with modifications for pupil segmentation. The imaging system had dual adaptive

elements, with a tunable liquid lens (Varioptic, Arctic 39N0) and a MEMS segmented

deformable mirror (PTT111, Iris AO, Inc.). The liquid lens was used to select the depth

focal position within the retina. The deformable mirror (DM) consisted of 37 individually

controllable mirror segments. The ability to tilt light from the segments, which were

mapped to the pupil, made the DM well suited for pupil segmentation. Single mirror-

segment beamlets were initially created to provide a densely sampled wavefront, however

the laser power at the sample was inadequate and the signal to noise ratio was poor.

Consequently, we created ‘beamlets’ by grouping 7 mirror segments (a central segment

surrounded by six neighbors); for the groups at the edge of the DM, 4-5 mirror segments

were used. To create a beamlet, a group of segments remained flattened on the DM,

directing light to the sample, while the remaining segments were tipped to deflect the beam

toward a spatial filter created with an iris diaphragm.

Page 58: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

38

Figure 4.1. Schematic of the Scanning Laser Ophthalmoscope: 488 nm laser; dichroic mirror (DC); deformable mirror (DM); variable lens (VL); galvanometers (GM); {L1,L2, L3,L4,L5,L6} = {200,200,150,100,50,19} mm. (a),(b) PS-AO on 6 µm fluorescent beads with aberration correction (AO On) and without (AO Off). These images are an average of 30 frames. Scale bar: 8 µm. (c) Wavefront aberration map. (d) Normalized intensity plotted at the location indicated by the dashed lines with a ~30% increase in the peak intensity after correction. (e) The Zernike coefficients for the corrected wavefront.

We acquired in vivo fluorescent mouse retinal images using a 488 nm laser

(Coherent Inc. 488 OBIS). The full beam diameter at the cornea was 0.89 mm, focused

by the mouse eye with an NA of ~0.23. In this chapter, the maximum NA was limited by

the stroke of the DM. The beam diameter of the central beamlets at the mouse pupil was

0.38 mm with an NA of ~0.1, while the diameter of the outer beamlets was approximately

0.27 mm. The size of the focal spot of an individual beamlet (~3 µm Airy disk radius at the

retina) constrained the lower limit of the dynamic range for the wavefront sensing, while

the upper limit was restricted by the largest image shift that could be detected. The

Page 59: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

39

fluorescence signal was focused through a 100 µm pinhole with a confocal aperture 6.5

times the Airy disk, detected by a photomultiplier tube (Hamamatsu H7827-002). A 1 kHz

bi-directional sinusoidal scan pattern was applied to the galvanometer mirror (6210H

Cambridge Technology Inc.), capturing data in both the forward and backward scan

directions. Phantom imaging was performed with a 3 mm objective lens in place of the

mouse eye, with an NA of 0.15 while acquiring 400 x 400 sampled images (ie. pixels). For

in vivo applications, the image size was 200 x 200 samples at 10 frames per second.

Image acquisition was performed with custom software [89].

In our initial experiments on PS-AO in vivo mouse retinal imaging, the image shift

introduced from sample motion impeded the image-based measurement of the wavefront

slope. In order to mitigate the effect of motion introduced from live animal retinal imaging,

we modified the PS-AO image acquisition process to collect a reference image in rapid

succession with each target image. For the deformable mirror used in this chapter, this

resulted in 36 reference images and 36 target images acquired at each of the pupil sub-

regions. The total of 72 images was acquired in ~7.2 seconds. To further minimize the

effect of motion from in vivo samples, the acquisition alternated between the reference

and target beamlets within a frame. We acquired sections of the frame corresponding to

50 lines at a time, alternating between reference and target regions on the segmented DM

every 25 ms.

The reference images were acquired with the central region of the pupil, selected

as the desired focal point in the image plane for the target beamlets. Aberrations in the

optical path caused a lateral shift between the reference and target images. A rigid

registration algorithm was used to find the shift between each pair of images. The

translation distance was used to calculate the local wavefront slope at each pupil segment.

Using this information, we then applied a modal wavefront reconstruction. The wavefront

was approximated by multiplying the slopes by the pseudoinverse of the Zernike gradients

to determine the Zernike coefficients. The set of Zernike coefficients up to the 15th mode

was applied to the DM for aberration correction. The final images with the aberration

correction were acquired with the full imaging pupil, and the deformable mirror shaped to

the phase conjugate of the overall wavefront. Modes 0, 1, 2 (piston, tip, tilt) were set to

zero [68].

Page 60: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

40

We demonstrated PS-AO imaging on fluorescent phantom samples before

performing in vivo experiments. The phantom samples were created using fluorescent

beads or lens tissue fibers labelled with fluorescein. Aberrations were induced by placing

a gel between two non-uniform plastic surfaces in front of the sample. In Fig. 4.1 the

aberration correction was performed on a 400 x 400 sample image with a field of view

(FOV) of ~80 µm. Fig. 4.2 shows a comparison between PS-AO and a hill-climbing

coordinate search algorithm [15] to investigate the amount of wavefront correction

achieved with each method.

Figure 4.2. Aberration correction performed with both hill-climbing and PS-AO. (a) Image without aberration correction. (b) Correction performed with hill-climbing. (c) Correction performed with PS-AO. (d) The Zernike coefficients for the corrected wavefronts.

An experiment was performed on the lens tissue phantom to investigate the

sensitivity of PS-AO to sample motion. Aberration correction was demonstrated three

times on a sample at nominally the same location. In the first scenario, the sample was

static, and the results from the aberration correction are shown in Fig. 4.3(a). Next, lateral

motion was manually introduced in the sample, approximating the motion observed due

to respiration in an anesthetized mouse. The results of PS-AO using only a single

reference frame (i.e. a total of 37 images) in Fig. 4.3(c) indicate that the aberration

correction was not successful. By using the intra-frame reference acquisition method

described above (i.e. a total of 72 images), the PS-AO results shown in Fig. 4.3(b),

demonstrate good aberration correction even in the presence of motion.

Page 61: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

41

Figure 4.3. PS-AO aberration correction on (a) static and (b, c) moving samples. The aberration correction was performed with (b) using the multiple intra-frame reference images and (c) a single reference image.

Next, PS-AO was performed to optimize the images of the capillaries for

fluorescein angiography in mouse retina in vivo. Wild type C57BL/6J mice (Jackson

Laboratories, Bar Harbor, ME) were used for imaging in this chapter. All imaging

experiments were conducted with the approval of the University Animal Care Committee

at Simon Fraser University while following the protocols compliant to the Canadian Council

on Animal Care. The mice were anesthetized with a subcutaneous injection of ketamine

(100 mg/kg of body weight) and dexmedetomidine (0.1 mg/kg of body weight) prior to the

imaging session. Following the injection, the eyes were dilated with a drop of topical

solution (Tropicamide, 1%). A contact lens (Cantor & Nissel Ltd, UK) was then applied to

protect the cornea from dehydration. The anesthetized mouse was placed on a translation

stage and aligned to the imaging beam without contact. When switching between imaging

with a fully illuminated pupil versus a pupil segment, the laser power was adjusted to not

exceed ~150 µW incident on the mouse cornea. After the experiment, the recovery of the

mice was induced with an injection of atipamezole (1.8 mg/kg of body weight). The PS-

AO process was performed on a small FOV (~200 µm) sampled with a 200 x 200 points.

The defocus term in the second panel of images in Fig. 4.4 was set to zero in order to

examine the same imaging plane before and after correction. Following aberration

correction, images were acquired at the same FOV, and then optically zoomed out to a

larger FOV (~400 µm). Each image in Fig. 4.4 is an average of 30 frames registered using

the Medical Image Registration Toolbox [69]. The images before and after correction were

selected at the best focus with the Varioptic lens prior to the acquisition. The results show

the improvement in the image brightness and sharpness after PS-AO.

Page 62: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

42

Figure 4.4. (a), (b) PS-AO for retinal fluorescein angiography with aberration correction (AO On) and without (AO Off) for two mice. In each panel, the top row of images (angular FOV 5.2°) is an optically zoomed in section of the bottom row of images (angular FOV 10.4°). Scale bars: 20 µm. (c) Zernike coefficients for the corrected wavefront. (d) On the top panel, the normalized intensity plot at the location indicated by the dashed lines had a ~30% increase in the peak intensity after correction, and (d) on the bottom panel, the wavefront aberration map.

4.3. Discussion

Wavefront sensorless AO alleviates some of the challenges with direct wavefront

sensing, facilitating the extension to applications of retinal imaging in small animals.

Strong specular reflections of a laser beacon from refractive elements, such as the cornea

or lens-based optics in the imaging system, can impede conventional SH-WFS

approaches to AO. The ‘small eye artifact’ is another source of potential error confounding

SH-WFS methods, particularly in albino animals. In this chapter, we have demonstrated a

Page 63: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

43

wavefront sensorless lens-based AO system that is robust to these issues by using the

fluorescence images for aberration correction.

A segmented deformable mirror was used in this chapter as it was naturally well

adapted for pupil segmentation. Each beamlet contained 4-7 mirror segments that were

used to sample the gradient of the wavefront, defining the reconstruction resolution of the

wavefront. Using one mirror segment per beamlet would improve sampling density as long

as the laser can deliver enough power to maintain good signal-to-noise images. Using

multiple segments to define a beamlet provided a smaller depth of focus (~50 µm) than a

single segment (~500 µm), reducing noise due to fluorescence from other layers in the

sample.

We demonstrated PS-AO as an approach to WSAO using 72 frames, which were

acquired in 7.2 s to measure the wavefront slope at 36 locations across the pupil. For

PS-AO, the limit to the number of correctable modes is dependent on the number of

samples taken across the wavefront. This upper limit can be determined by minimizing

the difference in the root-mean-square error between a measured and reconstructed

wavefront, and approximately corresponds to the sampling density across the wavefront

[39], or the number of beamlets created within the imaging pupil. Had the individual

mirror segments been used to form beamlets with our particular implementation of PS-

AO, up to 36 Zernike modes could be corrected while maintaining the same optimization

time. In this chapter, we performed aberration correction up to the 15th mode as our

previous work demonstrated that higher modes had less impact in correcting mouse

ocular aberrations [15]. Compared to our previous work using a modal hill-climbing

algorithm [15], this is ~5 times quicker for the optimization of the same number of

modes. However, the amount of modes that are required for aberration correction can

vary due to factors such as motion artifact, the quality of the mouse eye, or the

eccentricity of the region of interest. We showed that PS-AO has the potential to perform

well while requiring fewer images.

Page 64: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

44

4.4. Summary

In this chapter, we demonstrated aberration correction for fluorescein angiography

of retinal capillaries. PS-AO is susceptible to errors in the wavefront reconstruction due to

motion of the sample. We presented an implementation of PS-AO that minimizes the

effects of motion artifact, and demonstrated a viable technique for WSAO aberration

correction for in vivo retinal imaging in mice.

Page 65: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

45

Chapter 5. Adaptive optics in the mouse eye: Wavefront sensing based vs. image-guided aberration correction

5.1. Introduction

Retinal imaging with Adaptive Optics (AO) is necessary to allow reliable

visualization and monitoring of single retinal cell morphology in vivo by correcting for

ocular aberrations of the eye, which acts as the microscope objective. Animal models are

important for studying pathophysiology and treatment of many human diseases. This also

includes common eye diseases such as diabetic retinopathy, glaucoma, and age-related

macular degeneration, as well as rare genetic diseases such as retinitis pigmentosa.

Conventional ex vivo immunohistochemistry often used in these studies provides exquisite

cellular contrast and high cellular resolution of the retina, but only at a single point in time.

This results in studies with large cohorts of animals multiplied by the number of time points

that are needed. Non-invasive imaging of living animals enables the characterization of

the progression of pathology and the evaluation of new therapies for eventual use in

humans within a single mouse during longitudinal studies, greatly reducing numbers of

animals needed for the experiment [4]. Mouse models are widely used in preclinical

research partially due to the availability of transgenic strains, which includes mice with the

relevant cell classes labeled by fluorescent proteins. Imaging modalities such as Adaptive

Optics - Scanning Light Ophthalmoscopy (AO-SLO) employed with fluorescence detection

can be used to examine the structural and functional features [64,65] in the retina at

cellular resolution. A recent review article [3] further describes the significance of AO for

retinal imaging in vision science.

In comparison to primates, mice have eyes with a relatively high numerical

aperture that provides the potential for sub-micrometer diffraction limited resolution. There

are challenges associated with consistent high-resolution AO imaging for small animal

eyes that arise from sensitivity to alignment, and the length of the eye relative to the

thickness of the retina. The short focal length creates a large relative optical thickness of

the retina, requiring a large dioptric power to optically section through the retinal layers

[78], which is often also a source of aberrations.

Page 66: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

46

Accurate measurements with a Shack-Hartmann (SH) Wavefront Sensor (WFS)

require a well-defined reference light source or ‘guide star’, with sufficient signal above the

background noise photons from other scattering tissue for layer specific aberration

correction [90]. The so-called ‘small eye artifact’ [5] means that instead of a single

scattering reference for the SH-WFS as in the case in human retinal imaging, the various

strongly scattering retinal layers, such as the heavily pigmented Retinal Pigmented

Epithelium (RPE), Choroid, and Nerve Fiber Layer (NFL), have the potential to confound

wavefront sensor measurements. The difficulty of performing direct wavefront

measurements from the small animal eye further increases for albino strains where the

backscattered signal from the desired imaging target, such as RPE, is not present due to

lack of melanin and the WFS is overwhelmed by background scattering from the sclera.

The conventional method for aberration correction with AO requires closed-loop

feedback between residual aberrations measured by SH-WFS in response to changing

the shape of wavefront corrector. Multiple groups have demonstrated WFS AO to provide

high-resolution retinal images in the mouse eye, including [4,13,14] to list a few. However,

a direct measurement of the wavefront from the retinal volume of interest in small animal

eyes requires a complex system and may not always be possible due to the multi-layered

structure of the sample, in addition to mild cornea or intraocular lens opacities.

Wavefront Sensorless (WFS-less) AO is an alternative approach to the WFS

based AO method that uses an image-based optimization method to correct the

aberrations. Multiple approaches to WFS-less AO have been reported, including amongst

others: stochastic steepest gradient descent, simulated annealing, hill climbing modal

optimizations, and pupil segmentation [45,85,86]. WFS-less AO SLO has been

demonstrated for cellular resolution retinal imaging in mice [15,20,63], and WFS-less AO

has also been demonstrated to correct for non-common path errors in combination with

WFS AO [54]. WFS-less AO has the potential to alleviate the dependency on the WFS

alone and provide an alternative method for aberration correction at multiple depths in the

retina as well as simplifying the imaging system [16]. However, WFS-less AO requires an

optimization execution time several times longer, which can be disrupted by the motion

inherent in a living, breathing animal.

In this chapter, we demonstrate WFS-less AO for aberration correction that was

implemented in a state-of-the-art WFS AO SLO system design for mouse retinal imaging,

Page 67: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

47

and present comparisons of the image quality and wavefront measurements during AO

correction performed by each method. Confocal SLO back-scattering images of the mouse

photoreceptor mosaic are presented, representing a case where the WFS and WFS-less

AO are both using the same reference to guide the AO. As an example of a case where

the wavefront sensing and the imaging planes are different, images of mouse retinal

microglia labeled with fluorescent proteins are presented. We demonstrate that WFS-less

AO can correct the same aberrations that are measured by a SH-WFS and that WFS-less

AO can perform depth-resolved aberration correction resulting in reliable imaging focal

plane shift in a mouse eye.

5.2. Methods

Mouse retinal imaging was performed using a custom designed AO SLO system

that has previously been described [14,19]; please see Section 5.2.1 for details. The AO

SLO design was based on SH-WFS AO, using wavefront measurements from the mouse

eye to perform aberration correction. The differences from the previously reported

configuration include: the scanning mirrors were changed to galvanometer mounted

mirrors, and the optical detection path. For WFS AO, the deformable mirror (DM) control

was provided by the WFS measurement software, which allowed for the closed-loop

functionality. The SH-WFS sub-system is described in Section 5.2.2. Only software

changes were required to implement the image-based WFS-less AO approach; the WFS-

less AO software is described in Section 5.2.3. The WFS-less AO required the image

acquisition software to control the DM. The system recorded wavefront data from the

mouse eye as measured by the SH-WFS while performing WFS-less AO.

5.2.1. AO SLO system description

The AO SLO system was custom designed for reflectance and fluorescence

imaging of the mouse retina, further described in Zawadzki et al. [14,19]. Light from a

superluminescent diode (SLD, Superlum, SLD-26-HP) with a 663 nm center wavelength

was used for reflectance imaging, and as a beacon for wavefront sensing. The system

also used co-aligned light from a 488 nm laser (Coherent, OBIS 488 nm LX) for

fluorescence excitation. The laser power at the mouse eye was limited to 100 W for each

light source. The first pupil plane was defined by the continuous membrane DM (ALPAO,

Page 68: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

48

DM97-15). The optical plane was relayed to the horizontal galvanometer scanning mirror

and then to the vertical galvanometer scanning mirror (Cambridge, 6215H) using afocal

telescopes made from pairs of spherical mirrors. The pupil plane was further relayed with

a spherical mirror and a lens to the mouse eye with a final beam diameter of 2 mm to be

focused to the retina with the maximum available numerical aperture of ~0.5. The contact

lens was mounted at the last pupil plane for alignment of the mouse eye. The optical layout

of the system is shown in Fig. 5.1.

Figure 5.1. Adaptive Optics Scanning Laser Ophthalmoscopy (AO-SLO) system schematic. The layout is presented in a scale drawing. Abbreviations: L#, lens; F#, filter; BS#, beamsplitter; M, mirror; SM, spherical mirror; DM, deformable mirror; D#, dichroic mirror; Hsc, horizontal resonant scanner; Vsc, vertical scanner; PMT, photomultiplier tube; P (circled in blue) optical planes conjugate with the pupil; SLD, superluminescent diode. Collimated beams are marked as dashed lines and focusing beams are marked as solid lines. The on-axis beams are represented by red lines and scanned beams by green and blue. Image credit: Pengfei Zhang.

Table 5.1 lists the optical parameters of the important system components.

Table 5.1. Key optical parameters of the AO-SLO system components

Item BS1 BS2 BS3 D1 D2 F1 F2 DM

Type R:T= 50:50

R:T= 10:90

R:T= 30:70

T635lpxr-UF, Chroma

ZT488rdc-UF, Chroma

FF01-660/13, Semrock

FF01-525/45, Semrock

DM97-15, Alpao

Item SM1 SM2 SM3 SM4 SM5 SM6 SM7 L0

Focal length

900mm 1350mm 1350mm 375mm 150mm 150mm 762mm 400mm

Page 69: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

49

The back-scattered light from the 488nm laser was split and relayed with pairs of

lenses from the DM to the first photo multiplier tube (PMT1, Hamamatsu Photonics,

H7422-20). The fluorescence light was separated with a dichroic mirror and relayed to

PMT2 (Hamamatsu Photonics, H7422-40). The back-scattered light from the 663nm SLD

was split by BS3 (beamsplitter, R:T=30:70), the reflected portion was acquired by PMT3

(H7422-50), and the other portion went to the SH-WFS, which was created by a lenslet

array (Pitch = 150 μm, f = 6.43 mm) and a CMOS camera (UI306xcp-M; IDS Imaging

Development Systems GmbH). The back-scattered light created a 6 mm circular aperture

on the SH-WFS with a total of 1264 wavefront samples. The WFS-AO control is described

in Section 5.2.3.

The image acquisition program was developed using custom C/C++ for real-time

image display and to control the galvanometer mirrors. The current output from each PMT

was converted to voltages with transimpedance amplifiers (Femto, HCA-2M-1M-C) and

digitized with an analog-to-digital converter (NI PCIe-6363) capable of multi-channel

acquisition at 1.00 MSPS (mega samples per second). The frames were sampled at 400

x 200 pixels during acquisition, and the sampling density was reduced to 400 x 100 pixels

to increase the frame rate during WFS-less AO optimizations. The galvanometer mirrors

were scanned using a bi-directional pattern in a 1 kHz sinusoid that acquired data in the

forward and backward scan directions. The images were de-warped for display in real-

time.

5.2.2. WFS AO description

We used a custom control software to perform WFS measurements from the

sample and to control the DM for closed-loop AO aberration correction. This software was

provided by the University of Verona [91]. The WFS centroids were selected with a circular

aperture for wavefront reconstruction and display in Zernike modes. The WFS-AO was

activated by closing the control loop between the WFS and DM. In order to image different

retinal layers within the eye, the user could enter the desired amount of defocus. Under

this condition, the wavefront measurement would be relative to a reference with the

defocus value included. The WFS software could release the connection from the DM so

that the acquisition software could control the DM for WFS-less AO, yet allow the

wavefront measurements to still be recorded for analysis. The exposure time of the WFS

was set to capture light as the beacon was scanned across the sample, which

Page 70: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

50

accumulated light in the WFS from across an area within the eye. The camera exposure

time and wavefront reconstruction time limited the WFS AO system to 100Hz.

The AO control began with calibrated actuator settings or ‘system flat’, which

removed the system’s static aberrations measured by the WFS. The calibration procedure

is further described in Section 5.2.4

5.2.3. WFS-less AO algorithm.

We implemented a hill climbing Coordinate Search (CS) algorithm, which was

driven by either the reflectance images or the fluorescence images [15]. The CS algorithm

searched within the range of Zernike coefficients expected for a mouse eye. Modal control

of the DM was calibrated using the procedure further described in Section 5.2.4. The

optimization algorithm used a merit function for the highest image sharpness (Simg), which

was calculated by the sum of the pixel intensity squared on the entire image, Equation

5.1. This metric has been used extensively in implementations of WFS-less AO [92–95],

since it is easy to compute with good performance on both reflectance and fluorescence

SLO imaging of a variety of features in the retina [16].

𝑆𝑖𝑚𝑔 = ∑[𝐼(𝑥, 𝑦)]2

𝑥,𝑦

, (5.1)

where I(x,y) is the pixel intensity at the location x,y in the image.

The CS algorithm began with the system flat and then for a given mode (k), a

range of coefficients (±α) were applied to the DM. The coefficient (a*n) that corresponded

to the best image according to the image metric was applied to the DM and the algorithm

moved onto the next mode. Before recording the metric values used for the optimization

algorithm, extra imaging frames were included each time the DM returned to the best

coefficients to guarantee sufficient settling time between searched modes.

For the first iteration, the CS algorithm began by finding the best initial defocus (k

= 4) value, then the astigmatisms, and then continuing in ascending order to include up to

the 5th radial order of the Zernike polynomials (18 modes total). The Zernike polynomials

were ordered and reported using the mode number according to the OSA/ANSI standard

[68]. Further iterations always began with defocus and then the other 17 modes were

Page 71: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

51

optimized in a variety of sequences. Typically, we searched 3 to 4 iterations of n = 18

modes with m = 21 coefficients for each mode, which would require 20 seconds per

iteration. Note that the optimization speed was limited by the imaging system frame rate.

The number of coefficients and the number of iterations could be easily adjusted within

the same imaging session if required. We stopped iterating when the image quality metric

no longer significantly increased from the previous iteration. The following procedure

further explains each step in the CS algorithm.

1) Set the DM actuators values to the calibrated system flat and set the Zernike

coefficients to be 𝑎𝑘 = 0, for 𝑘 = 1, 2, … , 𝑛.

2) If this is not the first iteration, then the selected Zernike coefficients from the previous

optimization are applied to the DM.

3) For each Zernike mode 𝑘 for 𝑘 = 1, 2, … , 𝑛 starting with 𝑘 = 1.

a. Update DM shape using Zernike mode k with varying amplitude over

a range of ± ak,max by incrementing with 𝑚 evenly spaced steps:

𝑎𝑘,𝑚 = − 𝑎𝑘,𝑚𝑎𝑥 +2𝑎𝑘,𝑚𝑎𝑥

𝑚−1(𝑖 − 1), for 𝑖 = 1,2, … , 𝑚.

b. Calculate the merit function on the image, 𝑆𝑖𝑚𝑔, for each coefficient

𝑖 = 1,2, … , 𝑚 and select the coefficient with the highest value from

the search, 𝑎𝑘,𝑚∗ .

c. Apply the selected coefficient 𝑎𝑘,𝑚∗ .

d. Move optimization search to the next Zernike mode, 𝑘 = 𝑘 + 1.

4) After each iteration through the Zernike modes, the algorithm can repeat and search

around the chosen coefficient values from the previous iteration. The Zernike modes

can also be searched in a different sequence in further iterations.

The WFS-less could also be performed after the SH-WFS optimization. In this case,

the best corrected wavefront as determined using the SH-WFS software could be applied

as a starting point for the WFS-less algorithm to perform additional image improvement.

5.2.4. WFS and WFS-less AO system calibration

The system was calibrated for WFS AO by placing a model eye and using the

imaging light source for measurements. The model eye was constructed from a 100 mm

focal length achromatic doublet (AC254-100-A) that focused light onto a scattering sample

Page 72: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

52

such paper. Wavefront measurements were extracted from the SH-WFS camera image

feed. The incoming light passed through the Shack-Hartmann lenslet array, where each

lenslet focuses the incoming beam over a portion of the CMOS pixel array. The distance

from the measured position of the focused spot to the position of the ideal, non-aberrated

spot is related to the local wavefront gradient and is called slope [52]. The spot position

relative to each lenslet was determined with a Thresholded Weighted Center of Gravity

(TWCoG) algorithm. Then, the slopes are used to reconstruct the wavefront with Zernike

coefficients.

The influence of each actuator on the wavefront was measured by poking the

actuators and collecting the slope responses. Hadamard pattern was used to reduce

calibration time and maximize SNR [96]: Each poke pattern was a vector of half 0’s and

half 1’s, with an orthogonal column space generated from all the patterns. The slopes are

then measured after the actuators reach steady state. To account for stroke hysteresis,

the actuators are poked with the same patterns but with an inverted sign, and the average

between the two slopes responses is kept. The slopes are used to generate the actuator

control value space, obtained with the SVD of the calibrated influence matrix. The control

values are processed in an integrator array which asymptotically steers to remove the

difference between the aberrated wavefront and the DM shape compensating for it. With

the model eye in place, the actuator control values to remove the system’s static

aberrations were characterized by closing the WFS AO control loop using the calibrated

control matrix. Those values, called system flat, are saved for the WFS-less AO mouse

imaging.

For the WFS-less AO algorithm, the system was calibrated for modal control with

the Zernike polynomials. This was performed by stopping the scanning and illuminating

the DM and WFS in the reverse direction, starting at the sample pupil plane, illuminating

the sample with a ‘single pass’ of the DM. This was necessary in order to include the

Zernike modes tip and tilt in the actuator measurements. The wavefront influence of each

actuator was measured in Zernike modes, and each measurement was used as a column

in an influence matrix (A). Then, the pseudo-inverse was calculated so that the actuator

control values (c) for any Zernike vector (z) could be calculated by Equation 5.2.

𝐜 = 𝐀ϯ𝐳 . (5.2)

Page 73: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

53

5.2.5. Animal handling and image processing

The animal handling in this work was performed in accordance with guidelines of

the animal study protocol approved by the University of California Animal Care and Use

Committee (IACUC). The three strains of adult animals, pigmented (C57Bl/6), albino

(BALB/cJ), and Cx3cr1GFP/+ mice from Jackson Labs (2-6 months old, female, 5 for each

strain) were used in the experiments, including Cx3cr1GFP/+ mice strain had retinal

microglia cells labeled with Enhanced Green Fluorescent Protein (EGFP). During imaging,

the mice were anesthetized with isoflurane (2% in O2), and the eye was dilated with

tropicamide (1%) and phenylephrine (2.5%). The anesthetized mouse was aligned to a

zero Diopter contact lens (Unicon Corporation, Osaka, Japan) with a gel (GenTeal, Alcon,

Fort Worth, United States) placed between the lens and the cornea to prevent dehydration

and the development of cataract.

Motion within and between imaging frames was mostly caused by the respiration

of the anesthetized mouse [81]. Registering and aligning frames in post-acquisition

processing was required for averaging of frames to improve the image quality. Typically,

frames were recorded at 10 fps for 5 to 10 seconds for a total of 50 to 100 frames for

averaging.

The registration process began by manually selecting a single frame, ideally free

of motion artifact, as the template to align the other frames. The rigid registration process

included a global frame translation followed by the translation of image slices created in

the horizontal fast-scan direction. The global frame translation was determined by

maximizing the cross-correlation between the target frame and the template frame using

the fast Fourier transform. Frames that had a much lower cross-correlation value than

most other frames due to large amplitude of motion were discarded. The remaining frames

were broken up into horizontal strips of 3 vertical pixels. Each strip was translated

horizontally and then vertically with subpixel resolution to maximize the cross-correlation

with equivalently sized strips on the template. Finally, the translated strips were averaged

together and down-sampled to the original frame size. This method was effective on low

SNR images [97]. However, it would not be capable of correcting motion within the 1.5 ms

required to acquire 3 lines as well as rotational and torsional distortions.

Page 74: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

54

5.3. Results

5.3.1. WFS and WFS-less AO for phantom imaging, comparison of performance

We constructed an imaging phantom (a ‘model eye’) prior to in vivo mouse retinal

imaging in order to evaluate the AO performance in an ideal case. The model eye was

composed of a 100 mm focal length achromatic doublet and 30 µm diameter fluorescent

particles (Cat. No. 36-6, Ex/Em Max: 542/612 nm, Dry Fluorescent Particles, Thermo

Scientific, Waltham, US) on white paper. This enabled good WFS measurements for the

WFS-based AO and the fluorescence images provided well-defined features for the WFS-

less optimization using the fluorescent channel. Aberrations were loaded onto the DM

using a prior wavefront measurement from a mouse eye [19], which decreased the

fluorescence image quality as shown in Fig 5.2(a). First, we used WFS AO to correct the

aberrations; then we used WFS-less AO to correct the same aberrations. As described in

the Methods, the WFS-less AO correction proceeded for 4 iterations and increased the

sharpness metric during the optimization, as shown in Fig 5.2(b). The image quality metric

after WFS-less AO was 16% better than the image produced by the WFS-based AO.

The residual ocular wavefront was recorded before, during, and after aberration

correction using the slopes from the 1264 wavefront centroids, which was used to

reconstruct the first 300 Zernike modes and to calculate the RMS of the wavefront. The

WFS AO-corrected aberrations typically within the first 0.25 seconds, as shown in the

recorded wavefront in Fig 5.2(c). The WFS-less AO total execution time depended on the

optimization parameters selected for the CS algorithm, which could be modified in the

software user interface. In this optimization, we densely sampled 18 modes with 21

coefficients. This required about 80 seconds to decrease the WFS measurement to below

the diffraction limit according to the Maréchal criterion (λ/14), as shown Fig 5.2(d). The

optimization proceeded for 4 iterations, but it only took 2 iterations to get to 70% of the

final image metric value. Fig 5.2(e) compares the wavefront measurements in Zernike

modes before and after each method of AO to show that both methods effectively remove

aberrations.

The measured RMS of the residual wavefront after WFS AO was 0.023 ± 0.003

µm and after WFS-less AO was 0.047 ± 0.002 µm over 100 measurements (1 s of

Page 75: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

55

wavefront data) including defocus. The RMS after WFS-less AO was 0.029 ± 0.002 µm

when defocus was removed from the calculation. The defocus term could explain the

image quality improvement of the WFS-less image over the WFS image, which would be

a result from the WFS-less AO having an imaging plane at the middle of the fluorescent

beads and the WFS AO having an imaging plane on the paper. Tip/tilt aberrations caused

by scanning across the sample were removed from all RMS calculations.

Page 76: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

56

Figure 5.2. Phantom imaging of fluorescent beads and wavefront measurements during Wavefront Sensor Adaptive Optics (WFS AO) and Wavefront Sensorless Adaptive Optics (WFS-less AO). (a) Fluorescence images of 30 µm beads on white paper with a 100 mm focal length model eye before AO, after WFS AO, and after WFS-less AO. For the inset image before AO, the pixel intensity values were multiplied by 8, so the beads could be visualized. (b) The increase in the fluorescence image quality during the WFS-less AO optimization. (c) The wavefront RMS excluding defocus, tip and tilt during WFS AO correction. (d) The wavefront RMS excluding defocus, tip and tilt during WFS-less AO optimization. (e) The Zernike decomposition of the wavefront measured before and after each method of AO correction.

Page 77: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

57

5.3.2. WFS and WFS-less AO comparison on mouse photoreceptor mosaic

To perform in vivo imaging experiments, we first we used pigmented mouse strain

(C57BL/6J) and targeted the strongly scattering RPE/choroid layer for good WFS

measurements. The SLD source was used for imaging and wavefront measurements with

a Field of View (FOV) of ~70 µm. First, we corrected and recorded the aberrations from

the mouse eye with WFS AO and acquired reflectance images of the rod photoreceptor

mosaic after the aberration correction, shown in Fig. 5.3(a). After WFS AO correction, the

RMS was calculated to be 0.07 ± 0.02 µm over 100 measurements. The system flat was

applied to the DM and aberration correction was performed with the reflectance images to

drive the WFS-less AO. The images after optimization are shown in Fig. 5.3(b). The

optimization is shown Fig. 5.3(c) used 5 iterations; however, the image quality was 75%

of the final quality after only 3 iterations. Fig. 5.3(d) shows the aberrations removed from

the wavefront as the RMS of the wavefront decreased during the optimization from 0.80 ±

0.01 µm to 0.22 ± 0.01 µm over 100 measurements, which included defocus in the

calculation. Note, we did not separately correct non-common path aberrations although

based on Fig. 5.2 results, correcting them would have a negligible impact on the remaining

results presented this chapter. In this optimization, the Zernike modes were searched in

the same order for each iteration. During iteration 4, the motion of the mouse disrupted

the optimization but the algorithm recovered in the final iteration. Fig. 5.3(a,b) shows that

the image quality is similar after both methods of AO finish correcting residual waterfront

error and the image quality metric after WFS-less AO was 9% better than the image

produced by the WFS-based AO before post-processing. Fig. 5.3(e) shows the wavefront

represented by Zernike modes before and after each correction method.

Page 78: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

58

Figure 5.3. Imaging the mouse photoreceptor mosaic with Wavefront Sensor based Adaptive Optics (WFS AO) and Wavefront Sensorless Adaptive Optics (WFS-less AO). (a,b) Images after WFS AO and WFS-less AO. Scale bar: 10 µm. (c) The image quality improvement during WFS-less AO optimization. (d) The wavefront RMS during WFS-less AO optimization. (e) The Zernike decomposition of the wavefront measured before and after each method of AO.

We repeated this imaging experiment and measurements with different mouse

eyes with a variety of optimization parameters with similar imaging performance. Fig. 5.4

represent additional measurements for imaging the mouse retina photoreceptors. In both

of these cases, the WFS AO and the WFS-less AO demonstrated similar image quality

and measured aberrations were removed. In Fig. 5.4(a), after WFS AO correction, the

RMS was calculated to be 0.06 ± 0.02 µm. For the WFS-less optimization after the first

iteration, the search order of the Zernike modes was shuffled for each of the following two

iterations. The wavefront RMS decreased during the optimization from 0.93 ± 0.04 µm to

0.20 ± 0.01 µm. In Fig. 5.4(b), after WFS correction, the RMS was calculated to be 0.08 ±

0.03 µm. For the WFS-less optimization, the number of coefficients searched for each

Zernike mode was decreased from 21 to 11 and further iterations were included instead.

The wavefront RMS decreased during the optimization from 1.16 ± 0.02 µm to 0.24 ± 0.01

µm. (Note: The Zernike decomposition coefficient values of the measured wavefront after

WFS AO were too small to be well visualized on the same plot as the measured wavefront

before AO.)

Page 79: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

59

Figure 5.4. (a, b) Further mouse photoreceptor imaging with Wavefront Sensor Adaptive Optics (WFS AO) and Wavefront Sensorless Adaptive Optics (WFS-less AO). Images mouse photoreceptor mosaic after WFS AO and WFS-less AO. Scale bar 10 µm. The Zernike decomposition of the wavefront measured before and after each method of AO. The wavefront RMS during WFS-less AO optimization. The image quality improvement during WFS-less AO optimization.

Page 80: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

60

5.3.3. AO SLO reflectance imaging of an albino mouse strain

To demonstrate the potential advantage of WFS-less AO over classical AO, we

performed imaging of an albino mouse strain (BALB/cJ). Since these mice lack melanin,

which highly scatters light in the RPE and choroid, we expect the WFS measurements to

be greatly degraded due to lack of well-defined reference plane. The SLD source was also

used for imaging as well as wavefront measurements. The WFS centroids shown in Fig.

5.5(a) show the scattered light used for the WFS measurements in an albino mouse for

inner retinal imaging compared to the ideal case in a pigmented mouse and the centroids

from the albino mouse are larger and less sharp than the centroids from a pigmented

mouse. Fig. 5.5(b) shows the RMS excluding tip, tilt, and defocus of the measured

wavefront and Fig. 5.5(c) shows the improvement in the image quality metric during the

3-iteration WFS-less AO optimization on the NFL layer in an albino mouse. Defocus was

excluded from the RMS calculation since the imaging light was manually focused on the

inner retina. The optimization that had a 52% increase in the image quality metric.

However, the WFS could not measure the wavefront properly in albino. The attempted

WFS measurements reported an RMS change during the optimization from 0.47 ± 0.02

µm to 0.33 ± 0.01 µm with little response to the changes to the wavefront during the

optimization, especially in the high-order Zernike modes.

Figure 5.5. SH-WFS measurements from an Albino mouse strain (BALB/cJ) retina. (a) The SH-WFS centroids of an albino mouse compared to a pigmented mouse. (b) The RMS of the wavefront measurement without defocus. (c) The image quality metric during WFS-less AO optimization.

Fig. 5.6 shows WFS-less AO images of the blood vessels of the retina. The

optimization was performed on the NFL layer, and then the focus was incremented

through the other vascular layers of the inner retina to image the Nerve Fiber Layer (NFL),

Inner Plexiform Layer (IPL), and Outer Plexiform Layer (OPL). The WFS data was not

Page 81: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

61

acquired during this optimization; however, the optimization results demonstrated a 2.2-

fold increase in the image quality after WFS-less AO. The image brightness increased for

NFL, and the image pattern differed among layers, which indicates that the WFS-less AO-

corrected the aberrations to some extent. In contrast, WFS AO was not effective due to

its inability to perform wavefront measurement on the albino fundus.

Figure 5.6. Imaging the inner retinal of an Albino mouse (BALB/cJ) retina with Wavefront Sensorless Adaptive Optics (WFS-less AO). Images of the retina vasculature before and after WFS-less AO in the Nerve Fiber Layer (NFL), and after WFS-less AO in the Plexiform Layer (IPL), and Outer Plexiform Layer (OPL). Scale bar: 10 µm.

5.3.4. AO SLO fluorescence imaging of EGFP microglia cells

Fluorescence SLO imaging was demonstrated with both WFS and WFS-less AO

on EGFP labeled microglia cells (Cx3cr1GFP/+), which were typically found at varying

depths throughout the inner retina. Fig. 5.7(a) shows the WFS AO SLO reflectance image

of the 488 nm excitation light and the simultaneous fluorescence image of a microglia cell

in Fig. 5.7(b). The fluorescence image was superimposed in green on the reflectance

image in magenta in Fig. 5.7(c). Fig. 5.7(d) shows the RMS of the wavefront during the

WFS AO correction calculated with the defocus value excluded since the imaging system

was focused at the inner retina. Fig. 5.7(e) presents the Zernike decomposition of the

wavefront before and after WFS AO to show that measured aberrations are removed.

Page 82: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

62

Figure 5.7. Imaging EGFP labelled microglia with Wavefront Sensor Adaptive Optics (WFS AO). (a) Reflectance imaging in the inner retinal blood vessels. (b) Fluorescence imaging of EGFP labelled microglia. (c) The fluorescence image superimposed in green on the reflectance image in magenta. Scale bar: 20 µm. (d) The measured wavefront RMS during WFS AO without defocus. (e) The wavefront measurements in Zernike decomposition before and after the WFS AO aberration correction.

In the next example, the aberration correction was performed at the same region

of the retina twice, first by the SH-WFS AO, followed by WFS-less AO for comparison as

shown in Fig. 5.8. The WFS-less AO was optimized using the fluorescence images of

microglia and started from a system flat. The optimization was performed with a small FOV

(~40 µm across) and had a 3.5-fold improvement in the image quality after two iterations,

as shown in the before and after WFS-less images. Both methods of aberration correction

show a similar AO imaging performance in the over all image quality in Fig. 5.8(a). The

intensity line plot across the WFS and WFS-less AO images in Fig. 5.8(b) show small

differences in sharpness and brightness at the top and the bottom of the images, which

could be due to a small shift in the focal plane due to the WFS-less AO. However, the

overall resolution is similar.

Page 83: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

63

Figure 5.8. (a) Imaging EGFP labeled microglia within the inner retina of a mouse with Wavefront Sensor based Adaptive Optics (WFS AO) and Wavefront Sensorless Adaptive Optics (WFS-less AO). Fluorescence image with WFS AO aberration correction (left). Fluorescence image with WFS-less AO aberration correction (middle). Fluorescence images before and after WFS-less AO with a ~40 µm FOV (right). Scale bar: 20 µm. (b) The intensity line plot between the red arrows on the WFS AO image and between the blue arrows on the WFS-less AO image.

In some cases, during mouse retinal imaging, the wavefront aberrations are not

reliably measured for WFS AO as presented in Fig 5.7, so we used WFS-less AO to

provide additional aberration correction. In a representative case, we first used WFS AO

to measure and correct aberrations, which resulted in the left image in Fig. 5.9(a) where

the residual aberrations are apparent by the blurred image. The wavefront measured

before WFS AO as shown in Fig. 5.9(b), then WFS AO decreased the wavefront RMS

from 1.60 ± 0.01 µm to 0.06 ± 0.01 µm before manually shifting the imaging plane to the

microglia in the inner retina. The DM actuator values for the AO correction were loaded

as the starting point for the WFS-less AO to provide further improvements to the image

quality, with the results of the correction shown in the images (middle, right) of Fig.

5.9(a). The measured wavefront RMS (excluding defocus) increased to 0.18 ± 0.01 µm

after shifting the imaging plane and WFS-less AO optimization. However, the aberrations

in the inner retina were not correctly measured, as indicated by improvement in the

image quality despite the measured increase in aberrations. This optimization was

performed with two iterations on a smaller FOV (~80 µm across) and demonstrated that

WFS-less AO can correct for residual aberrations and improve upon WFS AO images

during in vivo retinal imaging sessions.

Page 84: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

64

Figure 5.9. (a) Imaging EGFP labeled microglia within the inner retina of a mouse with Wavefront Sensor based Adaptive Optics (WFS AO) and Wavefront Sensorless Adaptive Optics (WFS-less AO). Fluorescence image after WFS AO (left). Fluorescence image after WFS AO and WFS-less AO aberration correction of residual aberration (middle). Fluorescence image with a smaller FOV of microglia dendrites superimposed in green on the reflectance image of the retinal blood vessels in magenta (right). (b) The Zernike decomposition of the wavefront measured before WFS AO and after both methods of AO. Scale bar: 20 µm.

5.4. Discussion

Aberration correction in the living mouse eye presents challenges that include the

relatively thick retina, multiple scattering surfaces, the motion of the sample, and light-

sensitive tissue. This requires a WFS AO system with a fast and densely sampled WFS

and additional system complexity to make wavefront measurements that are suitable for

aberration correction. Alternatively, WFS-less AO can provide image-based aberration

correction at the cost of wavefront optimization time. In this chapter, we presented a

comparison between the imaging performance between the two AO techniques and

demonstrated their trade-offs. Our results include depth-resolved AO-SLO images at

multiple layers of the mouse retina, including reflectance imaging of the photoreceptor

layer of a pigmented mouse and vascular layers of an albino mouse, and fluorescence

imaging of various layers in the inner retina where microglia cells were located. The

wavefront was measured during the WFS-less optimizations to verify that the measured

aberrations from the mouse eye were removed. We also demonstrated that the WFS-less

AO could provide similar image resolution as WFS AO for imaging the rod photoreceptors

in the outer retina and fluorescently labeled microglia cells found in the inner retina of a

living mouse.

Page 85: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

65

The WFS AO was operated at 100 Hz and, unlike the WFS-less AO, was not

limited by the imaging frame rate and the aberrations were typically corrected within ~0.25

seconds. The WFS-less approach implemented in this chapter used a coordinate search

(CS) algorithm, requiring ~20 seconds per iteration through 18 Zernike modes. This makes

WFS AO advantageous in the presence of varying aberrations from the sample (in our

case, mostly due to eye movements) as the speed of WFS AO enables continuous

correction. However, WFS-less AO, as a method of correcting static aberrations, is valid

for in vivo mouse retinal imaging since the aberration correction can converge and data

can be acquired before the aberrations change (e.g. 5 minutes). The WFS AO was also

less affected by the amplitudes of the aberrations from the sample, as long as the

aberrations were within the dynamic range of the WFS. In contrast, the sample aberrations

in each case can affect the length and search space required for WFS-less AO. The results

presented in this report showed that WFS-less AO was able to provide similar or slightly

better imaging performance to the WFS-based AO. However, it may be required to search

higher-order Zernike modes to guarantee diffraction-limited performance in all cases [5]

or, alternatively, a more compact search space could be used [98].

The CS approach was used because it is straight forward to understand and

reproduce. Our optimization time could be reduced by calculating the best Zernike mode

amplitude based on fewer measurements [53]. Theoretically, as few as 2n+1

measurements for n Zernike modes could be used [86], but practically the convergence

of our algorithm would be limited by the image noise and motion artifact. Real-time image

registration and tracking on a region of interest could also be used to reduce erroneous

measurements due to the motion of the sample [71]. Other algorithms have also been

demonstrated such as Steepest gradient descent, and simulated annealing [79,80,99].

Alternatively, model-based approaches can theoretically converge even faster such as

‘sphere packing’ described by Booth et al. [100], and the DONE algorithm described by

Verstraete et al. [101]. Also, so-called pupil segmentation approaches have been

demonstrated [20,45]. Several reviews are available on the topic such as [86,102] just to

name a few. However, all of these methods will experience the same challenges to enable

rapid convergence in vivo.

Exclusively using WFS-less AO allows for more flexibility and simplicity when

designing the imaging system, such as our recent report on a compact WFSless AO

system [16]. The lens-based system with WFS-less AO allowed for a larger FOV and could

Page 86: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

66

potentially have better peripheral retinal imaging [103]. WFS-less AO also provides

flexibility with the imaging sample since often retinal feature in other layers such as the

large blood vessels can cause strong reflection that confounds the WFS measurement.

Although for both WFS and WFS-less AO, we must also consider the size of the

isoplanatic patch. If our FOV is too large when performing AO, we may only partially

correct for the aberrations.

A comparison can be made between AO for human retinal imaging versus other

applications, such as microscopy of small animal brain imaging in vivo. In the first case,

with a relatively long eye length and thin retina, the focal planes of the WFS beacon and

the imaging plane are nearly the same. In the latter case, backscatter from multiple depths

of the thick tissue layers impedes conventional WFS. The case of mouse retina AO is an

intermediate case between these extremes. Under ideal circumstances for mouse retinal

imaging, WFS works extremely well. When alignment to the mouse eye is near perfect,

and when the beacon and image focal planes are coincident, the convergence is rapid,

and diffraction limited imaging is quickly attained. For the more general case of mouse

retinal imaging, when the region of interest is outside of the mouse eye optical axis, or

when there are features (like blood vessels) that impede WFS measurement, the WFS-

less provides a solution.

The accuracy of WFS-less AO has been investigated by others including

Facomprez et al. [104] for microscopy and demonstrates that WFS-less approaches are

capable of diffraction limited imaging. WFS-less AO has been compared to WFS AO in

microscopy by Bourgenot et al. [105] demonstrating a benefit of an image-based approach

to AO. A comparison of imaging quality between WFS and WFS-less AO has also been

performed for human retinal imaging by Hofer et al. [80]. Although our results demonstrate

a similar image quality between each method in Fig 5.3 (photoreceptor imaging), the

residual wavefront RMS values are different. This may be due to small differences

between the focal planes of the beacon used for the WFS and the imaging plane, or

imaging an area larger than the isoplanatic patch where the aberrations are not uniform

across the entire FOV. Non-common path errors could also be a suspected source of

measurement errors; however, in this system, the phantom imaging in Fig 5.2 indicates

the amount of non-common path aberrations in the system. The phantom images revealed

a 16% improvement in the image quality in favor of the WFS-less AO but, the difference

in wavefront measurements between the WFS AO and WFS-less AO had an RMS of only

Page 87: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

67

~0.029 µm (excluding defocus). In the mouse eye, the difference between the wavefront

measurements after each method of AO was typically much larger. For example, the case

presented in Fig. 5.3 had a measurement difference of ~0.15 µm imaging the

photoreceptor layer and the case presented in Fig 5.9 had a measurement difference of

~0.17 µm imaging the inner retina. So, it is likely that the non-common path aberrations

as a source of discrepancy could be neglectable here.

WFS-less AO is advantageous in the optically thick small animal eyes where

aberration correction can be performed at the specific depths where the features of

interest are located, such as the microglia cells that are found at many layers throughout

the inner retina. The outer retinal layers typically provide the strongest scattering plane for

the WFS, and a large defocus is required to image the inner retinal layers. When the WFS

beacon and the imaging light share the same focal position at the inner retina, and the

scattering from the outer retina dominates and thus reduces the WFS spot quality. This is

the case represented in Fig. 5.9, where the WFS-less further improved the imaging quality

after the WFS AO. Another solution is to defocus the beacon relative to the imaging beam,

but this difference could also adversely affect the performance of the WFS.

Albinism in mice and other small animals is a common background phenotype for

transgenic strains. In the retina, the reflectivity characteristics are different in the albino

specimen due to the different amounts of pigments in the retina layers [106,107].

Performing retinal AO on albino mice would be more similar to the case of brain imaging

since there is no good intrinsic guide star for the WFS AO, and alternatives such as WFS-

less need to be considered.We have previously demonstrated that albinism does not affect

the ability of WFS-less OCT [61]. In this chapter, we have demonstrated that WFS-less

AO for SLO also provides the flexibility to include albino mice as samples using depth-

resolved back reflectance. Albino images were not presented with the SH-WFS mode of

operation since a good wavefront measurement is difficult due to enhanced scattering

from the choroid and sclera (due to the lack of pigment in the RPE and choroid).

The imaging applications presented in this chapter span the scenarios of having

the wavefront sensing plane and the imaging plane coincident and separated. For the

case where the WFS-less approach used the fluorescently labeled cells of interest as

guide stars, it is reasonable that the aberration correction would be better based on reports

in the Literature using guide stars, for example [90,108]. Alternatively, in cases where

Page 88: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

68

fluorescence guide stars are not available, methods such as coherence-gated wavefront

sensing [24,109] have been shown to provide depth-resolved measurements, but at the

expense of imaging system complexity.

5.5. Summary

In this chapter, we confirm previous reported WFS-less AO aberration correction

for imaging the eye using direct measurements. For retinal imaging in anesthetized and

stable small animals, our results indicate that exclusively using WFS-less methods is

reliable given enough time for the optimization method to find the best correction.

However, when imaging time is limited, WFS-based methods have a large advantage in

achieving the optimal aberration correction, especially for the correction of time-varying

aberrations. The ideal AO imaging for small animal imaging should use both methods in

order to find the best aberration correction across different mouse strains, different retinal

layers and eccentricities, and for different levels of sedation. Finally, our results also

suggest that AO without a well-defined guide star requires the use of WFS-less methods

for the optimal aberration correction.

Page 89: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

69

Chapter 6. Multi-modal imaging

6.1. Introduction

The study of small animal models of human diseases causing blindness is

important to understand the mechanisms of vision loss and to develop novel therapies.

Conventional histological approaches require sacrificing the animal at each study time

point. Non-invasive imaging is highly desirable for longitudinal studies, reducing the effects

of inter-animal variation and reducing the number of animals required for a study. There

would be potential benefits and advancements if more researchers had access to high

resolution in vivo imaging systems with the functional and structural detection capabilities

that were previously only attainable through histology [3]. Furthermore, in vivo imaging

allows for the study of physiological processes such as the dynamics of microglia [50,110].

Theoretically, the Numerical Aperture (NA) through the pupil of the mouse eye

permits sub-micrometer imaging of the retina. However, optical aberrations introduced by

the tear film, cornea, and lens reduce the actual resolution. In order to approach diffraction-

limited imaging, these aberrations can be corrected with Adaptive Optics (AO) using a

wavefront corrector such as a Deformable Mirror (DM) [4,5].

AO has been implemented in many ophthalmic imaging modalities such as Optical

Coherence Tomography (OCT), Scanning Laser Ophthalmoscopy (SLO), and fundus

photography, which have been well documented in References [2–4,51–53,111]. The

traditional approach to AO is to use a Wavefront Sensor (WFS) to measure the ocular

aberrations directly. For example, AO SLO has been demonstrated for in vivo imaging

with cellular resolution of Green Fluorescent Protein (GFP) labelled cells [4,14,62].

Performing accurate wavefront measurements for WFS AO imaging in a small animal

retina requires a high level of system complexity due to the short length of the eye creating

an optically thick sample with multiple scattering surfaces [5]. Alternatively, Sensorless

AO (SAO) has the potential to allow for systems to be compact, easily operated, robust,

and inexpensive. SAO does not require direct measurements of the optical wavefront but

instead uses an image-based aberration correction approach, such as a multi-dimensional

optimization or pupil segmentation [20,86]. SAO methods have the ability to provide depth

Page 90: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

70

resolved aberration correction by using images acquired at specific layers within the retina.

For example, AO OCT has been demonstrated using en face projections extracted from

three dimensional OCT volumes to drive the optimization algorithm on the selected retinal

layers [61].

The multi-modal system in this chapter was designed to incorporate SAO with

multiple modalities including Optical Coherence Tomography (OCT), OCT based

Angiography (OCT-A), confocal Scanning Laser Ophthalmoscopy (SLO), and

fluorescence detection. In this chapter, we present a compact lens-based design of a

imaging system for multi-purpose imaging of the small animal retina, which has

significantly improved performance and functionality since previous reports

[15,61,112,113]. The en face and cross-sectional imaging enable visualization of the

retinal structure while the fluorescence imaging has the ability to visualize the biological

function of the retina through labelled reporter cells. OCT and SLO can be combined to

employ a multi-modal system for simultaneous and co-localized structural and functional

imaging. We present representative images and analyses to demonstrate the

performance, versatility, and usability of the system for small animal imaging. Images

acquired prior to SAO aberration correction demonstrate the widefield and standard

resolution imaging in a mouse eye. After performing SAO optimization, our results

demonstrate high resolution imaging featuring in vivo volumetric and time-lapse imaging

of fluorescently labelled microglia.

6.2. Methods

6.2.1. Optical design

A schematic of the optical layout of the system is presented in Figure 6.1(a). The

system components were assembled with off-the-shelf optomechanics and custom

mounts designed with SolidWorks (Dassault Systèmes, France) as shown in the Figure

6.1(b). The light sources for the imaging system included a near infrared (NIR)

Superluminescent Diode (SLD, BLM2-D, Superlum Diodes Ltd., Ireland) for OCT using a

central wavelength of 840 nm with a spectral bandwidth of ~80 nm, and a 488 nm laser

(0488L-13A, Integrated Optics, Lithuania) for confocal SLO and fluorescence excitation.

Page 91: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

71

The OCT subsystem was based on a fiber optic Michelson interferometer. The

OCT light was split by a 70:30 single mode optical fiber coupler (AC Photonics Inc, CA,

USA). The 70% portion of the light was connected to a reference arm consisting of a fiber

collimator, a dispersion compensation block and a mirror. The OCT probe beam was the

30% portion of light from the coupler, which was launched from a reflective collimator

(RC04FC-F01, Thorlabs Inc., NJ, USA) and transmitted through a cold mirror (ZT670rdc-

xxrxt, Chroma Technology Corp, VT, USA) for combination with the 488 nm SLO light.

In the SLO subsystem, another reflective collimator (RC08FC-F01, Thorlabs Inc.,

NJ, USA) was used to launch the SLO light from a fiber with a polarization controller, such

that the horizontally polarized light was reflected from a Polarization Beam Splitter (PBS,

PBS251, Thorlabs Inc., NJ, USA). The light was then reflected from a dichroic mirror

(ZT405/488/561rpc-UF1, Chroma Technology Corp, VT, USA) to the cold mirror, and then

co-aligned with the OCT light.

The first pupil plane of both subsystems was defined by the Variable Focus Lens

(VFL, Arctic 39N0, Corning, NY, USA) with an aperture of 3.9 mm. The imaging beams

were relayed and magnified to a continuous membrane DM (DM69, Alpao, France) with a

10.5 mm aperture, and then to a mounted pair of Galvanometer-scanning Mirrors (GM,

6210H, Cambridge Technology Inc., MA, USA) with a clear aperture of 3.0 mm. Finally,

the light was reduced to a beam diameter of 1.0 mm to be focused by the mouse eye and

relayed from the GM to be scanned across the retina with a maximum scanning angle of

50 degrees. The optical relays were constructed using achromatic doublets with an

antireflection coating for both visible and near infrared light (VIS-NIR Coated Achromatic

Lenses, Edmund Optics). Each relay used two off-the-shelf lenses, except the final relay

to the mouse eye. See L5 and L6 in Figure 6.1. These elements were constructed from

two achromatic lenses that were placed symmetrically with a <1 mm air gap [114,115]

This design enabled shorter optical relay required for our desired scanning angles without

introducing significant aberrations.

Page 92: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

72

Figure 6.1. (a) Schematic of Optical Coherence Tomography (OCT) and confocal Scanning Laser Ophthalmoscopy (SLO) system. The cyan represents the beam path of only 488 nm light, the green represents the beam path of only the fluorescence emission and the red represents the beam path of only the SLD light. The pink represents the co-aligned beam path of the 488 nm light, fluorescence emission, and SLD light. System components: Superluminescent diode (SLD), fiber coupler (FC), polarization controller (PC), polarization beam splitter (PBS), dichroic mirror (DC), emission filter (EF), cold mirror (CM), variable focus lens (VFL), deformable mirror (DM), galvanometer-scanning mirrors (GM), quarter wave plate (QWP), photomultiplier tube (PMT), dispersion compensation block (DCB), mirror (M). Achromatic doublet lenses: L1=50mm, L2=150mm, L3=300mm, L4=75mm, L5=2x125mm, L6=2x50mm. (b) Computer simulation of optical layout on custom optical mounts using OpticStudio and SolidWorks.

Page 93: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

73

The final element before the mouse eye was a Quarter Wave Plate (QWP,

WPQ10E-488, Thorlabs Inc., NJ, USA), which rotated the polarization state of the back-

scattered SLO light from the sample so that the light would be transmitted by the PBS for

detection [15,56]. Although this technique does not entirely remove the strongest reflection

in the center of the back-scattered SLO images (see the Results section), this method

does remove other reflections from the optical elements. The optical design was simulated

in OpticStudio (Zemax, WA, USA) and Figure 6.2(a) presents the spot diagrams with an

ideal model eye. The expected resolution and calculations were performed for mouse eyes

using an NA of 0.25. The OCT and SLO spot diagrams are presented across a 15-degree

(~500 µm for mouse eyes) Field of View (FOV) with 0 D of vergence at the sample pupil

plane. The black circle in the top row represents the Airy disk with a ~2.1 µm radius for

820 nm, 840 nm, and 860 nm. The middle and bottom row have a 1.2 µm Airy disk radius

for 488 nm SLO light. The bottom row represents the 488 nm spots scanned across 7

degrees (~230 µm for mouse eyes) with a 20 D vergence at the pupil plane for the eye

produced by the simulated VFL. These FOVs (or smaller) are typically used for AO in the

mouse eye [4], whereas for imaging a larger FOV, it may not be necessary to have a spot

size on the order of microns. The axial resolution for OCT was estimated to be ~3 µm. For

the SLO axial resolution, the FWHM of the axial point spread function was ~18 µm. For

the OCT-A imaging, the system was reconfigured to have a smaller beam into the eye,

which reduced the NA to 0.15 and the Airy disk radius to 3.5 µm.

In order to reduce the overall size and simplicity of the system, we did not use an

optical relay between the fast and slow scanning mirrors, which prohibited perfect

conjugation to the pupil of the mouse eye with both scan axes. In Figure 6.2(b), the amount

of pupil wander in the mouse eye was simulated for the FOVs that are suitable for

diffraction-limited imaging.

Page 94: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

74

Figure 6.2. (a) Spot diagrams of the OCT light at 820 nm (red), 840 nm (pink) and 860 nm (purple) across a 15-degree FOV, where the black circle represents the Airy disk with a 2.1 µm radius. Spot diagrams of the 488 nm (blue) SLO light scanned across a 15-degree FOV with 0 D of vergence at the sample pupil plane and 7-degress with 20 D of vergence at the sample pupil plane where the black circle represents the Airy disk with a 1.2 µm radius. (b) The boundary of the imaging beam at the final pupil plane of the system. The black circle represents a 2 mm aperture. Each color represents a different scan position across a 15-degree and 7-degree FOV to simulate the pupil wander due to the space between the scanning mirrors in the optical design.

The back-scattered OCT light from the sample was recombined with the reference

arm light at the fiber coupler and directed to a spectrometer (Cobra 800, Wasatch

Photonics, NC, USA). The A-scans were acquired with a frame grabber (PCIe-1433,

National Instrument, Austin, TX) at 100 kHz and the OCT volumes were sampled at 1024

x 400 x 200 points. For OCT-A, two B-scans were acquired at the same location in the

slow scan direction to calculate changes due to blood flow. The OCT/OCT-A cross-section

and en face view were processed for real-time display using our custom GPU accelerated

acquisition software [89,112,113] written in C/C++.

The fluorescence emission was transmitted through the dichroic mirror, and the

emission filter (ZET405/488/561m-TRF, Chroma Technology Corp, VT, USA), and then

focused into a multimode fiber with a core diameter of ~2 Airy disk diameters (ADD) that

directed the light to a photo-multiplier tube (PMT, H10723-20, Hamamatsu Photonics K.

K., Japan). The back-scattered 488 nm laser light was reflected from the dichroic mirror,

transmitted through the PBS, focused into a multimode fiber with a core diameter ~5 ADD

or ~20 ADD, and detected by another PMT (H7827-002, Hamamatsu Photonics K. K.,

Japan). We used a 5 ADD fiber core when performing image-based SAO with the back-

Page 95: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

75

scattered images, or else we used a 20 ADD fiber core, which provided the higher SNR

for general navigation on the mouse retina. The PMT signal gain could be adjusted on the

power supply depending on the amount of signal from the sample. The digitization (PCIe-

6361, National Instrument, Austin, TX) of the PMTs was synchronized to the acquisition

of the OCT A-scans for simultaneous imaging, otherwise the SLO could be operated alone

at a 2 kHz line rate with a sampling density of 400 x 200 points.

6.2.2. Sensorless adaptive optics

The SAO could be performed on the en face projection of the OCT volumes, the

back-scattered confocal SLO, or the fluorescence SLO images. We implemented a hill

climbing Coordinate Search (CS) algorithm presented in Chapter 5, which provided an

exhaustive search to find the optimal Zernike coefficients. The merit function for

optimization was determined by the highest image sharpness (Simg) [94,95], defined by

the sum of the intensity squared of each image pixel I(x,y) in Equation 6.1.

𝑆𝑖𝑚𝑔 = ∑[𝐼(𝑥, 𝑦)]2

𝑥,𝑦

, (6.1)

The CS algorithm started with a flat wavefront with an RMS ~0.05 µm, which was

calibrated using a SH-WFS in the location of the GM scanners. Then, for the first mode

(𝑘) in a sequence, a range of coefficients (±α) was applied to the DM. The coefficient (𝑎𝑛∗ )

that corresponded to highest metric value was applied to the DM and the algorithm moved

onto the next mode. For the first iteration, the sequence of modes began with a defocus

(k = 4) value, then the astigmatisms, and continuing in ascending order up to the 21st mode

for a total of 18 modes. The Zernike polynomials were ordered and reported using the

mode number according to the OSA/ANSI standard [68]. The sequence of 18 modes was

usually repeated for multiple iterations, typically 2 or 3 times, until the metric value no

longer increased. Successive iterations would search coefficients ranges (±β) around the

previously selected coefficients. Between iterations, the imaging FOV or location could be

adjusted, as the features of interest became visible.

For high resolution imaging, SAO could be used to correct wavefront aberrations

from the mouse eye using the output from of the different imaging modalities for the image-

based optimization. During optimization, the sampling density of the OCT was decreased

Page 96: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

76

to 1024 x 400 x 20 which resulted in ~20 seconds for each iteration of the optimization.

When the SLO was used for optimization, the sampling was set to 400 x 100 and each

iteration took a total of 12 seconds.

6.2.3. Animal handling

The animal imaging sessions were performed under protocols compliant to the

Canadian Council on Animal Care and the approval of the University Animal Care

Committee at Simon Fraser University. The mice were anesthetized with a subcutaneous

injection of ketamine (100 mg/kg of body weight) and dexmedetomidine (0.1 mg/kg of

body weight). A drop of topical solution (Tropicamide, 1%) was applied to dilate the ocular

pupils. A rigid 0-Diopter contact lens was placed on the animal eyes to prevent the cornea

from dehydration and then the animal was aligned without any further contact to the

imaging system [19]. For fluorescein angiography, the mice were subcutaneously injected

with 100 µL of 100 mg/mL fluorescein. Mice were purchased from the Jackson Laboratory,

including wild type strain (C57BL/6J) and transgenic strain with Enhanced Green

Fluorescent Protein (EGFP) labelled retinal ganglion cells (Tg(Thy1-EGFP)MJrs/J) and

microglia (B6.129P-Cx3cr1tm1Litt/J).

For retinal imaging, the OCT imaging light was limited to 620 µW. The SLO imaging

light did not exceed 230 µW in this chapter and was limited to 100 µW when operating

simultaneously with the OCT.

6.2.4. Image processing

Images in this chapter were generated by standard post-processing techniques,

including steps to register and align frames to a template for averaging, using a

combination of Matlab (MathWorks Inc, MA, USA) and ImageJ (National Institutes of

Health (NIH), MD, USA) toolkits. The number of volumes and frames that were saved

could be easily changed in the acquisition software. For the images presented in this

chapter, we used the following parameters: for OCT, we recorded 5 volumes per

acquisition in 4 seconds; for OCT-A images, only one volume was recorded per acquisition

in 1.6 seconds; and for SLO, we recorded 50 to 100 frames per acquisition in 5 to 10

seconds. The OCT B-scans were aligned with a vertical translation to remove axial motion

Page 97: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

77

of the animal. Most of the B-scans presented in this chapter were an average of 5 adjacent

B-scans within one volume with an exception that is explained in the results section.

The en face OCT images were generated using a Maximum Intensity Projection

(MIP) between two manually selected horizontal lines corresponding to depths in the

retina. Then, the en face OCT projections and the SLO frames were processed with the

following procedure: 1) The registration process was initialized by manually selecting a

single frame as the template to align the other frames; 2) Each frame was globally

translated horizontally and vertically to maximize the cross-correlation with the template;

3) The frames were broken up into horizontal strips and each strip was translated

horizontally and vertically to maximize the cross-correlation with the template [70,97]; 4)

The frames were non-rigidly aligned to the template with a sum of squared differences

similarity metric along cubic B-splines using the Medical Image Registration Toolbox

(MIRT) [69]; 5) After registration, the frames were averaged and the black and white

thresholds were adjusted to enhance the image brightness and contrast for presentation.

All the B-scans in this chapter are presented in a linear intensity scale; 6) The images

were scaled so that the vertical and horizontal dimensions have the same scale.

SLO frames from the back-scattered and fluorescence channels were

simultaneously acquired, which would allow for co-registration if the fluorescence signal

was insufficient [24]. However, in this chapter, the fluorescence images had sufficient

signal to use directly for registration.

6.3. Results

6.3.1. Imaging without adaptive optics

For imaging large retinal features, a widefield image is preferred and it may not be

necessary to perform SAO. Figure 6.3(a) demonstrates a 50-degree OCT B-scan and a

44-degree en face projection of the Outer Plexiform Layer (OPL) of a wild type mouse

retina. Unlike the other B-scan images in this chapter, in Figure 6.3(a), the vertical

scanning was disabled and 200 B-scans were acquired, aligned, and averaged. In Figure

6.3(b) and (c), the sampling density is increased with a 22-degree FOV and the focus was

shifted with the VFL from the OPL to the Nerve Fiber Layer (NFL). The B-scans and en

Page 98: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

78

face images were registered and averaged as described in the previous section. The

location of the B-scan is indicated by the red dashed line.

Figure 6.3. (a) OCT B-scan across 50 degrees in the mouse retina and en face projection of the outer plexiform layer (OPL) across 44 degrees. The B-scan is an average of 200 consecutively acquired cross-sectional frames and the en face OCT image is an average of 5 frames. (b,c) Average of 5 adjacent OCT B-scans and an average of 5 en face OCT frames of the OPL. The B-scans are located at the position of the red dashed lines. Vertical scale bar: 50 µm. Horizontal scale bars: 100 µm.

After acquiring OCT volumes, the OCT was disabled in the software so that the

SLO could be operated at a faster speed. The structural SLO image in Figure 6.4(a) was

generated by the 488 nm back-scattered channel light from 50 averaged frames with the

imaging light focused on the NFL and a 5 ADD confocal aperture. Fluorescence SLO

images were generated from an average of 50 frames acquired a few minutes after a

fluorescein injection. Images were acquired from three different vascular layers in the inner

retina, including the NFL, inner plexiform layer (IPL), and OPL. These images were

combined with a MIP for presentation in Figure 6.4(b), which is demonstrated in further

detail in the following results with AO.

Page 99: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

79

Figure 6.4. Confocal SLO images of a mouse retina with 488 nm light. (a) Structural image of the nerve fiber layer from back-scattering. (b) Fluorescein angiography composited with a MIP from images of three different vascular layers. Scale bar: 100 µm.

OCT-A B-scans were created by calculating the difference between two intensity

B-scans in the same location. Figure 6.5(a) shows the en face OCT-A image generated

from the MIP of the OPL layer in the B-scans of a single volume. For comparison, Figure

6.5(b) shows the en face OCT intensity image that was generated from the same OPL

region. Figure 6.5(c) was created by coloring the en face OCT-A images that were

extracted from the OPL with red, the IPL with green, and the NFL with blue.

Figure 6.5. (a) En face OCT-A images of the OPL in a mouse retina. (b) En face OCT intensity image from the same image data. (c) En face OCT-A images that were generated from the OPL (red), IPL (green), and NFL (blue). Scale bar: 50 µm.

6.3.2. Structural imaging with sensorless adaptive optics OCT and SLO

For SAO OCT, the retinal layer of interest was selected and the image quality

metric was calculated on the en face image to drive the optimization. Figure 6.6 represents

an example of an imaging sequence. The imaging plane was focused on the OPL layer

with a FOV of ~250 µm. En face OCT images were used for optimization, and then OCT

Page 100: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

80

volumes were acquired for presentation. Figure 6.6(a) shows en face OCT images before

and after SAO at different focal planes. Figure 6.6(b) demonstrates the two-iteration

optimization with a plot of the image quality metric for each step in the optimization and

the coefficients selected for each iteration. Overall, there was a 1.9-fold improvement in

the image quality metric reported from the merit function of the optimization algorithm.

Figure 6.6. (a) En face images of the outer plexiform layer (OPL, top row, ~250 µm FOV) and nerve fiber layer (NFL, bottom row, ~280 µm FOV) retinal layers before and after Sensorless Adaptive Optics (SAO). SAO-OCT B-scans with the imaging focal plane on the OPL (red arrows) and NFL (blue arrows). (b) The normalized image quality for each step in the SAO optimization over two iterations and the Zernike coefficients selected for each iteration. Vertical scale bars: 50 µm. Horizontal scale bars: 20 µm.

Similarly, for SAO SLO, a typical imaging procedure is presented in Figure 6.7.

The confocal pinhole was 5 ADD for the structural SLO images in Figure 6.7(a). In this

case, the imaging light was focused on the NFL layer and the back-scattered SLO images

were used for the optimization metric. Figure 6.7(a) shows the averaged SLO images

before SAO and after SAO images of the NFL. Then the focus was shifted with the VFL

to image different retinal layers including the OPL as shown. Figure 6.7(b) demonstrates

the improvement in the image quality metric during the optimization. With a FOV of ~250

µm, the first iteration improved the image quality 2.2-fold. Then the FOV was reduced by

Page 101: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

81

half in the second iteration, which further improved the image quality 1.6-fold. The FOV

could be changed for the successive iterations because new image quality values would

be determined each iteration. Therefore, the plot of the image quality metric over the entire

optimization had to be normalized independently for the metric values in each of these

iterations. Figure 6.7(b) also presents the coefficients selected for each iteration.

Figure 6.7. (a) Confocal SLO images before and after Sensorless Adaptive Optics (SAO) of the nerve fiber layer (NFL) with a FOV ~250 µm. Images of the outer plexiform layer (OPL) after SAO. (b) The normalized image quality metric values for each step used for the SAO optimization for each iteration. The Zernike coefficients selected for each iteration. Scale bar: 20 µm.

For both SAO OCT (Figure 6.6) and SAO SLO (Figure 6.7), dark circular “holes”

are revealed in between the nerve fiber bundles after aberration correction similar to other

AO-SLO images [4], which are speculated to be retinal ganglion cell soma due to the size

and shape.

6.3.3. Fluorescence imaging with sensorless adaptive optics

The ability to image EGFP labelled cells with the SAO SLO fluorescence detection

further increases the functionality of the imaging system. The results in this section

demonstrate the SAO SLO image quality and the abilities of the fluorescence detection

Page 102: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

82

channel. The structural confocal SLO images in this section were acquired with the

detection pinhole ~20 ADD.

For Figure 6.8, a larger FOV (~750 µm) was used to locate the EGFP labelled

Retinal Ganglion Cell (RGC). The imaging FOV was reduced to ~250 µm to perform SAO

on the fluorescence imaging channel, then a second iteration was performed on a further

smaller FOV ~100 µm with only dendrites of the RGCs in view. Figure 6.8(a) presents a

comparison of the images acquired before and after SAO, including a line plot across

between the blue (before SAO) and red arrows (after SAO). Figure 6.8(b) presents the

structural images that were acquired when the imaging plane was focused on the RGC

axon. In the right column, the structural image was colored in magenta and the

fluorescence image was overlaid in green in order to better localize the RGC.

Figure 6.8. Confocal SLO images of a mouse retina with labelled retinal ganglion cells (Tg(Thy1-EGFP)MJrs/J). (a) Fluorescence images before and after Sensorless Adaptive Optics (SAO) and an intensity line plot between the blue arrows (before SAO) and red arrows (after SAO). (b) The left column presents structural images focused on the nerve fiber layer at a ~750 µm FOV (top) and ~230 µm FOV (bottom). The right column presents the structural image in magenta overlaid by the fluorescence image in green. The fluorescence image was composited from two different focal planes for the axon and the dendrites of the RGC. Scale bars: 50 µm.

We performed fluorescein angiography to demonstrate the confocal capability to

discriminate different layers in the inner retina. Figure 6.9 presents SAO images of three

distinct vascular layers, including the NFL, IPL and OPL. The images were composited

Page 103: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

83

using a MIP, with the NFL in red, IPL in green, and OPL in blue to show the vessel

connections in the axial direction.

Figure 6.9. Confocal SLO fluorescein angiography of a mouse retinal vasculature after Sensorless Adaptive Optics. Images (left to right) of the nerve fiber layer (NFL), inner plexiform layer (IPL), outer plexiform layer (OPL), and the MIP with the NFL in red, IPL in green, and NFL in blue. Scale bar: 50 µm.

To demonstrate the volumetric imaging ability of the system, we imaged a mouse

with EGFP labelled microglia, which are located in many retinal layers. SLO images were

acquired at 18 different focal positions between the OPL and NFL layer, and the axial

location of the fluorescence was determined by the structural images. The depth fly-

through of the back-scattered SLO images with the co-localized fluorescence SLO image

is presented in Visualization 1 of reference [16]. Figure 6.10 presents images from the

sequence, where the two right-most images are the structural and fluorescence images

from the NFL layer. The fluorescence image in the middle-left (Fig. 6.10) was located

immediately below the NFL layer and the image on the far-right was located deeper into

the retina at the OPL. The images were color-coded in depth from the OPL to the NFL and

presented with 3D shadowing that was rendered by ImageJ in Visualization 2 of reference

[16].

Page 104: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

84

Figure 6.10. Confocal SLO images with Sensorless Adaptive Optics of EGFP labelled microglia in the mouse retina (B6.129P-Cx3cr1{tm1Litt}/J) acquired at different focal position between the outer plexiform layer (OPL) and the nerve fiber layer (NFL) selected from Visualization 1 of reference [16]. The microglia images were color-coded in depth between the OPL and the NFL of the retina and rendered in 3D for Visualization 2 reference [16]. Scale bar: 20 µm.

Microglia are known to constantly survey the surrounding environment and time-

lapse imaging can reveal the dynamics of the cellular branches [110,116]. The microglia

in Figure 6.11(a) were located just below the NFL and these images were selected from

a 1-hour time-lapse video that acquired images in 20 second intervals. The SAO was

performed periodically throughout the imaging to ensure optimal image quality. For each

optimization, the FOV was reduced to 52 µm across, containing only the microglia

branches. The image in Figure 6.11(b) was annotated and color-coded at these time

points to highlight areas of growth and retraction. Visualization 3 in reference [16] presents

the entire time-lapse with the time stamp of acquisition.

Page 105: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

85

Figure 6.11. (a) Confocal SLO fluorescence images with Sensorless Adaptive Optics of EGFP labelled microglia in the mouse retina (B6.129P-Cx3cr1{tm1Litt}/J) from three time points in the time-lapse video from Visualization 3 reference [16]. (b) The microglia images color-coded with time. The white arrows 1-4 note areas of significant growth and retraction. Scale bar: 20 µm.

The central microglia in Figure 6.11(b) had a branch (white arrow #1) that retracted 24

µm, with an average velocity of 4.8 µm/min from minute 3 to minute 8, and a branch (white

arrow #2) that retracted 38 µm, with an average velocity of 1.3 µm/min during minute 19

to minute 49. The microglia branch on the right of the image (white arrow #3) generally

retracted over 50 minutes but also had periods of extension during that time. The microglia

branch on the left (white arrow #4) appears to move towards another microglia branch

(white arrow #2) at minute 24.

We performed further time-lapse imaging of microglia using the same methods for

SAO in order to investigate the potential effect of the imaging light. The 488 nm imaging

light was reduced to 100 µW for 39 minutes, then the exposure was increased to 230 µW

and imaging proceeded for another 50 minutes, as shown in Visualization 4 of reference

[16]. Figure 6.12(a) shows time-points before and after the laser power was increased.

The image in Figure 6.12(b) was annotated and color-coded image at these time points to

highlight areas of growth and retraction.

Page 106: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

86

Figure 6.12. (a) Confocal SLO fluorescence images with Sensorless Adaptive Optics of EGFP labelled microglia in the mouse retina (B6.129P-Cx3cr1{tm1Litt}/J) from three time points in the time-lapse video from Visualization 4 of reference [16] with an increase in laser power at 39 minutes. (b) The microglia images color-coded with time. The white arrows 1-2 note areas of significant growth and retraction. Scale bar: 20 µm.

6.4. Discussion

In this chapter, we have demonstrated a multi-modal en face imaging system with

diverse functionality for vision scientists needing a variety of imaging requirements. The

system imaging modalities include en face OCT, OCT-A, SAO OCT, as well as SLO and

SAO SLO with fluorescence detection. Our system uses lens-based optical relays

between the active elements, which include the VFL, the DM, and the GMs. Our results

demonstrate state-of-the-art AO imaging for the mouse retina and represent improvement

from our previous reported systems for each individual modality. Our results demonstrated

the variety of in vivo imaging abilities that included structural imaging, angiography,

volumetric and time-lapse imaging of microglia cells.

The imaging system primarily used an NA of 0.25 into the mouse eye, which only

represents about half of the theoretical maximum. However, with 488 nm light, this still

has a calculated resolution of ~1 µm. For the purpose of this chapter, the image quality

Page 107: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

87

was sufficient for clearly imaging the microglia branches and to report metrics, such as

movement speeds, yet maintains good quality imaging without requiring AO for imaging

large features. The system was initially designed and tested for mouse imaging; however,

it is also capable of imaging the rat retina as well, which is often required by many vision

researchers for longitudinal studies [117]. Since the rat eye is larger than the mouse eye,

this decreases the maximum attainable resolution. However, it was still beneficial to have

the SAO to correct for aberrations.

During the time-lapse imaging of microglia cells, we only illuminated the retina with

488 nm since we did not require the use of a beacon for WFS measurements. The

microglia time-lapse in Figure 6.11 appears to have more retraction than the microglia

time-lapse from Figure 6.12, despite the increase in laser intensity. It is possible that this

was normal microglia surveillance of a healthy retina or a response to the 488 nm imaging

light. If the 488 nm imaging light itself has an effect on microglia, then it may be difficult to

conclude the reason for a microglia response when investigating their role in immunity

studies. There is no established maximum permissible exposure (MPE) for the mouse

eyes; however, other groups have scaled the MPE for SLO in human eyes [4,64,97]. The

MPE for human SLO imaging decreases with imaging FOV [118], so as we image small

features in small animal experiments, it will be important to continue to consider laser

irradiance as a potential factor.

The imaging system was designed to be used by a non-specialist and future

improvements could improve the reliability and robustness of the SAO. For example, a

Region of Interest (ROI) within the display could be selected by the user instead of

reducing the entire imaging FOV, which further increases the exposure during the ~10 to

20 seconds required for the optimization iteration. Real-time image tracking on the ROI

would also enable the optimization algorithm to follow an object of interest or reject frames

with a large amount of motion artifact [71,85,119]. In this chapter, we were using a multi-

iteration exhaustive search, which was robust to the occasional motion artifact over the

~30 to 60 seconds required for the entire optimization. However, accurate image tracking

would encourage the use of faster optimization algorithms, such as model-based

approaches [100,101] that require much fewer measurements, thereby decreasing

optimization time. This would be advantageous to further reduce the exposure of the entire

imaging process and the potential for damage over time.

Page 108: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

88

In this chapter, we optimized up to the 21st Zernike mode for 18 modes in total.

The improvement in the image quality after each mode is optimized is represented in

Figure 6.7(b), which demonstrates that there is an increase in the metric value in the 5th

radial order in the first and second iteration. Using higher orders in the optimization

algorithm could improve results but it would come at the cost of algorithm time. Since time

is limited for in vivo imaging, the algorithmic execution time is better spent on further

iterations [120]. For example, the step sizes between coefficients can be reduced to

improve the wavefront correction. Furthermore, successive iterations have an improved

SNR, which will also improve the performance of the AO correction.

6.5. Summary

In conclusion, we have demonstrated a lens-based system, capable of high-

resolution en face small animal imaging with multiple modalities. The compactness and

simplicity of the system allow for the potential translation to vision scientists that require

tools for in vivo and longitudinal studies. Our results demonstrate the potential for studying

individual cells, such as RGCs and microglia, in healthy and diseased animal models.

Page 109: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

89

Chapter 7. Non-invasive cellular-resolution imaging of the retina with two-photon excited fluorescence

7.1. Introduction

Non-invasive retinal imaging is a valuable tool that is used in both clinical and

preclinical vision research to aid the development of novel therapies for preventing

irreversible vision loss. More sensitive assessment of the physiological and biochemical

processes within the retina could be used to detect earlier signs of disease in order to

preserve sight [12,121]. Fluorescence can be used to image many biomarkers, since

fluorophores can be added to the retina to provide contrast or fluorophores that are

intrinsic to the retina can be used for measurements. For example, measuring

autofluorescence from the Retinal Pigment Epithelium (RPE) is of high interest for

investigating diseases such as Age-Related Macular Degeneration (AMD) and Stargardt

disease [3,17,122].

For many fluorophores intrinsic to the retina, the single-photon excitation is in the

ultraviolet (UV) range and the fluorescence cannot be excited through the eye of many

species due to the ocular transmission window [65,81]. Imaging the retina non-invasively

with Two-Photon Excited Fluorescence (TPEF) could enable novel in vivo studies of

disease and retinal physiology [65,81,123,124]. Furthermore, the multiphoton process

suppresses out-of-focus background signal, which improves the axial sectioning without

a confocal aperture in the optical detection path. Imaging the retina with near-infrared

(NIR) light has advantages since the retina is less sensitive to NIR than visible light. NIR

light is also less scattered within biological tissue than the equivalent visible light required

to excite the same fluorophores [37,38].

The difficulty of retinal TPEF imaging in vivo is that high energy is typically required

to generate the TPEF, while minimizing the incident exposure energy is required for the

imaging to be non-invasive. An active area of research is the development of technology

that reduces the average laser power required for TPEF imaging in the eye [12,97,122].

The TPEF signal intensity is highly sensitive to the focused spot size, which leads to high

Page 110: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

90

sensitivity to aberrations [38]. It has been demonstrated that the TPEF signal for retinal

imaging can be increased by using Adaptive Optics (AO) for aberration correction

[24,97,123].

For high-resolution imaging of the mouse retina, optical aberrations introduced by

the eye must be corrected for diffraction-limited imaging. AO has been demonstrated to

correct ocular aberrations extensively [3,4,13,39]. The traditional approach to AO is to use

a Wavefront Sensor (WFS) to directly measure the aberrations. Alternatively, it has also

been demonstrated that Sensorless AO (SAO) can provide depth-resolved aberration

correction using an image-based approach. SAO has the advantage of avoiding the

system complexities that are required due to the short length of the mouse eye, which

creates an optically thick sample with multiple scattering surfaces [5].

In this chapter, we present a multi-modal imaging system that uses SAO and

Optical Coherence Tomography (OCT) to achieve non-invasive TPEF imaging of cellular

features in the retina. We used the same light source to simultaneously generate OCT

and TPEF, where both imaging modalities could be used to drive the SAO optimization

algorithm. We demonstrate the high-resolution OCT/TPEF system with a variety of

samples, which included fluorescein angiography (FA) and fluorescently labelled cells. We

also demonstrate the ability to non-invasively image intrinsic fluorescence from the RPE

of the mouse retina in various strains.

7.2. Methods

7.2.1. System setup

The SAO OCT TPEF imaging system used a femtosecond pulsed laser (Mai Tai

HP, Spectra-Physics, CA, USA) for both the OCT and TPEF excitation. The laser had a

tuning range from 690 nm to 1040 nm, where the central wavelength was chosen for each

of the sample fluorophores. Table 7.1 summarizes the calculations to estimate the axial

resolution of the OCT in tissue [125] and the maximum laser power used for each

fluorophore imaged in this report. The laser power could be adjusted with neutral density

filters. A schematic of the imaging system is presented in Figure 7.1.

Page 111: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

91

Figure 7.1. Schematic of the Sensorless Adaptive Optics (SAO) Optical Coherence Tomography (OCT) and Two-Photon Excitation Fluorescence (TPEF) imaging system. The imaging system was constructed with a pellicle beam splitter (PeBS), a variable focus lens (VFL), a deformable mirror (DM), a dichroic mirror (DcM), galvanometer-scanning mirrors (GM), emission filters (EF), a photo-multiplier tube (PMT), dispersion compensation (DC), and the following lenses: L1=100 mm, L2=300 mm, L3=400 mm, L4=100 mm, L5=2×125 mm, L6=2×50 mm. The reference arm denoted as a dashed line.

For the OCT/TPEF system, we used a pellicle beam splitter to separate the light

into the sample arm and the reference arm. In the sample arm, we used a variable focus

lens (VFL, A-39N1, Corning, NY, USA) to control the focal plane within the sample, a

continuous membrane deformable mirror (DM, DM-69, ALPAO, France) to correct the

wavefront aberrations, and XY mounted pair of galvanometer-scanning mirrors (GM,

6215H, Cambridge Technology Inc., MA, USA) to scan the light across the sample. The

scanning angles of the GMs could be adjusted to change the imaging field of view (FOV)

on the retina and the maximum FOV for this system was ~25 degrees (~850 µm). Three

lens-based telescopes were used to optically relay the pupil planes at the VFL, DM, and

GMs to the mouse eye. The telescopes were constructed with achromatic doublets that

had focal lengths listed in Figure 7.1. The imaging beam entered the mouse eye with an

NA of ~0.25. Table 7.1 summarizes the calculations to estimate the spot size in tissue

using the Airy disk radius and the FWHM of the axial point spread function for each center

wavelength [26]. The reference arm included dispersion compensating glass to match the

sample arm.

Page 112: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

92

Table 7.1. Laser specifications used for each fluorescent sample and the calculated resolution.

Fluorescent Sample

Center Wavelength

(nm)

Spectral Bandwidth

(nm)

Laser Power (mW)

Lateral Spot Size (µm)

TPEF Axial

Resolution (µm)

OCT Axial Resolution

(µm)

FA 800 13 < 3 2.0 29 16 GFP 910 16 < 9 2.2 33 17 YFP 940 15 < 3 2.3 34 19 RPE 740 8.2 < 8 1.8 26 22

For OCT, back-scattered light from the eye and the light returning from the

reference arm were combined at the PeBS and coupled into a single mode fiber. The

spectral interference signal was then detected by a custom-built spectrometer (1024

pixels, 100 kHz, Bioptigen Inc., NC, USA) with a spectral range from ~730 nm to ~995 nm.

The data was acquired through a Camera Link frame grabber board (PCIe-1433, National

Instrument, Austin, TX) and processed using a custom GPU-accelerated program for real-

time processing and display [89]. The two-dimensional transverse area (en face) was

scanned and sampled with 400 × 200 A-scans, which resulted in an acquisition rate of 1

volume per second. For OCT-based Angiography (OCTA), we used two intensity B-scans

(two BM-scans) to calculate the angiography at 400 × 200 (200 × 2) A-scans. During the

SAO-OCT optimization process, the density in the direction of the slow scan was reduced

to 400 × 20 A-scans for aberration correction using the en face OCT images.

TPEF emission from the sample was de-scanned by the GMs and reflected by the

dichroic mirror (DcM) to the photo-multiplier tube (PMT, H7422P-40, Hamamatsu

Photonics, Japan) detector. The short-pass filters (FF01-650/SP-25, FF01-720/SP-25,

IDEX Health & Science LLC, NY, USA), a lens, and an aperture were positioned before

the PMT to reject stray light. The electric current from the PMT was converted to a voltage

with a transimpedance amplifier (LCA-400K-10M, FEMTO Messtechnik GmbH, Germany)

and digitized by a DAQ device (PCIe-6361, National Instrument, TX, USA). The digitization

of the PMT signal was synchronized with the OCT A-scans for simultaneous operations,

which ensured that both OCT and TPEF images were co-registered. The TPEF could also

be operated without the OCT to acquire at 10 frames per second for averaging in post-

processing.

The SAO optimization used a hill-climbing coordinate-search algorithm, which was

recently reported for mouse retinal imaging [16,18]. In brief, the optimization used the merit

function defined by the image sharpness of the en face OCT images that were extracted

Page 113: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

93

at the user-selected depth or the TPEF images. The algorithm began by finding the best

initial defocus value, then the astigmatisms, and then continuing in ascending order for a

total of 18 modes. Multiple iterations could be performed to ensure optimal the aberration

correction. The optimization required ~20 seconds per iteration using either the OCT or

the TPEF images.

7.2.2. Animal handling and image processing

The mouse imaging was performed under protocols compliant to the Canadian

Council on Animal Care and the approval of the University Animal Care Committee at

Simon Fraser University. A subcutaneous injection of ketamine (100 mg/kg of body

weight) and dexmedetomidine (0.1 mg/kg of body weight) was used to anesthetize the

mouse prior to imaging. The pupils of the mouse were dilated with a drop of topical solution

(Tropicamide, 1%). A rigid 0-Diopter contact lens was placed on the mouse eyes to

prevent dehydration of the cornea. Then, the animal was aligned to the imaging system

without any contact [19]. All of the mice imaged in this work were purchased from The

Jackson Laboratory, ME, USA, which included B6 mice (C57BL/6J), albino B6 mice

(B6(Cg)-Tyrc-2J/J), mice with GFP labelled microglia (B6.129P-Cx3cr1tm1Litt/J), mice with

GFP labelled retinal ganglion cells (Tg(Thy1-EGFP)MJrs/J), mice with YFP labelled neural

cells (B6.Cg-Tg(Thy1-YFP)16Jrs/J), and mice with a mutated rpe65 gene (B6(A)-

Rpe65rd12/J). Table 7.2 summarizes the that mice used for the imaging presented in this

report. In addition, all the mice were female and weighed 25 – 35 grams. For fluorescence

angiography, the mice were anesthetized and then subcutaneously injected with 100 μL

of 100 mg/mL fluorescein.

Table 7.2. Summary of mice that were used in this report.

Mouse Strain Stock Number Pigmentation Number

C57BL/6J 000664 Pigmented 1

B6(Cg)-Tyr{c-2J}/J 000058 Albino 3

B6.129P-Cx3cr1{tm1Litt}/J 005582 Pigmented 3

Tg(Thy1-EGFP)MJrs/J 007788 Pigmented 1

B6.Cg-Tg(Thy1-YFP)16Jrs/J 003709 Pigmented 1

B6(A)-Rpe65{rd12}/J 005379 Pigmented 2

The images in this report were generated with post-processing steps that were

performed using Matlab (MathWorks Inc, MA, USA) and ImageJ (National Institutes of

Page 114: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

94

Health, MD, USA), which included steps to register, average, and contrast stretch the

images [16]. For OCT images, we recorded 5 volumes per acquisition. For TPEF images

of fluorescently labelled cells and FA, we recorded 100 frames per acquisition. For TPEF

images of the RPE layer, we recorded 300 to 900 frames per acquisition. The OCT images

presented in this chapter are either cross-sectional B-scans in the fast scanning direction,

or the en face view at a user-selected depth. The OCT B-scans were aligned to remove

axial motion with a vertical translation to maximize the cross-correlation between the

images. The B-scans presented in this report were an average of 5 adjacent B-scans

within one of the acquired volumes. The en face OCT and OCTA images were generated

with a Maximum Intensity Projection (MIP) between two manually selected axial positions

on the B-scans, which corresponded to the focal plane within the sample. From five en

face OCT/OCTA volumes that were acquired sequentially, one image was selected as the

template image and the other images were registered to the template image prior to

averaging. In the registration process, vertical and horizontal image translations were first

performed to maximize the cross-correlation between the template and moving image.

Then, each image was non-rigidly aligned to the template with a sum of squared

differences similarity metric along cubic B-splines using the Medical Image Registration

Toolbox (MIRT) [69]. For the TPEF images, the same rigid and non-rigid steps were used

to register each frame to the template. However, the unregistered average image was

used as a template [97]. After registration, the OCT/OCTA and TPEF images were

averaged, the image pixel intensity was scaled for presentation if required, and the images

were resized to have the same vertical and horizontal scale. The images that were

acquired at multiple depths from the same eye and the images before/after SAO were

processed identically for comparisons. We estimated the peak signal-to-noise ratio (SNR)

defined by Eq. 7.1 [126,127]:

𝑆𝑁𝑅 = 10 log

max (𝐼)2

𝜎𝑏2 , (7.1)

Where I is the image pixel intensities and σb2 is the variance, which was calculated from

a region within the image that only contained background noise.

Page 115: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

95

7.3. Results

The results are split into three parts for each type of fluorophore that was imaged

in the mouse eye. Please see Section 7.3.1 for fluorescein angiography (FA), where we

used OCT to guide the aberrations correction to reduce the laser power to < 3 mW. Section

7.3.2 we demonstrate the imaging performance with fluorescently labelled cells and

Section 7.3.3 we demonstrate imaging of the RPE layer with SAO in a variety of mouse

strains.

7.3.1. Fluorescein angiography

The murine vasculature of the inner retina is stacked in three distinct layers,

including the Outer Plexiform Layer (OPL), Inner Plexiform Layer (IPL), and Nerve Fiber

Layer (NFL). This hierarchical structure provides an opportunity to demonstrate the depth-

resolved aberration correction and the ability of TPEF to provide axial sectioning without

a confocal aperture. In Figure 7.2, the laser power was adjusted to 2.5 mW, and the focal

plane was positioned at the OPL of a 6-month-old albino B6 mouse (B6(Cg)-Tyrc-2J/J).

Then, the SAO wavefront optimization was performed using the en face OCT images of

the OPL, which were extracted from the B-scans between the two depth positions that are

marked by yellow arrows in Figure 7.2. The SAO optimization was performed with 2

iterations for a total algorithmic execution time of ~40 seconds. The top and bottom rows

of Figure 7.2 are before and after the aberration correction, respectively. The first column

shows the improvement in the OCT B-scans (linear scale) at the OPL layer. The OCT B-

scans were selected from the cross-section between the blue arrows on the en face OCT.

The middle column shows the improvement in the sharpness and brightness in the en face

OCT. The third column shows improvement of the TPEF images that resulted from the

wavefront optimization on the en face OCT images. The TPEF images were estimated to

have an SNR of 26.3 dB before aberration correction and 31.0 dB after OCT-guided

aberration correction using Equation 7.1.

Page 116: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

96

Figure 7.2. Optical Coherence Tomography (OCT) and Two-Photon Excited Fluorescence (TPEF) images of the mouse retina before (top row) and after (bottom row) OCT-guided Sensorless Adaptive Optics (SAO). The improvement in the OCT B-scan is shown in the left column, the improvement in the en face OCT is shown in the middle column, and the improvement in the TPEF is shown in the right column. The yellow arrows represent the imaging focal position and the line between the blue arrows represents the cross-sectional location of the OCT B-scans. Scale bars: 50 µm.

After SAO at the OPL, using the same aberration correction, OCT and OCTA

volumes were acquired while shifting the focus to the OPL, IPL, and NFL using the

tunable lens, as shown in Figure 7.3(a). The OCT B-scans in top row were from the

same location as in Figure 7.2, where the shift in focal plane can be seen by the change

in image intensities along the depth of the B-scan. The B-scans focused on the OPL,

IPL, and NFL were color-coded in depth and composited with a Maximum Intensity

Projection (MIP), as shown in Figure 7.3(a) (top right). Similarly, TPEF images were

acquired while shifting the focus with the tunable lens through the inner retinal layers

from the OPL to the NFL to create a depth-stack (z-stack). TPEF images were acquired

at 25 depth location with increments of ~5 µm through the inner retina. The bottom row

of Figure 7.3(a) shows the TPEF images focused at the OPL, IPL, and NFL. The TPEF

images were composited with color corresponding to the relative location between the

OPL and NFL (bottom right), which matched the color-coded OCT B-scan (top right).

The red arrows in Figure 7.3(a) point to a few connecting or diving vessels between the

Page 117: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

97

retinal layers [128,129], which appear clearly in the TPEF FA. However, these vessels

are not clearly identifiable as connecting vessels in the OCTA.

In order to demonstrate the effect of the SAO-OCT optimization on the TPEF

imaging in a cross-sectional view, TPEF z-stacks were resliced to generate a cross-

section in the same direction as the OCT B-scans (fast scan direction). Prior to the

extraction of TPEF cross-sectional images from the TPEF volume, the 25 TPEF images

were interpolated to 75 pixels in depth to scale with the OCT B-scans. Figure 7.3(b) shows

the TPEF cross-sections before and after SAO optimization. The image intensity was

plotted between the blue arrows and the red arrows in Figure 7.3(b) to show the

improvement due to the aberration correction. The TPEF cross-sectional slices presented

in Figure 7.3(b) had two distinctive blood vessels on top of each other, which

demonstrations the axial sectioning ability of the TPEF.

Page 118: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

98

Figure 7.3. (a) OCT B-scans (top row), OCTA en face (middle row), and TPEF (bottom row) with the focal plane at the Outer Plexiform Layer (OPL), Inner Plexiform Layer (IPL), and Nerve Fiber Layer (NFL). In the right column, the images of the vascular layers were composited with a MIP. The red arrows point out connecting vessels in the TPEF. (b) Cross-sectional TPEF images (left) of the inner retinal vasculature before and after Adaptive Optics (SAO) acquired with a 25-step z-stack that was interpolated to 75 image pixels. The axial intensity profile plot between the red and blue arrows of the TPEF cross-sectional images. Scale bars: 50 µm.

7.3.2. GFP and YFP labelled cells

GFP labelled cells have small features that can be used to demonstrate the

imaging resolution of the TPEF system. In this section, we demonstrate the ability to

visualize cellular features after SAO optimization.

Page 119: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

99

We imaged a mouse strain with GFP labelled microglia (B6.129P2(Cg)-

Cx3cr1tm1Litt/J) with a center wavelength of 910 nm. Figure 7.4(a) shows a single TPEF

image frame (left), and then an average of 100 frames (right). First, OCT B-scans were

used to navigate to the focal plane where we would expect to find the GFP labelled cells.

Then, as shown in Figure 7.4(b), we decreased the imaging FOV to ~70 µm to performed

the SAO optimization using the fluorescence images. The SAO optimization was

performed with 5 iterations for a total algorithmic execution time of ~100 seconds. The

example in Figure 7.4(b) demonstrates an overall image improvement from an SNR of

33.6 dB before aberration correction to 36.6 dB after aberration correction using Equation

7.1. There was also an improvement in the features of the microglia branches that are

visualized after aberration correction. The images in Figure 7.4(a) and 7.4(b) were from

two mice that were both 18 months of age. In Figure 7.4(c), we show another image after

aberration correction with a FOV that was ~100 µm from another mouse that was also 18

months of age.

Figure 7.4. TPEF imaging of GFP labelled microglia (B6.129P2(Cg)-Cx3cr1{tm1Litt}/J) in the mouse retina. (a) Single TPEF frame (left) and an average of 100 frames (right) at a ~0.8 mm FOV. The red square represents a 100 µm FOV to represent the scale of the microglia. Scale bar: 100 µm. (b) TPEF images of a GFP labelled microglia cells before (left) and after (right) Sensorless Adaptive Optics (SAO). (c) TPEF image after SAO. Scale bars: 20 µm.

We also demonstrate the ability to image GFP labelled retinal ganglion cells (RGC)

of a transgenic mouse strain (Tg(Thy1-EGFP)MJrs/J) with SAO-TPEF and we compared

Page 120: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

100

the images to single photon excited fluorescence (SPEF) with SAO. Figure 7.5 (left) shows

an image of a RGC from a 14-month-old mouse after SAO-TPEF with a FOV of ~200 µm.

Figure 7.5 (middle) shows the same cell imaged with SAO-SPEF with a matching FOV.

The larger FOV of ~1.3 mm is presented in Figure 7.5 (right) to show the location of the

RGC relative to the optic nerve head. The imaging system used for SPEF is described in

a recent report [16], which used a 488 nm laser for fluorescence excitation and the same

numerical aperture (NA) into the mouse eye. By comparing these results, it appears that

the TPEF imaging can resolve similar features in the lateral plane. However, the apparent

size of the soma in the SPEF image is 30% larger than the soma in the TPEF image. The

‘glow’ of the soma is reduced in the TPEF image due to improved TPEF axial sectioning

in this example.

Figure 7.5. Comparison of a GFP labelled retinal ganglion cell that was imaged using SAO TPEF (left) and using SAO SPEF with the same 200 µm FOV (middle). A SPEF image is also shown at a ~1.3 mm FOV (right), where the red square represents the 200 µm FOV that was used for the other images. Left scale bar: 20 µm. Right scale bar: 100 µm.

The Thy-1 YFP-16 Line (B6.Cg-Tg(Thy1-YFP)16Jrs/J) mouse retina was also

imaged using SAO-TPEF with only 2.5 mW at 940 nm . From a 7-month-old mouse, a z-

stack of TPEF images were acquired throughout the inner retina to visualize the

fluorescently labelled cells. In Figure 7.6, the OCT B-scans on the top row are presented

in linear scale to show the focal plane at the same depth location as the TPEF images in

the middle row, which are focused at the NFL, IPL, and OPL.

Page 121: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

101

Figure 7.6. OCT B-scans (top row) and TPEF (middle row) imaging with the focal plane at the Nerve Fiber Layer (NFL), Inner Plexiform Layer (IPL), and Outer Plexiform Layer (OPL) of a Thy-1 YFP-16 Line (B6.Cg-Tg(Thy1-YFP)16Jrs/J) transgenic mouse. The blue arrow and yellow arrow point at fluorescently labelled cell bodies. The red arrow points at fluorescently labelled axons. In the bottom row, the OCTA en face image (magenta) was composited with the TPEF image (green). Vertical scale bar: 50 µm. Horizontal scale bars: 20 µm.

As shown by the OCT B-scans, some of the fluorescently labelled cell bodies (blue

arrow) and axons (red arrow) appear to located near the NFL, whereas others cell bodies

(yellow arrow) appear to be located near the OPL. The fluorescently labelled cells in the

OPL could be a different type of neural cell, such as horizontal cells [130]. In the bottom

row of Figure 7.6, the OCTA en face images (magenta) were combined with the TPEF

(green) with the focal plane at the NFL (left) and the OPL (right) to co-localize of the

fluorescently labelled cells with the blood vessels

7.3.3. RPE imaging

In the following results (Figure 7.7 and 7.8), we demonstrate TPEF imaging of the

RPE with the assistance of SAO-OCT in an albino B6 mouse (B6(Cg)-Tyrc-2J/J). For Figure

7.7, we imaged a mouse that was 5.5 months of age. First, the OCT was used to ensure

Page 122: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

102

the focal plane was at the RPE layer of the retina. In Figure 7.7(a), the OCT B-scans (top

row) are presented in linear scale to show the intensity change as the focal plane was

shifted from the NFL (left), to the OPL (middle), and to the RPE (right). Here, we used the

TPEF images to perform one iteration of the SAO optimization for a total algorithmic

execution time of ~20 seconds. In Figure 7.7(b) and 7.7(c), there was an apparent

increase in signal after SAO and an improved image quality of the RPE mosaic, as shown

by the line plots between the blue arrows and the red arrows. From the line plot, we can

calculate a 25 - 30 µm spacing between peaks, which corresponds the spacing of the RPE

cells.

Figure 7.7. (a) The SAO-OCT B-scans in linear scale (top row) and the en face OCT (bottom row) with the focal plane at the Nerve Fiber Layer (NFL), Outer Plexiform Layer (OPL), and Retinal Pigment Epithelium (RPE) in the mouse retina. The en face OCT images were extracted between the cyan arrows (NFL), yellow arrows (OPL), and green arrows (RPE). The OCT B-scans were located between the red arrows on the en face OCT image. (b) TPEF images of the RPE of the mouse retina before and after SAO. (c) An intensity line plot between the blue arrows and the red arrows on the TPEF images of the RPE mosaic. Scale bars 50 µm.

Page 123: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

103

We imaged the RPE of one mouse (5.5 months of age) four days apart to

demonstrate the ability to perform non-invasive longitudinal imaging of the same eye over

time. This interval was sufficient time for photo-chemical and photo-thermal damage to

appear in the OCT [131,132]. The TPEF from the RPE and the OCT images are shown

on day 1 in Figure 7.8(a). The OCT B-scans are presented in log-scale and show that

there was no damage immediately after using an 8 mW exposure for a few minutes. In

Figure 7.8(b), we imaged the same area four days later, where there was no apparent

damage from the two imaging sessions. In Figure 7.8(c), we have digitally zoomed into a

small area of the TPEF images to show a similar RPE pattern on each day. TPEF images

from day 1 (green) and day 4 (magenta) were combined with a MIP.

Figure 7.8. (a) TPEF images of the RPE (left), en face OCT (middle), and OCT B-scans (right). (b) TPEF images of the RPE (left), en face OCT (middle), and OCT B-scans (right) from the same mouse four days later. (c) The digital enlargement of the TPEF images on day 1 (green) and day 4 (magenta), which were combined with a MIP. Scale bars 50 µm.

An advantage of working with mice is the well-established ability to manipulate

their genetics. However, here, we show differences in the TPEF images from the RPE

layer that resulted from a variety of mouse strains. In Figure 7.9, using 740 nm excitation

light, we imaged the popular pigmented B6 mouse (C57BL/6J) at 16 months of age, an

albino B6 mouse (B6(Cg)-Tyrc-2J/J) at 5.5 months of age, and a pigmented mouse strain

with a disruption of the rpe65 gene (B6(A)-Rpe65rd12/J)) at 2.5 months of age. The TPEF

images from the B6 mouse had an evenly distributed signal from the RPE with only faint

Page 124: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

104

structures visible, as compared the albino B6 mouse, where the RPE mosaic was clear.

In the pigmented rpe65 mouse, we can also visualize the RPE mosaic, which is likely due

to the build-up of retinyl esters storage particles from the dysfunction in the visual system

[123,124]. Fluorescent compounds in the RPE are known to accumulate with age [133],

which would increase the autofluorescence in older mice and could be an additional factor

that affects the visualization of the RPE mosaic.

Figure 7.9. TPEF from the RPE layer of the mouse retina in three different mouse strains, including a pigmented B6 mouse (C57BL/6J), an albino B6 mouse (B6(Cg)-Tyr{c-2J}/J), and a pigmented rpe65 mouse (B6(A)-Rpe65{rd12}/J). Scale bar 100 µm.

We further investigated the ability of TPEF to provide compound-specific contrast.

Figure 7.10 shows TPEF images at different central wavelengths from the pigmented

rpe65 mouse at 18 months of age. At 760 nm, we are able to visualize the fluorophores

that have accumulated near the cell membrane of the RPE cells. However, after 800 nm

the TPEF signal is dominated by a more evenly distributed fluorescence signal in the RPE

layer. These results agree with other Literature [81,133,134], which demonstrate that the

fluorescence from the retinyl ester storage particles is more efficient at 760 nm than at

longer wavelengths. At the longer wavelengths used in these results, the fluorescence

emission is largely due to the compound A2E in lipofuscin. Since autofluorescent

compounds, including A2E, accumulate in the RPE with age [133], there was likely more

fluorescence emission with the older mouse than the rpe65 mouse imaged for Figure 7.9.

Page 125: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

105

Figure 7.10. TPEF image of a pigmented rpe65 mouse (B6(A)-Rpe65{rd12}/J) with different central wavelengths, including 760 nm, 780 nm, 800 nm, and 820 nm. The red arrow highlights an RPE cell where the fluorescence near the cell membrane is reduced with longer wavelengths. Scale bar 50 µm.

7.4. Discussion

In this chapter, we have demonstrated OCT guided TPEF imaging with image-

based wavefront optimization to improve the fluorescence signal. Since the TPEF is very

sensitive to the focal spot size, aberration correction is crucial for high-resolution imaging

with minimal laser power. We used a tunable femtosecond laser source to enable retinal

imaging from a variety of fluorophores. The system also used a tunable lens to shift the

focal plane within the retina, and a deformable mirror for aberration correction with a hill-

climbing coordinate-search algorithm. Our results included high-resolution volumetric

fluorescein angiography of mouse retinal vasculature, sub-cellular imaging resolution of

fluorescently labelled cells, and intrinsic fluorescence imaging from the RPE. We have

demonstrated that some samples can be imaged with < 3 mW of laser power, including

YFP labelled cells and fluorescein angiography. We have also demonstrated RPE imaging

in various mouse strains and we have shown that the amount of pigment in RPE of the

particular strain can reduce the ability to visualize the RPE mosaic with TPEF.

The optical layout in this chapter was similar to the imaging systems for the mouse

retina [15,61] in Chapter 6, which we were able to compact into a 1.5x2 ft footprint, as

described in reference [16]. Using a similar design, the compact system would enable

easier collaboration with other vision scientists.

In the mouse model, common methods such as genetic manipulation or viral

injection can be used to label cells with GFP. In the Literature, TPEF imaging of GFP

Page 126: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

106

labelled cells without AO has been demonstrated [135,136], as well as with WFS-based

AO methods [65]. However, it can be difficult to accurately measure the aberrations from

a mouse eye without a high degree of system complexity, whereas SAO has the ability to

perform depth-specific aberration correction at the cost of algorithmic execution time and

the potential for errors due to sample motion [18]. This is important since high-resolution

GFP imaging with NIR light could be used to investigate the visual response of RGCs with

calcium indicators, such as GCaMP [64,65,135]. While the single photon excitation at 488

nm would provide a higher SNR, the NIR light used for TPEF is much less likely to activate

the visual system. In this chapter, we used 9 mW of laser power at 910 nm to image the

GFP labelled RGCs. It may be possible to measure a calcium response in the RGC by

only imaging soma [136]. This could be performed with a significant reduction in the laser

power, which may be similar or less than the 2.5 mW of laser power at 940 nm that was

required for imaging the YFP labelled cells.

For TPEF, SAO is well-suited for good aberration correction performance due to

the nonlinear relationship between the signal and the spot size [38]. Hence, the TPEF

signal strength is significantly degraded in the presence of aberrations. The overall

optimization convergence speed and performance could be improved by using a

combination of aberration correction with the OCT signal, increasing the TPEF signal

enough that it could be used for a second iteration of the SAO process for a more precise

correction. Furthermore, even in the cases where there are poor structural features in the

OCT at the location of the fluorophore, the SAO could potentially be improved with low-

order aberrations using the OCT images before further optimization using the TPEF.

Similarly, at a large imaging FOV, a faster initial optimization could have been used for

partial aberration correction before reducing the FOV to an area interest where the

aberration correction would have better performance [18].

The images of the RPE in this chapter had strong differences for each strain and

appear to be affected by the type of accumulated compounds, pigmentation, and age of

the mouse. In order to visualize the cellular mosaic of the RPE, the TPEF signal from the

features at the edges of the cells had to be greater than the background signal. This was

the case for the for the albino B6 mice and the pigmented rpe65 mouse, where fluorescent

compounds were visualized near the RPE cell membrane [123,124]. Palczewska et al.

[12] has recently reported results for imaging RPE of albino rpe65 mice with laser powers

down to 1 mW by averaging many frames, and using ultra short light pulses down to ~20

Page 127: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

107

femtoseconds into the eye. Our results use similar methods to average many frames, but

we could potentially visualize the RPE mosaic with similar power levels by imaging the

albino rpe65 mouse and by pre-compensating for the group velocity dispersion through

the optical components. However, we would still be limited by our source at 70 to 100 fs

without sacrificing tunability with a broader spectral bandwidth laser. As shown in this

chapter, the strain of mouse is important for visualizing the RPE mosaic at low power

levels. However, the lack of pigment in albino mouse strains could reduce the performance

of WFS-based aberration correction.

Reducing the average laser power on the retina is important not only for robust

longitudinal animal studies, but also for potential applications in human imaging [121,137].

Although our laser power levels for these results are not safe for human eye imaging,

mouse imaging can be used to further develop imaging techniques to enable TPEF at safe

power levels. Recently, it has been demonstrated that averaging hundreds of OCT

volumes can produce high-resolution retinal imaging of transparent features, including

RGCs [138,139]. This so called ‘super-averaging’ could enable all 100 – 900 TPEF frames

to be co-registered with the OCT, which could further reduce the TPEF signal

requirements.

7.5. Summary

In conclusion, we have demonstrated that SAO and OCT can be used to achieve

non-invasive cellular-resolution TPEF imaging from a variety of fluorophores in the retina.

The increase in TPEF signal after aberration correction enables imaging to be performed

with less laser power than without AO. Finally, if safe levels exposure levels are

determined, the techniques shown here could be developed for human imaging.

Page 128: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

108

Chapter 8. Future work and conclusion

8.1. Technology refinement

Improvements and refinement of the current imaging technology will further

increase the reliability and usability of the SAO imaging. In order to enable widespread

use SAO imaging, the future objectives are to increase the image optimization speed and

reduce motion artifacts. I envision that SAO aberration correction will be quick enough that

it is not be cumbersome to the user or even continuous. In this section, I will outline some

proposed solutions for future investigations.

Although SAO imaging has been demonstrated to be able to provide similar

performance as WFS-based AO, the major disadvantage is the time required for SAO. A

method for potentially improving the sensorless AO is to search an ideal set of basis

functions in the optimization algorithm. An idea basis can be generated from system

measurements in order to orthogonalize the basis used by corrective element to the effect

the modes have on the image quality metric [98]. Generating these modes has been

demonstrated to improved the performance of the sensorless optimization using a modal-

search, especially in single iteration [140,141]. Iterating multiple times through the Zernike

modes is the simple solution that we have implemented to the SAO algorithm, which also

benefits from the improved image SNR as the aberration correction proceeds. An

alternative basis was used for WFS-based AO and described in Section 5.2.4. The basis

was generated by the SVD of the influence matrix measured by the wavefront sensor and

could potentially also be use as modes for the sensorless AO algorithm.

Another solution is to use faster converging algorithms that require fewer imaging

frames to determine the optimal correction. The fastest algorithms develop a mathematical

model using the effect on the images due to known aberrations put on the deformable

element. These are known as model-based algorithms, which can converge is as few as

N+1 measurements, with N being the number of modes that are used in the model.

Examples of model-based algorithms include the DONE algorithm [21,101], and sphere

packing [100].

Page 129: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

109

Since motion artifact can be detrimental to the SAO performance and to the final

averaged image, real-time tracking systems could provide solutions [71,85,119]. We have

started with simple image-based registration, as mentioned in Section 6.4. For SAO, this

has allowed a region of interest (ROI) to be followed and used during the course of an

optimization. The next logical step is to have the imaging system respond to the registered

motion, potentially in three dimensions using the OCT for volumetric localization. For the

motion of an anesthetized mouse, it may be sufficient to follow the sample with the

galvanometer mirrors and tunable lens. Furthermore, this would reduce the reliance on

registration in post processing.

Imaging the human retina with two-photon excited fluorescence will not be possible

without reducing the laser exposures currently required for imaging. As discussed in

Section 7.5, further reducing the imaging power required for TPEF imaging of the mouse

RPE seems possible through the development image processing tools that allow the OCT

and TPEF to be concurrently registered and averaged. Preliminary results, shown in

Figure 8.1, demonstrate that registration and averaging of >150 OCT volumes acquired

from the mouse eye using the multi-modal system described in Chapter 5 can produce

high quality images the nerve fiber bundles.

Figure 8.1. Volumetric averaging of 150 OCT volumes. Scale bar: 50 µm.

8.2. Non-confocal Scanning Laser Ophthalmoscopy

For structural imaging purposes, non-invasive in vivo visualization of the cellular

mosaics in the outer retina in the mouse has been scarce in the Literature, most likely due

to difficultly correcting aberrations. Notable examples include, rod photoreceptor mosaic

Page 130: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

110

imaging reported by Y. Geng et al. [4], and P. Zhang et al. [19] using a confocal AO SLO

system and the RPE mosaic was imaged by Palczewska et al. and Stremplewski et al.

[12,81,123] using two-photon excited fluorescence. Another approach uses a split-

aperture detector that can image the photoreceptor mosaic and horizontal cell bodies of

the outer plexiform layer, demonstrated by A. Guevara-Torres et al. [130]. Other non-

confocal regimes are also possible in order to favor multiply scattered light and reject the

dominating back-scattered light. These methods have been demonstrated to enable the

visualization of transparent cells without fluorescent contrast agents.

A split-detector system or other non-confocal configuration replaces the confocal

aperture in an SLO with an annulus, a filament, or a knife edge. For the split-detector

configuration, the PSF is divided into two detectors with a knife edge in the detection path.

The intensity received in each detector is subtracted to reveal the asymmetric intensity

variations that arise in the PSF from the multiply scattered light.

Combining SAO with non-confocal SLO is an interesting prospect. However, the

image quality metric may require the center portion of the aperture in order to optimize the

image. The simplest solution is to allow the center of the aperture to transmit to one

detector and an outer annulus could be reflected to another device. Then, the confocal

channel could be used for sensorless AO methods. Other approaches could use a

configurable aperture, such as a spatial light modulator (SLM) or digital-micro-device

(DMD) in the detection path. This has been demonstrated by S. A. Burns et al. [142] with

an AO SLO system for enhanced retinal vasculature imaging. There are also the

possibilities to take advantage of multiple detectors for improving the speed and

performance of sensorless AO, which has been demonstrated by Pozzi et al. [92].

8.3. Extensions of two-photon excited florescence technology

Future technology developments of the two-photon imaging capabilities are

important because of the potential for the modality to provide functional insight to the

retina. Imaging with near infrared light is preferred since the retina is less sensitive to the

longer wavelengths.

Page 131: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

111

Genetically encoded calcium indicators, such as GCaMP, can be expressed in

retinal ganglion cells of mice. The fluorescence intensity from a labelled ganglion cell will

increase when the cell is activated by an electrochemical signal originating from stimulated

photoreceptors. Some studies have demonstrated that the activity of the cell expressing

a calcium indicator can be optically measured [64,65,136,143–145]. Imaging the calcium

indicators with infrared light would be advantageous, since the photoreceptors are less

likely to be activated by the imaging light. This could enable novel studies of the signal

pathways via the interconnected neural cells in the retina by the response to light in

different areas of retina [64,65].

For imaging autofluorescence of the retina, often the source of the fluorescence

intensity is a result of many fluorophores with overlapping emission spectra. A technique

known as fluorescence lifetime imaging microscopy (FLIM) has been combined with

ophthalmoscopy to produces contrast from the fluorescence decay rate, which is

characteristic of the fluorophore [146–148]. This enables fluorescence imaging to be

mostly independent of the concentration of the fluorophores, as well as the ability to

distinguish emission from different fluorophores are mixed together. Furthermore, FLIM

has been demonstrated to provide information about the health of the cells. Recently, J.

A. Feeks et al. [149] demonstrated two-photon excited FLIM with AO in the mouse retina.

Using two-photon excited FLIM with cellular resolution provides an opportunity to

distinguish intrinsic fluorophores that are important for visual function.

8.4. Conclusion

Imaging the mouse retina with optimized resolution presents challenges that

require the development of specialized equipment. This is a worthwhile endeavour since

the mouse model is used ubiquitously in biology and health research. This thesis has

presented novel imaging systems (with an emphasis on fluorescence detection) that are

capable of high-resolution imaging for mouse retina imaging using the flexibility provided

by Sensorless Adaptive Optics (SAO) for aberration correction.

The SAO optimization algorithm developed throughout this thesis used a multi-

iteration hill-climbing coordinate search algorithm, with a decreasing search range to

enable diffraction-limited imaging. Furthermore, pupil segmentation methods for image-

based adaptive optics were demonstrated to be feasible for imaging the retina.

Page 132: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

112

The work presented in this thesis has resulted in a compact multi-modal imaging

design, which included Optical Coherence Tomography (OCT), OCT-Angiography,

Scanning Laser Ophthalmoscopy (SLO), and fluorescence detection. The imaging

capability of these imaging modalities were demonstrated by high resolution fluorescence

and structural imaging of the mouse retina. Also based on the results in Chapter 7, it

seems straightforward to use the same compact design for Two-Photon Excited

Fluorescence and OCT imaging with a few modifications to the detection scheme. The

flexibility of the SAO has been demonstrated by a range of imaging scenarios that may be

desired by a vision scientist, including imaging the structures in the different layers of the

retina with OCT and SLO, fluorescein angiography, GFP labelled cells, and

autofluorescence of the RPE mosaic.

Page 133: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

113

References

1. "Blindness In Canada," https://cnib.ca/en/sight-loss-info/blindness/blindness-canada?region=bc. Accessed Feb. 2019.

2. D. R. Williams, "Imaging single cells in the living retina," Vision Res. 51(13), 1379–1396 (2011).

3. S. Marcos, J. S. Werner, S. A. Burns, W. H. Merigan, P. Artal, D. A. Atchison, K. M. Hampson, R. Legras, L. Lundstrom, G. Yoon, J. Carroll, S. S. Choi, N. Doble, A. M. Dubis, A. Dubra, A. Elsner, R. Jonnal, D. T. Miller, M. Paques, H. E. Smithson, L. K. Young, Y. Zhang, M. Campbell, J. Hunter, A. Metha, G. Palczewska, J. Schallek, and L. C. Sincich, "Vision science and adaptive optics, the state of the field," Vision Res. 132, 3–33 (2017).

4. Y. Geng, A. Dubra, L. Yin, W. H. Merigan, R. Sharma, R. T. Libby, and D. R. Williams, "Adaptive optics retinal imaging in the living mouse eye," Biomed. Opt. Express 3(4), 715–734 (2012).

5. Y. Geng, L. A. Schery, R. Sharma, A. Dubra, K. Ahmad, R. T. Libby, and D. R. Williams, "Optical properties of the mouse eye," Biomed. Opt. Express 2(4), 717–738 (2011).

6. M. Land, "Photoreception," Encycl. Br. (2018). https://www.britannica.com/science/photoreception. Accessed Feb. 2019.

7. L. Li, N. Eter, and P. Heiduschka, "The microglia in healthy and diseased retina," Exp. Eye Res. 136, 116–130 (2015).

8. "Webvision – The Organization of the Retina and Visual System," https://webvision.med.utah.edu/. Accessed Feb. 2019.

9. P. D. Kiser, M. Golczak, and K. Palczewski, "Chemistry of the retinoid (visual) cycle.," Chem. Rev. 114(1), 194–232 (2014).

10. R. Sharma, D. R. Williams, G. Palczewska, K. Palczewski, and J. J. Hunter, "Two-Photon Autofluorescence Imaging Reveals Cellular Structures Throughout the Retina of the Living Primate Eye.," Invest. Ophthalmol. Vis. Sci. 57(2), 632–646 (2016).

11. J. J. Hunter, B. Masella, A. Dubra, R. Sharma, L. Yin, W. H. Merigan, G. Palczewska, K. Palczewski, and D. R. Williams, "Images of photoreceptors in living primate eyes using adaptive optics two-photon ophthalmoscopy," Biomed. Opt. Express 2(1), 139–148 (2011).

12. G. Palczewska, P. Stremplewski, S. Suh, N. Alexander, D. Salom, Z. Dong, D. Ruminski, E. H. Choi, A. E. Sears, T. S. Kern, M. Wojtkowski, and K. Palczewski, "Two-photon imaging of the mammalian retina with ultrafast pulsing laser.," JCI insight 3(17), 121555 (2018).

13. J. Schallek, Y. Geng, H. Nguyen, and D. R. Williams, "Morphology and topography of retinal pericytes in the living mouse retina using in vivo adaptive optics imaging and ex vivo characterization," Invest. Ophthalmol. Vis. Sci. 54(13), 8237–8250 (2013).

14. R. J. Zawadzki, P. Zhang, A. Zam, E. B. Miller, M. Goswami, X. Wang, R. S.

Page 134: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

114

Jonnal, S. H. Lee, D. Y. Kim, J. G. Flannery, J. S. Werner, M. E. Burns, and E. N. Pugh Jr., "Adaptive-optics SLO imaging combined with widefield OCT and SLO enables precise 3D localization of fluorescent cells in the mouse retina," Biomed. Opt. Express 6(6), 2191–2210 (2015).

15. D. J. Wahl, Y. Jian, S. Bonora, R. J. Zawadzki, and M. V. Sarunic, "Wavefront sensorless adaptive optics fluorescence biomicroscope for in vivo retinal imaging in mice," Biomed. Opt. Express 7(1), 1–12 (2016).

16. D. J. Wahl, R. Ng, M. J. Ju, Y. Jian, and M. V. Sarunic, "Sensorless adaptive optics multimodal en-face small animal retinal imaging," Biomed. Opt. Express 10(1), 252–267 (2019).

17. L. L. Molday, D. Wahl, M. V Sarunic, and R. S. Molday, "Localization and functional characterization of the p.Asn965Ser (N965S) ABCA4 variant in mice reveal pathogenic mechanisms underlying Stargardt macular degeneration," Hum. Mol. Genet. 27(2), 295–306 (2018).

18. D. J. Wahl, P. Zhang, J. Mocci, M. Quintavalla, R. Muradore, Y. Jian, S. Bonora, M. V. Sarunic, and R. J. Zawadzki, "Adaptive optics in the mouse eye: wavefront sensing based vs. image-guided aberration correction," Biomed. Opt. Express (In Production) (2019).

19. P. Zhang, J. Mocci, D. J. Wahl, R. K. Meleppat, S. K. Manna, M. Quintavalla, R. Muradore, M. V. Sarunic, S. Bonora, E. N. Pugh, and R. J. Zawadzki, "Effect of a contact lens on mouse retinal in vivo imaging: Effective focal length changes and monochromatic aberrations," 172, 86–93 (2018).

20. D. J. Wahl, C. Huang, S. Bonora, Y. Jian, and M. V. Sarunic, "Pupil segmentation adaptive optics for invivo mouse retinal fluorescence imaging," Opt. Lett. 42(7), 1365–1368 (2017).

21. H. R. G. W. Verstraete, M. Heisler, M. J. Ju, D. Wahl, L. Bliek, J. Kalkman, S. Bonora, Y. Jian, M. Verhaegen, and M. V Sarunic, "Wavefront sensorless adaptive optics OCT with the DONE algorithm for in vivo human retinal imaging [Invited]," Biomed. Opt. Express 8(4), 2261–2275 (2017).

22. M. J. Ju, M. Heisler, D. Wahl, Y. Jian, and M. V. Sarunic, "Multiscale sensorless adaptive optics OCT angiography system for in vivo human retinal imaging," J. Biomed. Opt. 22(12), 1 (2017).

23. M. J. Ju, C. Huang, D. J. Wahl, Y. Jian, and M. V. Sarunic, "Visible light sensorless adaptive optics for retinal structure and fluorescence imaging," Opt. Lett. 43(20), 5162–5165 (2018).

24. M. Cua, D. J. Wahl, Y. Zhao, S. Lee, S. Bonora, R. J. Zawadzki, Y. Jian, and M. V. Sarunic, "Coherence-Gated Sensorless Adaptive Optics Multiphoton Retinal Imaging," Sci. Rep. 6(1), 32223 (2016).

25. D. J. Wahl, M. J. Ju, Y. Jian, and M. V. Sarunic, "Non-invasive cellular-resolution retinal imaging with two-photon excited fluorescence," Biomed. Opt. Express (In Production) (2019).

26. R. H. Webb, "Confocal optical microscopy," Reports Prog. Phys. 59(3), 427–471 (1996).

27. J. B. Pawley, ed., Handbook Of Biological Confocal Microscopy (Springer US, 2006).

Page 135: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

115

28. T. Wilson1 and A. R. Carlini, "Size of the detector in confocal imaging systems," Opt. Lett. 12(4), 227–229 (1987).

29. J. F. de Boer, C. K. Hitzenberger, and Y. Yasuno, "Polarization sensitive optical coherence tomography – a review [Invited]," Biomed. Opt. Express 8(3), 1838–1873 (2017).

30. Y. Jia, O. Tan, J. Tokayer, B. Potsaid, Y. Wang, J. J. Liu, M. F. Kraus, H. Subhash, J. G. Fujimoto, J. Hornegger, and D. Huang, "Split-spectrum amplitude-decorrelation angiography with optical coherence tomography," Opt. Express 20(4), 4710–4725 (2012).

31. I. Gorczynska, J. V. Migacz, R. J. Zawadzki, A. G. Capps, and J. S. Werner, "Comparison of amplitude-decorrelation, speckle-variance and phase-variance OCT angiography methods for imaging the human retina and choroid," Biomed. Opt. Express 7(3), 911–942 (2016).

32. A. Zhang, Q. Zhang, C.-L. Chen, and R. K. Wang, "Methods and algorithms for optical coherence tomography-based angiography: a review and comparison.," J. Biomed. Opt. 20(10), 100901 (2015).

33. M. S. Mahmud, D. W. Cadotte, B. Vuong, C. Sun, T. W. H. Luk, A. Mariampillai, and V. X. D. Yang, "Review of speckle and phase variance optical coherence tomography to visualize microvascular networks," J. Biomed. Opt. 18(5), 050901 (2013).

34. T. Liu, H. Jung, J. Liu, M. Droettboom, and J. Tam, "Noninvasive near infrared autofluorescence imaging of retinal pigment epithelial cells in the human retina using adaptive optics," Biomed. Opt. Express 8(10), 4348 (2017).

35. S. Schmitz-Valckenberg, D. Lara, S. Nizari, E. M. Normando, L. Guo, A. R. Wegener, A. Tufail, F. W. Fitzke, F. G. Holz, and M. F. Cordeiro, "Localisation and significance of in vivo near-infrared autofluorescent signal in retinal imaging.," Br. J. Ophthalmol. 95(8), 1134–1139 (2011).

36. U. Kellner, S. Kellner, B. H. F. Weber, B. Fiebig, S. Weinitz, and K. Ruether, "Lipofuscin- and melanin-related fundus autofluorescence visualize different retinal pigment epithelial alterations in patients with retinitis pigmentosa," Eye 23(6), 1349–1359 (2009).

37. F. Helmchen and W. Denk, "Deep tissue two-photon microscopy.," Nat. Methods 2(12), 932–940 (2005).

38. W. Denk, J. Strickler, and W. Webb, "Two-photon laser scanning fluorescence microscopy," Science 248(4951), 73–76 (1990).

39. J. Porter, H. Queener, J. Lin, K. Thorn, and A. Awwal, Adaptive Optics for Vision Science: Principles, Practices, Design and Applications (Wiley, 2006).

40. A. Roorda, F. Romero-Borja, W. Donnelly, III, H. Queener, T. Hebert, and M. Campbell, "Adaptive optics scanning laser ophthalmoscopy," Opt. Express 10(9), 405 (2002).

41. R. J. Zawadzki, S. M. Jones, S. S. Olivier, M. Zhao, B. A. Bower, J. A. Izatt, S. Choi, S. Laut, and J. S. Werner, "Adaptive-optics optical coherence tomography for high-resolution and high-speed 3D retinal in vivo imaging," Opt. Express 13(21), 8532–8546 (2005).

Page 136: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

116

42. D. T. Miller, J. Qu, R. S. Jonnal, and K. E. Thorn, "Coherence gating and adaptive optics in the eye," in V. V. Tuchin, J. A. Izatt, and J. G. Fujimoto, eds. (International Society for Optics and Photonics, 2003), 4956, pp. 65–72.

43. D. G. Voelz, Computational Fourier Optics: A MATLAB Tutorial (SPIE, 2011).

44. D. E. Milkie, E. Betzig, and N. Ji, "Pupil-segmentation-based adaptive optical microscopy with full-pupil illumination.," Opt. Lett. 36(21), 4206–4208 (2011).

45. N. Ji, D. E. Milkie, and E. Betzig, "Adaptive optics via pupil segmentation for high-resolution imaging in biological tissues.," Nat. Methods 7(2), 141–147 (2010).

46. C. Wang and N. Ji, "Pupil-segmentation-based adaptive optical correction of a high-numerical-aperture gradient refractive index lens for two-photon fluorescence endoscopy," Opt. Lett. 37(11), 2001 (2012).

47. B. C. Chauhan, K. T. Stevens, J. M. Levesque, A. C. Nuschke, G. P. Sharpe, N. O’Leary, M. L. Archibald, and X. Wang, "Longitudinal in vivo imaging of retinal ganglion cells and retinal thickness changes following optic nerve injury in mice," PLoS One 7(6), e40352 (2012).

48. L. Chen, P. Yang, and A. Kijlstra, "Distribution, markers, and functions of retinal microglia," Ocul. Immunol. Inflamm. 10(1), 27–39 (2002).

49. P. Zhang, A. Zam, Y. Jian, X. Wang, M. E. Burns, M. V. Sarunic, E. N. Pugh, and R. J. Zawadzki, "Multispectral scanning laser ophthalmoscopy combined with optical coherence tomography for simultaneous in vivo mouse retinal imaging," Proc. SPIE 9307, 93070H (2015).

50. C. Alt, J. M. Runnels, L. J. Mortensen, W. Zaher, and C. P. Lin, "In vivo imaging of microglia turnover in the mouse retina after ionizing radiation and dexamethasone treatment," Invest. Ophthalmol. Vis. Sci. 55(8), 5314–5319 (2014).

51. A. Roorda, "Adaptive optics for studying visual function: a comprehensive review," J. Vis. 11(5), 6 (2011).

52. R. Tyson, Principles of Adaptive Optics (CRC Press, 2010).

53. J. A. Kubby, Adaptive Optics for Biological Imaging (CRC Press, 2013).

54. Y. N. Sulai and A. Dubra, "Non-common path aberration correction in an adaptive optics scanning ophthalmoscope," Biomed. Opt. Express 5(9), 3059–3073 (2014).

55. J. W. Evans, R. J. Zawadzki, S. M. Jones, S. S. Olivier, and J. S. Werner, "Error budget analysis for an adaptive optics optical coherence tomography system," Opt. Express 17(16), 13768–13784 (2009).

56. F. Felberer, J. S. Kroisamer, C. K. Hitzenberger, and M. Pircher, "Lens based adaptive optics scanning laser ophthalmoscope," Opt. Express 20(16), 17297–17310 (2012).

57. A. Dubra and Y. Sulai, "Reflective afocal broadband adaptive optics scanning ophthalmoscope," Biomed. Opt. Express 2(6), 1757–1768 (2011).

58. M. J. Booth, "Adaptive optics in microscopy," Philos. Trans. A. Math. Phys. Eng. Sci. 365(1861), 2829–2843 (2007).

59. Y. Jian, R. J. Zawadzki, and M. V Sarunic, "Adaptive optics optical coherence tomography for in vivo mouse retinal imaging," J. Biomed. Opt. 18(5), 56007 (2013).

Page 137: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

117

60. K. S. K. Wong, Y. Jian, E. M. C. Cua, J. Xu, S. Bonora, R. J. Zawadzki, M. V. Sarunic, Y. Jian, M. Cua, S. Bonora, R. J. Zawadzki, and M. V. Sarunic, "In vivo imaging of human photoreceptor mosaic with wavefront sensorless adaptive optics optical coherence tomography," Biomed. Opt. Express 6(2), 580–590 (2015).

61. Y. Jian, J. Xu, M. A. Gradowski, S. Bonora, R. J. Zawadzki, and M. V. Sarunic, "Wavefront sensorless adaptive optics optical coherence tomography for in vivo retinal imaging in mice," Biomed. Opt. Express 5(2), 547–559 (2014).

62. D. P. Biss, D. Sumorok, S. A. Burns, R. H. Webb, Y. Zhou, T. G. Bifano, D. Côté, I. Veilleux, P. Zamiri, and C. P. Lin, "In vivo fluorescent imaging of the mouse retina using adaptive optics," Opt. Lett. 32(6), 659–661 (2007).

63. C. Alt, D. P. Biss, N. Tajouri, T. C. Jakobs, and C. P. Lin, "An adaptive-optics scanning laser ophthalmoscope for imaging murine retinal microstructure," Proc. SPIE 7550, 755019 (2010).

64. L. Yin, Y. Geng, F. Osakada, R. Sharma, A. H. Cetin, E. M. Callaway, D. R. Williams, and W. H. Merigan, "Imaging light responses of retinal ganglion cells in the living mouse eye," J. Neurophysiol. 109(9), 2415–2421 (2013).

65. R. Sharma, L. Yin, Y. Geng, W. H. Merigan, G. Palczewska, K. Palczewski, D. R. Williams, and J. J. Hunter, "In vivo two-photon imaging of the mouse retina," Biomed. Opt. Express 4(8), 1285–1293 (2013).

66. S. A. Burns, R. Tumbar, A. E. Elsner, D. Ferguson, and D. X. Hammer, "Large-field-of-view, modular, stabilized, adaptive-optics-based scanning laser ophthalmoscope," J. Opt. Soc. Am. A 24(5), 1313–1326 (2007).

67. M. Martínez-Corral, A. Doblas, E. Sánchez-Ortiga, J. Sola-Pikabea, and G. Saavedr, "Static real-time capture of 3D microscopy images," SPIE Newsroom (2015).

68. L. N. Thibos, R. A. Applegate, J. T. Schwiegerling, R. Webb, and VSIA Standards Taskforce Members, "Standards for reporting the optical aberrations of eyes," J. Refract. Surg. 18(5), 652–660 (2002).

69. A. Myronenko and X. Song, "Intensity-based image registration by minimizing residual complexity," IEEE Trans. Med. Imaging 29(11), 1882–1891 (2010).

70. A. Dubra and Z. Harvey, "Registration of 2D Images from Fast Scanning Ophthalmic Instruments," in Biomedical Image Registration, Vol. 6204 of Lecture Notes in Computer Science (Springer, 2010), pp. 60–71.

71. C. K. Sheehy, Q. Yang, D. W. Arathorn, P. Tiruveedhula, J. F. de Boer, and A. Roorda, "High-speed, image-based eye tracking with a scanning laser ophthalmoscope," Biomed. Opt. Express 3(10), 2611–2622 (2012).

72. O. P. Kocaoglu, R. D. Ferguson, R. S. Jonnal, Z. Liu, Q. Wang, D. X. Hammer, and D. T. Miller, "Adaptive optics optical coherence tomography with dynamic retinal tracking," Biomed. Opt. Express 5(7), 2262–2284 (2014).

73. W. Sun, N. Li, and S. He, "Large-scale morphological survey of mouse retinal ganglion cells," J. Comp. Neurol. 451(2), 115–126 (2002).

74. J. Coombs, D. van der List, G. Y. Wang, and L. M. Chalupa, "Morphological properties of mouse retinal ganglion cells," Neuroscience 140(1), 123–136 (2006).

Page 138: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

118

75. W. Zou, X. Qi, and S. A. Burns, "Wavefront-aberration sorting and correction for a dual-deformable-mirror adaptive-optics system," Opt. Lett. 33(22), 2602–2604 (2008).

76. C. Li, N. Sredar, K. M. Ivers, H. Queener, and J. Porter, "A correction algorithm to simultaneously control dual deformable mirrors in a woofer-tweeter adaptive optics system," Opt. Express 18(16), 16671–16684 (2010).

77. S. M. Jones, S. Olivier, D. Chen, S. Joeres, S. Sadda, R. J. Zawadzki, J. S. Werner, and D. T. Miller, "Adaptive optics ophthalmologic systems using dual deformable mirrors," Proc. SPIE 6467, 64670H (2007).

78. X. Zhou, P. Bedggood, and A. Metha, "Limitations to adaptive optics image quality in rodent eyes," Biomed. Opt. Express 3(8), 1811–1824 (2012).

79. S. Zommer, E. N. Ribak, S. G. Lipson, and J. Adler, "Simulated annealing in ocular adaptive optics," Opt. Lett. 31(7), 939–941 (2006).

80. H. Hofer, N. Sredar, H. Queener, C. Li, and J. Porter, "Wavefront sensorless adaptive optics ophthalmoscopy in the human eye," Opt. Express 19(15), 14160–14171 (2011).

81. G. Palczewska, Z. Dong, M. Golczak, J. J. Hunter, D. R. Williams, N. S. Alexander, and K. Palczewski, "Noninvasive two-photon microscopy imaging of mouse retina and retinal pigment epithelium through the pupil of the eye," Nat. Med. 20(7), 785–789 (2014).

82. Y. Geng, K. P. Greenberg, R. Wolfe, D. C. Gray, J. J. Hunter, A. Dubra, J. G. Flannery, D. R. Williams, and J. Porter, "In vivo imaging of microscopic structures in the rat retina," Invest. Ophthalmol. Vis. Sci. 50(12), 5872–5879 (2009).

83. S. Bonora, Y. Jian, P. Zhang, A. Zam, E. N. Pugh, R. J. Zawadzki, and M. V. Sarunic, "Wavefront correction and high-resolution in vivo OCT imaging with an objective integrated multi-actuator adaptive lens," Biomed. Opt. Express 23(17), 21931–21941 (2015).

84. A. Jesacher and M. J. Booth, "Sensorless Adaptive Optics for Microscopy," in Adaptive Optics for Biological Imaging, S. S. Olivier, T. G. Bifano, and J. A. Kubby, eds. (2013), pp. 177–190.

85. X. Zhou, P. Bedggood, B. Bui, C. T. O. Nguyen, Z. He, and A. Metha, "Contrast-based sensorless adaptive optics for retinal imaging," Biomed. Opt. Express 6(9), 3577–3595 (2015).

86. M. J. Booth, "Adaptive optical microscopy: the ongoing quest for a perfect image," Light Sci. Appl. 3(4), 165 (2014).

87. R. Navarro and M. A. Losada, "Aberrations and relative efficiency of light pencils in the living human eye," Optom. Vis. Sci. 74(7), 540–547 (1997).

88. C. Wang, R. Liu, D. E. Milkie, W. Sun, Z. Tan, A. Kerlin, T.-W. Chen, D. S. Kim, and N. Ji, "Multiplexed aberration measurement for deep tissue imaging in vivo.," Nat. Methods 11(10), 1037–40 (2014).

89. Y. Jian, K. Wong, and M. V Sarunic, "Graphics processing unit accelerated optical coherence tomography processing at megahertz axial scan rate and high resolution video rate volumetric rendering," J. Biomed. Opt. 18(2), 26002 (2013).

90. K. Wang, W. Sun, C. T. Richie, B. K. Harvey, E. Betzig, and N. Ji, "Direct

Page 139: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

119

wavefront sensing for high-resolution in vivo imaging in scattering tissue," Nat. Commun. 6(1), 7276 (2015).

91. J. Mocci, M. Quintavalla, C. Trestino, S. Bonora, and R. Muradore, "A multi-platform CPU-based architecture for cost-effective adaptive optics systems," IEEE Trans. Ind. Informatics 14(10), 4431–4439 (2018).

92. P. Pozzi, D. Wilding, O. Soloviev, H. Verstraete, L. Bliek, G. Vdovin, and M. Verhaegen, "High speed wavefront sensorless aberration correction in digital micromirror based confocal microscopy," Opt. Express 25(2), 949–959 (2017).

93. S. Bonora and R. J. Zawadzki, "Wavefront sensorless modal deformable mirror correction in adaptive optics: optical coherence tomography.," Opt. Lett. 38(22), 4801–4804 (2013).

94. R. A. Muller and A. Buffington, "Real-time correction of atmospherically degraded telescope images through image sharpening," J. Opt. Soc. Am. 64(9), 1200–1210 (1974).

95. D. Debarre, M. J. Booth, and T. Wilson, "Image based adaptive optics through optimisation of low spatial frequencies," Opt. Express 15(13), 8176–8190 (2007).

96. M. Kasper, E. Fedrigo, D. P. Looze, H. Bonnet, L. Ivanescu, and S. Oberti, "Fast calibration of high-order adaptive optics systems.," J. Opt. Soc. Am. A. Opt. Image Sci. Vis. 21(6), 1004–1008 (2004).

97. N. S. Alexander, G. Palczewska, P. Stremplewski, M. Wojtkowski, T. S. Kern, and K. Palczewski, "Image registration and averaging of low laser power two-photon fluorescence images of mouse retina," Biomed. Opt. Express 7(7), 2671–2691 (2016).

98. B. Wang and M. J. Booth, "Optimum deformable mirror modes for sensorless adaptive optics," Opt. Commun. 282(23), 4467–4474 (2009).

99. T. DuBose, D. Nankivil, F. LaRocca, G. Waterman, K. Hagan, J. Polans, B. Keller, D. Tran-Viet, L. Vajzovic, A. N. Kuo, C. A. Toth, J. A. Izatt, and S. Farsiu, "Handheld adaptive optics scanning laser ophthalmoscope," Optica 5(9), 1027–1036 (2018).

100. M. Booth, "Wave front sensor-less adaptive optics: a model-based approach using sphere packings," Opt. Express 14(4), 1339–1352 (2006).

101. H. R. G. W. Verstraete, S. Wahls, J. Kalkman, and M. Verhaegen, "Model-based sensor-less wavefront aberration correction in optical coherence tomography," Opt. Lett. 40(24), 5722–5725 (2015).

102. N. Ji, "Adaptive optical fluorescence microscopy," Nat. Methods 14(4), 374–380 (2017).

103. J. Polans, B. Keller, O. M. Carrasco-Zevallos, F. LaRocca, E. Cole, H. E. Whitson, E. M. Lad, S. Farsiu, and J. A. Izatt, "Wide-field retinal optical coherence tomography with wavefront sensorless adaptive optics for enhanced imaging of targeted regions.," Biomed. Opt. Express 8(1), 16–37 (2017).

104. A. Facomprez, E. Beaurepaire, and D. Débarre, "Accuracy of correction in modal sensorless adaptive optics," Opt. Express 20(3), 2598–2612 (2012).

105. C. Bourgenot, C. D. Saunter, G. D. Love, and J. M. Girkin, "Comparison of closed loop and sensorless adaptive optics in widefield optical microscopy," J. Eur. Opt.

Page 140: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

120

Soc. Rapid Publ. 8, 13027 (2013).

106. P. Zhang, M. Goswami, R. J. Zawadzki, E. N. Pugh, and Jr, "The Photosensitivity of Rhodopsin Bleaching and Light-Induced Increases of Fundus Reflectance in Mice Measured In Vivo With Scanning Laser Ophthalmoscopy.," Invest. Ophthalmol. Vis. Sci. 57(8), 3650–3664 (2016).

107. G. Huber, S. C. Beck, C. Grimm, A. Sahaboglu-Tekgoz, F. Paquet-Durand, A. Wenzel, P. Humphries, T. M. Redmond, M. W. Seeliger, and M. D. Fischer, "Spectral Domain Optical Coherence Tomography in Mouse Models of Retinal Degeneration," Investig. Opthalmology Vis. Sci. 50(12), 5888–5895 (2009).

108. X. Tao, O. Azucena, M. Fu, Y. Zuo, D. C. Chen, and J. Kubby, "Adaptive optics microscopy with direct wavefront sensing using fluorescent protein guide stars," Opt. Lett. 36(17), 3389–3391 (2011).

109. M. Rueckel, J. A. Mack-Bucher, and W. Denk, "Adaptive wavefront correction in two-photon microscopy using coherence-gated wavefront sensing.," Proc. Natl. Acad. Sci. U. S. A. 103(46), 17137–17142 (2006).

110. J. B. Schallek and A. Joseph, "Time-lapse imaging of retinal microglia in vivo show dynamic process motility at rest," Invest. Ophthalmol. Vis. Sci. 58(8), 316 (2017).

111. M. Pircher and R. J. Zawadzki, "Review of adaptive optics OCT (AO-OCT): principles and applications for retinal imaging [Invited]," Biomed. Opt. Express 8(5), 2536–2562 (2017).

112. J. Xu, K. Wong, Y. Jian, and M. V. Sarunic, "Real-time acquisition and display of flow contrast using speckle variance optical coherence tomography in a graphics processing unit," J. Biomed. Opt. 19(2), 26001 (2014).

113. J. Xu, S. Han, C. Balaratnasingam, Z. Mammo, K. S. K. Wong, S. Lee, M. Cua, M. Young, A. Kirker, D. Albiani, F. Forooghian, P. Mackenzie, A. Merkur, D.-Y. Yu, and M. V Sarunic, "Retinal angiography with real-time speckle variance optical coherence tomography," Br. J. Ophthalmol. 99(10), 1315–1319 (2015).

114. A. Negrean and H. D. Mansvelder, "Optimal lens design and use in laser-scanning microscopy," Biomed. Opt. Express 5(5), 1588–1609 (2014).

115. A. Nagler, "Plossl type eyepiece for use in astronomical instruments," U.S. patent 4482217 (February 18, 1983).

116. A. Nimmerjahn, F. Kirchhoff, and F. Helmchen, "Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo," Science 308(5726), 1314–1318 (2005).

117. M. V Sarunic, A. Yazdanpanah, E. Gibson, J. Xu, Y. Bai, S. Lee, H. U. Saragovi, and M. F. Beg, "Longitudinal study of retinal degeneration in a rat using spectral domain optical coherence tomography," Opt. Express 18(22), 23435–23441 (2010).

118. F. C. Delori, R. H. Webb, and D. H. Sliney, "Maximum permissible exposures for ocular safety (ANSI 2000), with emphasis on ophthalmic devices," J. Opt. Soc. Am. A 24(5), 1250–1265 (2007).

119. D. W. Arathorn, Q. Yang, C. R. Vogel, Y. Zhang, P. Tiruveedhula, and A. Roorda, "Retinally stabilized cone-targeted stimulus delivery," Opt. Express 15(21),

Page 141: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

121

13731–13744 (2007).

120. D. J. Wahl, P. Zhang, Y. Jian, S. Bonora, R. J. Zawadzki, and M. V. Sarunic, "Image-based adaptive optics compared to wavefront sensing methods for retinal imaging," Proc. SPIE 10502, 105020G (2018).

121. C. Schwarz, R. Sharma, W. S. Fischer, M. Chung, G. Palczewska, K. Palczewski, D. R. Williams, and J. J. Hunter, "Safety assessment in macaques of light exposures for functional two-photon ophthalmoscopy in humans.," Biomed. Opt. Express 7(12), 5148–5169 (2016).

122. D. Merino and P. Loza-Alvarez, "Adaptive optics scanning laser ophthalmoscope imaging: technology update," Clin. Ophthalmol. 10, 743–755 (2016).

123. P. Stremplewski, K. Komar, K. Palczewski, M. Wojtkowski, and G. Palczewska, "Periscope for noninvasive two-photon imaging of murine retina in vivo.," Biomed. Opt. Express 6(9), 3352–3361 (2015).

124. Y. Imanishi, M. L. Batten, D. W. Piston, W. Baehr, and K. Palczewski, "Noninvasive two-photon imaging reveals retinyl ester storage structures in the eye.," J. Cell Biol. 164(3), 373–383 (2004).

125. W. Drexler and J. G. Fujimoto, Optical Coherence Tomography: Technology and Applications, 2nd ed. (Springer, 2015).

126. M. Heisler, S. Lee, Z. Mammo, Y. Jian, M. Ju, A. Merkur, E. Navajas, C. Balaratnasingam, M. F. Beg, and M. V. Sarunic, "Strip-based registration of serially acquired optical coherence tomography angiography," J. Biomed. Opt. 22(3), 036007 (2017).

127. D. C. Adler, T. H. Ko, and J. G. Fujimoto, "Speckle reduction in optical coherence tomography images by use of a spatially adaptive wavelet filter," Opt. Lett. 29(24), 2878–2880 (2004).

128. M. Salas, M. Augustin, L. Ginner, A. Kumar, B. Baumann, R. Leitgeb, W. Drexler, S. Prager, J. Hafner, U. Schmidt-Erfurth, and M. Pircher, "Visualization of micro-capillaries using optical coherence tomography angiography with and without adaptive optics," Biomed. Opt. Express 8(1), 207–222 (2017).

129. M. T. Bernucci, C. W. Merkle, and V. J. Srinivasan, "Investigation of artifacts in retinal and choroidal OCT angiography with a contrast agent," Biomed. Opt. Express 9(3), 1020–1040 (2018).

130. A. Guevara-Torres, D. R. Williams, and J. B. Schallek, "Imaging translucent cell bodies in the living mouse retina without contrast agents," Biomed. Opt. Express 6(6), 2106–2119 (2015).

131. J. I. W. Morgan, J. J. Hunter, B. Masella, R. Wolfe, D. C. Gray, W. H. Merigan, F. C. Delori, and D. R. Williams, "Light-Induced Retinal Changes Observed with High-Resolution Autofluorescence Imaging of the Retinal Pigment Epithelium," Investig. Opthalmology Vis. Sci. 49(8), 3715–3729 (2008).

132. W. T. Ham, H. A. Mueller, J. J. Ruffolo, and A. M. Clarke, "Sensitivity of the retina to radiation damage as a function of wavelength," Photochem. Photobiol. 29(4), 735–743 (1979).

133. G. Palczewska, T. Maeda, Y. Imanishi, W. Sun, Y. Chen, D. R. Williams, D. W. Piston, A. Maeda, and K. Palczewski, "Noninvasive multiphoton fluorescence

Page 142: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

122

microscopy resolves retinol and retinal condensation products in mouse eyes," Nat. Med. 16(12), 1444–1449 (2010).

134. S. He, C. Ye, Q. Sun, C. K. S. Leung, and J. Y. Qu, "Label-free nonlinear optical imaging of mouse retina," 6(3), 1055–1066 (2015).

135. T. Kamali, J. Fischer, S. Farrell, W. H. Baldridge, G. Zinser, and B. C. Chauhan, "Simultaneous in vivo confocal reflectance and two-photon retinal ganglion cell imaging based on a hollow core fiber platform," J. Biomed. Opt. 23(9), 091405 (2018).

136. A. S. Bar-Noam, N. Farah, and S. Shoham, "Correction-free remotely scanned two-photon in vivo mouse retinal imaging," Light Sci. Appl. 5(1), 16007 (2016).

137. C. Schwarz, R. Sharma, S. K. Cheong, M. Keller, D. R. Williams, and J. J. Hunter, "Selective S Cone Damage and Retinal Remodeling Following Intense Ultrashort Pulse Laser Exposures in the Near-Infrared," Invest. Ophthalmol. Vis. Sci. 59(15), 5973–5984 (2018).

138. Z. Liu, K. Kurokawa, F. Zhang, J. J. Lee, and D. T. Miller, "Imaging and quantifying ganglion cells and other transparent neurons in the living human retina.," Proc. Natl. Acad. Sci. U. S. A. 114(48), 12803–12808 (2017).

139. Z. Liu, J. Tam, O. Saeedi, and D. X. Hammer, "Trans-retinal cellular imaging with multimodal adaptive optics.," Biomed. Opt. Express 9(9), 4246–4262 (2018).

140. P. Pozzi, O. Soloviev, D. Wilding, G. Vdovin, and M. Verhaegen, "Optimal model-based sensorless adaptive optics for epifluorescence microscopy," PLoS One 13(3), e0194523 (2018).

141. A. Thayil and M. J. Booth, "Self calibration of sensorless adaptive optical microscopes," J. Eur. Opt. Soc. 6, 11045 (2011).

142. K. A. Sapoznik, T. Luo, A. de Castro, L. Sawides, R. L. Warner, and S. A. Burns, "Enhanced retinal vasculature imaging with a rapidly configurable aperture," Biomed. Opt. Express 9(3), 1323–1333 (2018).

143. S. N. Blandford, M. L. Hooper, T. Yabana, B. C. Chauhan, W. H. Baldridge, and S. R. M. Farrell, "Retinal Characterization of the Thy1-GCaMP3 Transgenic Mouse Line After Optic Nerve Transection," Investig. Opthalmology Vis. Sci. 60(1), 183 (2019).

144. W. Wei, J. Elstrott, and M. B. Feller, "Two-photon targeted recording of GFP-expressing neurons for light responses and live-cell imaging in the mouse retina," Nat. Protoc. 5(7), 1347–1352 (2010).

145. B. G. Borghuis, L. Tian, Y. Xu, S. S. Nikonov, N. Vardi, B. V Zemelman, and L. L. Looger, "Imaging light responses of targeted neuron populations in the rodent retina.," J. Neurosci. 31(8), 2855–67 (2011).

146. D. Schweitzer, M. Hammer, F. Schweitzer, R. Anders, T. Doebbecke, S. Schenke, E. R. Gaillard, and E. R. Gaillard, "In vivo measurement of time-resolved autofluorescence at the human fundus," J. Biomed. Opt. 9(6), 1214 (2004).

147. C. Dysli, M. Dysli, V. Enzmann, S. Wolf, and M. S. Zinkernagel, "Fluorescence Lifetime Imaging of the Ocular Fundus in Mice," Investig. Opthalmology Vis. Sci. 55(11), 7206–7215 (2014).

148. C. Dysli, S. Wolf, M. Y. Berezin, L. Sauer, M. Hammer, and M. S. Zinkernagel,

Page 143: Aberration correction with sensorless adaptive optics for imaging the mouse retinasummit.sfu.ca/system/files/iritems1/19394/etd20402.pdf · 2020-05-18 · compactable system layout

123

"Fluorescence lifetime imaging ophthalmoscopy," Prog. Retin. Eye Res. 60, 120–143 (2017).

149. J. A. Feeks and J. J. Hunter, "Adaptive optics two-photon excited fluorescence lifetime imaging ophthalmoscopy of exogenous fluorophores in mice," Biomed. Opt. Express 8(5), 2483 (2017).