81
AUTHORIZATION TO LEND AND REPRODUCE THE THESIS As the sole author of this thesis, I authorize Brown University to lend it to other institutions or individuals for the purpose of scholarly research ____________ __________________________ Date Ahbid Zein-Sabatto, Author I further authorize Brown University to reproduce this thesis by photocopying or other means, in total or in part, at the request of other institutions or individuals for the purpose of scholarly research ____________ __________________________ Date Ahbid Zein-Sabatto, Author

AUTHORIZATION TO LEND AND REPRODUCE THE THESIS

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

AUTHORIZATION TO LEND AND REPRODUCE THE THESIS

As the sole author of this thesis, I authorize Brown University to lend it to other

institutions or individuals for the purpose of scholarly research

____________ __________________________

Date Ahbid Zein-Sabatto, Author

I further authorize Brown University to reproduce this thesis by photocopying or other

means, in total or in part, at the request of other institutions or individuals for the purpose

of scholarly research

____________ __________________________

Date Ahbid Zein-Sabatto, Author

Utilizing Three-Dimensional, Noninvasive, and Label-Free

OCT Imaging to Detect Cellular Viability in Neurospheroids

By

Ahbid Zein-Sabatto

B.E., Vanderbilt University, 2016

Submitted in partial fulfillment of the requirements for the degree of Master of Science in

the School of Engineering at Brown University

PROVIDENCE, RHODE ISLAND

May 2019

ii

This thesis by Ahbid Zein-Sabatto is accepted in its present form by the School of

Engineering as satisfying the thesis requirements for the degree of Master of Science.

____________ _____________________________

Date Jonghwan Lee, Ph.D., Advisor

____________ _____________________________

Date Diane Hoffman-Kim, Ph.D., Reader

____________ _____________________________

Date Celinda Kofron, Ph.D., Reader

Approved by the Graduate Council

____________ _____________________________

Date Andrew G. Campbell, Ph.D.,

Dean of the Graduate School

iii

Curriculum Vitae

Ahbid Zein-Sabatto | [email protected]

Education

Master of Science, Sc.M. Biomedical Engineering May 2019

Brown University: Providence, Rhode Island

Bachelor of Engineering, B.E. Biomedical Engineering May 2016

Vanderbilt University: Nashville, Tennessee

Research

Graduate Researcher, Biomedical Optics and Neuroengineering Aug 2017-Present

Department of Biomedical Engineering

Advisor: Dr. Jonghwan Lee in collaboration with Dr. Diane Hoffman-Kim

Brown University: Providence, Rhode Island

▪ Evaluating the use of high-throughput, noninvasive, and label-free OCT imaging

techniques to assess cellular viability of tissue engineered neurospheroids

▪ Validating the use of OCT imaging to longitudinally monitor viability in ischemic

stroke models

Methods: optical coherence tomography, Matlab, Fluor-4 calcium imaging, lactate

dehydrogenase assay, live/dead assay, confocal microscopy, immunohistochemistry

Independent Researcher, Idea2Impact Summer Research Jun 2015-Aug 2015

Department of Biomedical Engineering

Advisors: Dr. Matthew Walker III, Dr. Amanda Lowery, Dr. Christina Marasco

Vanderbilt University: Nashville, Tennessee

▪ Designed and developed a medical device that utilizes low level light therapy to

treat chronic diabetic foot ulcers

Methods: printed circuit board design, low-level-light therapy, surface mount circuit

assembly

Undergraduate Student Researcher, Bone Biomechanics Jan 2011-Aug 2013

Department of Orthopedic Surgery

Advisor: Dr. Jeffry S. Nyman

Vanderbilt University: Nashville, Tennessee

▪ Investigated the effects of age, disease, genetics, and other factors on bone

toughness

▪ Advanced the understanding of bone fractures and their predictors in high risk

patients

iv

Methods: µ-computed tomography, nuclear magnetic resonance imaging, Raman

spectroscopy, 3-point-bend-test, osteoblast cell culture, light microscopy, metallurgic

sample preparation

Publications

1. Makowski AJ, Pence IJ, Uppuganti S, Zein-Sabatto A, Huszagh MC, Mahadevan-

Jansen A, & Nyman JS. Polarization in Raman Spectroscopy Helps Explain Bone

Brittleness in Genetic Mouse Models. Journal of Biomedical Optic.

2014;19(11):117008 DOI: 10.1117/1.JBO.19.11.117008.

2. Nyman JS, Gorochow LE, Horch RA, Uppuganti S, Zein-Sabatto A, Manhard

MK, & Does MD. Partial Removal of Pore and Loosely Bound Water by Low-

Energy Drying Decreases Cortical Bone Toughness in Young and Old

Donors. Journal of the Mechanical Behavior of Biomedical Materials.

2013;22:136–145 DOI: 10.1016/j.jmbbm.2012.08.013.

3. Zein-Sabatto A, Horch RA, Murry M, & Nyman JS. The Significance of ATF-4

and TGF-β to Bound Water and Bone Fracture Resistance. Young Scientist.

2012;2:51-53.

Patents

1. Hicks A, Fross B, Mendoza J, Piper L, Schlunk S, Stedman E, Zein-Sabatto A.

Systems, Devices, and Methods for Administering Low-Level-Light Therapy. U.S.

Patent Application 2018/0304094 A1. Publication Date: Oct. 25, 2018.

Poster Presentations (‘*’ designates the presenters)

1. Fross B*, Mendoza J*, Piper L*, Schlunk S*, Stedman E*, Zein-Sabatto A*, Hicks

A. Diabetic Foot Ulcer Treatment Utilizing a Low-Level Light Therapy Device.

Poster presentation at the 2016 Vanderbilt Senior Design Day, Nashville, TN, USA;

April 2016

2. Makowski AJ*, Uppuganti S, Zein-Sabatto A, Huszagh M, Granke M, & Nyman

JS. Loss in Tissue Heterogeneity Corresponds with the Age-Related Decrease in

Fracture Toughness of Bone. Poster presentation at the 2014 annual meeting of the

Orthopedic Research Society, New Orleans, LA, USA; March 2014

3. Makowski AJ*, Uppuganti S, Zein-Sabatto A, Whitehead J, Granke M,

Mahadevan-Jansen A, & Nyman JS. Measuring Bone Heterogeneity with Raman

v

Spectroscopy to Explain Aging Differences in Human Fracture Toughness. Poster

presentation at the 2013 annual meeting of the American Society for Bone and

Mineral Research, Baltimore, MD, USA; October 2013

4. Granke M*, Uppuganti S, Makowski AJ, Zein-Sabatto A, Schultze AK,

Whitehead J, & Nyman JS. Does Reference Point Indentation Assess the Fracture

Toughness of Human Cortical Bone? Poster presentation at the 2013 annual

meeting of the American Society for Bone and Mineral Research, Baltimore, MD,

USA; October 2013

5. Zein-Sabatto A*, Huszagh M, Makowski AJ, Uppuganti S, Mahadevan-Jansen A,

& Nyman JS. Polarization Raman Spectroscopy Mapping Signifies Age-Related

Changes in Bone Fracture Resistance. Poster presentation at the 2013 VUSE

Summer Research Program symposium, Nashville, TN, USA; September 2013

6. Zein-Sabatto A*, Horch RA, Murry M, Esparza J, Rowland B, Makowski AJ, &

Nyman JS. The Effects of ATF4 and TGF-β Mediation on Bone Integrity. Poster

presentation at the 2011 School for Science and Math at Vanderbilt symposium,

Nashville, TN, USA; July 2011

7. Zein-Sabatto A*, Kumar V*, Das S*, & Konjeti R*. The Effects of Surface Area,

Distance, and Voltage on the Production of Hydrogen by Means of Short-Pulse

Hydrolysis. Presentation at the 2011 annual meeting of the Tennessee Junior

Academy of Science, Nashville, TN, USA; April 2011

Teaching and Mentorship Experience

Professional Tutor, Bridges to Belmont Jan 2017-Apr 2017

Supervisor: Dr. Nadi Bishop

Belmont University: Nashville, Tennessee

▪ Tutored undergraduate students in various subjects

▪ Assisted students in developing studying and note-taking techniques

Lab Assistant, Vanderbilt Student Volunteers for Science Aug 2012-May 2016

Supervisor: Patricia C. Tellinghuisen

Vanderbilt University: Nashville, Tennessee

▪ Maintained and replenished science experiment kits used by middle school

students in the Nashville metropolitan area and surrounding counties

▪ Conducted in-class science experiments to enrich teaching curriculum

▪ Directed and trained peers to enhance work flow

vi

Honors and Awards

Best in Healthcare Award Apr 2016

Vanderbilt University, Senior Design Day

Nashville, Tennessee

Thomas G. Arnold Prize for Biomedical Engineering System Design May 2016

Vanderbilt University, Department of Biomedical Engineering

Nashville, Tennessee

Affiliation & Leadership

Muslim Students Association: Vanderbilt University

▪ Member Aug 2012-May 2016

▪ Treasurer Aug 2013-Apr 2016

Middle Eastern Student Association: Vanderbilt University

▪ Treasurer Aug 2015-Apr 2016

Languages

▪ Arabic: fluent

▪ French: beginner proficiency

vii

Acknowledgments

I am blessed to have been surrounded by wonderful people at Brown, for they

have contributed greatly to my research and accomplishments.

First, I would like to thank Dr. Jonghwan Lee for welcoming me to his lab,

guiding me through my research, providing continuous words of encouragement, and

exemplifying all the characteristics of a great advisor.

I would like to thank Dr. Diane Hoffman-Kim for collaborating with our lab and

providing invaluable insight to my research. None of this would not have been possible

without your support. I would also like to thank you for participating in my thesis

committee.

I would like to thank Dr. Celinda Kofron for participating in my thesis committee

and for fulfilling the role of Master’s Program Director during a time of need.

Thank you to the members of the Lee Lab for your support, friendship, and advice

throughout the years. A special thanks to Dr. Julia Lee for mentoring me, teaching me

how to use the OCT imaging system, and being there when things did not go as planned.

You have shown me how great an amazing mentor can be. I would also like to extend a

special thanks to Jess Sevetson from the Hoffman-Kim lab for being the backbone of this

research. Thank you for sharing your neurospheroids, answering my many questions, and

helping me with all the IHC and calcium imaging.

Finally, thank you to my family and friends for your unconditional love and for

making me the man I am today. I am eternally grateful for having you in my life.

viii

Table of Contents

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

Chapter 1: Taking Neurons to the Third Dimension .................................................... 1

1.1 Introduction ............................................................................................................... 1

1.2 Three-Dimensional Neurospheroids ......................................................................... 2

1.3 Applications of 3D Neural Cultures ......................................................................... 3

1.4 Cellular Viability Analysis ....................................................................................... 4

1.4.1 Traditional Cellular Viability Assays ................................................................ 4

1.4.2 Challenges Associated with Traditional Methods ............................................. 6

Chapter 2: Optical Coherence Tomography Imaging and System .............................. 9

2.1 Introduction ............................................................................................................... 9

2.2 Optical Coherence Tomography for Viability Imaging .......................................... 10

2.2.1 Advantages of Utilizing Optical Coherence Tomography Imaging ................ 10

2.2.2 The Link Between Cellular Viability and Optical Coherence Tomography ... 12

2.3 Optical Coherence Tomography Parameters .......................................................... 14

2.3.1 Intensity ............................................................................................................ 14

2.3.2 Decorrelation .................................................................................................... 15

2.3.3 Diffusion Coefficient ....................................................................................... 17

2.3.4 Surface Area-to-Volume-Ratio ........................................................................ 19

2.3.5 OCT Viability Imaging in Literature ............................................................... 20

Chapter 3: Optical Coherence Microscopy Revealed Time-Dependent Changes in

Cellular Dynamics in Response to Ethanol Treatment ............................................... 23

3.1 Abstract ................................................................................................................... 23

3.2 Introduction ............................................................................................................. 24

3.3 Methods .................................................................................................................. 25

ix

3.3.1 Neurospheroid Cell Isolation ........................................................................... 25

3.3.2 Forming Three-Dimensional, Self-Aggregating Neurospheroids .................... 26

3.3.3 Calcium Imaging .............................................................................................. 27

3.3.4 Immunohistochemistry .................................................................................... 28

3.3.5 ClearT2 Optical Clearing Protocol .................................................................... 28

3.3.6 OCM Imaging System and Scanning Protocol ................................................ 29

3.3.7 OCM Data Processing ...................................................................................... 29

3.3.8 Cellular Viability Manipulation ....................................................................... 31

3.3.9 Statistical Analysis ........................................................................................... 31

3.4 Results ..................................................................................................................... 32

3.4.1 OCM Imaging Revealed Cells Through Contrast in Intracellular Motility ..... 32

3.4.2 OCM Intensity Signal Displayed Unique Time-Dependent Responses to

Ethanol Treatment ..................................................................................................... 34

3.4.3 Ethanol Disrupted Membrane Integrity Through Oncotic Necrosis ................ 36

3.5 Discussion ............................................................................................................... 37

3.6 Conclusion .............................................................................................................. 41

Chapter 4: High-throughput, Multi-Metric, and 3D Assessment of Neurospheroids

........................................................................................................................................... 42

4.1 Abstract ................................................................................................................... 42

4.2 Introduction ............................................................................................................. 42

4.3 Methods .................................................................................................................. 44

4.3.1 Neurospheroid Cell Isolation ........................................................................... 44

4.3.2 Forming Three-Dimensional, Self-Aggregating Neurospheroids .................... 45

4.3.3 SD OCT Imaging System and Scanning Protocol ........................................... 46

4.3.4 Cellular Viability Manipulation ....................................................................... 47

4.3.5 OCT Data Processing ....................................................................................... 47

x

4.3.6 Statistical Analysis ........................................................................................... 48

4.4 Results ..................................................................................................................... 49

4.4.1 Ethanol Elicited Dose and Time-Dependent Changes in OCT Intensity Signal

................................................................................................................................... 49

4.4.2 Ethanol Did Not Elicit Dose and Time-Dependent Changes in OCT

Decorrelation ............................................................................................................. 50

4.4.3 Ethanol Elicited a Slight Increase in Diffusion Coefficient ............................. 51

4.5 Discussion ............................................................................................................... 53

4.6 Conclusion .............................................................................................................. 56

References ........................................................................................................................ 57

xi

List of Figures

Figure 1. A schematic of a typical SD OCT system is shown. Light paths are represented

as arrows. The source, sample, and reference light paths are colored red, blue, and gray,

respectively. ...................................................................................................................... 10

Figure 2. (A) Transmission electron micrographs of a normal neuron, (B) an apoptotic

neuron, (C and D) and necrotic neurons. Scale bar 2.5 µm. Abbreviations: N, nucleus;

PM, plasma membrane; Mi, mitochondria. Adapted from Won et al., 2002.................... 13

Figure 3. A diagram explaining the steps and equations used to derive the intensity for a

continuous sample using SD OCT. E0 is the equation for the electric field of the source

light, R is the sample’s field reflectivity, c is the speed of light, ε is the permittivity of

free space, k is the wave number, z is the sample depth, ω is the angular frequency, σ is

the standard deviation of the center angular frequency, CC is the complex conjugate, and

ζ is the Fourier transformed wave number........................................................................ 15

Figure 4. (A) An illustration of the cross-section of a voxel (blue square) containing light

scattering particles (colored circles) before and after Δt time has passed. Black arrows

signify the direction of particle movement, and white arrows indicate the reflected light’s

path. (B) The resulting OCT signal produced by the combined interference of all

scattering particles present within the voxel (i.e. speckle). (C) The equation used to derive

decorrelation where D is the decorrelation, R is the field reflectivity, and t is time. ....... 16

Figure 5. (A) An illustration of the particles contributing to the DLS OCT signal in a

given voxel. (B) A model of the autocorrelation function on a complex plane. The

magnitude of MS and MF are proportional to the static and flowing particles, respectively.

xii

Particles with larger cross-sections contribute more to the DLS OCT signal. Adapted

from Lee et al., 2012. ........................................................................................................ 18

Figure 6. The process of acquiring and analyzing OCM data is described. (Top left) The

imaging plane, identified by the shaded blue region, was acquired along the central plane

of the neurospheroid. (Top right) A 2D OCM decorrelation map was then reconstructed

at the focal plane. (Bottom right) Multiple longitudinal images were acquired at a fixed

FOV for one hour. (Bottom left) OCM-derived viability data can then be plotted

longitudinally for each spheroid. ...................................................................................... 30

Figure 7. A visual representation of how the shell (region between the dotted red and

blue line) and core (region within the dotted blue line) are designated when analyzing

OCM data. Scale bar 100 µm. ........................................................................................... 31

Figure 8. (Left to right) IHC, calcium, decorrelation, and intensity images of

neurospheroids at 14 DIV at various depths. Cell nuclei and neuronal microtubules were

labeled using DAPI (blue) and β-III-tubulin (red), respectively. Calcium images were

acquired with OGB indicator (green). Scale bars 100 µm. ............................................... 33

Figure 9. OGB calcium (left) and OCM decorrelation (right) en face images of

neurospheroids at 9 and 14 DIV, respectively. Magnified views are designated by the red

squares............................................................................................................................... 34

Figure 10. Raw and normalized mean intensity and decorrelation parameters were

longitudinally measured in 4 M etOH treated neurospheroids (n=4; DIV: 5, 7, 19, and

28). Error bars represent the standard deviations. Either the one-way repeated measures

ANOVA with Tukey test or the Friedman test with Dunn’s post hoc tests were conducted

depending on the normality and variance of the data. Statistical results compared the 10-

xiii

minute value with all other time points and are reported in the order of core, shell, and

total regions from left to right. P < 0.05 *, < 0.01 **, and < 0.001 †. .............................. 35

Figure 11. Z-projections were acquired from neurospheroids at 9 DIV using OGB

calcium indicator after 4 M etOH treatment. Projections were created from the average of

the top 100 slices. Images A-F were taken at pretreatment, 1, 10, 30, 45, and 60 minutes

after treatment, respectively. Image G is of a 60-minute control spheroid. Images A-F are

from the same spheroid. Scale bars 100 µm. .................................................................... 36

Figure 12. Z-projections were acquired from neurospheroids at 14 DIV using the top 100

slices. Astrocytes (GFAP, gray) and cell nuclei (DAPI, blue) were stained. Images were

taken at 10 min control (A), 10 min treatment (B), 60 min control (C), and 60 min

treatment (D). Scale bars 100 µm. .................................................................................... 37

Figure 13. (Left) En face intensity image of multiple neurospheroids within a gel. The

spheroid centered in the red square appears in various orthogonal planes (center and

right). The red ellipses in the orthogonal views are fitted for each selected spheroid. ..... 48

Figure 14. Raw and normalized mean intensity were longitudinally plotted for

neurospheroids in the 4 M etOH (n spheroids = 85; n gels = 6; n batches = 3; DIV 7 and

14) 2 M etOH (n spheroids = 45; n gels = 3; n batches = 2; DIV 21), and PBS (n

spheroids = 60; n gels = 4; n batches = 2; DIV 14 and 21) treatment groups. The core,

shell, and total intensity parameters were plotted individually to visualize any regional

changes. One-way repeated measures ANOVA/Friedman test with Tukey/Dunn’s post

hoc tests were conducted to determine statistical significance (for normally or non-

normally distributed data, respectively). The statistical significance between the initial

xiv

condition and all other time points was reported in the order of core, shell, and total

region from left to right. P values < 0.05 *, < 0.01 **, and < 0.001 †.............................. 49

Figure 15. Raw and normalized mean decorrelation were longitudinally plotted for

neurospheroids in the 4 M etOH (n spheroids = 85; n gels = 6; n batches = 3; DIV 7 and

14) 2 M etOH (n spheroids = 45; n gels = 3; n batches = 2; DIV 21), and PBS (n

spheroids = 60; n gels = 4; n batches = 2; DIV 14 and 21) treatment groups. The core,

shell, and total intensity parameters were plotted individually to visualize any regional

changes. One-way repeated measures ANOVA/Friedman test with Tukey/Dunn’s post

hoc tests were conducted to determine statistical significance (for normally or non-

normally distributed data, respectively). The statistical significance between the initial

condition and all other time points were reported in the order of core, shell, and total

region from left to right. P values < 0.05 *, < 0.01 **, and < 0.001 †.............................. 51

Figure 16. Changes in diffusion coefficient (D) were plotted longitudinally for

neurospheroids exposed to 4 M etOH (n spheroids = 12; n gels = 2; n batch = 1; DIV 7),

2 M etOH (n spheroids =18; n gels = 3; n batches = 2; DIV 21), and PBS (n=18; n gels =

3; n batches = 2; DIV 14 and 21). One-way repeated measures ANOVA/Friedman test

with Tukey/Dunn’s post hoc tests were conducted to determine the statistical significance

between the initial condition and all other time points. P values were reported in the order

of core, shell, and total region from left to right. P values < 0.05 *, < 0.01 **, and < 0.001

†......................................................................................................................................... 52

1

Chapter 1: Taking Neurons to the Third Dimension

1.1 Introduction

Cell cultures have proven to be an invaluable research tool used to advance the

understanding of complex biological systems. Two-dimensional (2D) in vitro models (i.e.

cells cultured on flat petri dishes and well plates) enable researchers to investigate

various cellular mechanisms in a low-cost, controlled, and high-throughput environment

prior to initiating more rigorous studies. For example, 2D in vitro models are often used

in preclinical studies to predict the safety and efficacy of novel drug candidates before

advancing to expensive and time-consuming clinical trials. Successful screenings help

pharmaceutical companies focus their resources on the most promising drugs to reduce

attrition rates. A recent study has found that only 13.8% of all drug development

programs progressed from phase 1 to final approval 1. Pharmaceutical companies are

constantly searching for more relevant in vitro models to help improve the efficiency of

their drug pipelines. Cell cultures are also heavily used in academia for various

applications including cellular biology, developmental biology, and disease pathogenesis.

It is clear that our understanding of cellular mechanisms and physiology is largely due to

the advancements made using cell cultures.

Nevertheless, multiple studies have found that 2D in vitro models inaccurately

replicate in vivo conditions and cellular behavior. Cells are responsive to their

microenvironments and rely on cell-cell and cell-matrix interactions for normal

physiological function and differentiation 2-4. Cells grown on a petri dish experience an

unnaturally stiff attachment surface and are forced into a planar morphology which

2

affects cell adhesion, signaling, and organization 4,5. Neurons and cells of the central

nervous system (CNS) are particularly affected by these limitations since their native

environment is drastically altered when grown on a flat surface. Under healthy in vivo

conditions, neurons in the brain create intricate three-dimensional (3D) networks with

neighboring glia and neurons that are compromised in 2D cultures. As a result, the

inaccurate culturing environment can challenge the relevance of neuronal cell cultures

and their ability to properly model the in vivo response. The recent development of 3D in

vitro models aims to improve upon traditional 2D techniques by fostering an environment

that more accurately replicates in vivo conditions 6.

1.2 Three-Dimensional Neurospheroids

Multiple neuronal cell lines have been successfully translated into 3D constructs

using a variety of methods 7; however, the present study focuses on neurospheroids

developed by the Hoffman-Kim lab which are derived from postnatal cortical rat tissue.

These neurospheroids are scaffold-free and self-aggregating; as a result, they do not

require unnaturally high concentrations of scaffolding proteins, growth factors, or

extracellular matrix (ECM) to promote adhesion and proliferation 8-10. Furthermore, the

neurospheroids could be produced with high batch consistency and in large quantities

without the need for specialized equipment which enables the use of high-throughput

testing. Depending on the days in vitro (DIV) and seeding density, neurospheroids can

have diameters between 100-300 µm 10. Cells within the neurospheroids successfully

replicated various in vivo cortical tissue characteristics including cell diversity,

morphology, electrophysiology, and biomechanical properties. Immunohistochemistry

(IHC) coupled with confocal imaging revealed the presence of neurons, astrocytes,

3

oligodendrocytes, microglia, and progenitor cells at relevant densities 10. In addition, the

neurospheroids innately produced the necessary ECM, in the form of laminin, without the

need for exogenous factors 10. The dispersion of glia and formation of neurites that

extended in all directions were morphologically similar to in vivo tissue as well 10.

Neurons at 14 and 21 DIV were capable of firing spontaneous and evoked action

potentials with robust excitatory and inhibitory synaptic activity as determined by whole

cell patch-clamps 10. Finally, the neurospheroids had an elastic modulus within the range

of neonatal brain tissue 10. These findings promote the use of neurospheroids as a robust

and reliable testing platform with versatile applications that bridge the gap between

traditional in vitro and in vivo models.

1.3 Applications of 3D Neural Cultures

Neurospheroids and other 3D neural cultures can be utilized in a variety of studies

due to their high versatility. In some studies, 3D cultures provided a testing platform that

was more clinically relevant than in vivo animal models. Microcephaly pathogenesis is

normally studied in vivo using mice; however, mice failed to exhibit the characteristic

reduction of brain size seen in human patients 11. Meanwhile, 3D cerebral organoids

derived from human pluripotent stem cells (hPSCs) successfully replicated fundamental

neurodevelopmental processes associated with healthy and diseased brains 11. In other

studies, neurospheroids derived from hPSCs expressed the hallmark pathologies of

Alzheimer’s disease through β-amyloid and tau accumulation 12-14. The establishment of

more relevant in vitro disease models will help researchers develop the most accurate

understanding of the underlying biological mechanisms associated with disease

pathogenesis.

4

3D neural tissues also provide an advantageous, high-throughput testing platform

for the use in cytotoxicity and drug screening studies. The 3D in vitro model provides

more relevance than its 2D counterpart while mitigating the ethical concerns associated

with using animal models, especially when a large sample size is required. In one study,

neurospheroids were used to evaluate the cytotoxicity of perfluorooctanoic and

perfluorooctanesulfonic acid which were commonly widespread industrial compounds 15.

Neurospheroids have also been used to study the effects of regional glioblastoma tumor

heterogeneity and its impact on drug efficacy 16-18. Sufficiently large tumor spheroids

inherently reproduce the hypoxic, acidic, and low nutrient cores of tumors found in situ;

as a result, they can more accurately predict the true tumor drug response 16-18. The drug

screening capabilities of 3D in vitro models was further validated in studies that analyzed

Zika-infected 3D neural tissue and the effectiveness of various drugs to counteract the

infection 19,20. As highlighted by these studies, 3D neural cultures enable researchers to

utilize powerful high-throughput approaches to examine the effects of development,

diseases, and drugs on the CNS.

1.4 Cellular Viability Analysis

1.4.1 Traditional Cellular Viability Assays

Cellular viability is an example of a critical analytical assessment used for in vitro

studies. As the name suggests, cellular viability assays determine the health of cells

within a sample. Techniques including live/dead assays, colorimetric assays, fluorescence

imaging, and IHC are utilized often 21-24. Live/dead assays such as calcein AM, ethidium

homodimer, propidium iodide, and trypan blue rely on the ability of cells with intact

5

membranes to exclude or uptake certain dyes to differentiate them from dead cells. The

sample can be analyzed by a flow cytometer or fluorescence microscope to measure

cellular viability. Colorimetric assays such as the MTT [3-(4,5-dimethylthiazol-2-yl)-2,5-

diphenyltetrazolium bromide] and lactate dehydrogenase (LDH) assays rely on the

interaction of cellular enzymes with a color changing substrate. Viability can then be

analyzed utilizing a plate reader which monitors the change in optical absorption as a

result of the color change. Fluorescence imaging overlaps with some tags used in

live/dead assays and IHC; however, it can also take the form of calcium imaging.

Calcium homeostasis is an integral part of neuronal growth, and the disruption of

intracellular calcium levels leads to cell death 25-27. As a result, intracellular calcium

measurements can be used in conjunction with other assays to determine cellular

viability. Other assays, such as terminal deoxynucleotidyl transferase dUTP nick end

labeling (TUNEL) which binds to fragmented DNA and annexin V which binds to

extracellular phosphatidylserine, can be coupled with fluorescent markers to detect

specific death pathways such as apoptosis. IHC utilizes antibodies, typically attached to a

fluorescent marker, to detect certain antigens within tissue. If structural proteins such as

the cytoskeleton are tagged, then dead cells can be identified due to the morphological

changes associated with cell death. In addition, this technique could be used to

differentiate between various cell death pathways occurring within a sample based on

their unique morphological “fingerprints”. Cell death pathways will be discussed in more

detail in section 2.2.2.

6

1.4.2 Challenges Associated with Traditional Methods

Despite the advancements made to produce improved 3D in vitro models, there is

an evident lack of noninvasive, label-free, and nondestructive evaluation techniques for

3D tissue cultures 17. In many cases, traditional techniques have been developed and

optimized in 2D cultures; therefore, they do not translate well into the 3D construct. One

main concern of utilizing these methods involves the addition of foreign reagents or

labeling molecules to the sample. These assays analyze viability as a “snapshot” at

certain time points due to the terminal and cytotoxic nature of the added reagents. In

addition, some procedures (e.g. IHC, TUNEL, and propidium iodide) require the sample

to be fixed or optically cleared/sectioned. This prevents the true longitudinal assessment

of viability within the same sample. Another concern associated with adding labeling

molecules to the cell media involves the disproportionate tagging of cells closer to the

surface. This makes it difficult to analyze cellular viability within the deeper regions of

3D cultures due to the limited diffusion of labeling molecules that depends on their initial

concentration, size, and cell permeability 21. Live florescence imaging can eliminate the

use of cytotoxic reagents; however, fluorescent labels are susceptible to photobleaching

and phototoxicity which reduces their longitudinal capabilities 28. Imaging with chemical

fluorophores is also affected by the concentration of labeling molecules, the duration of

incubation, and 3D culture size (i.e. limitations in label diffusion). Recommended

concentrations and incubation times provided by the manufacturer are generally

determined for 2D cultures which forces researchers to independently run optimization

experiments tailored for their 3D setups. This is time consuming and necessitates the

need for standardized methods. Genetically encoded fluorescent tags overcome the

7

aforementioned limitations; however, they are costlier and can be conducted readily in a

small selection of animal models.

Challenges also arise with the utilization of certain analytical equipment. Plate

readers are typically used with colorimetric assays using specially made well plates

containing a known density of cells. A substrate solution that is either clear or a specific

color is added to the cells in the well plate. As the enzymes begin to react with their

substrate, a color change occurs which is measured as a change in absorbance at a given

wavelength. The change in absorbance is then compared to a standard curve and

correlated with viability. These assays are sensitive to both the viability and quantity of

cells present within a sample. It is more difficult to estimate the number of cells present

in 3D than 2D cultures due to increased batch heterogeneity (i.e. differences in spheroid

quantity, size, and composition). This makes it difficult to optimize the protocol to

achieve reliable and repeatable results. In addition, depending on seeding format, there

are generally fewer cells present in a single gel of neurospheroids than a confluent layer

on a well plate. The smaller quantity of cells may not be sufficient to produce a

noticeable color change in the assay which results in false negatives. To overcome these

challenges, 3D culture experiments utilizing a plate reader should include multiple

positive and negative controls as well as a robust standard curve for effective data

analysis.

Flow cytometers are also optimally designed for the use in 2D cultures. Flow

cytometry generally involves the measure and classification of cells within a suspension

based on their fluorescence. For example, cells can be simultaneously labeled using two

fluorescent markers. Each label should have a unique emission spectrum which allows

8

the flow cytometer to quantifiably plot the intensity distribution of multiple wavelengths

as it measures the fluorescence of cells passing a detector. This technique requires the

disaggregation of cells which eliminates any spatially relevant data present in 3D

cultures. Moreover, there are fewer cells fluorescently tagged in 3D cultures due to the

limited diffusion of the labeling molecules which could be potentially analyzed as false

negatives. As a result, plate readers and flow cytometers are most effectively used with

2D cultures.

9

Chapter 2: Optical Coherence Tomography Imaging and

System

2.1 Introduction

Optical coherence tomography (OCT) is a high resolution, label-free, and

noninvasive imaging modality with micrometer spatial resolution and millimeter imaging

depth 29. OCT imaging fundamentally relies on low coherence interferometry to analyze

sample tissue. Source light from a superluminescent diode is split into a reference and

sample arm. Light traveling through the reference arm is reflected off a mirror and

redirected to a detector. Meanwhile, the beam in the sample arm interacts with the

sample, and the backscattered light travels to the detector. The interference pattern

created from the two light paths creates an OCT signal related to the sample’s optical

properties. Broadband light sources are used to maintain low coherence such that only

light from a certain focal plane is resolvable. Adjusting the position of the reference

mirror shifts the focal plane allowing for 3D imaging of the sample throughout its depth.

Spectral Domain OCT (SD OCT) utilizes a similar imaging protocol but replaces the

detector with a spectrometer and does not require adjusting the reference mirror to obtain

3D images. Instead, the inverse Fourier transform is used to resolve the depth encoded

OCT signal from the sample at faster acquisition rates than traditional OCT. Figure 1

provides a basic schematic of the light paths and components of a SD OCT imaging

system. The SD OCT system can also be coupled with a variety of objective lenses to

obtain 3D images with different spatial resolutions. A SD OCT imaging system was

exclusively used in this study.

10

2.2 Optical Coherence Tomography for Viability Imaging

2.2.1 Advantages of Utilizing Optical Coherence Tomography Imaging

OCT and its various forms have multiple advantages over current cellular

viability imaging modalities. It most notably eliminates the need for foreign labels or

cytotoxic protocols which permits longitudinal, paired sample analysis. Theoretically,

OCT can be used to indefinitely image the same sample for the duration of the sample’s

usability. Current techniques preclude the true longitudinal assessment of disease

pathogenesis. Analyzing data from disease models like Alzheimer’s disease and cerebral

reperfusion injury is difficult with terminal assays since the disease symptoms can

progress at different rates across samples. In addition, the time intervals used for terminal

Figure 1. A schematic of a typical SD OCT system is shown. Light paths

are represented as arrows. The source, sample, and reference light paths

are colored red, blue, and gray, respectively.

11

assays are arbitrarily decided. This makes the continuous and longitudinal assessment of

a sample using OCT imaging more ideal for tracking the progression of diseases than

traditional methods.

Paired sample analysis, which is enabled by OCT imaging, is preferable over

unpaired analysis due to the reduction of variance within a sample. The reduction in

sample variance enables researchers to use smaller sample sizes while maintaining

statistical significance. This is critical for applications involving personalized medicine

since tissue biopsies can be limited. As a result, OCT viability imaging has the potential

to effectively implement personalized medicine and improve patient response to

treatment.

OCT’s penetration capability enables it to acquire images throughout 3D tissue

with high fidelity. Depending on the field of view (FOV), an entire spheroid could be

mapped which allows for the unique analysis of 3D regions (i.e. shell and core

measurements). This approach enables researchers to analyze the effects of hypoxic, low-

nutrient cores on treatment efficacy independently from the nutrient rich surface.

Arguably, confocal and two-photon microscopy can be used to generate 3D stacks of a

sample; however, OCT imaging does not require substantial time-intensive sample

preparation such as mixing reagents, sample incubation, or optical clearing. Multiple

spheroids can also be mapped in a relatively short time period due to OCT’s fast image

acquisition rates. A large FOV and fast imaging speeds are necessary components for an

efficient high-throughput imaging system. Drug screening studies can greatly benefit

from these features. Coupling a high-throughput imaging system with the ability to

12

analyze region-specific changes in viability can be applied to study the effects of 3D

tumor heterogeneity and its impact on drug resistance.

Finally, neuronal tissue has been successfully imaged using OCT in ex vivo and in

vitro environments with high spatial resolution and in close accordance with ‘gold

standards’ of the field 30-32. Based on these findings, it is expected that OCT can

successfully image neurospheroids. Overall, the combined advantages of the OCT system

make it a potentially powerful cellular viability imaging tool for neurospheroids.

2.2.2 The Link Between Cellular Viability and Optical Coherence Tomography

OCT-based viability imaging is dependent on the changes in optical scattering

properties between viable and dead cells. For the purposes of this study, cell death will be

categorized through apoptosis or oncotic necrosis. This classification is overly simplified

since there exist other cell death pathways; however, the two pathways mentioned are

expected to be the most commonly found in this study. Apoptosis occurs when the cell

actively initiates signaling cascades resulting in the degradation and condensation of the

nucleus and cytoplasm into dense apoptotic bodies 33,34. Oncosis is characterized by

cellular and organelle swelling, increased membrane permeability, and blebbing that

ultimately leads to necrosis 34. Transmission electron micrographs of single neurons (Fig.

2) clearly display the unique morphological differences between apoptotic and necrotic

cells 35. These distinct morphological characteristics impact the optical scattering profile

of cells that is detectable by OCT imaging; therefore, it is reasonable to believe that

apoptotic neurons will produce an OCT “signature” that is uniquely different from those

13

produced by healthy or oncotic neurons. This approach can be utilized for a variety of

cell types since cells exhibit similar morphological changes after undergoing apoptosis or

oncotic necrosis.

Intracellular motility (ICM), defined here as the ATP-dependent motion of

organelles within a cell, could also be used to determine viability. The active transport of

organelles like the mitochondria is a well-established phenomenon that occurs in healthy

Figure 2. (A) Transmission electron micrographs of a normal neuron, (B) an apoptotic

neuron, (C and D) and necrotic neurons. Scale bar 2.5 µm. Abbreviations: N, nucleus;

PM, plasma membrane; Mi, mitochondria. Adapted from Won et al., 2002.

14

neurons 36-39. Meanwhile, dead or metabolically compromised neurons are incapable of

sustaining these motions. It is expected that live cells will have more ICM than dead

cells. OCT imaging can detect the motion of scattering particles in each voxel which can

be used to measure ICM and differentiate between live and dead cells. The impact of

different cell death pathways on ICM will be discussed in more detail in sections 2.3.2

and 2.3.3.

2.3 Optical Coherence Tomography Parameters

Multiple OCT-derived parameters can be obtained using different scanning

protocols and post-processing procedures. This allows for a multi-metric approach to

analyze cellular viability within a sample. Parameters can be assessed individually or

collectively to obtain the best measurement of cellular viability. In particular, intensity,

decorrelation, diffusion coefficient, and surface area-to-volume-ratio (SATVR) will be

discussed.

2.3.1 Intensity

Intensity is an OCT parameter that quantifies the amount of light, in mW/cm2,

that is backscattered to the detector/spectrometer. Regions with higher reflectivity are

capable of backscattering more light than their surroundings and appear brighter on OCT

images. However, the intensities of deeper sample regions become attenuated due to the

absorption and decreased penetration of light. The process for deriving the intensity of a

sample is outlined in Figure 3. Since intensity is sensitive to the field reflectivity of a

sample, physical changes to the density and structure of the intracellular scattering

particles will affect the intensity profile of the cell. It can be hypothesized that the

15

intensity of a cell undergoing apoptosis will increase due to the compaction of organelles

(i.e. more optically dense); meanwhile, the intensity will decrease in oncotic cells due to

the dispersal and swelling of organelles (i.e. less optically dense). Samples that are

optically denser are capable of reflecting more light and therefore have a higher intensity.

2.3.2 Decorrelation

Decorrelation measures the change in the complex-valued OCT signal (i.e.

speckle) per given voxel over a short period of time. The associated imaging protocol

repetitively scans along the y-axis through a B-scan. The time between B-scans,

Figure 3. A diagram explaining the steps and equations used to derive the intensity

for a continuous sample using SD OCT. E0 is the equation for the electric field of the

source light, R is the sample’s field reflectivity, c is the speed of light, ε is the

permittivity of free space, k is the wave number, z is the sample depth, ω is the

angular frequency, σ is the standard deviation of the center angular frequency, CC is

the complex conjugate, and ζ is the Fourier transformed wave number.

16

identified as Δt, can be optimized based on the speed of moving particles within a

sample. In general, a longer Δt allows for the capture of slow-moving particles while

shorter periods are more sensitive to fast-moving particles. Stationary particles will

exhibit no decorrelation; meanwhile, the light scattered from particles entering, leaving,

and moving within a voxel collectively contribute to the speckle (Fig. 4A). Nevertheless,

Figure 4. (A) An illustration of the cross-section of a voxel (blue square) containing

light scattering particles (colored circles) before and after Δt time has passed. Black

arrows signify the direction of particle movement, and white arrows indicate the

reflected light’s path. (B) The resulting OCT signal produced by the combined

interference of all scattering particles present within the voxel (i.e. speckle). (C) The

equation used to derive decorrelation where D is the decorrelation, R is the field

reflectivity, and t is time.

C

𝐷 ≡ |�̃�(𝑡0 + ∆𝑡) − �̃�(𝑡0)|2

A

B

17

particle size, particle depth, and the wavelength of incident light determine which

particles contribute to the speckle. The difference in speckle in the complex plane is then

used to derive the decorrelation (Fig. 4B and 4C). Decorrelation can be used to measure

cellular viability since the movement of scattering particles can be attributed to ICM.

Necrotic cells are expected to have decorrelations measurements similar to that of the

extracellular space. This is expected because the damaged cell membrane can no longer

isolate the cytoplasm from the cell’s surroundings. Early apoptotic cells may exhibit

higher decorrelation since the process of nucleic and cytoplasmic fragmentation involves

the dynamic restructuring of organelles.

2.3.3 Diffusion Coefficient

The diffusion coefficient is similar to decorrelation in that it is capable of

measuring ICM. Unlike decorrelation, the diffusion coefficient is a standardized, metric

unit-based parameter that is measured in µm2/s. Standardized metrics can be compared

across imaging systems; however, arbitrary metrics (e.g. relative intensity and

decorrelation) cannot be cross-analyzed reliably. Acquiring the diffusion coefficient

requires an alternate scanning protocol termed dynamic light scattering (DLS). DLS OCT

repetitively acquires axial line scans, also known as A-scans, with a speed of 47,000 A-

scans/s 40. The brief fluctuations in scattered light captured by the DLS OCT system are

used to analyze the motion of particles within a voxel (Fig. 5) 40. The OCT signal from a

voxel with mixed particle dynamics is then fitted using the autocorrelation function to

derive the diffusion coefficient using the following equation:

18

Equation 1.

𝑔(𝒓, 𝜏) =⟨𝑅∗(𝒓, 𝑡)𝑅(𝒓, 𝑡 + 𝜏)⟩𝑡

⟨𝑅∗(𝒓, 𝑡)𝑅(𝒓, 𝑡)⟩𝑡=

𝑀𝑠(𝒓) + 𝑀𝐹(𝒓)𝑒−ℎ𝑡2𝑣𝑡

2(𝒓)𝜏2−ℎ2𝑣𝑧2(𝒓)𝜏2

𝑒−𝑞2𝐷(𝒓)𝜏𝑒𝑖𝑞𝑣𝑧(𝒓)𝜏 + [1 − 𝑀𝑠(𝒓) − 𝑀𝐹(𝒓)]𝛿(𝜏)

The Ms term relates to the ratio of static particles, MF to the moving particles, vt is the

transverse velocity, vz is the axial velocity, and D is the diffusion coefficient 40. The

exponential components of the equation are related to the velocity-dependent term,

diffusion-oriented decay, and axial velocity-dependent term in the order they appear in

Equation 1 40. Unlike other ICM based metrics, the diffusion coefficient is less sensitive

to noise and errors caused by the presence of mixed particle dynamics within the same

voxel. Instead, the DLS OCT mathematical model accounts for the different particle

dynamics (Eq. 1). As a result, the diffusion coefficient can directly measure ICM since

intracellular dynamics can be modeled by Brownian motion 41. Necrotic cells that are no

longer encapsulated by a membrane will exhibit diffusive motions that are

Figure 5. (A) An illustration of the particles contributing to the DLS OCT signal in a

given voxel. (B) A model of the autocorrelation function on a complex plane. The

magnitude of MS and MF are proportional to the static and flowing particles,

respectively. Particles with larger cross-sections contribute more to the DLS OCT

signal. Adapted from Lee et al., 2012.

19

indistinguishable from the extracellular space. Early apoptotic cells with intact

membranes will sustain intracellular dynamics that differ from their backgrounds.

2.3.4 Surface Area-to-Volume-Ratio

Geometry and bulk characteristics of 3D tissues can be used to indirectly predict

viability. As previously mentioned, cells undergoing apoptosis or oncotic necrosis will

experience morphological changes that impact their structural dimensions. These changes

are difficult to measure from a single cell; however, the collective changes from multiple

cells within a spheroid can be detected using a sufficiently large FOV. Analyzing

geometric measurements (e.g. surface area, diameter, volume, etc.) individually is not

adequate since these parameters can vary greatly among spheroids of the same DIV and

seeding density. Instead, the SATVR can be used to minimize the variance associated

with batch heterogeneity. Despite this, the SATVR parameter is driven by a r-1

dependency (Eq. 2), and spheroids with a larger radius will inherently have a smaller

SATVR regardless of viability.

Equation 2.

𝑆𝐴𝑇𝑉𝑅 =4𝜋𝑟2

43 𝜋𝑟3

∝1

𝑟

𝐴𝑑𝑗𝑢𝑠𝑡𝑒𝑑 𝑆𝐴𝑇𝑉𝑅 = 4𝜋𝑟2

(43 𝜋𝑟3)

23

∝ 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡

This can be accounted for by raising the denominator (i.e. volume) to the 2/3 power. This

eliminates the parameter’s dependency on the spheroid’s radius, thus enabling the

analysis of spheroids across different sizes.

20

2.3.5 OCT Viability Imaging in Literature

Initial applications of OCT involved diagnostic imaging of the eye 42,43. OCT’s

penetration depth, high resolution, noninvasive nature, and functionality (i.e. different

imaging modes) makes it an ideal imaging system for the detection of structural damage

to the retina and its associated vasculature 44. Although biomedical applications of OCT

are predominantly ophthalmic, recent applications have extended to neuroimaging,

endoscopic imaging, and various forms of structural imaging 45-47. OCT applications

involving cellular viability imaging of 3D tissues have also emerged with promising

results.

One of the earliest applications of OCT cellular viability imaging involved

measuring temporal changes in speckle from rat osteogenic sarcoma tumor spheroids 48.

Jeong et al. calculated the speckle’s standard deviation at each voxel from multiple

successive images and normalized it to the mean intensity 48. This approach requires that

viable cells produce time-varying speckle maps while dead cells remain static. They

demonstrated a decrease in speckle fluctuations associated with treating their tumor

spheroids with cytoskeletal anti-cancer drugs (e.g. nocodazole, colchicine, and

paclitaxel). These anti-cancer drugs disrupt microtubule polymerization to ultimately

prevent cell replication and cause cell death 48. They also took advantage of OCT’s

imaging capabilities to generate motility maps at depths of 600 µm from the spheroid’s

surface 48. In a similar study, Martucci et al. treated 3D liver spheroids with various

concentrations of acetaminophen to mimic drug-induced liver injury and measured the

change in speckle-variance distribution. They observed a decrease in speckle variance

associated with an increase in treatment concertation that was consistent with the results

21

from a CellTiter-Glo cellular activity assay 49. In another study, Dunkers et al. utilized

image cross-correlation to determine the number of features a final image had in common

with an earlier image taken at the same location. Their findings demonstrated that the

number of common features in dead cells approached 100% after treatment with

ultraviolet light which can be associated with the termination of ICM 50. Other studies

that measured the phase change in OCT signal observed a similar reduction in

intracellular fluctuations associated with cell death 51,52.

Nevertheless, correlating ICM with fluctuations in speckle has its challenges.

Speckle fluctuations provide an indirect and qualitative measure of ICM since it

fundamentally relies on healthy cells being “noisier” than dead cells. Furthermore,

measurements of speckle are considerably affected by system noise, vibrations, sample

depth, and the passive drifting of cells within the FOV. The effects of different cell death

pathways were also not addressed in the studies mentioned. An alternate approach,

utilized by Farhat et al., measured the decorrelation time extracted from the intensity

autocorrelation function which provided a better measurement of ICM. They induced

apoptosis in acute myeloid leukemia cells using cisplatin and observed an increase in

ICM (i.e. a decrease in decorrelation time) 53. These findings are consistent with the

hypothesis that ICM in apoptotic cells is higher than viable cells due to ongoing cell

fragmentation 53.

OCT-based parameters involving backscattered intensity and scattering

attenuation have also been used to measure cellular viability. In a different study

conducted by Farhat et al., the backscattered intensity of control, apoptotic, and necrotic

cells was measured. Cells undergoing apoptosis and mitotic arrest experienced an

22

increase in integrated intensity while a decrease was observed in necrotic cells 54.

Nevertheless, no consistent changes in intensity were observed in liver spheroids by

Martucci et al. after treatment with acetaminophen which may suggest cell-dependent

differences in the parameter’s ability to measure viability 49. Scattering attenuation,

which is closely related to backscattered intensity, measures the decay of scattered light

as a function of sample depth 55. In a study conducted by van der Meer et al., the

attenuation coefficient increased in apoptotic cells and decreased in necrotic cells 55.

Huang et al. utilized a similar approach to differentiate cells in the necrotic core from the

remaining viable cells in glioblastoma spheroids 17. Interestingly, both studies correlated

regions of higher scattering attenuation with apoptotic cell death as confirmed by either

an Annexin V or TUNEL assay 17,55. These results support the hypothesis that apoptotic

cells become more optically dense due to cell fragmentation and condensation;

meanwhile, oncotic cells become less dense due to the dispersal of internal scattering

organelles and particles.

Correlating cellular viability with intensity-based parameters has its challenges as

well. Since backscattered intensity is largely dependent on the incident light, factors

including the inherent fluctuations in source light, angle of objective lens, depth of the

sample, presence of debris above the sample, and changes in the refractive index

surrounding the spheroid could have an impact on intensity measurements.

23

Chapter 3: Optical Coherence Microscopy Revealed Time-

Dependent Changes in Cellular Dynamics in Response to

Ethanol Treatment

3.1 Abstract

Three-dimensional neurospheroids provide a robust in vitro platform that

improves upon traditional 2D cultures. Nevertheless, conventional cell viability assays

are not fully optimized for the use in 3D cultures. OCT is a label-free, 3D, and high-

resolution imaging modality that has a potential for nondestructive cell viability

assessment. In this study, a SD OCM system was utilized to assess the viability of

neurospheroids derived from postnatal cortical rat tissue. Neurospheroids at various DIV

were treated with 4 M etOH to induce oncotic necrosis and repeatedly imaged for one

hour. SD OCM imaging was capable of visualizing cell bodies, based on contrast

produced by intracellular motility, with greater penetration depth than confocal

microscopy. Prolonged exposure to etOH resulted in the steady decrease in the OCM

intensity signal while no significant changes occurred in decorrelation. The decrease in

intensity was attributed to the disruption and degradation of the neuronal cell membrane

resulting in oncotic necrosis as visualized by immunohistochemistry and calcium

imaging. These results support the feasibility of nondestructive, multi-parametric OCT-

based cell viability imaging in 3D tissue which is capable of tracking underlying cell

death processes. This approach can be implemented to a variety of in vitro disease and

drug studies.

24

3.2 Introduction

Currently, cellular viability can be determined utilizing a variety of techniques

that assess various aspects of cellular function. Membrane integrity, metabolic activity,

DNA damage, and cytoskeletal morphology form the basis of some assays; and, despite

their differences, they can all be correlated with viability 21,56. Nevertheless, there are

multiple limitations associated with utilizing traditional techniques including their

inability to longitudinally assess viability and inability to effectively evaluate 3D tissue

throughout its depth. Cellular viability assays with enhanced analytical power, ease of

use, and 3D capabilities have the potential to revolutionize drug discovery, tissue

engineering, and other fields that rely on cell culture as a cornerstone of research.

OCT imaging provides a promising method for the label-free, noninvasive, and

3D evaluation of cellular viability in tissue spheroids. Another advantage of OCT

imaging involves its ability to output multiple parameters (e.g. intensity, decorrelation,

diffusion coefficient, and spheroid geometry) with relative ease. This allows for a multi-

metric approach for determining cellular viability that is not currently possible using a

single traditional method alone. Nonetheless, OCT parameters must accurately detect

changes in cellular viability in a manner that is both reproducible and efficient. It is also

necessary to identify the biological mechanism(s) linked to the change in viability that is

detected by the OCT parameters. These parameters can be analyzed individually or

collectively to best evaluate the viability of sample tissue.

Optical coherence microscopy (OCM) utilizes a typical OCT system coupled with

a high-power objective lens to create images with micrometer resolution. It is necessary

25

to validate the changes in OCM parameters with a change in viability at the cellular level.

Once an OCM parameter is validated, more studies can be conducted to better understand

how the parameter responds to multiple death pathways and cell types. In this study, a

commercially available SD OCT system was coupled with a 40x objective lens to image

neurospheroids extracted from primary postnatal cortical rat tissue. Neurospheroids were

exposed to high concentrations of ethanol (etOH) to elicit oncotic necrosis. The cytotoxic

impact of etOH treatment is well documented by many in vitro studies involving a variety

of cell types. Depending on the concentration, cell type, and exposure period, etOH can

elicit cell death through multiple mechanisms including DNA synthesis inhibition 57,58,

DNA fragmentation 59,60, induction of oxidative stress 61,62, and reduction of metabolic

function 63,64. Ethanol is also capable of disordering the cell membrane and, at high

concentrations, cause the solubilization of the cell membrane leading to necrosis 60,65-67.

Pure 200-proof etOH was utilized to induce oncotic necrosis in the tissue sample.

Ethanol was directly added to the media to achieve a final concentration of 4 M. The

etOH was allowed to diffuse throughout the sample while roughly maintaining a

temperature of 37ºC. The spheroids were imaged longitudinally over the course of one

hour to track the changes in viability. IHC and calcium images were utilized to determine

the underlying biological mechanism(s) responsible for the changes in OCM parameters.

3.3 Methods

3.3.1 Neurospheroid Cell Isolation

The neurospheroids used in this study were derived from primary cortical tissue

extracted from female postnatal Charles River rats. Neonatal rats were wrapped in a layer

26

of gauze followed by plastic wrap and placed inside an ice bucket for 30-40 minutes to

induce hypothermia. Enough gauze was used to ensure a painless cooling (i.e. providing

indirect contact with the ice bed). The plastic wrap was left open at the ends to allow air

flow but prevent any ice or water from entering the bundle. After inducing hypothermia,

a pair of sharpened dissection scissors were used to decapitate the rat below the ears. The

scissors were then used to make a sagittal cut from the base of the head to the skull to

expose the brain. The brain was then placed into a petri dish with Neurobasal A/B27

media without calcium, and the cortex was isolated. Cortical tissue was then placed in a

conical tube and submerged with 2 mg/mL papain solution dissolved in Hibernate A

(BrainBits, LLC) without calcium for 30 minutes at 30°C. The papain solution was

removed, and the tissue was triturated with a fire-polished Pasteur pipette in Hibernate A

buffer supplemented with 1 x B27 supplement (Invitrogen) and 0.5 mM GlutaMAX

(Invitrogen). The cell solution was then centrifuged at 150g for 5 minutes, and the

supernatant was removed. The cell pellet was resuspended in Neurobasal A/B27 media

using the following components: Neurobasal A media (Invitrogen) supplemented with

1xB27, 0.5 mM GlutaMAX, and 1x Penicillin-Streptomycin (Invitrogen). The cell

suspension was filtered using a 40 µm cell strainer to remove debris and washed once by

repeating the centrifuging, resuspending, and filtering steps. Cortical cells were seeded at

a density of 8,000 cells/well in 2% agarose microarrays.

3.3.2 Forming Three-Dimensional, Self-Aggregating Neurospheroids

Microarrays were created from molten 2% agarose (Invitrogen) by pouring the

solution onto a mold with rounded 400-µm diameter pegs (#24-96-small, MicroTissues,

Inc.). Newly created gels were equilibrated in cell media using three media exchanges

27

over a 48-hour period. Media was aspirated from the gels before seeding, and a cell

solution of 75 µL/gel was added directly onto the gel. The gels have a small recess that

prevents the cell suspension from flowing out. The cell suspension was allowed to settle

in the microwells for 30 minutes followed by the addition of 1 mL of Neurobasal A/B27

media in a 24-well plate. Neurospheroids were incubated at 37ºC with 5% CO2, and the

media was changed every 3-4 days.

3.3.3 Calcium Imaging

Intracellular calcium levels were measured using OGB calcium indicator. 30 µL

of 4% F-127 (diluted using DMSO) was added to one vial of OGB1 powder. The solution

was sonicated in the dark for 15 minutes followed by the addition of 30 µL of media and

an additional 5 minutes of sonication. Roughly 8 µL of OGB1 solution was added to a

single well of a 24-well plate which contained about 1 mL of media and the

neurospheroids. The plate was incubated in the dark for 20-25 minutes at 37ºC and 5%

CO2. After incubation, the well was washed twice with media to remove excess dye. The

neurospheroids were then removed from the agarose gel by gently pipetting up and down

over the wells. The extracted neurospheroids were transferred to a 35 mm petri dish with

a 25 mm glass bottom for confocal imaging. Images were acquired using a high

sensitivity resonant scanner in an Olympus FV3000RS with a laser intensity of 0.6%, HV

from 400-500, gain of 1, and offset of 6. Cells were kept at 37ºC throughout the imaging

session. Images were converted to a .tif format and collapsed into a single plane for

viewing using Fiji ImageJ software.

28

3.3.4 Immunohistochemistry

Neurospheroids were washed three times using Neurobasal A/B27 media. The

media was removed, and the cells were fixed using 4% vol/vol paraformaldehyde for at

least 24 hours. The fixative was removed, and the samples were washed with 8% wt/vol

sucrose in phosphate-buffered saline (PBS) three times. Following fixation,

neurospheroids were extracted from the gels and placed into Eppendorf tubes. A blocking

solution comprised of 10% normal goat serum, 4% bovine serum albumin, and 1%

Triton-X 100 in PBS was added to the tubes for 2-3 hours. Primary antibody was diluted

using PBT solution of 0.2% Triton-X 100 in PBS and added to the tube. Primary

antibody solution was removed and neurospheroids were washed using PBT twice prior

to the addition of the secondary antibody solution. Spheroids were optically cleared using

ClearT2 protocol prior to imaging using the Olympus FV3000RS. Images were prepared

in Fiji ImageJ software.

3.3.5 ClearT2 Optical Clearing Protocol

ClearT2 optical clearing protocol was used to enhance IHC imaging of

neurospheroids. Solutions which consisted of 20% and 40% wt/vol polyethylene glycol

(PEG) were made in water, and 50% vol/vol formamide (Sigma Aldrich) solution was

made in PBS. Neurospheroids were first placed in 25% formamide/10 % PEG for 10

minutes, then 50% formamide/20% PEG for 5 minutes, and finally 50% formamide/20%

PEG for 60 minutes. The neurospheroids were kept in the final solution and transferred to

a glass-bottom confocal dish for IHC imaging.

29

3.3.6 OCM Imaging System and Scanning Protocol

A commercially available Telesto III SD OCT system (Thorlabs, Newton, NJ,

USA) was used with a center wavelength of 1310 nm and bandwidth of 170 nm. The SD

OCT system was coupled with a 40x objective lens (1-U2M587, Olympus America, Inc.)

for microscopic imaging and a 2048-pixel line-scan camera capable of 47,000 A-scans/s.

A FOV capturing 832 x 832 x 256 pixels was obtained with a lateral and axial resolution

of 0.5 μm and 3.5 μm, respectively, resulting in an imaging volume of 416 μm x 416 μm x

896 μm. A time delay of 10 ms between B-scans was used to create 3D decorrelation maps

of the neurospheroids. Custom LabVIEW software was used to acquire OCM data. Gels

containing neurospheroids were transferred from a 24 well-plate to a 35 mm petri dish to

allow the objective lens easier access to the spheroids. Prior to the transfer, molten agarose

was spread across the bottom of the petri dish. The gel was then placed over the agarose

(without any media) and allowed to solidify in place for 1-2 minutes. Once set in place, 3

mL of media was added to the petri dish. Neurospheroids were kept on a heating pad

(PhysioSuite, Kent Scientific, Torrington, CT, USA) to maintain a temperature of 37°C

throughout the imaging session. Images were acquired at 10, 30, 45, and 60 minutes after

the addition of etOH.

3.3.7 OCM Data Processing

OCM image reconstruction was accomplished using Matlab (MathWorks, Natick,

MA, USA). Image enhancements such as median filtering, motion correction, dispersion

compensation, interpolation, and volume averaging were utilized to improve overall image

quality and resolution. Ultimately, 3D intensity and decorrelation maps were produced for

the neurospheroid throughout the FOV. Analysis of the neurospheroids was conducted on

30

2D en face slices extracted from the 3D maps at the focal plane near the center of the

neurospheroid (Fig. 6). Further processing of the en face images involved categorizing the

neurospheroid into core and shell regions (Fig. 7). The radius of each spheroid in the en

face image was determined and the core was characterized as the inner half of the radius

while the outer half was designated as the shell (Fig. 7).

Figure 6. The process of acquiring and analyzing OCM data is described. (Top left)

The imaging plane, identified by the shaded blue region, was acquired along the

central plane of the neurospheroid. (Top right) A 2D OCM decorrelation map was

then reconstructed at the focal plane. (Bottom right) Multiple longitudinal images

were acquired at a fixed FOV for one hour. (Bottom left) OCM-derived viability data

can then be plotted longitudinally for each spheroid.

31

Figure 7. A visual representation of how the shell (region between the dotted red and

blue line) and core (region within the dotted blue line) are designated when analyzing

OCM data. Scale bar 100 µm.

3.3.8 Cellular Viability Manipulation

In order to cause a decrease in cellular viability, 200-proof etOH was used to

induce oncotic necrosis in the neurospheroids. Ethanol was directly pipetted into the petri

dish until a final concentration of 4 M was achieved. The etOH was not directly added

over the gel, to prevent the dislodging of neurospheroids, and allowed to passively

diffuse throughout the media for 10 minutes prior to acquiring the first image.

Consecutive images were acquired at 30, 45, and 60 minutes.

3.3.9 Statistical Analysis

Raw and normalized mean intensity and decorrelation values were reported for

the core, shell, and total regions of the neurospheroid alongside the standard deviation.

SigmaPlot 12.5 software (Systat Software Inc., San Jose, CA, USA) was used to create

figures and conduct one-way repeated measures ANOVA alongside Tukey post hoc tests

to determine statistical significance. Data sets that failed normality and equal variance

assumptions were analyzed using the Friedman test (i.e. non-parametric equivalent of the

32

one-way repeated measures ANOVA) with Dunn’s post hoc test. Statistical analysis was

only conducted on the raw means.

3.4 Results

3.4.1 OCM Imaging Revealed Cells Through Contrast in Intracellular Motility

One of the primary advantages of utilizing OCM involves its ability to image cells

without the need for labeling molecules. This becomes a crucial feature when imaging

large spheroids and other 3D tissue since cells in deep regions are not exposed to the

same concentration of labeling molecules as those closer to the surface 68. The depth-

dependent tagging of cells causes disproportionate intensity profiles which can skew

results. To demonstrate this advantage, neurospheroids were imaged using OCM and

confocal microscopy at depths of 30 µm, 90 µm, and 150 µm from the spheroid’s surface

(Fig. 8). IHC and calcium images were acquired from separate spheroids while OCM

intensity and decorrelation maps were from the same spheroid at identical locations. Cell

nuclei, neuronal cytoskeletal networks, and cell bodies were identified using DAPI, β-III-

tubulin, and OGB calcium indictor, respectively. OCM decorrelation maps revealed cell-

like features that are consistent with those found in IHC and calcium images (Fig. 8 and

9). Cellular features in OCM decorrelation maps had larger decorrelation values than

their surroundings which indicated regions with higher motility. Interestingly, many

bodies featured darker center regions with lower motility which may indicate cell nuclei.

Confocal image intensity began to noticeably attenuate at depths of 90 µm while OCM

images resolved the neurospheroid all throughout the depths. The cell-like bodies present

in the OCM decorrelation maps were compared to cell bodies found in the calcium

33

images (Fig. 9). Upon visual inspection, cell geometry and morphology were in close

accordance between the two image types.

Figure 8. (Left to right) IHC, calcium, decorrelation, and intensity images of

neurospheroids at 14 DIV at various depths. Cell nuclei and neuronal microtubules

were labeled using DAPI (blue) and β-III-tubulin (red), respectively. Calcium images

were acquired with OGB indicator (green). Scale bars 100 µm.

34

3.4.2 OCM Intensity Signal Displayed Unique Time-Dependent Responses to

Ethanol Treatment

2D en face OCM intensity and decorrelation maps were obtained for

neurospheroids treated with 4 M etOH (n=4; DIV: 5, 7, 19, and 28). The intensity and

decorrelation time courses were obtained by averaging their values over the core, shell,

and total regions of the neurospheroid (Fig. 10). Statistical tests revealed that the mean

intensities at 30, 45, and 60 minutes were significantly different from the intensity at 10

minutes for the core region alone. In addition, the mean intensity measurements of the

core were significantly different from the shell across all time intervals. No significant

changes were observed in the decorrelation parameter (Fig. 10). Both intensity and

decorrelation were unchanged under control conditions (results not shown).

Figure 9. OGB calcium (left) and OCM decorrelation (right) en face images of

neurospheroids at 9 and 14 DIV, respectively. Magnified views are designated by the

red squares.

35

Figure 10. Raw and normalized mean intensity and decorrelation parameters were

longitudinally measured in 4 M etOH treated neurospheroids (n=4; DIV: 5, 7, 19, and

28). Error bars represent the standard deviations. Either the one-way repeated

measures ANOVA with Tukey test or the Friedman test with Dunn’s post hoc tests

were conducted depending on the normality and variance of the data. Statistical results

compared the 10-minute value with all other time points and are reported in the order

of core, shell, and total regions from left to right. P < 0.05 *, < 0.01 **, and < 0.001 †.

36

3.4.3 Ethanol Disrupted Membrane Integrity Through Oncotic Necrosis

Neurospheroids were imaged using OGB calcium indicator to reveal changes in

cellular membrane integrity in response to etOH treatment. Neuronal cell bodies can be

easily identified in calcium images due to their distinct intracellular calcium signatures.

Within one minute, membrane disruption was observed through the dissolution of some

cell bodies near the spheroid periphery and increased fluorescence intensity associated

with the influx of calcium down its concentration gradient (Fig. 11B). Within 10 minutes,

cell membranes were disrupted throughout the spheroid (Fig. 11C-F). Cell nuclei and

astrocytes remained intact despite 60 minutes of exposure to 4 M etOH as seen in IHC

images (Fig. 12).

Figure 11. Z-projections were acquired from neurospheroids at 9 DIV using OGB

calcium indicator after 4 M etOH treatment. Projections were created from the

average of the top 100 slices. Images A-F were taken at pretreatment, 1, 10, 30, 45,

and 60 minutes after treatment, respectively. Image G is of a 60-minute control

spheroid. Images A-F are from the same spheroid. Scale bars 100 µm.

37

3.5 Discussion

The primary objective of this chapter was to assess and validate the use of OCM-

based cellular viability parameters. A high magnification 40x objective lens was coupled

with a SD OCT setup to image neurospheroids with cellular resolution. Cell specific

changes could be observed using this approach to correlate the changes in OCM

parameters with their associated biological mechanism(s). OCM decorrelation en face

and cross-sectional images of cell bodies were similar to those found in confocal images

using OGB calcium indicator (Fig. 8 and 9). The cell bodies could be resolved deeper

within and throughout the neurospheroid using OCM which highlighted the advantages

Figure 12. Z-projections were acquired from neurospheroids at 14 DIV using the top

100 slices. Astrocytes (GFAP, gray) and cell nuclei (DAPI, blue) were stained. Images

were taken at 10 min control (A), 10 min treatment (B), 60 min control (C), and 60

min treatment (D). Scale bars 100 µm.

38

of using a label-free imaging system (Fig. 8). Interestingly, many cell bodies could be

identified with regions of lower decorrelation near their center. It is likely that the regions

with higher decorrelation represent the cytoplasm while the central region with lower

decorrelation is the cell nucleus. These findings were consistent with a previous study

conducted using DLS OCT on rat cerebral cortex 69. When particles in a given voxel

move over a short time period, they produce a fluctuating scattering profile that is

measurable by an increase in decorrelation. The contrast in decorrelation between the

cytoplasm and nucleus could be explained by the presence of faster moving particles in

the cytoplasm. It is also possible that the scattering particles within the nucleus are too

small to be resolved by the current SD OCM imaging system.

The ability to acquire longitudinal intensity and decorrelation data from the same

samples highlighted another strength of the OCM imaging system. This method provided

a more detailed approach to analyzing viability within 3D tissues. Statistical analysis

revealed that the core, but not shell or total intensity, was significantly affected by etOH

treatment starting at 30 minutes. Furthermore, one-way repeated measures ANOVA and

Friedman tests revealed that the core and shell intensities were significantly different

throughout the experiment. These findings indicated that regional and time-dependent

differences to etOH treatment could be distinguished (Fig. 10). It was expected that

significant changes in shell intensity would occur with treatment. This assumption is

supported by Figure 11 were clear disruption in cell membrane integrity could be

observed in the periphery. It is possible that changes to the shell intensity occurred

quickly and leveled out within 10 minutes. Acquiring and tracking OCM data at earlier

time points with greater temporal resolution should reveal the presence of fast dynamics.

39

In addition, increasing the sample size will greatly benefit the performance of the

statistical tests.

Differences between intensity and decorrelation trends suggest that the parameters

measured different processes within cells. Nevertheless, no statistically significant

changes in decorrelation were observed. In a previous study, a decrease in OCT speckle

variation, an analog of decorrelation, was observed in liver spheroids after treatment with

acetaminophen 49. Similarly, a decrease in decorrelation time was observed when acute

myeloid leukemia cells were treated with cisplatin 53. The contradiction between results

from the present and previous studies may be explained by the time-dependent effects of

different treatments on ICM. Ethanol treatment may have induced rapid changes to ICM-

related decorrelation across spheroid regions that were not captured by the earliest time

interval utilized in this study. This assumption is supported by Figure 11 since the most

significant changes to cell membrane integrity occurred within the first 10 minutes of 4

M etOH exposure. Future decorrelation images should be acquired within 0-10 minutes

to track these major changes. Furthermore, lower concentrations of etOH could be used

to elicit slower changes in viability.

A microscopic approach was utilized in this study to determine the underlying

biological mechanisms associated with the changes in OCM parameters. In particular, the

steady decrease in core intensity should correlate with changes to biological factors

associated with cellular viability. The major disruption of membrane integrity suggested

that oncotic necrosis occurred within the sample (Fig. 11). The presence of intact nuclei

after 60 minutes of etOH exposure confirmed that apoptotic condensation of nuclei had

not occurred; therefore, oncotic necrosis was the primary pathway of cell death (Fig. 12).

40

It can also be seen in Figure 12 that astrocytes were not primarily affected by the etOH

treatment; therefore, neuronal membrane integrity was the most likely source for the

decrease in intensity. The disruption of neuronal membrane integrity can impact intensity

in two ways. First, neurons undergoing oncosis swell which causes their volumes to

increase and organelles to become more disperse. This can result in a decrease in

intensity since the probability of light backscattering due to organelles decreases. Another

explanation involves the inability of the membrane to keep organelles constrained within

the neuron. Neurons undergoing oncotic necrosis spew their contents into the

extracellular space which also causes a decrease in optical density resulting in a decrease

in intensity.

In this study, cellular viability was monitored using OCM imaging in response to

etOH treatment. Additional viability manipulators should be tested to assess how OCM

parameters respond to various cell death processes. For instance, methods that induce

apoptosis should be tested and compared to those that induce oncosis. Theoretically, the

differences in morphology and optical scattering properties between cells undergoing

different cell death pathways could be differentiated by OCM imaging. In the present

study, neurospheroids of various DIV were imaged and analyzed; however, age, size, and

gender could impact their response to the etOH treatment. Future studies could address

this by increasing sample size and comparing OCM parameters across the various groups.

Given the label-free and noninvasive nature of OCM imaging, sample viability could be

monitored well over the one-hour period measured in this study. Ultimately, longitudinal

dose response curves could be created to assess cytotoxic properties of drugs or monitor

treatment of injury and/or disease models utilizing 3D neurospheroids.

41

3.6 Conclusion

In conclusion, the validity of utilizing OCM-based viability imaging was

examined in this study. Neurospheroids were treated with 4 M etOH to induce oncotic

necrosis and imaged longitudinally using OCM. Cell bodies can be identified in the OCM

decorrelation images without the use of labels and at depths that exceed conventional

confocal microscopy. Differences between the intensity and decorrelation trends may

indicate their ability to measure different biological mechanisms associated with cell

death (i.e. disruption of cellular morphology and ATP-dependent ICM). Further studies

are warranted to examine OCM’s ability to track the effects of other cell death pathways.

In addition, it would be advantageous to examine the effects of spheroid age, gender, and

various etOH concentrations to fully validate the OCM viability imaging system.

42

Chapter 4: High-throughput, Multi-Metric, and 3D

Assessment of Neurospheroids

4.1 Abstract

High-throughput drug screenings provide a powerful tool to investigate the

efficacy of various drug treatments. 3D in vitro models help improve the overall clinical

relevance of these tests; however, current high-throughput imaging systems are not

optimized for the 3D construct. OCT is a label-free, nondestructive, and 3D imaging

modality that can address the limitations faced with traditional methods. A low

magnification objective lens was utilized to obtain images with large FOVs to capture

multiple spheroids. Intensity, decorrelation, and diffusion coefficient parameters were

measured longitudinally to track cell viability in neurospheroids. Neurospheroids were

exposed to final concentrations of 4 M etOH, 2 M etOH, and PBS sham groups to

determine the interaction between treatment dose and exposure time on viability. A

significant decrease in intensity was observed in the 4 M etOH group starting at 45

minutes. The diffusion coefficient had a slight yet significant increase in the 4 M group at

60 minutes and in the 2 M group at 45 and 60 minutes. Overall, the use of macroscopic

OCT-based cellular viability imaging is promising for high-throughput applications.

4.2 Introduction

High-throughput drug screening studies form a vital component of the drug

discovery pipeline. Early preclinical studies help determine the safety and efficacy of

novel drugs to predict their performance at treating various diseases. Despite the

implementation of early preclinical tests, only 13.8% of drugs are approved 1. One step

43

that can be taken to improve overall drug success rate is to enhance the performance of

preclinical in vitro models at predicting clinical outcomes. Recent studies have shown

that 3D in vitro cultures more accurately model the in vivo response to treatment 70-73. As

a result, adopting 3D in vitro models in high-throughput studies will increase their overall

clinical relevance. Nonetheless, there are challenges associated with transitioning to 3D

models. Currently, high-throughput confocal imaging systems are used to measure

viability across multiple samples and treatment groups. As previously shown, confocal

fluorescence imaging is not optimized for the evaluation of 3D tissues. OCT is a label-

free, nondestructive, and 3D imaging system with cellular viability applications capable

of replacing the confocal setup 17,18,48-55.

Thus far in this study, a microscopic approach was utilized to correlate OCT-

based parameters with changes in viability. High resolution imaging allows for the cell-

specific tracking of viability which enables the visual identification of the cell death

pathway occurring within a sample. Nevertheless, only a single spheroid can be imaged

along a very narrow focal plane using higher magnifications. This can make it difficult to

obtain an adequate sample size to establish statistical significance. Microscopic imaging

is also sensitive to noise and external vibrations that can affect image quality and

analysis. On the other hand, macroscopic imaging lacks cell specificity while

substantially increasing the number of spheroids within the FOV. The entirety of multiple

spheroids can be imaged and analyzed simultaneously using lower magnifications. SD

OCT’s rapid image acquisition rates and imaging depth make it a promising candidate for

high-throughput viability imaging. If coupled with a motorized stage, an OCT imaging

system could be utilized in a similar manner to a high-throughput confocal setup 74.

44

The ability of the SD OCT system to detect changes in viability using a low

magnification objective lens was examined. Previous studies have shown that

macroscopic parameters like the scattering attenuation and SATVR correlate well with

tumor spheroid viability and traditional viability assays 17,18. In this study, intensity,

decorrelation, and diffusion coefficient were monitored longitudinally in neurospheroids.

Two concentrations of etOH were used to determine the dose-dependent response to

treatment. One-way repeated measures ANOVA and the Friedman test were used to

determine if treatment concentration and duration significantly affected OCT-derived

viability parameters (e.g. intensity, decorrelation, and diffusion coefficient).

4.3 Methods

4.3.1 Neurospheroid Cell Isolation

The neurospheroids used in this study were derived from primary cortical tissue

extracted from female postnatal Charles River rats. Neonatal rats were wrapped in a layer

of gauze followed by plastic wrap and placed inside an ice bucket for 30-40 minutes to

induce hypothermia. Enough gauze was used to ensure a painless cooling (i.e. providing

indirect contact with the ice bed). The plastic wrap was left open at the ends to allow air

flow but prevent any ice or water from entering the bundle. After inducing hypothermia,

a pair of sharpened dissection scissors were used to decapitate the rat below the ears. The

scissors were then used to make a sagittal cut from the base of the head to the skull to

expose the brain. The brain was then placed into a petri dish with Neurobasal A/B27

media without calcium, and the cortex was isolated. Cortical tissue was then placed in a

conical tube and submerged with 2 mg/mL papain solution dissolved in Hibernate A

45

(BrainBits, LLC) without calcium for 30 minutes at 30°C. The papain solution was

removed, and the tissue was triturated with a fire-polished Pasteur pipette in Hibernate A

buffer supplemented with 1 x B27 supplement (Invitrogen) and 0.5 mM GlutaMAX

(Invitrogen). The cell solution was then centrifuged at 150g for 5 minutes, and the

supernatant was removed. The cell pellet was resuspended in Neurobasal A/B27 media

using the following components: Neurobasal A media (Invitrogen) supplemented with

1xB27, 0.5 mM GlutaMAX, and 1x Penicillin-Streptomycin (Invitrogen). The cell

suspension was filtered using a 40 µm cell strainer to remove debris and washed once by

repeating the centrifuging, resuspending, and filtering steps. Cortical cells were seeded at

a density of 8,000 cells/well in 2% agarose microarrays.

4.3.2 Forming Three-Dimensional, Self-Aggregating Neurospheroids

Microarrays were created from molten 2% agarose (Invitrogen) by pouring the

solution onto a mold with rounded 400-µm diameter pegs (#24-96-small, MicroTissues,

Inc.). Newly created gels were equilibrated in cell media using three media exchanges

over a 48-hour period. Media was aspirated from the gels before seeding, and a cell

solution of 75 µL/gel was added directly onto the gel. The gels have a small recess that

prevents the cell suspension from flowing out. The cell suspension was allowed to settle

in the microwells for 30 minutes followed by the addition of 1 mL of Neurobasal A/B27

media in a 24-well plate. Neurospheroids were incubated at 37ºC with 5% CO2, and the

media was changed every 3-4 days.

46

4.3.3 SD OCT Imaging System and Scanning Protocol

A commercially available Telesto III SD OCT system (Thorlabs, Newton, NJ,

USA) was used with a center wavelength of 1310 nm, bandwidth of 170 nm, and an axial

resolution of 3.5 µm. The SD OCT system was coupled with a low magnification LSM03

objective lens (Thorlabs, Newton, NJ, USA) for macroscopic imaging and a 2048-pixel

line-scan camera capable of 47,000 A-scans/s. A FOV capturing 512 x 512 x 1024 pixels

was obtained, resulting in an imaging volume of 5 mm x 5 mm x 3 mm (in the x,y,z

directions) which contained the 3D intensity and decorrelation maps. A time delay of 10

ms between B-scans was used to derive the decorrelation.

The diffusion coefficient was acquired using DLS OCT imaging protocol. A FOV

capturing 128 x 256 x 1024 pixels was obtained, resulting in an imaging volume roughly

1 mm x 2.5 mm x 3 mm (in the x, y, and z directions). A total of 256 volume scans were

taken along the x direction with a time delay of 3.9 ms between each line scan.

Custom LabVIEW software was used to acquire OCT data. Gels containing

neurospheroids were transferred from a 24 well-plate to a 35 mm petri dish to allow the

objective lens easier access to the spheroids. Prior to the transfer, molten agarose was

spread across the bottom of the petri dish. The gel was then placed over the agarose

(without any media) and allowed to solidify in place for 1-2 minutes. Once set in place, 3

mL of media was added to the petri dish. Neurospheroids were kept on a heating pad

(PhysioSuite, Kent Scientific, Torrington, CT, USA) to maintain a temperature of 37°C

throughout the imaging session. Images were acquired at 15, 30, 45, and 60 minutes after

the addition of etOH.

47

4.3.4 Cellular Viability Manipulation

In order to cause a decrease in cellular viability, 200-proof etOH was used to

induce oncotic necrosis in the neurospheroids. The interaction between treatment dose

and duration was examined by comparing spheroids exposed to 4 M etOH, 2 M etOH

(50% vol/vol in PBS), and PBS sham treatment. The appropriate volume of etOH or PBS

was directly pipetted into the petri dish. Care was taken to not directly add liquids over

the gel to prevent the dislodging of neurospheroids. The treatment administered was

allowed to passively diffuse throughout the media for 15 minutes prior to acquiring the

first image. Consecutive images were acquired at 30, 45, and 60 minutes.

4.3.5 OCT Data Processing

OCT image reconstruction was accomplished using Matlab (MathWorks, Natick,

MA, USA). Image enhancements such as median filtering, motion correction, dispersion

compensation, interpolation, and volume averaging were utilized to improve overall

image quality and resolution. A total of 10-15 healthy spheroids near the center of the gel

were selected and tracked longitudinally for the intensity and decorrelation parameters.

Each spheroid was fitted into a 3D ellipsoid from which all the voxels were grouped into

the core, shell, and total regions and averaged (Fig. 13). Individual spheroid means were

averaged again for the gel at each time point. This process was repeated for the diffusion

coefficient except it involved selecting six spheroids and gel averages were not calculated

(i.e. statistical analysis was conducted on spheroid, and not gel, averages due to limited

replicates).

48

4.3.6 Statistical Analysis

Raw and normalized mean intensity, decorrelation, and diffusion coefficient (only

raw) were calculated along with the standard deviation. SigmaPlot 12.5 software (Systat

Software Inc., San Jose, CA, USA) was used to create figures and conduct one-way

repeated measures ANOVA alongside Tukey post hoc tests for normally distributed data

with equal variance. Data sets that failed these assumptions were analyzed using the

Friedman test with Dunn’s post hoc tests to determine statistical significance. These tests

were used to determine if statistically significant changes occurred due to time or

spheroid region within each treatment group. Regular one-way ANOVA with

Tukey/Dunn’s post hoc tests were used to determine the statistical differences between

treatment groups within the same time interval. Statistical tests were only conducted on

the raw means.

Figure 13. (Left) En face intensity image of multiple neurospheroids within a gel. The

spheroid centered in the red square appears in various orthogonal planes (center and

right). The red ellipses in the orthogonal views are fitted for each selected spheroid.

49

4.4 Results

4.4.1 Ethanol Elicited Dose and Time-Dependent Changes in OCT Intensity Signal

Raw and normalized mean intensity were plotted as a function of time for the 4 M

etOH (n spheroids = 85; n gels = 6; DIV 7 and 14) 2 M etOH (n spheroids = 45; n gels =

3; DIV 21), and PBS (n spheroids = 60; n gels = 4; DIV 14 and 21) treatment groups

(Fig. 14). The intensity significantly decreased in the 4 M etOH treatment group at 45

Figure 14. Raw and normalized mean intensity were longitudinally plotted for

neurospheroids in the 4 M etOH (n spheroids = 85; n gels = 6; n batches = 3; DIV 7

and 14) 2 M etOH (n spheroids = 45; n gels = 3; n batches = 2; DIV 21), and PBS (n

spheroids = 60; n gels = 4; n batches = 2; DIV 14 and 21) treatment groups. The core,

shell, and total intensity parameters were plotted individually to visualize any regional

changes. One-way repeated measures ANOVA/Friedman test with Tukey/Dunn’s post

hoc tests were conducted to determine statistical significance (for normally or non-

normally distributed data, respectively). The statistical significance between the initial

condition and all other time points was reported in the order of core, shell, and total

region from left to right. P values < 0.05 *, < 0.01 **, and < 0.001 †.

50

and 60 minutes which resulted in a 20% (p<0.05) and 25% (p<0.01) drop from the initial

value, respectively. No statistically significant changes were seen in the 2 M or PBS

groups. One-way repeated measures ANOVA/Friedman test revealed statistical

differences between the core and shell regions within all treatment groups except for the

4 M group at 45 minutes and 2 M group at 60 minutes. Similarly, one-way ANOVA

revealed that there were statistically significant differences in intensity between the 2 M

and 4 M groups as well as the 4 M and PBS groups across all time intervals expect at 60

minutes were no differences were found between the 4 M and 2 M groups.

4.4.2 Ethanol Did Not Elicit Dose and Time-Dependent Changes in OCT

Decorrelation

Data for the decorrelation metric was longitudinally plotted as well (Fig. 15). The

number of spheroids, gels, and batches were identical with the intensity data sets. No

statistically significant differences in decorrelation were observed in any of the treatment

groups. The core and shell decorrelation were only significantly different in the 2 M

group at 15 and 30 minutes. There were no differences between treatment types at each

time interval.

51

4.4.3 Ethanol Elicited a Slight Increase in Diffusion Coefficient

While the decorrelation relates to relative changes in the OCT signal, the

diffusion coefficient is measured using standardized units that do not require

normalization. Changes in diffusion coefficient due to 4 M etOH (n spheroids = 12; n

batch = 1; DIV 7), 2 M etOH (n spheroids = 18; n batches = 2; DIV 21), and PBS (n

spheroids = 18; n batches = 3; DIV 14 and 21) were plotted against time (Fig. 16). There

Figure 15. Raw and normalized mean decorrelation were longitudinally plotted for

neurospheroids in the 4 M etOH (n spheroids = 85; n gels = 6; n batches = 3; DIV 7

and 14) 2 M etOH (n spheroids = 45; n gels = 3; n batches = 2; DIV 21), and PBS (n

spheroids = 60; n gels = 4; n batches = 2; DIV 14 and 21) treatment groups. The core,

shell, and total intensity parameters were plotted individually to visualize any regional

changes. One-way repeated measures ANOVA/Friedman test with Tukey/Dunn’s post

hoc tests were conducted to determine statistical significance (for normally or non-

normally distributed data, respectively). The statistical significance between the initial

condition and all other time points were reported in the order of core, shell, and total

region from left to right. P values < 0.05 *, < 0.01 **, and < 0.001 †.

52

was a slight but significant increase in the core diffusion coefficient at 60 minutes from

0.2065 to 0.2139 µm2/s (p < 0.001). In addition, there were statistically significant

increases in diffusion coefficient with the 2 M group across all spheroid regions at 45 and

60 minutes (p<0.001). Interestingly, there were statistically significant changes in

diffusion coefficient within the PBS treated group. Core and shell diffusion coefficients

were significantly different in the 4 M group at 15 minutes, in the 2 M group at 60

minutes, and in the PBS group across all time intervals. Finally, one-way ANOVA with

Tukey/Dunn’s post hoc tests revealed significant differences between the 4 M and 2 M

treatment groups as well as the 4 M and PBS groups across all time intervals.

Figure 16. Changes in diffusion coefficient (D) were plotted longitudinally for

neurospheroids exposed to 4 M etOH (n spheroids = 12; n gels = 2; n batch = 1; DIV

7), 2 M etOH (n spheroids =18; n gels = 3; n batches = 2; DIV 21), and PBS (n=18; n

gels = 3; n batches = 2; DIV 14 and 21). One-way repeated measures

ANOVA/Friedman test with Tukey/Dunn’s post hoc tests were conducted to

determine the statistical significance between the initial condition and all other time

points. P values were reported in the order of core, shell, and total region from left to

right. P values < 0.05 *, < 0.01 **, and < 0.001 †.

53

4.5 Discussion

The ability of the SD OCT imaging system to track changes in cellular viability at

a macroscopic level was examined in this chapter. Three treatment groups (i.e. 4 M etOH,

2 M etOH, and PBS sham) were imaged using a low magnification LSM03 objective

lens. Multiple spheroids from a single FOV were analyzed for changes in intensity,

decorrelation, and diffusion coefficient at the core, shell, and total regions of the

neurospheroids.

Based on the results from the previous higher magnification approach, it was

expected that intensity would decrease in a time-dependent manner in the 4 M etOH

group. Indeed, this was observed with the macroscopic approach (Fig. 14). It was also

expected that a smaller, delayed decrease in intensity would be observed in the 2 M etOH

group; however, no statistically significant changes occurred. Interestingly, the mean

intensities of the 2 M and 4 M etOH groups were statistically different at each time point

except at 60 minutes. This suggested that the effects of 2 M etOH on intensity may

require exposure periods in excess of 60 minutes to manifest. Another factor that could

impact the results is the difference in spheroid age between the treatment groups.

Spheroids in the 4 M etOH group were considerably younger than those in the 2 M group

(DIV 7 and 14 vs. 21, respectively). In addition, only three gels were analyzed for the 2

M group which could impact the performance of the statistical tests. Future studies

should test later time points, adjust the disparity in spheroid age, and include more

replicates to improve the reliability of the results.

54

The decorrelation results from the macroscopic approach were similar to those

found using microscopic imaging. There were no statistically significant changes in

decorrelation in any of the treatment groups (Fig. 15). As previously mentioned, the lack

of significant changes in the 4 M group could be due to the inadequate temporal window

used in this study. It is reasonable to believe that 4 M etOH treatment elicits rapid

changes in decorrelation that occur soon after the etOH is added which plateau by 15

minutes. It is unclear from the results whether the 2 M etOH treatment elicits

decorrelation changes within the same time period as the 4 M group or if it is not

concentrated enough to cause a change in decorrelation. Furthermore, different

concentrations of etOH have been shown to elicit different effects on cells 57-67. This

could mean that 2 M etOH treatment elicited a different type of cell death that does not

induce a change in decorrelation. Acquiring OCT images immediately after the treatment

is added to the spheroids, with an appropriate temporal resolution, should help address

these concerns.

It was expected that ICM decreases with cell viability since dead cells are no

longer capable of sustaining metabolic functions. Nevertheless, the diffusion coefficient

increased in both the 4 M and 2 M etOH treatment groups (Fig. 16). These results

suggested that measurements of ICM are overwhelmed by other diffusive motions.

Similar results were observed in a recent study which reported a decrease in decorrelation

time (which is comparable to an increase in diffusion coefficient) when oncotic necrosis

was induced in breast cancer cells 75. The exact process responsible for the increase in

diffusion coefficient is not yet identified. Interestingly, some significant changes in

diffusion coefficient could also be observed in the PBS sham treatment group (Fig. 16).

55

This suggests that PBS may not serve as an ideal sham treatment since the diffusive flow

within the cells was affected. It can also be assumed that the change in temperature,

concentration of nutrients, and fluid densities caused by the addition of PBS to the media

elicited some effects on intracellular transport within the cells. Future studies can address

these concerns by replacing PBS with media sham treatment.

Overall, the results from this chapter are promising regarding the ability of a low

magnification OCT system to analyze cellular viability in a high-throughput manner.

Despite this, more studies need to be conducted to fully validate the use of OCT for

viability imaging. Most notably, the effects of other cell death pathways must be

examined using the OCT system. It would also be beneficial to induce the same type of

cell death pathway using alternate methods (e.g. temperature, osmolarity, glucose

deprivation, etc.) to compare the changes in OCT parameters. Ideally, the same cell death

pathway should affect the OCT parameters in the same way regardless of treatment

method. Future studies should account for the effects of DIV on viability by testing

spheroids from the same age. In addition, changes to OCT parameters within the first 10-

15 minutes should be investigated more thoroughly especially for the decorrelation and

diffusion coefficient. Finally, more parameters should be tested such as the SATVR and

scattering attenuation. Testing additional parameters allows for a more comprehensive

and holistic approach to determining viability. Ultimately, an algorithm utilizing multiple

OCT-derived parameters will be created through cluster and correlation analysis to

measure the viability within a sample in a high-throughput, label-free, and nondestructive

manner.

56

4.6 Conclusion

The recent emergence of in vitro tissue spheroids for applications in drug

screening studies necessitates the need for more efficient high-throughput imaging

modalities. OCT is label-free, nondestructive, and capable of imaging multiple spheroids

throughout their depths. In this chapter, macroscopic changes in intensity, decorrelation,

and diffusion coefficient were measured in response to 4 M etOH, 2 M etOH, and PBS

sham treatment. Significant changes in intensity and diffusion coefficient were observed

with a change in viability. More studies need to be conducted to fully validate the effects

of viability on OCT-based parameters. Finally, the impacts of spheroid age, treatment

type, and temporal resolution on OCT viability imaging need to be determined.

57

References

1. Wong CH, Siah KW, Lo AW. Estimation of clinical trial success rates and related

parameters. Biostatistics. 2018;20(2):273-86.

2. Ma W, Tavakoli T, Derby E, Serebryakova Y, Rao MS, Mattson MP. Cell-extracellular

matrix interactions regulate neural differentiation of human embryonic stem cells.

BMC Developmental Biology. 2008;8:90.

3. Togashi H, Sakisaka T, Takai Y. Cells adhesion molecules in the central nervous

system. Cell Adhesion and Migration. 2009;3(1):29-35.

4. Baker BM, Chen CS. Deconstructing the third dimension: how 3D culture

microenvironments alter cellular cues. Journal of Cell Science. 2012;125(13):3015-24.

5. Cukierman E, Pankov R, Stevens DR, Yamada KM. Taking cell-matrix adhesion to

the third dimension. Science. 2001;294(5547):1708-12.

6. Edmondson R, Broglie JJ, Adcock AF, Yang L. Three-dimensional cell culture

systems and their applications in drug discovery and cell-based biosensors. Assay and

Drug Development Technology. 2014;12(4):207-18.

7. Hopkins AM, DeSimone E, Chwalek K, Kaplan DL. 3D in vitro modeling of the

central nervous system. Progress in Neurobiology. 2015;125:1-25.

8. Lin RZ, Chang HY. Recent advances in three-dimensional multicellular spheroid

culture for biomedical research. Biotechnology Journal. 2008;3(9-10):1172-84.

9. Fennema E, Rivron N, Rouwkema J, van Blitterswijk C, de Boer J. Spheroid culture

as a tool for creating 3D complex tissues. Trends in Biotechnology. 2013;31(2):108-

15.

58

10. Dingle YT, Boutin ME, Chirila AM, Livi LL, Labriola NR, Jakubek LM, Morgan JR,

Darling EM, Kauer JA, Hoffman-Kim D. Three-dimensional neural spheroid culture:

an in vitro model for cortical studies. Tissue Engineering Part C Methods.

2015;21(12):1274-83.

11. Lancaster MA, Renner M, Martin C, Wenzel D, Bicknell LS, Hurles ME, Homfray T,

Penninger JM, Jackson AP, Knoblich JA. Cerebral organoids model human brain

development and microcephaly. Nature. 2013;501(7467):373-379.

12. Choi SH, Kim YH, Hebisch M, Sliwinski C, Lee S, D’Avanzo C, Chen H, Hooli B,

Asselin C, Muffat J, Klee JB, Zhang C, Wainger BJ, Peitz M, Kovacs DM, Woolf CJ,

Wagner SL, Tanzi RE, Kim DY. A three-dimensional human neural cell culture

model of Alzheimer’s disease. Nature. 2014;515(7526):274–8.

13. Jorfi M, D’Avanzo C, Tanzi RE, Kim DY, Irimia D. Human Neurospheroid Arrays

for In Vitro Studies of Alzheimer’s Disease. Scientific Reports. 2018;8:2450.

14. Lee H-K, Sanchez CV, Chen M, Morin PJ, Wells JM, Hanlon EB, Xia W. Three-

dimensional human neuro-spheroid model of Alzheimer’s disease based on

differentiated induced pluripotent stem cells. PLoS One. 2016;11(9):e0163072.

15. Choi S, Kim J, Park J, Lee K, Kim E, Jeon WB. Cytotoxicity and inhibition of

intracellular interaction in N2a neurospheroids by perfluorooctanoic acid and

perfluorooctanesulfonic acid. Food and Chemical Toxicology. 2013;60:520-9.

16. Khaitan D, Chandna S, Arya MB, Dwarakanath BS. Establishment and

characterization of multicellular spheroids from a human glioma cell line; implication

for tumor therapy. Journal of Translational Medicine. 2006;4:12.

59

17. Huang Y, Wang S, Guo Q, Kessel S, Rubinoff I, Chan LL, Li P, Liu Y, Qiu J, Zhou

C. Optical coherence tomography detects necrotic regions and volumetrically

quantifies multicellular tumor spheroids. Cancer Research. 2017;77(21):6011-20.

18. Jung Y, Klein OJ, Wang H, Evans CL. Longitudinal, label-free, quantitative tracking

of cell death and viability in a 3D tumor model with OCT. Nature. 2016;6:27017.

19. Watanabe M, Buth JE, Vishlaghi N, de la Torre-Ubieta L, Taxidis J, Khakh B,

Coppola G, Pearson CA, Yamauchi K, Gong D, Dai X, Damoiseaux R, Aliyari R,

Liebscher S, Schenke-Layland K, Caneda C, Huang EJ, Zhang Y, Cheng G,

Geschwind DH, Golshani P, Sun R, Novitch BG. Self-organized cerebral organoids

with human specific features predict effective drugs to combat Zika virus infection.

Cell Reports. 2017;21(2):517-32.

20. Garcez PP, Loiola EC, de Costa RM, Higa LM, Trindade P, Delvecchio R,

Nascimento JM, Brindeiro R, Tanuri A, Rehen SK. Zika virus impairs growth in

human neurospheres and brain organoids. Science. 2016;352(6287):816-18.

21. Aras MA, Hartnett KA, Aizenman E. Assessment of Cell Viability in Primary

Neuronal Cultures. Current Protocols in Neuroscience. 2008;44(1):7.18.1-7.18.15.

22. Brana C, Benham C, Sundstorm L. A method for characterizing cell death in vitro by

combining propidium iodide staining with immunohistochemistry. Brain Research

Brain Research Protocols. 2002;10(2):109-14.

23. Landry S, McGhee P, Girardin R, Keeler W. Monitoring live cell viability:

Comparative study of fluorescence, oblique incidence reflection and phase contrast

microscopy imaging techniques. Optics Express. 2004;12(23):5754-9.

24. Gerlich D, Ellenberg J. 4D imaging to assay complex dynamics in live specimens.

Nature Cell Biology. 2003;Suppl:S14-9.

60

25. Ringler SL, Aye J, Byrne E, Anderson M, Turner CP. Effects of disrupting calcium

homeostasis on neuronal maturation: early inhibition and later recovery. Cellular and

Molecular Neurobiology. 2008;28(3):389-409.

26. Kim JW, Oh HA, Lee SH, Kim KC, Eun PH, Ko MJ, Gonzales ELT, Seung H, Kim

S, Bahn GH, Shin CY. T-Type Calcium Channels Are Required to Maintain Viability

of Neural Progenitor Cells. Biomolecules and Therapeutics. 2018;26(5):439-445.

27. Turner CP, Connell J, Blackstone K, Ringler SL. Loss of calcium and increased

apoptosis within the same neuron. Brain Research. 2006;1128(1):50-60.

28. Ettinger A, Wittmann T. Fluorescence Live Cell Imaging. Methods in Cell Biology.

2014;123:77-94.

29. Huang D, Swanson EA, Lin CP, Schuman JS, Stinson WG, Chang W, Hee MR,

Flotte T, Gregory K, Puliafito CA. Optical coherence tomography. Science.

1991;254:1178–81.

30. Tamborski S, Lyu HC, Dolezyczek H, Malinowska M, Wilczynski G, Szlag D, Lasser

T, Wojtkowski M, Szkulmowski M. Extended-focus optical coherence microscopy

for high resolution imaging of the murine brain. Biomedical Optics Express.

2016;7(11):4400-14.

31. Magnain C, Augustinack JC, Konukoglu E, Frosch MP, Sakadžić S, Varjabedian A,

Garcia N, Wedeen VJ, Boas DA, Fischl B. Optical coherence tomography visualizes

neurons in human entorhinal cortex. Neurophotonics. 2015;2(1):015004.

32. Bizheva K, Unterhuber A, Hermann B, Považay B, Sattmann H, Drexler W. Imaging

ex vivo and in vitro brain morphology in animal models with ultrahigh resolution

optical coherence tomography. Journal of Biomedical Optics. 2004;9(4):719-724.

61

33. Fink SL, and Cookson BT. Apoptosis, Pyroptosis, and Necrosis: Mechanistic

Description of Dead and Dying Eukaryotic Cells . Infection and Immunity.

2005;73(4):1907-1916. doi:10.1128/IAI.73.4.1907-1916.2005.

34. Majno G, and Joris I. Apoptosis, oncosis, and necrosis. An overview of cell death.

The American Journal of Pathology. 1995;146(1):3-15.

35. Won SK, Kim DY, Gwag BJ. Cellular and molecular pathways of ischemic neuronal

death. Journal of Biochemistry and Molecular Biology. 2002;35(1):67-86.

36. Mironov SL. ADP Regulates Movements of Mitochondria in Neurons. Biophysical

Journal. 2007;92(8):2944-2952. doi:10.1529/biophysj.106.092981.

37. Xiao Q, Hu X, Wei Z, and Tam KY. Cytoskeleton Molecular Motors: Structures and

Their Functions in Neuron. International Journal of Biological Sciences.

2016;12(9):1083-1092. doi:10.7150/ijbs.15633

38. Appert-Rolland C, Ebbinghaus M, and Santen L. Intracellular transport driven by

cytoskeletal motors: General mechanisms and defects. Physics Reports. 2015;593:1-

59. doi:10.1016/j.physrep.2015.07.001.

39. Lin MY, Sheng ZH. Regulation of mitochondrial transport in neurons. Experimental

Cell Research. 2015;334(1):35-44.

40. Lee J, Wu W, Jiang JY, Zhu B, Boas DA. Dynamic light scattering optical coherence

tomography. Optics Express. 2012;20(20):22262-77.

41. Joo C, Evans CL, Stepinac T, Hasan T, de Boer JF. Diffusive and directional

intracellular dynamics measured by field-based dynamic light scattering. Optics

Express. 2010;18(3):2858-71.

62

42. Swanson EA, Izatt JA, Hee MR, Huang D, Lin CP, Schuman JS, Puliafito CA,

Fujimoto JG. In vivo retinal imaging by optical coherence tomography. Optics

Letters. 1993;18(21):1864-66.

43. Izatt JA, Hee MR, Swanson EA, Lin CP, Huang D, Schuman JS, Puliafito CA,

Fujimoto JG. Micrometer-scale resolution imaging of the anterior eye in vivo with

optical coherence tomography. Archives of Ophthalmology. 1994;112(12):1584-9.

44. Fercher AF, Drexler W, Hitzenberger CK, Lasser T. Optical coherence tomography-

principles and applications. Reports on Progress in Physics. 2003;66(2):239-303.

45. Men J, Huang Y, Solanki J, Zeng X, Alex A, Jerwick J, Zhang Z, Tanzi RE, Li A,

Zhou C. Optical coherence tomography for brain imaging and developmental biology.

IEEE Jornal of Selected Topics in Quantum Electronics. 2015;22(4):6803213.

46. Gora MJ, Suter MJ, Tearney GJ, Li X, Endoscopic optical coherence tomogprahy:

techniques and clinical applications. Biomedical Optics Express.

2017;8(5):2405:2444.

47. Kashani AH, Chen CL, Gahm JK, Zheng F, Richter GM, Rosenfeld PJ, Shi Y, Wang

RK. Optical coherence tomography angiography: a comphrensive review of current

methods and clinical applications. Progress in Retinal Eye Research. 2017;60:66-100.

48. Jeong K, Turek JJ, Nolte DD. Volumetric motility-contrast imaging of tissue response

to cytoskeletal anti-cancer drugs. Optics Express. 2007;15(21):14057-64.

49. Martucci NJ, Morgan K, Anderson GW, Hayes PC, Plevris JN, Nelson LJ,

Bagnaninchi PO. Nondestructive optical toxicity assays of 3D liver spheroids with

optical coherence tomography. Advanced Biosystems. 2018;2(3):1700212.

63

50. Dunkers JP, Lee YJ, Chatterjee K. Single cell viability measurements in 3D scaffolds

using in situ label free imaging by optical coherence tomography. Biomaterials.

2012;33(7):2119-26.

51. Holmes C, Tabrizian M, Bagnaninchi PO. Motility imaging via optical coherence

phase microscopy enables label-free monitoring of tissue growth and viability in 3D

tissue-engineering scaffolds. Journal of Tissue Engineering and Regenerative

Medicine. 2015;9(5):641-5

52. Bagnaninchi PO, Drummond N, Holmes C, Daoud J, Tabrizian M. Two-dimensional

and three-dimensional viability measurements of adult stem cells with optical

coherence phase microscopy. Journal of Biomedical Optics. 2011; 16(8):0860035.

53. Farhat G, Mariampillai A, Yang VX, Czarnota GJ, Kolios MC. Detecting apoptosis

using dynamic light scattering with optical coherence tomography. Journal of

Biomedical Optics. 2011;16(7):070505.

54. Farhat G, Yang VX, Czarnota GJ, Kolios MC. Detecting cell death with optical

coherence tomography and envelope statistics. Journal of Biomedical Optics.

2011;16(2):026017.

55. van der Meer FJ, Faber DJ, Aalders MC, Poot AA, Vermes I, van Leeuwen TG.

Apoptosis- and necrosis-induced changes in light attenuation measured by optical

coherence tomography. Lasers in Medical Science. 2009;25(2):259-67.

56. Riss TL, Moravec RA, Niles AL, Duellman S, Benink HA, Worzella TJ, Minor L.

Cell viability assays. In: Sittampalam GS, Coussns NP, Brimacombe K, et al., editors.

Assay Guidance Manual. Bethesda, MD: Eli Lilly & Company and the National

Center for Advancing Translational Sciences; 2004.

https://www.ncbi.nlm.nih.gov/books/NBK144065/; accessed March 10, 2019.

64

57. Yeo EJ, Lim HK, Park SC. Effect of short-term ethanol on the proliferative response

of Swiss 3T3 cells to mitogenic growth factors. Experimental and Molecular

Medicine. 2000;32:161-9.

58. Carter EA, & Wands JR. Ethanol inhibits hormone stimulated hepatocyte DNA

synthesis. Biochemical and Biophysical Research Communication. 1985;128:767-74.

59. DeVito WJ, Stone S, Shamgochian M. Ethanol increases the neurotoxic effect of

tumor necrosis factor-alpha in cultured rat astrocytes. Alcoholism, Clinical and

Experimental Research. 2000;24:82-92.

60. Castilla R, Gonzalez R, Fouad D, Fraga E, Muntane J. Dual effect of ethanol on cell

death in primary culture of human and rat hepatocytes. Alcohol and Alcoholism.

2004; 39: 290-6.

61. Kurose I, Higuchi H, Miura S, Saito H, Watanabe N, Hokari R, Hirokawa M,

Takaishi M, Zeki S, Nakamura T, Ebinuma H, Kato S, Ishii H. Oxidative stress-

mediated apoptosis of hepatocytes exposed to acute ethanol intoxication. Hepatology.

1997;25:368-78.

62. Higuchi H, Kurose I, Kato S, Miura S., Ishii H. Ethanol-induced apoptosis and

oxidative stress in hepatocytes. Alcoholism, Clinical and Experimental Research.

1996;20:340-6.

63. Devi BG, Henderson GI, Frosto TA, Schenker S. Effect of acute ethanol exposure on

cultured fetal rat hepatocytes: relation to mitochondrial function. Alcoholism, Clinical

and Experimental Research. 1994;18:1436-42.

64. Cunningham CC, Coleman WB, Spach PI. The effects of chronic ethanol

consumption on hepatic mitochondrial energy metabolism. Alcohol and Alcoholism.

1990;25:127-36.

65

65. Chang SW, Chou SF, Wang YH. Ethanol treatment induces significant cell death in

porcine corneal fibroblasts. Cornea. 2006;25:1072-9.

66. Baker RC, Kramer RE. Cytotoxicity of short-chain alcohols. Annual Review of

Pharmacology and Toxicology. 1999;39:127-50.

67. Lyon RC, McComb JA, Scherurs J, Goldstein DB. A relationship between alcohol

intoxication and the disordering of brain membranes by a series of short-chain

alcohols. Journal of Pharmacology and Experimental Therapeutics. 1981;218:669-

75.

68. Zanoni M, Piccinini F, Arienti C, Zamagni A, Santi S, Polico R, Bevilacqua A, Tesei

A. 3D tumor spheroid models for in vitro therapeutic screening: a systematic

approach to enhance the biological relevance of data obtained. Scientific Reports.

2016;6:19103.

69. Lee J, Radhakrishnan H, Wu W, Daneshmand A, Climov M, Ayata C, Boas DA.

Quantitative imaging of cerebral blood flow velocity and intracellular motility using

dynamic light scattering-optical cohere tomography. Journal of Cerebral Blood Flow

and Metabolism. 2013;33(6):819-25.

70. Sant S, Johnston PA. The production of 3D tumor spheroids for cancer drug

discovery. Drug Discovery Today: Technologies. 2017;23:27–36.

71. Smith SJ, Wilson M, Ward JH, Rahman CV, Peet AC, Macarthur DC, Rose FR,

Grundy RG, Rahman R. Recapitulation of tumor heterogeneity and molecular

signatures in a 3D brain cancer model with decreased sensitivity to histone

deacetylase inhibition. PLoS One. 2012;7(12):e52335.

66

72. Karlsson H, Fryknas M, Larsson R, Nygren P. Loss of cancer drug activity in colon

cancer HCT-116 cells during spheroid formation in new 3-D spheroid cell culture

system. Experimental Cell Research. 2012;318(13):1577-85.

73. Khawar IA, Park JK, Jung ES, Lee MA, Chang S, Kuh HJ. Three Dimensional

Mixed-Cell Spheroids Mimic Stroma-Mediated Chemoresistance and Invasive

Migration in hepatocellular carcinoma. Neoplasia. 2018;20(8):800–12.

74. Huang Y, Guo Q, Zou J, Wang X, Titus S, Ferrer M, Zhou C. High-throughput

optical coherence tomography imaging for drug screening with 3D tumor spheroids.

Biophotonics Congress: Biomedical Optics Congress 2018. 2018;OTh2D.5.

75. Arezza NJ, Razani M, Kolios MC. Dyanmic light scattering optical coherence

tomography to probe motion of subcellular scatterers. Journal of Biomedical Optics.

2019;24(2):025002.