micro- and nano-scale approaches for disease detection and ... · ii micro- and nano-scale...

120
Micro- and Nano-Scale Approaches for Disease Detection and Characterization by Adam Hunter Mepham A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Institute of Biomaterials and Biomedical Engineering University of Toronto © Copyright by Adam Hunter Mepham 2019

Upload: others

Post on 12-Jul-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Micro- and Nano-Scale Approaches for Disease Detection and Characterization

by

Adam Hunter Mepham

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy

Institute of Biomaterials and Biomedical Engineering University of Toronto

© Copyright by Adam Hunter Mepham 2019

ii

Micro- and Nano-Scale Approaches for Disease Detection and

Characterization

Adam Hunter Mepham

Doctor of Philosophy

Institute of Biomaterials and Biomedical Engineering

University of Toronto

2019

Abstract

Devices that diagnose and characterize disease have the potential to greatly improve healthcare

worldwide. This thesis explores a number of different elements pertaining to device design and

application. A microfluidic device for the capture and profiling of circulating tumor cells (CTCs)

is tested against a rabbit model of cancer. This device demonstrates both an increase in CTC load

and aggressiveness which correlates with traditional computed tomography measurements. CTC

biology is also shown to differ markedly from tumour precursor cells. Next, a study of gold

microelectrode architecture is performed with the aim of improving performance towards

biosensing. A unique regime of gold ion concentration, applied voltage, and electrolyte viscosity

is determined which drives the assembly of a highly structured morphology. Further studies

illustrate growth mechanisms and the sensitivity of the electrode towards biomolecule detection.

Additionally, a microfluidic device for instrument-free manipulation of microscopic fluid

quantities is developed. This design allows the metering and dispensing of reagents in an intuitive

manner by combining a series of capillary valves and a simple push-button. This “Digit Chip” is

applied to the detection of antibacterial susceptibility alongside a simple smart-phone based

fluorimeter. Together these studies explore the application of electrochemical and microfluidic

modalities to the realm of disease monitoring.

iii

Acknowledgments

I would like to begin by thanking Professor Shana Kelley for her support throughout the tenure of

my PhD. Her unique blend of creativity and practicality helped me to pursue interesting projects

which nonetheless had clear and important applications in the world of biotechnology. Moreover,

her ability to assemble a group of friendly and intelligent individuals helped to create a warm and

interesting environment in which to practice research.

I would further like to thank the members of my advisory committee, Professor Ted Sargent,

Professor Aaron Wheeler, and Professor Axel Guenther. They all helped to keep me motivated

and see the larger picture in my work. Their guidance served me well in planning out my projects

and keeping my destination clear.

Next I would like to thank the innumerable labmates who helped to make my rough days tolerable

and my good days even better. My appreciation goes out to Dr. Jagotamoy Das, Dr. Mahmoud

Labib, Dr. Brenda Green, Dr. Libing Zhang, Surath Gomis, Fan Xia, Bill Duong, Zongjie Wang,

Dr. Ivaylo Ivanov, Laili Mahmoudian, Dr. David Tulumello, Dr. Brian Lam, Dr. Mahla Poudineh,

Dr. Andrew Sage, Dr. Sae Rin Jean, Dr. Simon Wisnovsky, Dr. Leyla Kermanshah, Dr. Yige Zhou,

Dr. Ying Wan, Dr. Sam Chang, Dr. Julie Shi, Alexandre Zaragoza, Thy Vu, and Thaddeus “Thad”

Gibbs. Special thanks to those who provided invaluable assistance in my publications, Dr. Reza

Mohamadi, Dr. Sharif Ahmed, Dr. Justin Besant, Dr. Ian Burgess, Dr. Sahar Mahshid and Dr. Sara

Mahshid. Further thanks to those members who keep the Kelley lab running smoothing;

specifically, Dr. Mark Pereira, Bob Christensen and especially Barbara Alexander. Finally, a shout

out to those who helped keep my days fun and interesting, Dr. Wendi Zhou, Peter Aldridge,

Wenhan Liu, Carine Nemr, Tanja Sack, Jenise Chen, David Philpott, Xiaolong Yang, Cindy Ma,

Hanie Yousefi, Dr. Eric Lei, , Dr. Tina Saberi Safaei, and Dr. Sarah Smith.

Last but not least, I would like to thank my family for all of their support over the years. I would

not have survived without the steady supply of peanut butter and almonds which fueled my body

and mind. I want to thank my parents, Peter and Linda, who have both encouraged me to pursue

what I enjoy. My brother, who called me a nerd. And my entire extended family, who pretend that

they understand when I ramble on about my work.

iv

Table of Contents

Acknowledgments.......................................................................................................................... iii

Table of Contents ........................................................................................................................... iv

List of Tables ............................................................................................................................... viii

List of Figures ................................................................................................................................ ix

List of Abbreviations .................................................................................................................... xii

Introduction .....................................................................................................................1

1.1 Point-of-Care Diagnostics ....................................................................................................1

1.2 Electrochemistry ..................................................................................................................3

1.3 Electrode Properties and Design ..........................................................................................6

1.4 Microfluidics ........................................................................................................................7

1.5 Scope of Thesis ....................................................................................................................9

Sorting of Circulating Tumor Cells during Disease Progression in an Animal Model

...................................................................................................................................................11

2.1 Abstract ..............................................................................................................................12

2.2 Introduction ........................................................................................................................12

2.3 Results and Discussion ......................................................................................................14

2.3.1 Animal study and microfluidic chip design ...........................................................14

2.3.2 Characterization of device function .......................................................................16

2.3.3 Measurement of tumor volumes and analysis of metastatic lymph nodes.............17

2.3.4 Correlation of CTC levels with disease progression..............................................18

2.3.5 Analysis of CTC sorting patterns during tumor growth ........................................20

2.3.6 Discussion ..............................................................................................................22

2.4 Conclusion .........................................................................................................................24

2.5 Methods..............................................................................................................................25

v

2.5.1 Animal model.........................................................................................................25

2.5.2 Tumor cell line propagation ...................................................................................25

2.5.3 CT imaging and image analysis .............................................................................25

2.5.4 Histopathological evaluation .................................................................................26

2.5.5 Capture of cell lines ...............................................................................................26

2.5.6 Immunocytochemistry of cell lines ........................................................................26

2.5.7 Isolation and fluorescent staining of CTCs ............................................................26

2.5.8 Capture efficiency of white blood cells .................................................................27

2.5.9 Image scanning and analysis ..................................................................................27

2.5.10 Flow cytometry ......................................................................................................27

2.5.11 Microchip fabrication.............................................................................................27

Mechanistic Control of the Growth of Three-Dimensional Gold Sensors ....................29

3.1 Abstract ..............................................................................................................................30

3.2 Introduction ........................................................................................................................30

3.3 Results and Discussion ......................................................................................................32

3.3.1 Outline of experiments ...........................................................................................32

3.3.2 Effects of solution conditions of 3D nanostructured microelectrode growth ........33

3.3.3 Nucleation and growth mechanism of 3D gold structures .....................................39

3.3.4 Study of current-time transients .............................................................................39

3.3.5 Analysis of i2/im2 vs t/tm .........................................................................................41

3.3.6 Surface area of nanostructured 3D microelectrodes ..............................................45

3.3.7 Study of DNA hybridization efficiency on the surface of 3D microelectrodes .....46

3.4 Conclusion .........................................................................................................................48

3.5 Methods..............................................................................................................................48

3.5.1 Chip fabrication .....................................................................................................48

vi

3.5.2 Electrodeposition ...................................................................................................48

3.5.3 Surface characterization .........................................................................................49

3.5.4 Nucleation and growth model ................................................................................49

3.5.5 Electrode functionalization ....................................................................................50

3.5.6 Sensor measurements .............................................................................................50

Power-free, digital and programmable dispensing of picoliter droplets using a Digit

Chip ...........................................................................................................................................51

4.1 Abstract ..............................................................................................................................52

4.2 Introduction ........................................................................................................................52

4.3 Results and discussion .......................................................................................................54

4.3.1 Overview of the digit chip .....................................................................................54

4.3.2 Bursting pressure model and design principles .....................................................56

4.3.3 Optimization of device geometry...........................................................................57

4.3.4 Designing a user-friendly interface ........................................................................59

4.3.5 Generation of a discrete concentration gradient ....................................................62

4.3.6 A low-cost platform for rapid determination of bacterial antibiotic susceptibility

................................................................................................................................63

4.3.7 Discussion ..............................................................................................................65

4.4 Conclusion .........................................................................................................................67

4.5 Methods..............................................................................................................................67

4.5.1 Digit chip fabrication .............................................................................................67

4.5.2 Fabrication of user-friendly interface ....................................................................67

4.5.3 Contact angle measurements..................................................................................67

4.5.4 Bursting pressure measurements............................................................................68

4.5.5 Measurements of applied pressure using the elastomeric button ...........................68

4.5.6 Chamber filling percentage measurements ............................................................68

vii

4.5.7 Generation of a discretized concentration gradient ...............................................68

4.5.8 Antibiotic susceptibility testing .............................................................................69

4.5.9 Fluorescence image acquisition and analysis ........................................................69

Conclusions and Future Outlook ...................................................................................70

5.1 Thesis Findings ..................................................................................................................70

5.2 Future Outlook ...................................................................................................................71

References ......................................................................................................................................74

Single Cell Capture Device .......................................................................................83

A.1 Background ........................................................................................................................83

A.2 Device Design ....................................................................................................................84

Supporting Information ...........................................................................................104

Supporting Information for Chapter 2 .............................................................................104

Supporting Information for Chapter 4 .............................................................................107

viii

List of Tables

Table A-1 Truth table for the multiplexer. .................................................................................... 98

ix

List of Figures

Figure 2-1 Overview of the study monitoring CTC heterogeneity in a rabbit cancer model. ...... 14

Figure 2-2 Characterization of capture. ........................................................................................ 16

Figure 2-3 VX2 tumor growth in New Zealand rabbits. ............................................................... 17

Figure 2-4 CTC sorting and analysis. ........................................................................................... 19

Figure 2-5 Comparison of sorting profile ..................................................................................... 22

Figure 3-1 Schematic of gold microelectrode experiemnts .......................................................... 32

Figure 3-2 Effects of concentration, viscosity, and voltage on electrodeposition of gold

microsensors are studied using scanning electron microscopy (SEM). ........................................ 35

Figure 3-3 Analysis of the interior structures of electrodeposited gold using FIB. ...................... 38

Figure 3-4 I-t curves during electrodeposition.............................................................................. 40

Figure 3-5 Nucleation during electrodeposition ........................................................................... 42

Figure 3-6 Two-dimensional time-varying simulation results for deposition of Au .................... 44

Figure 3-7 Effect of deposition overpotential on surface nanostructuring. .................................. 45

Figure 3-8 DNA detection assay based on 3D gold microsensors................................................ 47

Figure 4-1 Overview of the Digit Chip. ........................................................................................ 55

Figure 4-2 Experimental investigation of the device geometry and its optimization. .................. 58

Figure 4-3 The Digit Chip interface for controlled dispensing of droplets. ................................. 60

Figure 4-4 Generation of a discretized concentration gradient. .................................................... 62

Figure 4-5 Testing of antibiotic susceptibility. ............................................................................. 63

x

Figure A-1 Structure of a microfluidic weir. ................................................................................ 85

Figure A-2 A microfluidic weir with a parallel shunt channel. .................................................... 86

Figure A-3 Equivalent circuit diagram for a sequence of weir-shunt pairs. ................................. 87

Figure A-4 Encouraging flow to enter the weir. ........................................................................... 87

Figure A-5 Reducing shunt resistance by translating cells across streamlines. ............................ 88

Figure A-6 Design of the gravity driven cell alignment and capture device, which is operated in

the vertical orientation. ................................................................................................................. 90

Figure A-7 Function of the bleeding flow motif. .......................................................................... 91

Figure A-8 Two stage cell aligner. ............................................................................................... 92

Figure A-9 First “bleeding flow” element. ................................................................................... 93

Figure A-10 End of the first “bleeding flow” element. ................................................................ 93

Figure A-11 Reintroduction of removed fluid. ............................................................................. 94

Figure A-12 Second “bleeding flow” element. ............................................................................. 95

Figure A-13 The cell alignment module interfaced with the weir and shunt capture module. ..... 96

Figure A-14 Single cell captured in weir trap. .............................................................................. 96

Figure A-15 Photograph of integrated devce including membrane deflection valves. ................. 97

Figure A-16 Complete design of integrated devce. ...................................................................... 97

Figure A-17 Pressure manifold with three 4-way valves (left) and two 3-way valves (right). .... 99

Figure A-18 Pressure manifold feeding pressurized water reservoirs. ......................................... 99

Figure A-19 Cell capture region of the devce including all 8 valves.. ....................................... 100

xi

Figure A-20 A single trap in capture mode. ............................................................................... 101

Figure A-21 A single trap in incubation mode. .......................................................................... 102

Figure A-22 A single trap in release mode. ................................................................................ 103

Figure B-1 Nanoparticle characterization. .................................................................................. 105

Figure B-2 Cell line flow cytometry. .......................................................................................... 106

Figure B-3 VX2 EpCAM flow cytometry. ................................................................................. 106

Figure B-4 The critical advancing contact angle of PBS on PDMS as a function of oxygen plasma

exposure time. ............................................................................................................................. 107

Figure B-5 Accuracy of filling as a function of chamber size. ................................................... 107

Figure B-6 A replication of Figure 4-2 F-H using buttom pressess rather than using a hydraulic

pump, .......................................................................................................................................... 108

xii

List of Abbreviations

BSA: Bovine serum albumin

CT: Computed tomography

CTC: Circulating tumor cell

DI: Deionized

EMT: Epithelial-to-mesenchymal transition

EpCAM: Epithelial cell adhesion molecule

FDA: Food and drug administration

FIB: Focused ion beam

H&E: Hematoxylin and eosin

HBSS: Hank’s balanced salt solution

HCG: Human chorionic gonadotropin

LED: Light-emitting diode

MACS: Magnetic-activated cell sorting

MCH: Mercaptohexanol

PBS: Phosphate-buffered saline

PDMS: Polydimethylsiloxane

POC: Point-of-care

RGB: Red, green, blue

SEM: Scanning electron microscope

SWV: Square-wave voltammetry

TCEP: Tris(2-carboxyethyl)phosphine

VOI: Volume of interest

WBC: White blood cell

1

Introduction

1.1 Point-of-Care Diagnostics

There is currently a shift occurring in the realm of disease diagnostics. This transformation is

harnessing the power of innovative, more precise technologies to provide deeper and more

personalized insight into the disease state and to facilitate the development of treatments more

closely tailored to the unique needs of individual patients. Moreover, these approaches

significantly expedite testing and allow for more direct communication between clinical physicians

and patients. This represents a marked shift from previously recognized models of medical care.

The gold standard for medical testing has historically been laboratory testing.1 This approach

served the needs of patients well throughout the 20th century, as the application of medicine

became more systematic. Traditional laboratory tests were performed by trained professionals in

a controlled setting, and results were later communicated to clinicians and, in turn, patients. This

approach allowed the powerful techniques developed in the laboratory, such as bacterial culture

and genetic testing, to be brought to bear on the problems of identifying illnesses and determining

prognosis and appropriate treatments. This paradigm allowed for batch testing of large numbers of

samples, and provided accurate and actionable information.

Despite its utility, the traditional laboratory-based model suffers from a number of critical flaws.

Firstly, these tests often take a considerable amount of time. For example, traditional approaches

to identifying bacterial infections require a number of steps which can take multiple days to

complete.2 These typically include a culture stage, which serves to grow bacteria to a suitable

number, as well as multiple additional tests such as disc-diffusion assays and mass-spectrometric

testing. In critical situations, such as bacterial sepsis, such a delay can have a severe impact on

patient mortality and morbidity.

2

Secondly, testing is performed at a central laboratory, physically removed from the patient and

clinician. This has two major drawbacks. First, additional time is wasted as the samples must be

transferred to the central location prior to testing, and, in turn, the results must be communicated

to the patient. Additionally, due to the requirement of a large, centralized laboratory, this approach

is not feasible in regions where existing infrastructure is unavailable, such as low-resource

settings.3

Finally, laboratory testing is devised to have utility for the mass population, and as such is typically

not designed to address the idiosyncrasies of particular patients. As the importance of subtle

differences in diseases such as cancer have become more apparent, the need for tests that can give

deeper insight into each patient and individualize patient care has become more concrete.4 These

weaknesses in the traditional approach to probing disease have motivated the development of novel

diagnostic approaches.

Flaws in the existing paradigm for diagnostic testing are being addressed by the development of

point-of-care (POC) devices. POC devices are defined to be diagnostic devices which are used at

or near the location where patient care is administered. This approach has a number of benefits

when compared to centralized laboratory testing. POC testing is typically much faster than

laboratory assays, reducing the time required for diagnosis by as much as an order of magnitude.1

Such devices are also designed to require minimal external equipment, allowing for their use in

low-resource settings. Furthermore, a number of POC devices are designed to detect rare genetic

or protein markers, which allows treatment to be customized on a patient-by-patient basis.5

Most POC devices are examples of biosensors. A biosensor is defined as an apparatus that couples

a biological sensing element to a transducer in order to detect or quantify chemical species,

typically in a sample of biological origin.6 There are a wide variety of useful sensing elements and

transducers, allowing for a large number of possible device designs. The most important property

of a sensing element is the ability to recognize the chemical species of interest, especially in the

presence of a complex sample matrix. As such, two prominent classes of sensing elements are

proteins and nucleic acids. Among proteins, antibodies and enzymes are the most used, due to their

tendency to bind/catalyze a specific target or family of targets.7,8 For nucleic acids, DNA is often

chosen due to its ability to form a duplex with complementary strands of interest.9 DNA has a

number of excellent properties for a biosensor. It can detect both other DNA strands (e.g. genomic

3

DNA) and RNA (e.g. mRNA). It is also highly sensitive to nucleotide sequence, allowing for the

discrimination of mutant strands. Finally, it is stable and can be easily modified with chemical

handles, allowing its incorporation into diverse device architectures.10 In addition, a multitude of

transducers exist which lend themselves to varied applications. Each different transducer type is

tied to a different physicochemical phenomenon and provides a different method of signal readout.

Optical biosensors rely on changes in absorbance, fluorescence or luminescence, for detection

either by a photodetector or the human eye.11 Mechanical biosensors exploit changes of properties

such as resonance frequencies or exerted forces upon the binding of target molecules.12

Electrochemical biosensors measure changes in voltage, impedance or current produced by the

presence of the target.13

One POC biosensor that illustrates the utility of this modality is the home pregnancy test.14 This

test detects the presence in urine of human chorionic gonadotropin (HCG), a peptide hormone

produced by the placenta. Devices employ an antibody specific against HCG. In a traditional

pregnancy test, this antibody is tethered to an enzyme which catalyzes the reaction of a

chromogenic substrate. The presence of HCG localizes the antibodies to a particular strip on the

device, in turn localizing the chromogenic reaction and allowing the visual readout of a colored

band. The pregnancy test is a lateral flow assay, with the flow of urine through the device driven

by capillary action through a porous flow pad. This approach allows for the passive movement of

fluid along a device without the need for an active pumping mechanism.

Another POC biosensor is the handheld glucose monitor.15 This device quantifies the levels of

glucose in the blood, primarily for use by diabetic patients. The sensing element employed is an

enzyme, typically either glucose oxidase or glucose dehydrogenase. These enzymes catalyze the

oxidation of glucose and the concomitant reduction of an electrochemically-active mediator. This

mediator is in turn detected at an electrode, with the current produced ultimately proportional to

the glucose concentration. This electrochemical device is one of a burgeoning number of

biosensors that utilize electrochemical events as output.

1.2 Electrochemistry

Electrochemistry is a field that exists at the intersection of electricity and chemistry. Specifically,

it involves the study of the influence of electrical phenomena on chemical reactions, and,

conversely, the production of electrical current and potentials as a result of chemical reactions.

4

The underlying reason for this interrelation is the fact that all matter is, at the atomic level,

comprised of charged particles. Consequently, electric fields can influence the energy of different

chemical states, and therefore the application of electric fields can cause molecular composition

and molecular distribution to change accordingly.

The fundamental reaction underlying electrochemistry is the redox reaction. In this reaction, one

of the components of a system gains electrons (and is thus reduced), while the other loses electrons

(and is thus oxidized). This reaction can occur homogenously- for example, between two different

solute molecules within a solution- or heterogeneously, between molecules and an electrode. In

either case, electrons will flow in the direction that allows them to reach a lower energy state.

Most electrochemistry is centered on the idea of the electrochemical cell. An electrochemical cell

is comprised of at least two electrodes and, typically, a conductive electrolyte solution. Such cells

can behave in either a galvanic or electrolytic mode. A galvanic cell is one in which a spontaneous

chemical reaction produces electrical energy, which can in turn be measured or employed to

perform work. An example of a galvanic cell is a traditional battery. By contrast, an electrolytic

cell is one in which an externally generated potential is applied in order to promote chemical

reactions that would otherwise not be spontaneous. The electrolysis of water to produce hydrogen

and oxygen gas is a classic example of an electrolytic cell.

Electrochemical cells must contain at least two electrodes due to the paired nature of

electrochemical reactions: if electrons are flowing out of one electrode to reduce one chemical

species, an equal number of electrons must be flowing into a second electrode as a second chemical

species is oxidized. The electrode through which electrons are entering the electrolyte is known as

the cathode, whereas the electrode via which electrons are leaving the electrolyte is known as the

anode.

The majority of activity in electrochemical cells occurs at the interface between phases, such as

the interface between the electrode and the electrolyte. This interface marks the division between

two difference conductors, one electronic and one ionic. Due to the markedly different chemical

environments, the energy of an electron in each of the phases will tend to be different. Hence, if

the interface is to be at equilibrium, a potential difference must arise which offsets this chemical

difference in electronic energy and allows the total electrochemical energy (electrical + chemical

energy) to be equal. This potential difference is the driving force for the flow of electrons. In the

5

two-electrode cell, the potential between the two electrodes will be comprised of two parts: the

potential between the first electrode and the electrolyte, and the potential between the electrolyte

and the second electrode. This potential difference is produced by an electrical double layer which

exists at the electrode/liquid interface. At the electrode surface, at excess or dearth of electrons

produces a net charge. This charge is balanced by a diffuse layer of oppositely charged ions in

solution. Specifically adsorbed ions may also contribute charge to the interface. Changes in

potential modify both the charge density of the interface and the quantity/type of ions adsorbed,

leading to a transient capacitive current. Since such currents do not involve the transfer of charge

across the interface, they are known as non-faradaic currents, in contrast with the faradaic current

of typical redox reactions.

The paired nature of electrodes introduces a difficult problem for the study of electrochemistry.

Most of the time, it is only the chemical reaction that is occurring at one of the electrodes that is

the subject of interest (the “working” electrode). However, the potentials and currents involved are

typically dependent on nature of both of the electrodes present. In fact, the potential of an electrode

can only be defined in relation to a second electrode. This has necessitated the development of

reference electrodes. Reference electrodes are electrodes with a carefully controlled chemical

composition. Since the potential difference between the electrolyte and the electrode is a function

of the presence and arrangement of molecules, this serves to keep this potential difference fixed.

Consequently, any potential differences between the reference electrode and the electrode of

interest can be ascribed to changes at the second electrode/electrolyte interface.

When measurements are being performed at equilibrium, such a simple two electrode system is

functional. However, complications arrive in non-equilibrium states when current is flowing. Since

charge conservation necessitates that current flow through the reference electrode as well as the

working, the potential of the reference electrode must shift from its equilibrium value. In general,

reference electrodes are designed to be non-polarizable. This means that they are capable of

supplying large quantities of current in either direction without a significant change in voltage.

However, in order to remove this effect entirely, a three-electrode system can be employed. This

setup introduces a counter electrode, in addition to the working and reference electrodes. The

voltage is controlled between the working electrode and the reference electrode; however, no

current flows through the reference electrode. Instead, the counter electrode passes an equal but

opposite current to the working electrode, completing the circuit. In general, the reaction occurring

6

at the counter electrode is not controlled, and as such may involve any of the chemical species

present.

Three-electrode systems can be used for a wide variety of different experiments. In potentiometry,

the potential between electrodes is measured. In voltammetry, the potential is controlled (either

held constant or varied) and the resultant current is measured. When the potential is fixed, the

technique is known as amperometry. By contrast, potential can be varied in a large number of

ways, both cyclic and acyclic. One such technique is square-wave voltammetry (SWV).16 The

potential in SWV is a superposition of a staircase voltage and a square wave. In each period two

measurements of current are made; one towards the end of the positive voltage step and one

towards the end of the negative voltage step. Delaying the measurement serves to separate the two

types of current, faradaic and non-faradaic. Since non-faradaic processes are generally transient,

they decay much more rapidly than faradaic ones. As such, waiting for non-faradaic currents to

decay to negligible values (using on the order of milliseconds) allows faradaic currents to

dominate. Generally, forward and reverse measurements are subtracted from one another to give

the final trace. This subtraction serves two purposes: it increases the magnitude of the peak due to

the difference in sign of the forward and reverse currents, and it helps to remove the interference

of faradaic currents from contaminants in the solution.

1.3 Electrode Properties and Design

Metal electrodes are widely employed as working electrodes, due to their excellent conductivity

and ease of manufacture. A number of properties of metal electrodes can strongly influence their

behavior, including electrode size/shape,17 material,18 and crystallinity19. The size of an electrode

dictates a number of important properties. In recent years, there has been a push towards producing

smaller electrodes, so-called ultramicroelectrodes.20 Such electrodes deliver enhanced

performance by reducing double-layer charging time, minimizing ohmic loss of potential, and

allowing for a hemispherical diffusion pattern to improve mass delivery. They are ideal for POC

devices, where a small footprint is important and the ability to concentrate analytes of interest into

a small region serves to improve sensitivity. Structure at a finer scale can also serve important

function, and three-dimensional electrodes can interact with a greater amount of solution than flat,

two-dimensional electrodes. Dendritic forms on the nano- or micro-scale serve to enhance surface

area without required enlarged footprint, and may also dictate the geometry of molecular binding.21

7

The material used strongly influences the nature of the interface. The kinetics of redox reactions

are a strong function of electrode material, and the use of certain metals over others can control

the relative rate of reactions in a manner largely independent of thermodynamic favourability. For

example, the reduction of carbon monoxide can proceed along a large number of different routes,

producing a variety of carbon-based structures. Which molecules are formed and in what

proportions is dictated primarily by the nature of the electrode, with judicious selection of electrode

material allowing for enrichment in molecules of interest.22 Similarly, the metal employed

influences which surface modifications can be made, such as the tethering of biological

molecules.23 This phenomenon further extends to the crystal facet of the metal.24 Since it is the

outer surface of the metal which interacts with the solution, it is the properties of this facet rather

than the bulk material which dictate behaviour. Different crystal planes have different surface

energies and different affinities for adsorbents. In the case of polycrystalline electrodes, the

observed effect is the superposition of all of the composite facets.

1.4 Microfluidics

One technology that has been applied extensively in diagnostic devices is microfluidics.

Microfluidic devices are those that serve to manipulate fluids and have a least one critical

dimension on the micrometre scale. Fluid manipulation in this regime is marked by a number of

deviations from macroscopic behaviour. In general, microfluidic devices have a very low Reynolds

number.25 The Reynolds number is an indicator of the ratio of inertial forces to viscous forces. Due

to the relatively low inertia of the flow, the fluid moves as a series of layers or lamina (laminar

flow). Consequently, convective mixing is largely absent in these systems and diffusion is the

primary mechanism by which mixing occurs.26 This allows for the very precise manipulation of

fluid lamina and molecules/particles present therein, facilitating controlled movement of

substances.27 Furthermore, due to the small volumes of fluid being manipulated, forces which scale

with surface area have proportionally greater impact. Surface tension is one such force.28 Surface

tension arises as a consequence of the increase in free energy at the interface between difference

phases. This tension acts within the plane of the interface (tangentially) and acts to minimize the

area of contact. When multiple phases exist in concert, a complicated series of forces arises which

seeks to reduce the interphase between low affinity phases and increase the interface between high

affinity phases. Although this phenomenon can produce unwanted consequences, careful design

can take advantage of these naturally arising forces to drive desired behaviors. These approaches

8

can often allow for more passively actuated devices, whose behaviours are guided by the chemical

nature of the components.29

A large number of techniques exist to produce microfluidic devices. The preferred technique

depends on the requisite spatial resolution as well as the number of devices required and concurrent

cost constraints. One of the most common techniques for microfluidic fabrication is

photolithography. Photolithography depends at its core on the ability of certain chemicals

(photoresists) to undergo a chemical change upon exposure to particular wavelengths of light.

Positive photoresists are those whose chemical bonds are weakened by light exposure, whereas

negative resists are strengthened. By passing light through a patterned photomask, the pattern can

be transferred to the photoresist and revealed by chemical development. Devices with very fine

features may require e-beam lithography, a technique which boasts resolutions as fine as 10 nm.30

Looser precision can be achieved by performing optical photolithography on silicon or glass

substrates.31 If resolutions >5 μm are adequate, soft lithography is a very useful technique.32 In

soft lithography, a mold is made using traditional lithographic techniques, typically using a

negative photoresist such as SU-8. A silicone elastomer, such as PDMS, is then poured onto the

mold and allowed to cure. This elastomer device can then be peeled off the mold and adhered to a

substrate, allowing for repeated use of the mold. PDMS has been widely employed in microfluidic

devices for biological samples due to its biocompatibility.33

The scale of microfluidic devices makes them ideally suited for the manipulation of cells (~10 μm

in diameter). This size compatibility, along with the convenience of laminar flow for deterministic

particle movement, has encouraged the development of a wide range of devices designed for cell

samples. Two such categories are devices for cell separation and devices for single cell isolation.

Cell separation devices serve to remove cells of interest from a sample matrix that contains

extraneous cells or other unwanted contaminants. These depend on exploiting characteristics of

the target population which are not shared by confounding populations, such as physical properties

(size, shape, deformability, etc.) or biochemical properties (surface markers, mRNA, etc.).34 In the

case of discrimination by surface markers, cells are often labelled by antibodies or aptamers which

are tethered to a convenient tag. These tags may be fluorescent, magnetic, dielectric, or actuated

by another means.35 By either active or passive separation the cells bearing this marker can then

be spatially separated from those without. This approach has wide applicability across biology,

allowing for the purification of stem cells and circulating tumor cells, among other rare cell

9

populations.36 Microfluidics also enables the fine positioning of cells, allowing for controlled

treatment and culture of cells isolated on an individual basis.37 Such devices have gained increased

interest given recent paradigm shifts in cell biology, driven by the recognition of subtle

heterogeneities within cell populations.38–40 These differences are especially pertinent in tumour

biology, where the genotypic/phenotypic differences between cells can have drastic effects on their

pathogenic potential.41,42

1.5 Scope of Thesis

This thesis seeks to investigate a number of topics pertaining to the development of devices as

disease diagnostics and as tools for the study of disease biology. It spans microfluidic design and

application, the fine tuning of electrodes towards the detection of biomolecules, and the first steps

towards a POC devices for use in low-resource settings.

In chapter 2, a microfluidic device is described that monitors the properties of circulating tumor

cells in an animal model of cancer. This device captures and sorts CTCs, profiling them on the

basis of the EpCAM surface marker as a surrogate for cell invasiveness. The results show a marked

increase in the number and aggressiveness of CTCs, in good agreement with CT scans of the

primary tumor and secondary metastases. We also illustrate the temporary reprieve following the

resection of the main tumor, alongside the return to an aggressive state as secondary tumors re-

establish the metastatic state.

In chapter 3, we explore the mechanisms driving the morphogenesis of gold microelectrodes

during electroplating and the requisite parameters to fine-tune electrode shape and performance.

The effects of gold ion concentration, applied voltage, and electrolyte viscosity are tested and the

regime for a high surface area structure is established. Further investigations into crystal structure

and theoretical modelling help to elucidate the underlying mechanisms. Finally, the electrodes are

challenged with the detection of DNA strands and performance is tested as a function of

morphology.

In chapter 4, our attention shifts to the design of a device for microfluidic manipulation without

the need for external instrumentation. This device, termed the “Digit Chip”, is based on the

underlying technology of the capillary valve. First we delve into the parameters that control the

reliability and bursting pressure of the valves, determining an optimal geometry. We then devise

10

a simple “push button” control which allows the user to actuate the valves in an intuitive manner

using only their finger. This completed platform is first applied to a simple task of producing a

concentration gradient. Following this, a more advanced design is used in conjunction with a cheap

cell phone-based fluorescence detector to perform antibiotic susceptibility testing of bacteria.

Finally, in chapter 5, we discuss the possible future directions of the research. The potential for

the combination of the various platforms is examined and the position of this work in the field as

a whole is outlined.

11

Sorting of Circulating Tumor Cells during Disease Progression in an Animal Model

In this chapter we explore the application of a microfluidic device previously developed in our lab

to monitoring the progression of cancer in a rabbit model. This microfluidic device extracts CTCs

from rabbit blood samples by labelling them with magnetic nanoparticles and applying a strong

external magnetic field. By varying the drag force present along the length of the device, it is tuned

so that cells with a large amount of the epithelial marker EpCAM are captured in earlier zones,

and cells with a lower amount of EpCAM are captured in later zones. Since reduced epithelial

expression has been correlated with more aggressive cancer, we are able to observe changes in

cancer biology. Over the course of the experiment the tumor growth is monitored and eventually

the tumour is excised, allowing us to monitor changes in CTC number and distribution across the

natural course of disease development as it might occur in a patient.

Reprinted with permission from Muhanna N.*, Mepham A.*, Mohamadi R. M., Chan H., Khan

T., Akens M., Besant J. D., Irish J., Kelley S. O., “Sorting of Circulating Tumor Cells during

Disease Progression in an Animal Model” Nanomedicine: Nanotechnology, Biology and Medicine,

2015, 11(7) 1613-20.

Link to publication online: https://doi.org/10.1016/j.nano.2015.04.017

Disclosure of work within this manuscript: A.M., N.M., and R.M.M. designed the experiments.

A.M performed microfluidic experiments and NM performed traditional hospital testing with the

assistance of T.K, H.C. and M.A. Data analysis and manuscript writing were performed by A.M.,

N.M., and T.K. with assistance from R.M.M., J.D.B., J.I. and S.O.K.

12

2.1 Abstract

Circulating tumor cells (CTCs) can be used as markers for the detection, characterization, and

targeted therapeutic management of cancer. We recently developed a nanoparticle-mediated

approach for capture and sorting of CTCs based on their specific epithelial phenotype. In the

current study, we investigate the phenotypic transition of tumor cells in an animal model and show

the correlation of this transition with tumor progression. VX2 tumor cells were injected into

rabbits, and CTCs were evaluated during tumor progression and correlated with computerized

tomography (CT) measurements of tumor volume. The results showed a dramatic increase of

CTCs during the four weeks of tumor growth. Following resection, CTC levels dropped but then

rebounded, likely due to lymph node metastases. Additionally, CTCs showed a marked loss of the

epithelial cell adhesion molecule (EpCAM) relative to precursor cells. In conclusion, the device

accurately traces disease progression and CTC phenotypic shift in an animal model.

2.2 Introduction

Current cancer staging methods inadequately predict tumor prognosis and response to therapy,

thus underscoring the need for new tumor characterization approaches.43,44 The heterogeneous

nature of tumors and difficulty in identifying metastases during early stages of cancer confound

the ability to predict tumor prognosis and determine appropriate therapy.45,46 A promising

approach, which may serve to address these problems, is the analysis of circulating tumor cells

(CTCs).

CTCs are putative precursors of metastases.47 Specifically, they are cells which are released from

the perimeter of the tumor and intravasate into the blood stream.48 These cells then circulate until

they encounter an appropriate niche, at which time they may extravasate into the surrounding

tissue. Rapid division then establishes a secondary tumor, which can ultimately produce its own

CTCs and continue this process.

Since CTCs are derived from primary tumors and appear to be the cells that establish metastatic

sites, they can provide a wealth of information regarding specific tumor biology and the driving

factors behind invasive disease.49,50 Furthermore, numerous studies have shown that CTCs in

blood samples may be used as a marker to predict survival and prognosis in metastatic cancer

13

patients.51,52 CTC levels appear to be correlated with disease spread,51–53 and elevated CTCs are

associated with poor prognosis and increased probability of metastasis.54

Importantly, CTCs represent the biological state of the evolving disease. A significant aspect of

CTC biology, which has gained considerable traction in recent years, is the loss of epithelial

character accompanying cancer progression during a process known as epithelial-to-mesenchymal

transition (EMT).55 This process involves a cellular reprogramming event that causes CTCs to

assume a more invasive phenotype relative to the epithelial phenotype prominent in the primary

tumor.48 As a consequence, epithelial surface markers such as EpCAM are down-regulated.56

Although the consequences of this transition are not well understood, it has been implicated in

increased metastatic potential, possibly stemming from increased cell mobility due to loss of

adhesion molecules.57 Furthermore, the population of CTCs with low levels of epithelial marker

expression, including EpCAM, has been shown to peak during times of disease progression as

compared to treatment response.55 Thus monitoring changing levels of epithelial markers, the most

prominent of which is EpCAM, should provide valuable insight into cancer progression and

metastasis.

Numerous studies have successfully captured CTCs from diverse tumor origins including lung,58

prostate,58 head and neck,59 melanoma,60 gastric and pancreatic cancers.58,61 However, despite

numerous advances in CTC capture techniques, the analysis of CTCs is still not part of routine

tumor staging in clinical practice. In fact, the CellSearch system is the only existing Food and Drug

Administration (FDA) cleared platform.62 Furthermore, the vast majority of CTC detection

methods simply count the absolute number of CTCs, without distinguishing between varieties of

CTC subpopulations. Consequentially, potentially valuable information may be overlooked.

Recently, we developed a new approach that provides a means to capture and classify of CTCs

with high sensitivity and selectivity using immunomagnetic nanoparticles captured within a novel

microfluidic device.63,64 Our CTC isolation technique depends on antibodies against EpCAM

attached to magnetic nanoparticles allowing for capture using a magnetic field. This device

spatially sorts CTCs on the basis of EpCAM expression, thus providing insight into differential

expression of epithelial markers. EpCAM is expressed by a wide variety of epithelial tumors and

is a generally accepted marker of CTCs,58 and is one of the markers known to show a significant

decrease during EMT, thus establishing it as a surrogate marker of this process.55 Sorting cells

14

according to EpCAM expression therefore provides a means to monitor phenotypic changes in

CTCs. This device was shown to allow the profiling of CTC subpopulations with differing

epithelial character in samples collected from prostate cancer patients. Here, we report the

application of this device to an animal model of cancer. Using a rabbit host and the VX2 tumor

model, we monitored the epithelial character of CTCs during tumor growth and following

resection of the tumor. A significant change in CTC profile is observed from more epithelial to

less epithelial as tumors progress. This is the first study to monitor these dynamics in an animal

model of cancer.

2.3 Results and Discussion

2.3.1 Animal study and microfluidic chip design

The overall timeline of the study is shown in Figure 2-1 A. Following tumor induction, CT scans

and CTC analysis were performed bi-weekly until the 4th week post-tumor induction. At this point,

the primary tumor was resected and CT scan/CTC analysis continued until terminal surgery during

the 6th week.

Figure 2-1 Overview of the study monitoring CTC heterogeneity in a rabbit cancer model.

(A) Timeline showing tumor induction, survival surgery and terminal surgery. (B) Microfluidic

CTC capture and sorting strategy. Cells are labeled with paramagnetic nanoparticles conjugated

to anti-EpCAM antibodies. Labeled cells are introduced into the microfluidic device, where they

are captured in low velocity regions (velocity valleys, shown in blue) adjacent to X structures.

15

(C) A device capable of capturing and sorting CTCs based on expression levels consists of 4

zones with increasing cross-sectional area allowing for capture of cells with high EpCAM

expression in the earlier zones and low EpCAM expression in the later zones. Cells are captured

and specifically identified as CTC by immunostaining. (D) The CTC sorting approach is

hypothesized to produce a profile of CTC subpopulations which can be monitored as a tumor

progresses, enabling evaluation of CTC heterogeneity during tumor growth.

The microfluidic chip employed for CTC analysis demonstrates a number of key design features

that allow for capture as a function of EpCAM expression. Cells are first tagged with magnetic

nanoparticles modified with anti-EpCAM antibodies (Figure 2-1 B). The average hydrodynamic

radius of the magnetic nanoparticles used in this study was 70-80 nm. Further characterization of

the nanoparticles was also performed (Appendix Figure B-1). There are numerous advantages to

the use of nanoparticles in this application. Firstly, due to their small size they are able to diffuse

much more quickly than larger particles. This reduces the incubation time considerably and allows

for more reproducible binding coverage. Secondly, the nanoparticles have limited steric hindrance,

which means that the number of particles binding to a cell is representative of the number of

surface markers. Finally, unbound particles are less likely to coalesce and thus are unlikely to

obstruct fluid flow. Following nanoparticle binding and cell introduction, the chip is sandwiched

between an array of neodymium magnets and flow is induced using a syringe pump. Cruciform

structures within the device serve as pockets of minimum flow (velocity valleys), which allow for

localized capture of CTCs.

Cells within the device experience two main forces (neglecting gravity at the microscopic scale),

namely, a drag force and the magnetophoretic force. The magnetophoretic force is proportional to

the number of nanoparticles on the surface of the cell as well as the intensity of the magnetic field,

which is held approximately constant by the particular arrangement of magnets. The drag force is

proportional to the fluidic flow rate. This is the basis of the geometric design of the chip. Notably,

the number of parallel chambers increases across the length of the device, in a stepwise manner

from 1, to 2, to 4, to 8, which effectively divides the chip into 4 different zones (Figure 2-1 C).

Cells are captured in a region where the magnetophoretic force and the drag force are equal in

magnitude but opposite in direction. For cells with a large number of bound nanoparticles, this

will occur in the first zone. For cells with a small number of bound nanoparticles, this will occur

in the later zones. Since the number of nanoparticles bound is a function of EpCAM expression,

16

cells with high nanoparticle binding (and thus high levels of EpCAM) are captured in the earlier

zones, whereas cells with low nanoparticle binding (and thus low levels of EpCAM) are captured

in the later zones. It was our hypothesis that a change in sorting pattern will be revealed as the

tumor develops and becomes more metastatic (Figure 2-1 D).

2.3.2 Characterization of device function

In order to test the prediction that the microfluidic device facilitates sorting by relegating low

EpCAM cells to later zones, a characterization of the device was performed. SKBR3 and MDA-

MB-231, two well-established cell lines with high and low EpCAM expression respectively, were

chosen. EpCAM expression levels were first characterized using flow cytometry, which confirmed

the large difference in EpCAM expression between the two cell lines (Appendix Figure B-2). Cells

were then subjected to capture in the microfluidic device at a flow rate of 600 μL/h, the same flow

rate used for all capture experiments.

Figure 2-2 Characterization of capture. (A) Capture distribution across the four zones for a

cell line with high EpCAM expression (SKBR3, black) and low EpCAM expression (MDA-MB-

231, gray). The lower EpCAM cells (see inset of cells stained for EpCAM, scale bar = 10 μm)

are captured in the later zones. (B) Capture sensitivity and specificity demonstrated by the

excellent capture of SKBR3 cells and the very low capture of non-specific white blood cells.

*0.04%

High EpCAM Low EpCAM

17

As shown in Figure 2-2 A, SKBR3 cells, which express high levels of EpCAM, were captured

predominantly in zones 1 and 2, whereas MDA-MB-231 cells, which express low levels of

EpCAM, were captured in zones 3 and 4. Furthermore, immunocytochemistry shows the typical

EpCAM expression pattern on cells captured in the early and later chambers of the device, which

further illustrates the capacity of the device to sort based on EpCAM (Figure 2-2 A, inset). Finally,

Figure 2-2 B shows the capture efficiency of the device towards target SKBR3 cells and

nonspecific WBCs. The device demonstrates both high sensitivity and specificity, highlighting its

applicability to in vivo testing.

2.3.3 Measurement of tumor volumes and analysis of metastatic lymph nodes

VX2 cells were injected into the thigh muscles of the rabbit cohort in order to seed the growth of

a tumor. Following this injection, tumor size and morphology were monitored by digital palpation

and CT scan biweekly.

Figure 2-3 VX2 tumor growth in New Zealand rabbits. (A) Representative CT images

demonstrating the progression of primary tumor and (B) lymph node metastasis at 5, 13 and 22

days post induction. The red boxes highlight the region of tumor growth. (C) Tumor volume for

18

all rabbits included in the study. (D) Lymph node metastasis for Rabbit 2. (E) Tumor histology

visualized by staining with (i) H&E and (ii) pan-CK. Lymph nodes stained with (iii) H&E and

(iv) pan-CK, confirming its metastatic status.

All 6 injected rabbits exhibited tumors that increased volume over the course of the study (Figure

2-3 A). On average, tumors typically became detectable within 7-10 days post-injection and over

time, increased exponentially in volume (Figure 2-3 C). Two weeks after injection, the tumors

spread to loco-regional lymph nodes in all 6 rabbits (the volume of the lymph node tumor is shown

for one of the rabbits in Figure 2-3 B, D). At four weeks post injection, survival surgery was

performed to resect the tumor and malignant lymph nodes. At four weeks post-resection, terminal

surgery was performed. Following terminal surgery, residual disease and metastatic lymph nodes

were histologically analyzed. Staining with H&E and pan-CK confirmed the tumors and lymph

nodes to be malignant in all rabbits (Figure 2-3 E).

2.3.4 Correlation of CTC levels with disease progression

In order to determine if the induced tumors would express sufficient EpCAM for on-chip capture,

flow cytometry experiments were conducted to analyze EpCAM levels on VX2 cells. These trials

revealed considerable expression of EpCAM by the VX2 cells, indicating that this surface antigen

was a suitable capture agent (Appendix Figure B-3).

19

Figure 2-4 CTC sorting and analysis. (A) Fluorescence micrographs showing immunostaining-

based identification of white blood cells and captured cancer cells at 50× magnification, scale bar

= 10 μm. (B) Number of CTCs (DAPI+, CK+, CD45−) identified in each zone of the chip over

the 6 week study, averaged over 6 rabbits. Gray bars represent average tumor volume. Notably,

the number of CTCs in zone 3 and 4 shows a more pronounced increase than those in zones 1

and 2. Following surgery, CTC numbers decrease rapidly before rebounding, likely due to

release of CTCs from secondary tumors.

Furthermore, VX2 cells spiked into blood were used to confirm the efficacy of our

immunostaining approach (Figure 2-4 A). Specifically, anti-CK stain was used to identify cancer

cells, as cytokeratins are expressed by a wide variety of epithelial cells but are typically not

expressed by blood cells. CD45, a pan-leukocyte marker, was used as a counterstain to identify

20

contaminating white blood cells. This collection of stains enables the clear distinction of cancer

cells from white blood cells.

Over the course of the study, CTCs were captured and analyzed multiple times per week. Figure

2-4 B summarizes the CTC capture data averaged over the entire rabbit cohort (n = 6). The

utilization of nanoparticles allowed for an effective capture of CTCs even using a relatively small

volume of blood (1 mL). The CTCs in each zone are shown separately to elucidate the shifting

values of individual CTC subpopulations. Importantly, prior to tumor injection, all rabbits showed

very few or no CTCs, indicating a low rate of false positives. Levels of CTCs became significant

one week after tumor induction and continued to gradually increase for four weeks (Figure 2-4 B).

The increase in CTCs was correlated with growth of the primary tumor, which reached a maximum

at four weeks, just prior to survival surgery. Furthermore, CT imaging determined the development

of a secondary metastatic disease in most of the rabbits during the same timeline. These findings

suggested a positive correlation between the presence of CTCs and the appearance of metastatic

tumors, with a rise in CTCs preceding a detectable metastasis. At four weeks after tumor induction,

survival surgery for tumor resection was performed. Following this surgery, a drastic reduction in

CTC levels was observed, which suggests that the primary tumor was the major source of CTCs.

However, two weeks after surgery, a second increase in CTC levels was observed, indicating a

rebound in metastatic activity. This may be explained by the presence of metastatic lymph nodes

in the pelvis as revealed by CT scanning. This implies that CTCs might have originated not only

from the primary tumor but also from metastatic lymph nodes, and indicates a correlation between

CTCs and lymph node metastasis. Furthermore, this observation implies that surgical resection of

the primary tumor was insufficient in limiting disease spread. Taken together, these results suggest

that an increase in the number of CTCs is an indicative biomarker for metastatic tumors derived

from VX2 cancer cells.

2.3.5 Analysis of CTC sorting patterns during tumor growth

The sorting of CTCs on the basis of EpCAM revealed a notable capture trend over the progression

of the disease (Figure 2-4 B). Within our device, cells captured in Zone 1 represent those with high

levels of EpCAM, while those in Zone 2, 3, and 4 have decreasing levels of this epithelial marker.

In weeks 2, 3, and 4, the number of CTCs in zones 3 and 4 increased more rapidly than those in

zones 1 and 2. In fact, these cells account for most of the increase in the total number of captured

21

CTCs, which indicates a loss of EpCAM in the CTCs with tumor progression. This finding is

interesting, but reasonable considering that the reduction in epithelial cell adhesion molecules is a

well-established fact that is known to occur with tumor maturation.48,56 Notably, CTCs in zones 3

and 4 also showed the greatest reduction following surgery, which may reflect that these cells

originated from the primary tumor. However, CTCs in zones 3 and 4 also increase most drastically

in week 6, presumably as the lymph node metastases becomes increasingly dominant. The cells

with the lowest amount of EpCAM appear to be the most informative indicator to the state of the

disease. This is of particular interest as these are the most difficult to capture and as such are the

subpopulations most likely to be missed in conventional capture methods.

It is useful to benchmark the evolution of the surface expression profile reflected in the sorting

pattern we monitor against the sorting profile of the VX2 cells that are the precursor of the tumors

studied. In order to determine if the capture profile of the seeded VX2 cells differed from that of

the captured CTCs, experiments were performed where VX2 cells were spiked into healthy rabbit

blood. The majority of cells were captured, indicating that EpCAM expression was adequate for a

sufficient number of nanoparticles to be bound. Interestingly, a large number of VX2 cells were

captured in zone 1, with a steadily decreasing rate of capture in zones 2, 3 and 4 (Figure 2-5). This

is indicative of a high level of EpCAM expression, which is a reasonable result given the epithelial

origin of VX2 tumors.

22

Figure 2-5 Comparison of sorting profile for (i) VX2 cells and (ii) CTCs captured 4 weeks

after tumor induction, showing a prominent shift towards later zones as tumor growth progresses

with lower level of EpCAM expression.

Perhaps the most important revelation is the marked difference between the capture trends

exhibited by VX2 cells compared to that of CTCs collected after 4 weeks of tumor growth (Figure

2-5). Whereas the spiked VX2 cells were captured primarily in the earlier zones, the CTCs were

isolated primarily in the later zones. This indicates a dramatic reduction in the abundance of

EpCAM expression on the surface of CTCs. We speculate that the development of the tumor within

the host triggers the abrupt shift in the biology of the VX2 cells, possibly initiating a pronounced

loss of epithelial character.

2.3.6 Discussion

It is well recognized that CTCs are precursors to metastasis and can serve as an integral component

of tumor staging criteria. Furthermore, previous studies involving breast cancer patients have

shown that CTC data can be the best indicator of disease state and can provide improvement to the

staging process.65 Because blood collection is simple and minimally invasive, CTCs can

potentially be used as a real-time marker to monitor disease progression. They also possess the

23

potential to guide therapeutic management, indicating therapy effectiveness or necessity, even in

the absence of detectable metastases.66

Despite this utility, the clinical application of CTCs remains minimal. One reason for this is the

inability of commercially-available systems to detect CTCs with low levels of EpCAM.56 Other

limitations include negligible EpCAM expression levels in subpopulations of CTCs. This results

in missing significant information regarding the tumor cell differentiation level. In this study, we

demonstrated the ability of a microfluidic device to capture and sort CTCs with varying levels of

EpCAM, and infer a correlation with tumor progression.

Our investigation shows that elevated numbers of CTCs correlate with disease progression and

lymph node metastasis. Specifically, the number of CTCs peaked at the same time that the tumor

reached its maximum size. This is expected given that previous studies have demonstrated that

high CTC numbers correlate with aggressive disease, increased metastasis, and decreased time to

relapse.51,52,54 The dramatic reduction in CTCs post-surgery provided real-time information on

treatment effectiveness, which could potentially be useful in the assessment of surgery or

chemotherapeutic efficacy. Since chemotherapy regimens have highly variable efficacies for

different patients, rapid determination of the effectiveness of treatment using CTC markers could

allow for rapid treatment modulation.

The recurrence of CTCs after the tumor resection suggests that the CTCs might also originated

from metastases in lymph nodes, which further supports the theory that secondary tumors can

become a source for CTCs, and helps to explain the recurrence of metastatic disease in patients

that have undergone surgical resection. This is in agreement with the ability of many cancers to

recur despite complete surgical resection of the primary tumor.66,67

There is ongoing research regarding whether tumor cells undergo a reduction in epithelial character

during dissemination, resulting in a more mesenchymal or even more stem cell-like phenotype.

Our results corroborate this theory; over time, in addition to the increase in their numbers, we

inferred that CTCs likely exhibited a reduction in EpCAM. In this respect, the use of nanoparticles

is important due to the minimal influence of steric effects on bead coating efficiency. Since

EpCAM levels reflect the epithelial nature of a cell, they can serve as a surrogate marker for this

process, indicating an increase in invasive and migratory properties. This is consequential because

the loss of epithelial character is believed to be the gatekeeper under which tumor cells intravasate

24

into the blood.68 Furthermore, it has been suggested that CTCs with the highest plasticity tend to

be the most aggressive.69

An interesting insight provided by the present study was the difference in EpCAM expression

levels between the VX2 cell line used to produce the tumors and the CTCs seeded from these

tumors. This is an aspect that would typically be neglected by detection methods that capture all

CTCs in the same fashion, regardless of their biological state and their EpCAM expression level.

This reveals a phenotypic change occurring immediately upon insertion of cells into the biological

environment of the rabbit host, suggesting a phenotypic transition that can occur in cells on the

periphery of the tumor. Interestingly, this is similar to the result shown by Steinert et al, which

demonstrated a rapid reduction in EpCAM expression upon colorectal cancer-derived CTC

compared to the corresponding tumor tissue.70

As a future step a comprehensive study with large patient cohorts using different clinical subtypes

of cancer, with standardized detection and characterization of CTCs would be desirable to evaluate

the performance of the device for human blood samples. Further studies, incorporating a suite of

biomarkers indicative of cell phenotype changes would allow one to determine the prognostic

significance of different categories of CTCs in cancer patients and aid in the clinical management

of these patients. Finally, since the device allows for the sorting of CTCs, which remain viable,

this opens the door for genomic, proteomic and transcriptomic investigations of these cells in

various phenotypic states. CTC genomics is still in its infancy, which is primarily due to the lack

of technologies that are capable of capturing sufficient numbers of CTCs to analyze somatic

mutations.49,50 The next stage in the investigation of CTCs is their characterization using these

‘omics’ techniques, which can be used to identify numerous characteristics of tumors, allowing

for targeted therapeutic approaches.55,71,72 These techniques could find application with the current

device.

2.4 Conclusion

CTCs are an important class of circulating cancer markers that may enable noninvasive

determination of prognosis. This study demonstrated the successful capture of CTCs in an animal

model and the sorting of these CTCs on the basis of their EpCAM expression. The numbers of

isolated CTCs were positively correlated with tumor growth while CTC sorting profiles

concurrently indicated a shift towards reduced expression of EpCAM. This indicates that both

25

number and phenotypic distribution of CTCs may possess clinical relevance. Future studies

including a greater diversity of cancer subtypes, cell surface markers, and genomic approaches are

to be investigated.

2.5 Methods

2.5.1 Animal model

Experiments were performed using 6 New Zealand white rabbits weighing 2.5-3.0 kg. All animal

studies were performed in accordance with the University Health Network/University of Toronto

guidelines for the humane use of animals. Care, handling and maintenance of all animals used in

this study were conducted in a humane manner, as per the animal care experimental protocol

approved by the institutional Animal Care and Use Committee of University Health Network,

University of Toronto. Male rabbits (Charles River, Wilmington, Massachusetts) were injected

with 300 μL of a high-density (approximately 5 × 106/mL) cell suspension of VX2 squamous cell

carcinoma of the rabbit into the thigh muscles (quadriceps). Tumor development and lymph node

metastases were monitored using computed tomography (CT) images and clinical evaluation bi-

weekly, and were resected 4 weeks after tumor induction. Tumors and enlarged lymph nodes were

sent for pathology and immunostaining analysis. Blood samples for CTCs (2-3 mL), hematology

and comprehensive biochemistry analysis were collected biweekly, pre and post-operation.

2.5.2 Tumor cell line propagation

The VX2 tumor cell line is maintained in small tumor pieces that are frozen at −80 °C. Tumor cells

were propagated by injecting 500 μL of VX2 tumor into the quadriceps of propagating rabbits

(different from rabbits used for the CTC study) and were harvested after approximately 3 weeks.

The harvested tumor was placed in Hanks balanced salt solution (HBSS) in a sterile 100 mL

container. Prior to tumor induction in rabbits, the tumor pieces were thawed and cut into small

pieces using a sterile scalpel and subsequently placed on to a 70-μm-cell strainer sitting on a 50

mL tube (BD Falcon brand). A syringe plunger was used to mince the cells and ~500 μL HBSS

was used to suspend the cells in the strainer (repeated several times).

2.5.3 CT imaging and image analysis

CT imaging and image analysis CT imaging (Locus Ultra, GE Healthcare, Milwaukee, Wisconsin,

USA) was performed biweekly pre and post-surgical tumor resection (80 kVp, 50 mA). All CT-

26

based image analysis was performed using Microview (GE Healthcare, Milwaukee, Wisconsin,

USA) and custom in-house program written using MATLAB (MathWorks®, Natick,

Massachusetts, USA). The tumor volumes were contoured using a semi-automated threshold based

method. The mean and standard deviation of the voxel signal distribution within each VOI were

calculated.

2.5.4 Histopathological evaluation

Tumor and lymph node tissue samples were fixed in formalin after resection, embedded in paraffin

blocks, cut and stained with hematoxylin and eosin (H&E) and pan-cytokeratin (AE1/AE3), the

intermediate filaments of epithelial cells. All histopathology images were analyzed using

ImageScope (Leica Biosystems, Wetzlar, Germany) after scanning.

2.5.5 Capture of cell lines

100 SKBR3 or MDA-MB-231 cells in 100 μL were incubated with 10 μL of MACS anti-EpCAM

nanoparticles (130-061-101) for 10 min. Cells were then introduced into the chip at a rate of 600

μL/h for 10 min. A 200 μL PBS 1× rinse was then added followed by 100 μL of PBS-4%

formaldehyde and 100 μL PBS–0.2% Triton. Cells were stained with anti-pan-CK Alexa Fluor

488 from eBioscience (53-9003-82) and 10% DAPI.

2.5.6 Immunocytochemistry of cell lines

Cells were incubated in 1% BSA PBS for 10 min to prevent non-specific binding. Cells were then

stained with Alexa Fluor 647 anti-human CD326 (EpCAM) antibody (Biolegend, 324212) for 1

h. Cells were rinsed twice using 1× PBS before DAPI staining in 0.1%Triton-PBS.

2.5.7 Isolation and fluorescent staining of CTCs

In order to detect and characterize CTCs, we used an immunomagnetic system targeting

EpCAM.21 Captured cells then underwent immunocytological staining to confirm their identity

as tumor cells. CTCs were identified as EpCAM isolated cells that stained positive for DAPI and

pan-cytokeratin and negative for CD45. The number of detected CTCs obtained per mL of blood

was recorded for correlation with clinical parameters such as tumor size, lymph node metastasis

and duration from tumor inoculation. CTCs in each of the 4 zones were counted separately.

27

1 mL of rabbit blood was incubated with 10 μL of MACS anti-EpCAM nanoparticles (130-061-

101) for 10 min. This blood was then introduced into the chip, which had been pretreated with 1%

pluronic acid, at a rate of 600 μL/h for 100 min. 200 μL of PBS-EDTA was introduced through to

rinse out red blood cells. Subsequently, 100 μL of PBS-4% formaldehyde was added to fix the

cells. Next, 100 μL of 0.2% Triton in PBS was added to permeabilize cells. 100 μL of a CTC-

specific antibody (1 μL of anti-pan-CK Alexa Fluor 488 from eBioscience (53-9003-82) and 1 μL

of anti-CD45 APC from AbdSerotec (MCA1114F) in 98 μL of PBS-1% BSA) was added for 1 h

at 100 μL/h for immunostaining. Afterwards, 100 μL of PBS with 10%DAPI solution was added

at 600 μL/h for 10 min to stain nuclei. Finally, 200 μL of PBS was added to remove excess non-

specifically bound antibody.

2.5.8 Capture efficiency of white blood cells

Capture and staining were performed as described above, with white blood cells considered those

cells which were DAPI+/ CD45+/CK−. Capture efficiency was calculated assuming 106 WBC/ml

blood.

2.5.9 Image scanning and analysis

After immunostaining, chips were scanned using a 10× objective and a Nikon Eclipse Ti

microscope equipped with an automated stage controller and a cooled CCD (Hamamatsu,

Hamamatsu, Japan). Images were acquired with NIS Element software (Nikon, Tokyo, Japan).

Red, green and blue fluorescence images were recorded. The captured images were then analyzed

in NIS Elements and target cells were enumerated.

2.5.10 Flow cytometry

VX2 cells were prepared in suspension as for tumor propagation. SKBR3 and MDA-MB-231 cells

were collected from culture. Cells were incubated in 1% BSA blocking buffer for 30 min to prevent

non-specific binding. Anti-pan-CK Alexa Fluor 488 from eBioscience (53-9003-82) was added

and incubated for 1 h prior to cytometry. Cytometry was performed using the BD FACSCanto

(Becton Dickinson, Franklin Lakes, New Jersey, USA) flow cytometer with 488 nm laser

excitation and 530/30 nm detection.

2.5.11 Microchip fabrication

28

Microchips were fabricated using poly-dimethylsiloxane (PDMS) after production of an SU-8

master by lithography on silicon wafers, with a height of 80 nm (University Wafer, Massachusetts,

USA). A PDMS (Dow Chemical, Michigan, USA) copy of the master was produced and peeled

off the wafer. Holes were pierced at the inlet and outlet. PDMS and a glass slide were treated with

1 min of plasma and attached to form a permanent bond. Silicon tubing was inserted into the holes

at the outlet and inlet. Prior to use, chips were incubated with Pluronic F68 acid to reduce non-

specific binding. During cell capture, arrays of NdFeB N52 magnets (KJ Magnetics, Pennsylvania,

USA) were placed above and below the chip.

29

Mechanistic Control of the Growth of Three-Dimensional Gold Sensors

In the previous chapter, we explored the application of a microscale technology towards a

particular disease. In this chapter, we are continuing to examine the role of microscale technology

for disease probing, but are investigating a more fundamental topic; the development of novel

electrode architectures to improve biomarker sensing. Gold electrodeposition is a well-established

technique for electrode synthesis, due to the inert nature of gold electrodes and the proclivity of

gold to encourage the formation of monolayers of thiolated molecules. Here we test the effects of

a number of adjustable parameters (gold concentration, voltage, and electrolyte viscosity) and

observe the fundamentally different architectures that result. We further investigate the

mechanisms of the underlying growth patterns, and ultimately challenge the final electrodes with

biomolecular detection.

Reprinted with permission from Mahshid S.*, Mepham A.*, Mahshid S. S., Burgess I. B., Safaei

T.S., Sargent E.H., Kelley S.O., “Mechanistic Control of the Growth of Three-Dimensional Gold

Sensors” Journal of Physical Chemistry C, 2016, 120(37) 21123–21132. Copyright 2016

American Chemical Society.

Link to publication online: https://doi.org/10.1021/acs.jpcc.6b05158

Disclosure of work within this manuscript: A.M., S.M., S.S.M. and I.B.B. designed the

experiments. A.M. and S.M. performed experiments. Data analysis and manuscript writing were

performed by A.M., S.M., S.S.M. and I.B.B. with assistance from T.S.S., E.H.S., and S.O.K.

30

3.1 Abstract

Three-dimensional (3D) electrodes with large surface areas are highly effective biomolecular

sensors. These structures can be generated via the electrodeposition of gold inside microscale

apertures patterned on the surface of a microelectronic chip. Such electrodes enable the

ultrasensitive analysis of nucleic acids, proteins, and small molecules. Since the performance of

these electrodes is directly related to their surface area, the ability to control their microscale

morphology is critical. Here, we explore an electrochemical model based on the theory of

nucleation and growth to better understand how to control the morphology of these electrodes. The

insights gained from this model enabled us to create preferential conditions for the formation of

different morphological features. We demonstrate for the first time that electrodeposition of 3D

nanostructured microelectrodes inside a microscale aperture is governed by two stages of

nucleation and growth. The first stage involves the creation of primary nuclei at the bottom of the

aperture. The second stage features the generation of new nuclei upon exposure to the bulk

solution. Depending on the overpotential, the deposition is then continued by either rapid growth

of the original nuclei or fast growth of new nuclei. Faster electrodeposition at high overpotentials

promotes directional growth, generating spiky structures. More isotropic growth is observed with

low overpotentials, generating rounder features. Ultimately we determine the efficiency of DNA

hybridization on a variety of structures and identify the optimal morphologies for rapid

DNA−DNA duplex formation.

3.2 Introduction

Engineering three-dimensional (3D) structures on the micro and nanometer scales is of importance

for the production of high-performance materials for electronics and biological applications.73–77

Bottom-up fabrication via self-assembly lends itself to the creation of complex 3D architectures in

a variety of material systems.78–81 This approach relies on minimal direct control, relying instead

on pattern formation on the basis of thermodynamic equilibria or instabilities.82 In such a system,

the assembly process and the resultant material morphology can be manipulated via any parameters

that affect kinetic and thermodynamic properties.83

One of the materials that is used extensively for generation of 3D micro- and nanostructures is

gold.84 Gold structures exhibit a variety of morphologies depending on the method of synthesis.85

Furthermore, 3D gold structures are outstanding candidates for electrochemical biosensing

31

applications86 and catalyze a number of important chemical reactions.87,88 The specificity of

catalysis and the degree to which the reaction rates are amplified can be linked directly to the

morphological properties of 3D gold structures.89

Moreover, the crystal structure of gold is also important in determining its performance as a sensor

or catalyst. Different crystal facets of gold promote distinct interactions with molecular substrates

and targets.90 Thus, the capacity to modulate the chemical reactivity of different facets during the

assembly process is advantageous for the design of optimized structures. As such, methods that

control the number and orientation of grain boundaries are of significant interest.

A variety of methods have been used to synthesize gold structures, including sputtering, e-beam

lithography, and chemical and electrochemical approaches.91–94 Although each of these methods

has its advantages, none offers access to the suite of synthetically accessible parameters that are

available with electrochemical synthesis.94 We have previously developed electrodeposited, 3D

gold microelectrodes and found that they represent a promising platform for electrochemical

biomolecular detection.95–100 By creating sensors with large surface areas that protrude into

solution, collisional frequencies for biomolecular targets are enhanced.96 Also, the electrochemical

currents generated by these structures are amplified due to the efficient transport of redox-active

reporters via radial diffusion.97 We have further demonstrated that the introduction of

nanostructured roughness on the surface of these microelectrodes enhances their biosensing

performance.98 While this system has been applied successfully to the detection of cancer

biomarkers,99 infectious pathogen identification,95 and organ transplant assessment,100 we know

little about how the 3D features of these sensors influence their performance.

Here, we explore the growth mechanism of 3D gold microsensors and identify factors that enable

precise control of morphology and crystallinity. We demonstrate that by changing the gold ion

concentration, we can preferentially grow different structures and control directional growth of

spiky structures. By increasing the viscosity of the electrolyte solution, we can suppress the growth

and generate finer spiky structures. Additionally, by increasing the deposition overpotential, we

can also generate finer spiky structures through promoting multiple nucleation and rapid growth

of the nuclei clusters. The collection of these parameters enables a high degree of control over the

microscale morphology and crystallinity of gold assemblies. We further explore the effects of

voltage in more depth to visualize processes related to three dimensional nucleation and diffusion-

32

controlled growth.101,102 We ultimately investigate the efficiency of hybridization of DNA

duplexes on the surfaces of sensors with differing morphologies and identify the optimal

morphology for this type of application. The most effective sensors are generated with high

growth-to-nucleation ratios during electrodeposition.

3.3 Results and Discussion

3.3.1 Outline of experiments

Our approach to generating 3D microelectrodes using gold electrodeposition benefits from micro-

patterned apertures fabricated via photolithography on the surface of a glass chip. A schematic

outlining the system used in this study is shown in Figure 3-1; the application of this

microelectrode system for the analysis of specific DNA sequences is also depicted.

Figure 3-1 Schematic of gold microelectrode experiemnts (A) Gold 3D microelectrodes are

grown using electrodeposition on a gold substrate with 10 μm apertures as a template. In this

study, we explore parameters that could generate different sensor morphologies to determine the

optimal properties for biomolecular detectors. (B) The parameters of gold ion concentration,

Gold substratePhotoresist

Glass

target DNA

probe DNA

A B

C

33

solution viscosity, and applied voltage are varied to explore which regions encourage the growth

of particular morphologies. (C) 3D microsensors are tested for DNA hybridization. Blue strand:

capture probe that promotes sequence-specific binding of a target sequence. Red: Target DNA

strand complementary to probe. Target binding can then be read out using redox-active reporters.

The electrodeposition is carried out on a glass wafer coated with a thin layer of gold, which is

topped with a passivating layer of SU-8 photoresist (Figure 3-1 A). The small (10 μm) apertures

in this passivating layer expose the gold substrate to the electrolyte solution, allowing for localized

deposition of gold structures. This template-based approach allows for a hemispherical diffusion

pattern on top of the aperture that facilitates rapid growth and the faster emergence of structural

features compared to the growth achieved on larger planar surfaces.

The three major parameters explored in this study are the concentration of gold ions (AuCl4-), the

viscosity of the electrolyte solution, and the applied voltage. The concentration of gold ions plays

a critical role in the kinetics and thermodynamics of the transition from a dissolved state to a solid

state. Higher concentrations increase the capacity of the system to transfer sufficient quantities of

gold to the growing deposit and enhance electrodeposition kinetics. Viscosity affects the kinetics

of the deposition reaction by impeding replenishment of fresh gold ions. Applied voltage controls

the relative energies of the solid and dissolved states of gold. As the potential of the cathode is

made increasingly negative, the reduction of gold ions to metallic gold becomes more favorable.

By varying these three parameters, a phase space is generated featuring regions with different

growth regimes and different morphologies (Figure 3-1 B).

For applications where dilute diffusible molecular targets like DNA are being detected (Figure 3-1

C), it is critical to precisely control the morphology and surface area, as both factors influence

collisional frequencies as well as binding affinities. To control the formation of 3D structures

generated via gold electrodeposition, it is important to understand the influence of nucleation and

growth processes. We therefore not only investigate the influence of solution parameters on 3D

sensor electrodeposition but also investigate the mechanism of nucleation/growth and seek to link

it to the geometrical properties of the resultant structures.

3.3.2 Effects of solution conditions of 3D nanostructured microelectrode growth

34

The effects of varying the concentration and viscosity of the electrolyte solution are shown in

Figure 3-2 A and Figure 3-2 B. For each condition, a pair of SEM images are shown at a low (5

000×) and high (50 000×) magnification. This allows for the elucidation of changes in the structure

at two different scales; both the microscale, so that the overall shape of the structure can be

evaluated, and the nanoscale, which allows a detailed analysis of the morphological changes.

35

Figure 3-2 Effects of concentration, viscosity, and voltage on electrodeposition of gold

microsensors are studied using scanning electron microscopy (SEM). (A) SEM images for

structures generated with varied concentrations of gold ions. Varying the concentration of gold

reveals a number of different morphologies. A low (1 mM) concentration produces nanoscale

spike-like structures, whereas a 5 mM concentration produces leaf-like structures (inset shows

structure at 1000× to visualize entire structure) and higher concentrations promote the growth of

needles. (B) Increasing viscosity creates rounded structures without visible facets or needles. (C)

A

B

C

(1)(2) (3)

(4)(5) (6)

(7)(8)

(9)

(10)

(11)

(12)

36

Increasing deposition potential changes the structures from isotropic rough hemispheres to

highly anisotropic structures with a large population of needles. White scale bars are 20 μm,

yellow are 2 μm, and red are 40 μm. The potentials used for electrodeposition are displayed

above each SEM image and numbered for cross-referencing to Figure 3-4.

The first parameter explored was the concentration of the gold solution (Figure 3-2 A). A wide

range of concentrations were tested, ranging from 1 mM to 500 mM, with a roughly logarithmic

distribution (each sample approximately 3× more concentrated than the previous). The different

concentrations produced at least three qualitatively different growth modes. At the lowest

concentration tested (1 mM), the overall morphology features a rounded structure containing a

number of distinct lobes on the order of 20−30 μm in size. A study of the surfaces of these

structures reveals that the gold assumes a propensity toward nanoscale spike-like structures and is

highly porous.

At a 5-fold higher concentration (5 mM) of gold ions, the structure assumes a morphology that is

remarkably different from that at 1 mM. Instead of featuring rounded lobes, the structure is

dominated by a small number of large leaf-like structures. Upon closer inspection at a higher

magnification, these structures demonstrate remarkable geometric patterns. Each leaf has a central

ridge that runs lengthwise along the middle of the leaf from stem to tip. From this ridge, a number

of smaller veins extend. These veins in turn have smaller sub-veins. This fractal structure indicates

that each leaf is either a single crystal domain or a number of domains with fixed relative

orientations.

As the gold ion concentration is further increased to 16 mM, a new structural regime is

encountered. Leaf-like structures are reduced in number, and there appears a new structural

feature: needles. At this concentration, the needles are highly ridged, exhibiting a series of peaks

and valleys along their length. These needles do not have a consistent cross-sectional shape and

appear to be incompletely formed. As the concentration is further increased from 16 mM to 50 and

500 mM, these needles persist. However, at higher concentrations, the ridges are significantly

reduced and the needles develop a consistent pentagonal cross section. These needles are of

particular interest due to their unique shape, one that has been recognized before in gold

structures.103 Interestingly, higher gold ion concentrations do not necessarily create a greater

37

abundance of needles but rather promote thicker and smoother shapes. As such, the region for

creating the greatest number of needles lies in the concentration range between 50 and 500 mM.

In order to change the viscosity of the solution, varying amounts of glycerol were introduced into

the solution (Figure 3-2 B). Glycerol was chosen due to its high viscosity and its miscibility in

water. As the glycerol concentration is increased, the structures become more compact and dense

and the local features are distorted. The sharp edges and facets are replaced with much more

rounded features. The most profound change occurs at 75% glycerol. Here, the needles are not

present and are replaced with a largely amorphous arrangement that displays weaker directional

preference than the highly crystalline material. This experiment indicates that there is another

important ingredient for the promotion of needle morphology; namely, low viscosity. The

associated high mobility of ions seems integral to allow needles to dominate over largely

amorphous, many-grained structures.

The influence of applied potential was initially studied using linear voltammetry (Figure 3-2 C).

Variation of applied potential also creates a demonstrable effect on the final morphology (see SEM

insets of Figure 3-2 C). At very low overpotentials (+500mV) the growth is hemispherical overall,

with no noticeable anisotropy. At the microscale, the structures are disordered, with a large number

of grains of different sizes and orientations. As the magnitude of the potential is increased (i.e., the

cathode is made more negative), the structures gain a pronounced anisotropy, with regions of the

structure extending farther into the surrounding solution. At a potential of about +150 mV, the

needle morphology reemerges, although it is clearly not pentagonal at this voltage and is decorated

with a large number of offshoots. Approaching 0 mV, the needles display a full pentagonal

geometry. Therefore, to promote the formation of the needles and generate finer spiky structure, a

high overpotential (i.e., −250 mV) is needed in addition to a high gold concentration (SEM insets

of Figure 3-2 C). It is only in the combinatorial application of these parameters that this particularly

interesting morphology becomes the most stable manifestation of gold. Increasing overpotential

to more negative values (above −500 mV) causes the spiky structures to be replaced by a flaky

morphology with fine structuring.

Images of the exterior of the structures provide insight into the growth mechanism but do not

elucidate the extent of crystallinity of the structures. To investigate the internal structures of our

3D sensors more directly, FIB sectioning was performed and SEM images were collected of the

38

resulting cross sections. Using this approach, different crystal grains are revealed within the

structures.

Figure 3-3 Analysis of the interior structures of electrodeposited gold using FIB. (A, B) The

structures generated under high viscosity conditions shows grains with a wide variety of sizes

and shapes. (C, D) The leaf structure has a clear crystal twinning about the central plane,

indicated by two different shades of gray. (E, F) The needle-like structures have pentagonal cross

sections with a 5-fold twinning about the center axis. All scale bars are 2 μm.

The structures grown in 75% glycerol are shown in Figure 3-3 A and again in cross section in

Figure 3-3 B. In these amorphous structures there is no conserved orientation or direction to the

grains. Furthermore, the grains are of vastly different sizes and shapes, with no discernible pattern.

The leaf-like structures have a fundamentally different inner morphology than those generated

under high viscosity conditions. Figure 3-3 B shows an example of this type of structure displaying

a series of veins along its surface and a bilaterally symmetrical shape. A FIB cross section is shown

in Figure 3-3 E. Notably, this structure is again composed of multiple grains; however, it has a 2-

fold symmetry rather than the disordered arrangement seen previously. This symmetry explains

why the top and bottom of the leaf are similar in geometry. This grain boundary is likely a

consequence of crystal twinning about a common central plane.104

Figure 3-3 C shows a high-resolution SEM image of a well-formed needle. This structure has well-

defined facets that are modified by a series of small ridges perpendicular to the length of the needle.

A B C

D E F

39

FIB sectioning of this structure reveals a clear pentagonal symmetry (Figure 3-3 F). Moreover,

this symmetry is seen to stem from a 5-fold arrangement of unique crystal grains about a common

center point, with distinct grain boundaries existing between them.

3.3.3 Nucleation and growth mechanism of 3D gold structures

In most electrodeposition processes, the first phase of structural growth is characterized by rapid

nucleation.101 These nuclei then enter a growth phase, with additional nuclei forming at a reduced

rate. As such, crystalline structure is largely established at the time of nucleation. It is apparent

from the images presented in Figure 3-3 that at relatively low concentrations (around 5 mM), nuclei

are formed with a 2-fold twinning structure. By contrast, at higher concentrations, there is a shift

such that the 5-fold twinned structure is preferable. The appearance of some leaf-like structures

even at higher concentrations can also be attributed to local depletion of gold ions.

In the case of increased viscosity, the growth rate is severely attenuated relative to the rate of

nucleation. Additionally, the geometric entities that represent the lowest energy state under other

conditions are no longer preferred, and there is instead a largely random process of deposition.

This leads to a loss of directionality that manifests both as an increase in isotropy and a less faceted

surface at the local level. Evidently, in order to promote needle growth, it is critical to create

conditions where 5-fold twinned nuclei are the most favorable.

Although FIB imaging provides qualitative insights into the mechanisms of growth, a more

comprehensive understanding of this phenomenon was desired. Due to the wide array of different

structures that can be produced (bulky, spiky, flaky) within a small window of potentials, we

elected to explore the effects of electrodeposition voltage in further depth.

3.3.4 Study of current-time transients

The early stage of electrodeposition in our system is associated with a three-dimensional

nucleation process, where the number of nuclei and the rate of nucleus formation are strongly

governed by the deposition overpotential. We investigated the I−t current transients during

deposition to establish a relationship between deposition overpotential and the nucleation and

growth mechanisms in 3D microelectrodes (Figure 3-4).

40

Figure 3-4 I-t curves during electrodeposition (A) I−t curves for gold electrodeposition within

the potential range of +500 to −750 mV. Each trace is labeled with a number that corresponds to

a voltage specified in Figure 3-2. These traces deviate from linearity under certain conditions as

shown by the region “e” specified with a dotted line. (B) Electrodeposition data for low-to-

moderate overpotentials (500 to 50 mV). (C) Electrodeposition data for higher overpotentials (0

to −750 mV).

During deposition inside the microscale aperture, nucleation and growth are strongly affected by

the diffusion of ions around the edges. However, as the deposition proceeds outside the aperture,

the geometry of the 3D electrode increases and consequently provides a larger surface for electron

transfer over time.

From the I−t curves (Figure 3-4), we can distinguish several deposition phenomena that occur

within our microscale apertures. The different phases of growth can be visualized by examining

1st nucleation phase

2nd nucleation phase

A

B C

41

the I−t curves generated with low overpotentials, +500 and +400 mV (Figure 3-4 B, curves 1 and

2). The first rise is related to the double-layer formation (region a). A sharp drop in current then

occurs due to the depletion of ions following the nucleation process on the planar Au at the bottom

of the aperture (region b). Immediately thereafter, steady-state growth fills the aperture (region c).

The isotropic growth then proceeds with most of the electrodeposited gold growing outside the

aperture and the 3D electrodes expand (region d) to form a bulky structure (SEM images 1 and 2

in Figure 3-2 C). When the overpotential is increased to +300, +150, and +50 mV, a sharp rise in

the I−t curve is observed at the beginning of region d (Figure 3-4 B, curves 3−5). Here the

deposition front undergoes a second nucleation stage, with new nuclei clusters forming on top of

the deposit as it confronts the bulk solution outside the aperture. This behavior appears to be

responsible for the initiation of needle-like structures in the deposit as shown in the SEM images

(5 and 6) of Figure 3-2 C. At more negative potentials (0 to −450 mV), the second nucleation

process is gradually merged with the first (Figure 3-4 C, curves 6− 12).

Similar types of behavior have been previously reported for metallic structures electrodeposited

from microscale/nanoscale pores.103–106 However, extrapolation of the I−t responses in Figure 3-4

demonstrates a unique phenomenon in our system related to the deposition of 3D electrodes at

higher overpotentials (0 to −450 mV). Within region “e” highlighted in Figure 3-4, the current

traces exhibit a second rise as deposition proceeds beyond 20 s (see curves 6−10). This may stem

from the production of additional small spikes “budding” off the side of the larger ones, a

phenomenon which appears to occur in this same potential range (Figure 3-2, C).

3.3.5 Analysis of i2/im2 vs t/tm

The I−t response curves revealed that two stages of nucleation exist: first, the creation of nuclei on

the planar Au at the bottom of the aperture, and second, the generation of new clusters on top of

the deposition front caused by exposure to the solution outside of the aperture. The first stage of

deposition takes place on a planar electrode located at the bottom of a 10 μm aperture (as shown

in Figure 3-5 A). All the area around this planar electrode is made of nonconductive amorphous

glass forcing the deposition to initiate at the bottom of aperture.

Figure 3-5 illustrates the relationship between the deposition overpotential and the nucleation

process by comparing current responses in a dimensionless i2/im2 vs t/tm plot specified by Equation

3.1 (See Methods). Here, the instantaneous nucleation is compared to progressive nucleation.

42

Figure 3-5 Nucleation during electrodeposition (A) Schematic of the chip-based templates

used for electrodeposition, where a 10 μm aperture provides a confined area for the growth of a

3D gold structure. The dimensionless i2/im2 vs t/tm responses of microelectrodes electrodeposited

at (B) +500, (C) +300, (D) +150, (E) +50, (F1) 0, and (F2) −550 mV and the corresponding

B C D

E F

A

43

theoretical dimensionless models for instantaneous and progressive nucleation. White scale bars

are 20 μm, and red scale bars are 40 μm.

At low overpotentials of +500 mV (Figure 3-5 B) and +400 mV (not shown), nucleation follows

the progressive model. The low overpotential allows diffusion from the bulk solution into the

aperture to readily replenish the consumed ions, resulting in formation of random-size nuclei at all

the possible active sites of planar Au (schematic cross section). This results in the formation of the

bulky structure shown in the SEM image.

At higher overpotentials such as +300 mV (Figure 3-5 C), after the first stage of nucleation inside

the aperture (described by the progressive model), a second nucleation occurs as soon as the

deposit is exposed to the bulk solution. This nucleation follows the instantaneous model. The

schematic cross section represents the sudden creation of nuclei (orange clusters) on top of the

random-sized progressive nuclei (black clusters). This results in the formation of two different

microstructures merged together (SEM image of Figure 3-5 C).

Further increases in the overpotential result in a gradual change in the first nucleation stage from

progressive to instantaneous, while the second nucleation remains instantaneous. At +150 mV

(Figure 3-5 D), a mixture of both the instantaneous and progressive behavior occurs for the first

nucleation stage, with the dimensionless current located between the two extreme forms of

nucleation. The resulting cross section and the SEM image also show the presence of a bulkier

structure underneath, which turns into thick needles as a result of the second instantaneous

nucleation (orange clusters in schematic). However, at +50 mV, instantaneous behavior is the only

dominant regime in the first and second nucleation stages (Figure 3-5 E). This instantaneous

regime in the second nucleation stage promotes a number of smaller nuclei clusters and as a result

reduces the thickness of needles in the final morphology.

At negative potentials (0 to −700 mV), the nucleation and growth follows the instantaneous model

as shown in Figure 3-5 F. From the SEM images (parts E, F1, and F2 of Figure 3-5), the deposits

tend to have finer needles and the thick ones disappear completely. This is in principle related to

the faster nucleation rate on the deposition front, which results in the creation of smaller nuclei on

a large number of active sites (Figure 3-5 F1 at 0 mV). The size of nuclei clusters continues to

shrink with increasing overpotential. At overpotentials above −450 mV, the possible sites for

44

nucleation appear on the edges and along the wall of the aperture (schematic of Figure 3-5 F2).

This results in the formation of a flaky structure.

We used 2D finite-element numerical simulations in COMSOL to model the profile of the

deposition front in our system for both the minimum and maximum overpotentials.

Figure 3-6 Two-dimensional time-varying simulation results for deposition of Au at (Top)

+500 mV and (Bottom) −500 mV. The COMSOL geometry consists of an aperture (10 μm wide,

1 μm deep) surrounded by an electrolyte in a 100 μm × 40 μm rectangular space. The electrolyte

had a diffusion coefficient of D = 1 × 10−9 m2/s with initial concentration of c = 50 mM. The

model is based on a Nernst−Planck equation with electric potential boundary conditions. The

simulation results show the deposition front progression from the bottom of the aperture for the

minimum and maximum overpotential in our system. At the right, the profiles of deposited layers

on the vertical wall of the aperture are shown at +500 mV (top) and −500 mV (bottom). All scale

bars are 20 μm.

The time-series simulation results of Figure 3-6 A reveal the slow progression of the deposition

front from the bottom of the aperture and confirm the formation of bulky structures at very low

overpotentials (+500 mV). Likewise, Figure 3-6 B demonstrates the deposition profile at −500

mV, where the deposition front moves quickly along the wall of the aperture to the edge. This

results in the formation of flaky structures at extreme negative overpotentials. Figure 3-6 C and

Figure 3-6 D also demonstrate the profile of the deposit thickness along the vertical wall of the

aperture for the bulky and flaky structures, respectively.

45

3.3.6 Surface area of nanostructured 3D microelectrodes

We also investigated variations in surface area for structures generated with different deposition

overpotentials by monitoring cyclic voltammograms generated in sulfuric acid107 (Figure 3-7 A).

Figure 3-7 Effect of deposition overpotential on surface nanostructuring. (A) Surface area of

3D structures deposited at different overpotentials (from +500 mV to −750 mV) measured in 50

mM H2SO4. Insets: SEM images of structures generated with deposition potentials of (1) +150

mV, (2) 0 mV, (3) −150 mV, (4) −750 mV. (B) Schematic of the individual spikes with different

textures deposited in the range of (1) +150 to +50 mV, (2) 0 to −50 mV,( 3) −150 to −250 mV.

All scale bars are 2 μm.

We observe that (not surprisingly) the structures generated with low overpotential (+500 mV)

exhibit the smallest surface areas (∼1200 μm2). In contrast, spiky structures generated at higher

overpotentials boast a much larger surface area. Interestingly, a 3-fold increase is observed for the

spiky structures deposited at −150 mV relative to those deposited at +50 mV, despite a similar

overall footprint. High-resolution SEM images demonstrate that the needles within these structures

have finer nanoscale features that are produced only at higher overpotentials. This finding is also

46

in accordance with the model proposed in Figure 3-5, which predicts that increasing the

overpotential results in the generation of smaller clusters of nuclei with finer spiky morphology.

As the potential increases above −450 mV, the needles are replaced by fine flakes and the surface

area decreases to ∼2600 μm2 at −750 mV (SEM image, Figure 3-7 A, inset 4). These structures

have a different overall morphology, with nucleation occurring on the edge of the aperture

followed by planar growth (described in Figure 3-5 and Figure 3-6). As a result, fine flakes with

smooth texture are formed, limiting surface area.

3.3.7 Study of DNA hybridization efficiency on the surface of 3D microelectrodes

We compared the efficiency of DNA hybridization on a subset of the structures generated in our

mechanistic study. The efficiency and kinetics of DNA hybridization depend strongly on the

density and accessibility of DNA probes attached to a surface.108–110 In general, a higher density

of probe strands attached to the surface produces larger signals upon hybridization to redox-tagged

complementary strands. However, extremely high probe densities can diminish the rate of

hybridization and the efficiency of detection. This is due to the accumulation of more negative

charges and the steric hindrance of the populated capturing stands on the surface, which limits the

ability of target molecules to easily access and hybridize to their complements.108

To investigate the efficiency of DNA hybridization on our 3D sensors, we immobilized a closely

packed monolayer of thiolated single-stranded DNA probes on their surface. A complementary

DNA strand labeled with methylene blue is then used to monitor the kinetics of hybridization

(Figure 3-8 A).

47

Figure 3-8 DNA detection assay based on 3D gold microsensors. (A) Capture probe (blue) is

attached to the sensor. The complementary DNA strand (red) carries methylene blue redox tag

(MB) to the electrode surface, which generates an electrochemical signal upon hybridization of

DNA strands. (B) Kinetic responses corresponding to DNA−DNA hybridization at sensors

deposited with different overpotentials. (C) The hybridization rate and (D) the corresponding t1/2

values show the pronounced variation in the efficiency of hybridization induced by different

sensor morphologies.The kinetic responses (B) and the corresponding rate measurements (C)

reveal that the rate of DNA−DNA hybridization exhibits a strong dependence on sensor

morphology. Spiky structures with finer nanostructuring provide more accessibility for the target

hybridization and accelerate the rate of DNA−DNA hybridization. However, on the flaky

structures, the rate of hybridization is reduced, likely due to a blocking effect. The structure of

the flakes creates hidden sites that may trap probe strands during overnight immobilization but

have limited accessibility during the hybridization time. The calculated time t1/2 (the time needed

for half of the capturing strands to hybridize to signaling strands) is also reported in D.

target DNA

probe DNA

AMB

B

C D

48

3.4 Conclusion

We conducted a detailed study of the electrodeposition of three-dimensional gold microsensors

with varied morphologies. This study revealed the influence of a number of important parameters

on the assembly of gold both on a nano- and microscopic scale. This allowed for the customization

of parameters to encourage the growth of 5-fold twinned needles which are of particular promise

due to their proven utility in biosensing applications. At low overpotentials (+500 to +300 mV)

the progressive nucleation and low growth rate encourage the formation of bulky structures. By

increasing the overpotential, two discrete stages of nucleation resulted that promoted the creation

of spiky structures (+150 to −250 mV). The more instantaneous the nucleation processes

(encouraged by higher overpotential) are, the finer is the nanostructuring of the spiky structures.

At overpotentials above −450 mV, flaky structures are formed as a result of nucleation on the edge

of the aperture. In order to understand the importance of morphology in DNA-based biosensing

applications, we studied the rate of hybridization on the surface of these electrodes. Optimal

structures that promoted highly efficient hybridization kinetics were identified.

3.5 Methods

3.5.1 Chip fabrication

Gold-coated glass wafers (with chrome adhesion layer and positive photoresist coating) were

purchased from Telic Company (Valencia, CA). Gold was patterned to create a series of seven

leads using standard lithography of the photoresist followed by etching of the gold and chrome

layers. Negative photoresist (SU-8 2002) was applied by spin coating and developed using

photolithography to create single 10 μm apertures on each lead.

3.5.2 Electrodeposition

Electrodeposition of 3D microelectrodes was performed with a BASI Epsilon potentiostat in DC

voltammetry mode. A three-electrode setup was employed with voltage measured relative to an

Ag/AgCl reference electrode, and a platinum wire was used as the counter electrode. The

deposition time was adjusted for each particular structure in order to keep the quantity of deposited

gold roughly constant (i.e., the total coulombs of charge transferred). For the concentration series,

the solution was between 1 mM and 500 mM HAuCl4 in a 0.5 M HCl supporting electrolyte. The

voltage applied was 0 mV relative to Ag/AgCl. At 50 mM, the time of deposition was 100 s and

49

other deposition times were adjusted accordingly. For the voltage series, the concentration of

HAuCl4 was 50 mM and the voltage was varied from +500 mV to −700 mV. For the viscosity

series glycerol was added to between 0% and 75% v/v (HCl final concentration was maintained at

0.5 M), with the voltage set to 0 mV and the HAuCl4 concentration maintained at 50 mM. The

surface area of 3D microelectrodes was calculated by measuring the area of the Au reduction peak

(∼0.80 V vs Ag/ AgCl) in 10 mM H2SO4 solution. The cyclic voltammetry studies were carried

out in the range of 0−1.5 V, and the reduction charge was obtained by integrating the reduction

peak. The result was then divided by 500 μC/geometric cm2 to obtain a geometric estimation of

the surface area.107

3.5.3 Surface characterization

Scanning electron microscopy (SEM) images were acquired on a Quanta FEG 250 ESEM. The

instrument was used in the high vacuum mode, with typical parameters being 10 kV bias and a 2.5

nm spot size. Magnifications between 1000× and 50 000× were employed. Focused ion beam (FIB)

imaging was performed using the Hitachi nanoDUE’T NB5000. Samples were first grounded

using a thin layer of silver paste. A thin layer of protective tungsten was then applied to the area

of interest using the machine’s built-in deposition capacity. Sectioning was performed using high-

energy ions to reveal a cross section of the structure. SEM imaging was then done using the built-

in SEM module.

3.5.4 Nucleation and growth model

The study of electrochemical three-dimensional nucleation processes involved correlating the

current to the number of clusters on the electrode surface.101–106 As the nucleation rate per active

site on the surface, A, and the number density of active sites for nucleation, N0, vary with

overpotential of deposition, the potentiostatic current transients establish the relationship between

deposition overpotential and the kinetics of nucleation:101,102

𝑖 = (

𝑧𝐹𝐷1/2𝑐

𝜋1/2𝑡1/2) (1 − 𝑒𝑥𝑝{−𝑁0𝜋𝑘𝐷[𝑡 − (1 − 𝑒−𝐴𝑡)/𝐴]}), (3.1)

where i is the current density; z, distance normal to the plane; F, the Faraday constant; D, diffusion

coefficient; c, bulk concentration; t, time; N0, number density of active sites; k, dimensionless

constant for growth rate of diffusion zones. Considering the im as the maximum current taking

50

place at tm, Equation 3.1 can be presented in a dimensionless form by simply plotting i2/im2 vs t/tm

for different values of the dimensionless parameter α = N0πkD/A. The two extreme forms of

nucleation are defined for small α or fast nucleation on a limited number of active sites,

“instantaneous” nucleation, and large α or slow nucleation on a large number of active sites,

“progressive” nucleation.

3.5.5 Electrode functionalization

Thiolated-DNA strands (0.1 mM) were incubated with TCEP (10 mM) for 1 h to reduce the

disulfide bonds. The solution was then diluted to 100 nM PBS 1×, pH 7.0. Glass chips having a

set of seven microelectrodes were incubated in 100 μL of 100 nM solutions of thiolated strands

overnight. Chips were then rinsed with DI water and incubated in 100 μL of 3 mM MCH in buffer

for another 3 h to displace nonspecifically adsorbed DNA and passivate the remaining electrode

area. After thoroughly rinsing with DI water, chips were stored in buffer. The thiolated strand

surface density (i.e., the number of DNA molecules per unit area of the surface) was determined

to be ∼6 × 1012 strands/ cm2 by measuring peak current.108

3.5.6 Sensor measurements

Electrochemical measurements were performed at room temperature using an EmStatMUX

potentiostat multiplexer (PalmSens Instruments, The Netherlands) and a standard three-electrode

configuration containing a platinum counter electrode wire, Sigma-Aldrich and an Ag/AgCl (3 M

NaCl) reference electrode (CHI). Experimental data were collected using square wave

voltammetry from −0.05 to −0.45 V in increments of 0.001 V vs Ag/AgCl, with an amplitude of

50 mV and a frequency of 60 Hz. Peak currents were fitted using the manual fit mode in the

PSTrace software (PalmSens). All measurements were taken immediately after adding the reagents

to the solution (100 nM of signaling strands) to measure the kinetics of hybridization.

51

Power-free, digital and programmable dispensing of picoliter droplets using a Digit Chip

So far this thesis has illustrated a number of elements which have application in a POC setting, but

which still rely on external instrumentation. In this chapter, we explore a device architecture

designed to be used with a minimum of external equipment and to perform a number of simple

assays in an intuitive and low-cost manner. The primary motivation for this device is the ability to

move microscopic quantities of fluid in a controlled manner without requiring external pumping

infrastructure, since the manipulating of fluid is the most fundamental operation of a microfluidic

device. Fluid movement is actuated by the press of a button with a single finger (digit), allowing

intuitive operation. This macroscopic operation is converted into discrete microscopic flow

through the integration of a series of capillary valves which break upon each instance of applied

pressure. This button-valve pairing is applied in the production of a concentration gradient and, in

conjunction with a cheap smartphone-based fluorimeter, the measurement of bacterial antibiotic

susceptibility. This assay requires no additional components and can be performed with minimal

training.

Reprinted with permission from Mepham A., Besant J.D., Weinstein A. W., Burgess I.B., Sargent

E.S., Kelley S.O., “Power-free, digital and programmable dispensing of picoliter droplets using a

Digit Chip” Lab on a Chip, 2017, 17 1505-1514. Reproduced by permission of The Royal Society

of Chemistry.

Link to publication online: https://doi.org/10.1039/C7LC00199A

Disclosure of work within this manuscript: A.M., J.D.B. A.W.W. and I.B.B. designed the

experiments. A.M, J.D.B. and A.W.W. performed experiments. Data analysis and manuscript

writing were performed by A.M. and J.D.B. with assistance from A.W.W, I.B.B., E.H.S., and

S.O.K.

52

4.1 Abstract

There is a growing need for power-free methods to manipulate small volumes of liquids and

thereby enable use of diagnostic assays in resource-limited settings. Most existing self-powered

devices provide analog manipulation of fluids using paper, capillary or pressure-driven pumps.

These strategies are well-suited to manipulating larger micro- and milliliter-scale volumes at

constant flow rates; however, they fail to enable the manipulation of nanoliter and picoliter

volumes required in assays using droplets, capillary sampling (e.g. finger prick), or expensive

reagents. Here we report a device, termed the Digit Chip, which provides programmable and

power-free digital manipulation of sub-nanoliter volumes. The device consists of a user-friendly

button interface and a series of chambers connected by capillary valves that serve as digitization

elements. Via a button press, the user dispenses and actuates ultra-small, quantitatively-

programmed volumes. The device geometry is optimized using design models and experiments

and precisely dispenses volumes as low as 21 pL with 97% accuracy. The volume dispensed can

be tuned in 10 discrete steps across one order-of-magnitude with 98% accuracy. As a proof-of-

principle that nanoliter-scale reagents can be precisely actuated and combined on-chip, we deploy

the device to construct a precise concentration gradient with 10 discrete concentrations.

Additionally, we apply this device alongside an inexpensive smartphone-based fluorescence

imaging platform to perform a titration of E. coli with ampicillin. We observe the onset of bacterial

death at a concentration of 5 μg mL-1, increasing to a maximum at 50 μg mL-1. These results

establish the utility of the Digit Chip for diagnostic applications in low-resource environments.

4.2 Introduction

Despite recent advances in microfluidics for diagnostics, many of the most sensitive tests remain

unavailable in resource limited settings due to the prohibitive cost and power requirements of the

necessary instrumentation.111 Most microfluidic systems require costly electrically-powered

pumps for fluidic actuation, and this limits the deployment of these technologies in the challenging

field conditions of developing world environments, where in many cases the requirements of

portability and remote location may limit access to reliable sources of electric power.

53

To increase access to point-of-care diagnostic technologies, low-cost and low-power fluidic

actuation systems are needed.111,112 Recently, compelling strategies for passive fluid control have

been reported that have included lateral flow paper microfluidics,113–115 vacuum pumps,116,117

pressure pumps,118,119 fibres,120 and capillary pumps.29,121 Sophisticated sample manipulations are

possible using these power-free actuation systems, and multi-step assays have been

demonstrated.116,122–126 However, these analog approaches to fluid manipulation are typically

optimized to regulate the flow rate of bulk fluids, and have not yet been demonstrated to be well-

suited for complex manipulations of very small reagent volumes.

Many diagnostic assays require the use of small samples or reagent volumes, including those that

sample blood from a finger prick and those that require expensive reagents. As a result, several

chip-based techniques for digital manipulation of small volumes have been developed including

droplet127–129 and digital microfluidics.130–132 In both cases, droplet actuation requires a power

source such as an electric pump (droplet microfluidics) or a high voltage source (digital

microfluidics). These requirements limit the feasibility of these methods in resource-limited

environments.

Developing power-free platforms for digital manipulation of small sample volumes poses two

distinct challenges not present in their analog counterparts: 1) dispensing mechanisms need to be

very precise and optimized for small volumes;133,134 2) sample controls need to be modular and

compatible with a user-friendly interface. Innovative architectures for self-powered,113 or

manually-powered135 fluid manipulation, including the SlipChip136–139 have recently overcome a

number of key challenges; however, thus far these self-powered techniques only allow a limited

number of sequential manipulations and lack the modularity and programmability of powered

techniques such as digital microfluidics. Devices driven by finger pressing have also been

developed,140,141 however, these devices typically make use of one directional flow valves and as

such are considerably more difficult to manufacture. Moreover, these devices cannot manipulate

picoliter volumes of fluid. Finally, these devices typically metre only a single, fixed volume of

fluid.

Here, we develop a user-friendly interface for power-free and digital manipulation of small

volumes using capillary valve- dispensers as modular elements, which are linked to pressure-

regulated buttons. The device, termed the Digit Chip, uses the pressure applied from pressing

54

buttons manually to break individual capillary valves and move liquids in fixed-volume

increments. Using theory and experiments, we optimize the geometry and surface properties of the

valves and pressure-regulated buttons, and show that droplets can be dispensed and manipulated

accurately across a wide range of volumes down to the picolitre scale. Dispensing, actuating, and

mixing reagents are essential components of many important biological and chemical assays. To

demonstrate that small volumes can be precisely actuated and mixed, we use the Digit Chip to

create a precise, discretized concentration gradient with nanoliter volumes. We also illustrate how

this architecture enables a convenient and low-cost technique to measure the susceptibility of

bacteria to antibiotics. This is an important capability that could allow small collections of bacteria

to be assessed for drug resistance via phenotypic testing without the need for any type of traditional

lab infrastructure. The Digit Chip could also be applied in a variety of biological and chemical

assays which depend on reagent dilutions and chemical gradients including generation of standard

curves, optimization of reaction conditions,142 and chemotaxis.143

4.3 Results and discussion

4.3.1 Overview of the digit chip

The Digit Chip consists of a series of chambers connected by capillary valves that serve as

digitization elements (Figure 4-1 A).

55

Figure 4-1 Overview of the Digit Chip. (A) Schematic illustrating precise and user-

programmable dispensing of ultra-low volumes using a Digit Chip Chambers, which serve as the

digital elements, are connected in series by capillary valves. After each button press, an

additional chamber is filled with fluid. (B) A schematic illustrating the spontaneous filling of a

chamber after pressure-induced bursting of a capillary valve. (C) Theoretical bursting pressure as

a function of capillary valve width with a 90° valve expansion angle and 50 μm channel height.

(D) Theoretical bursting pressure as a function of valve expansion angle for various contact

angles assuming a 50 μm valve width and a 50 μm channel height.

Upon manual application of a pressure via a button, the fluid bursts through the first capillary valve

and enters the adjacent chamber (Figure 4-1 B). Through capillary pressure, the chamber fills

spontaneously until the fluid reaches the subsequent capillary valve and the flow is arrested. The

user can opt to fill the next chamber by re-pressing the button. The solution volume dispensed is

programmed by the number of times the user applies pressure to the button.

The principle underlying the design of the device relies on the function of capillary valves created

within the fluidic structure.144 At the interface between the narrow and wide regions, the liquid–

air interface becomes pinned and requires a certain threshold pressure to resume movement. A

valve exists between the narrow linear channel and the larger circular chamber, and the pressure

is supplied by the press of a button. Once the valve bursts, the fluid will spontaneously fill the

56

circular chamber. As long as the user only presses the button for a short period of time, the pressure

is relieved before the fluid reaches the subsequent valve. As such, the pressure is below the critical

pressure upon arrival at the next chamber, and motion ceases.

4.3.2 Bursting pressure model and design principles

Each valve must have a bursting pressure low enough that it can be readily applied by a human

finger, but large enough that the valve does not burst spontaneously. To calculate the capillary

valve bursting pressure, we use the following previously described equation for the maximum

sustainable pressure across the meniscus in a rectangular configuration of the capillary burst

valve:144

𝛥𝑃 = −2𝜎 (

cos(𝜃𝐼)

𝑤+

𝑐𝑜𝑠(𝜃𝐴)

ℎ), (4.2)

where ΔP is the pressure difference across the liquid–air interface, w is the valve width, h is the

channel height, σ is the surface tension of the liquid–air interface (72.9 mN m-1), θI is the contact

angle with the initial side-wall prior to the expansion, and θA is the critical advancing contact angle.

Upon the application of pressure, the meniscus will bulge until the contact angle with the new

sidewall after the valve reaches the critical advancing contact angle, θA. This occurs when θI = θA

+ β, where β is the valve expansion angle, or when θI is greater than 180°, the maximum sustainable

contact angle.144 Thus the valve will burst when θI equals the lower of θA + β or 180°. These

equations assume that the fluid is dispensed from an infinite reservoir and do not consider the

receding interface.

Using this equation, we modeled the bursting pressure as a function of valve width (Figure 4-1 C)

and valve expansion angle (Figure 4-1 D) for a variety of contact angles. As expected, the bursting

pressure increases with narrower valves and larger valve expansion angles. The calculations

indicate that a valve expansion angle of approximately 90° or higher is optimal as the bursting

pressure rapidly drops off for angles less than 90°. At low contact angles, the bursting pressure

approaches zero for valve widths around 25 μm, thus valves narrower than 25 μm are ideal.

It is critical that the device surface be hydrophilic so that the chambers fill spontaneously via the

capillary pressure after the valves burst. Thus, the contact angle of the fluid should be less than

57

90°. On the other hand, Figure 4-1 C and Figure 4-1 D suggest that at low contact angles, the

valves will burst spontaneously for most valve widths and valve expansion angles. Thus, the

surface should be engineered to be only slightly hydrophilic. We measured the critical advancing

contact angle of PBS on PDMS for a variety of oxygen plasma treatment times (Appendix Figure

B-4). We found that with a 30 s oxygen plasma exposure, the contact angle of PBS on PDMS is

70° which is compatible with the Digit Chip. PDMS is a useful material for this device because of

its compatibility with rapid prototyping. While its elastomeric properties are useful in the user

interface, they are not essential for the functioning of the wells. The stability of PDMS surface

chemistry was sufficient for our device to function properly within the same day of plasma treating.

However, using different materials for the wells (e.g. glass) that have a more stable surface

chemistry may be more advantageous if the device were to be mass-produced and stored for longer

periods before use.

4.3.3 Optimization of device geometry

Motivated by these calculations, we fabricated an array of devices with various capillary valve

widths and expansion angles to further refine the design. Devices were fabricated by pouring

PDMS on a 50 μm tall SU-8 master mold patterned using standard photolithography. After curing,

the PDMS was plasma treated and bonded to a glass substrate.

Figure 4-2 A shows the measured bursting pressure as a function of valve width and Figure 4-2 B

shows the bursting pressure as a function of valve expansion angle.

58

Figure 4-2 Experimental investigation of the device geometry and its optimization. (A)

Experimentally measured bursting pressure as a function of valve width. The dotted line

represents the calculated bursting pressure assuming a 70° contact angle and 90° valve expansion

angle. (B) Experimentally measured bursting pressure as a function of valve expansion angle.

The dotted line represents the calculated bursting pressure assuming a 70° contact angle and 50

μm valve width. (C) The measured bursting pressure as sequential chambers are filled. (D) The

measured volume dispensed as a function of chamber volume. (E) Images illustrating the

spontaneous filling of a 12 nL chamber after the valve is burst by a user-applied pressure.

Sequential dispensing of volumes in approximately (F) 12 nL, (G) 580 pL and (H) 140 pL

increments. Errors bars represent standard error.

We compared our measured values to the theoretical predictions and in both cases, we find good

agreement with the theory. In all of our designs, the channels connecting the wells were sufficiently

short to ensure that there was not enough buildup of flow momentum between wells to break a

capillary valve. Figure 4-2 C shows the bursting pressure as a function of the number of valves

filled. On average, we observe only a 5% increase in bursting pressure after each sequential valve

is filled.

59

We studied the accuracy of filling as a function of chamber size (Figure 4-2 D). A series of devices

were designed with chamber diameters ranging from 55 μm to 800 μm and channel heights ranging

from 5 μm to 50 μm tall. Since the capillary valve bursting pressure depends on the width of the

valve and the expansion angle, the sizes of the circular chambers can be freely adjusted to allow

for different volumes while performing in the same manner. The volume of these chambers ranged

from approximately 21 pL to 24 nL. The valve widths scale with the chamber size and range from

2.5 to 20 μm wide. We found that all chambers and wells, including those as small as 21 pL, could

be filled with at least 97% accuracy. The remaining chamber sizes filled with high accuracy

(Appendix Figure B-5). A small amount of error is caused by the incomplete filling of some

chambers due to the occasional formation of small bubbles. However, the valves were stable, not

spontaneously breaking even after many minutes. A small amount of evaporation does occur if the

device is left unattended for a manner of minutes, but this does not destabilize the air–liquid

interface.

This smallest dispensing volume achieved approaches the minimum achievable limit for our

design, which we estimate to be 15–20 pL. The fabrication tolerance of our photolithography

process creates minimum lateral sizes of the channel and chamber (to maintain the near 90°

spreading angle). The valve design places an additional fundamental constraint on the valve aspect

ratio, which limits how small the channel height can be. Since the free-energy barrier encountered

at the side walls in the chamber opening must always exceed the favourable free-energy change

associated with wetting the top and bottom walls, the width:height aspect ratio in the channels

connecting the chambers must stay below a maximum value. This maximum aspect ratio is

derivable from Equation 4.2 and given by:

[𝑤/ℎ]𝑚𝑎𝑥 = −

𝑐𝑜𝑠(𝜃𝐴 + 𝛽)

cos(𝜃𝐴). (4.3)

A contact angle of 70° and a spreading angle (β) of 90°, gives a maximum aspect ratio is

2.75, limiting how small the channel height can be as a function of the width. This limit was in

agreement with experiments showing that further reduction of the channel height for our smallest

chamber size led to spontaneous breaking of the valves.

4.3.4 Designing a user-friendly interface

60

We sought to design an interface that allows the user to easily apply the appropriate pressure to

break one capillary valve. The interface consists of two hollow PDMS chambers connected in

series to the sample loaded in the Digit Chip (Figure 4-3 A and Figure 4-3 B).

Figure 4-3 The Digit Chip interface for controlled dispensing of droplets. (A) Schematic

illustrating the user-friendly interface for precise application of pressure. The interface consists a

button and a pressure regulation chamber patterned in PDMS. When the user depresses the

button with a thumbpress, the applied pressure is controlled by the size of the pressure regulation

chamber. Pressure is vented through a small hole when the user releases the button. (B) Image of

the device with a nickel shown for scale. (C) The applied pressure as a function of the ratio of the

size of the button and pressure regulation chambers. (D) Measurement of the volume displaced

when the button is depressed. (E) The number of chambers filled as a function of the size of the

pressure regulation chamber. Single chambers could be accurately filled when the button

chamber was 15% the volume of the pressure regulator. (F) Images of solution dispensed in 24

nL increments using the PDMS button. (G) Accuracy of filling in 24 nL increments after each

sequential button press using the user-interface. Solution was dispensed more than 10 times

sequentially. Errors bars represent standard error.

When the user fully compresses the first chamber, which serves as the button, the gas within the

chamber is displaced and a pressure is applied to the sample in the Digit Chip. The applied pressure

61

is regulated by tuning the size of the second pressure-regulation chamber. The pressure is vented

through a small hole when the user releases the button.

The applied pressure difference generated by pressing the button can be approximated using

Boyle's law:

𝛥𝑃𝑎𝑝𝑝𝑙𝑖𝑒𝑑 = 𝑃𝑎𝑡𝑚 (

𝑉𝑏 + 𝑉𝑠

𝑉𝑠− 1), (4.4)

where ΔPapplied is the applied pressure difference, Patm is the atmospheric pressure of 101 325 Pa,

Vb is the volume of the button chamber and Vs is the volume of the regulation chamber and the

connective tubing. This equation assumes the button is fully compressed when pressed. This two-

chamber design limits the maximum pressure a user can apply to the valve (achieved when the

button is fully compressed) and therefore ensures that the user cannot press too hard.

We measured the applied pressure as a function of the ratio of the size of the button and pressure

regulation chambers (Figure 4-3 C). The measured pressures are lower than the theoretical

predictions (Figure 4-3 D), suggesting that the button chamber retains about 20% of its volume

when fully depressed.

To test the efficacy of this interface, we connected various designs in series to a Digit Chip with

20 μm wide valves and 24 nL chambers. We measured the number of chambers filled per button

press (Figure 4-3 E). We found that when using an interface with a ratio of the two chambers of

15%, we could accurately fill a single chamber per button press. The applied pressure from this

interface design is 11 kPa, which is higher than the approximately 6 kPa capillary valve bursting

pressure predicted by Equation 4.2. This is expected as the pressure generated by the user must be

greater than the valve bursting pressure due to the pressure drop along the tubing, fluid reservoir,

and channel and the fact that the flexible PDMS chambers expand under pressure. Figure 4-3 F

shows images of dispensing fluid in 24 nL increments using this interface. Using the interface, we

found that we could accurately dispense liquid in 24 nL increments over 10 times sequentially.

The volume dispensed ranged from 24 nL for 1 button press to over 300 nL for 13 button-presses

(Figure 4-3 G) and the chambers filled with over 98% filling accuracy and less than 3% standard

deviation (Figure 4-3 G). The volumes dispensed in Figure 4-2 F–H were further replicated using

the button to show its applicability to a variety of volumes (Appendix Figure B-6).

62

4.3.5 Generation of a discrete concentration gradient

Concentration gradients and reagent dilutions are important in a variety of biological and chemical

assays. As a demonstration that solutions can be accurately dispensed and combined, we used the

Digit Chip to mix small solution volumes in precise ratios (Figure 4-4 A).

Figure 4-4 Generation of a discretized concentration gradient. (A) Schematic illustrating on-

chip dispensing and mixing of reagents using the Digit Chip. After the sample is loaded, both

solutions are dispensed in various ratios in 24 nL increments. The two solutions were sent to a

central mixing chamber by manually injecting air using a syringe. (B) Air channels were valved

using a screw to depress the PDMS and block the channel. (C) On-chip generation of a

discretized concentration gradient using the Digit Chip. Errors bars represent standard error. (D)

Images acquired with an optical microscope after mixing the blue and yellow dye in various

ratios with the Digit Chip.

In this device, two separate solutions can be dispensed up to 10 times in increments of 24 nL. After

dispensing the liquid, the fluid is actuated towards the central mixing chamber by manually

injecting air using a syringe. The air channels are closed during dispensing by screw valves,145

which depress the channel (Figure 4-4 B).

We dispensed 10 different ratios of blue and yellow dyes in 24 nL increments and mixed the dyes

in the central chamber on chip. For each concentration, the sum of the number of both droplets

63

dispensed was kept constant at 10, corresponding to a total volume after mixing of 240 nL. Figure

4-4 C shows the measured ratios of the two solutions mixed on chip. The r2 value of the fit to the

line representing the expected ratio of the dyes is 0.98. The deviation in our measured values is

due to the compounded error of filling 10 wells and the fact that some liquid remains trapped in

the chambers after injecting air. Figure 4-4 D shows images of the resulting colors generated from

mixing the two dyes.

4.3.6 A low-cost platform for rapid determination of bacterial antibiotic susceptibility

In order to demonstrate the usefulness of the Digit Chip for a practical application, a device was

developed which tests the effect of antibiotic concentration on bacterial growth.146

Figure 4-5 Testing of antibiotic susceptibility. (A) Schematic illustrating architecture of

device. Insets show Digit Chip motif (i) and capillary valve between bacterial metering channel

(yellow) and growth chamber (ii). (B) Simple, smartphone-base fluorescence platform allowing

for excitation with green light and imaging of resultant red fluorescence. (C) Time series

demonstrating the conversion of non-fluorescent resazurin to highly fluorescent resorufin by

bacterial metabolism. (D) Curve illustrating the effect of ampicillin titration on the viability of E.

coli.

64

The chip is comprised of a pair of channels leading into a central growth chamber, each with a

respective air inlet (Figure 4-5 A). A single outlet is present on the far side of the growth chamber.

The top channel is the standard Digit Chip motif, with 10 chambers separated by capillary valves

(Figure 4-5 A, i). By filling the desired number of chambers with antibiotic solution and expelling

the metered antibiotic into the growth chamber, between 1 and 10 equivalents of antibiotic are

introduced. The side channel is a single bacterial metering channel separated from the growth

chamber by a capillary valve (Figure 4-5 A, ii). This chamber is filled with a fixed volume of

bacterial solution, which can be transported by air into the growth chamber where it mixes

diffusively with the antibiotic solution. This device can then be submerged in a 37°C water bath

to allow for bacterial growth.

To observe the viability of the bacteria, resazurin dye was added to the bacterial medium. Only in

the presence of actively metabolizing bacteria will the resazurin be reduced to its fluorescently

active product, resorufin. In order to measure the fluorescence of this product, a simple smart

phone-based imaging platform was devised (Figure 4-5 B). The excitation system consists of a

high-powered green LED light source, along with a collimating lens to focus the excitation light

and a plastic green filter to remove extraneous wavelengths. This light is focused onto the growth

chamber of the chip from an oblique angle to minimize detection of the excitation light. The

emitted light is passed through a 589 nm bandpass filter and imaged using an LG G3 smartphone

camera. To magnify the image of the growth chamber, a PDMS lens was fabricated and affixed

directly onto the back of the smartphone, covering the camera lens. The resultant RGB image was

then split into its component red, green and blue channels, and the average intensity of the red

channel was measured. In keeping with the intended use of the Digit Chip in low resource settings,

the entire platform including all components can be purchased and constructed for under 60 USD.

In order to confirm the function of the platform, a high concentration (1.0 × 108) of bacteria was

introduced into the growth chamber in the absence of antibiotic and the fluorescence was imaged

every 45 minutes for 3 hours. The resultant images display monotonically increasing fluorescence

as the bacteria replicate and metabolize (Figure 4-5 C). Evidently, the system is more than adequate

for detecting the conversion of resazurin to resorufin and thereby measuring bacterial viability.

The susceptibility of E. coli to ampicillin was chosen as a suitable test case to confirm the efficacy

of the Digit Chip. A wide range of concentrations (0, 1, 2, 5, 10 and 50 μg mL-1) of antibiotic were

65

introduced using the discretized measurement chambers and two stock solutions (100 μg mL-1 and

1000 μg mL-1). These were mixed with bacterial solution (12.5 × 106 cfu mL-1) and incubated for

8 h prior to measurement of fluorescence. The results are illustrated graphically in Figure 4-5 D.

The graph follows a standard sigmoidal shape, with bacterial death beginning at a concentration

of about 5 μg mL-1 and increasing until 50 μg mL-1. These results are in good agreements with past

studies, which have shown a minimum inhibitory concentration of 2 μg mL-1147 and complete

inhibition of growth at 50 μg mL-1.148 Evidently, the ability of the Digit Chip to create titrations

with a wide range of concentrations is ideally suited to investigating bacterial antibiotic

susceptibility. Moreover, due to the extremely small volumes used (on the order of 100 nL) very

small quantities of bacterial sample and antibiotic solution are required. This is particularly

important in situations where the sample is limited by biological or financial constraints.

Furthermore, this device is amenable to any combination of bacterial species and antibiotic,

allowing for widespread implementation.

4.3.7 Discussion

This type of self-powered digital fluidic device may be useful for a wide class of assays in low-

resource settings that require low sample or reagent volumes (e.g. pin-prick assays), or that require

complex multi-step manipulations or precise timing that are difficult to automate using existing

passive analog fluidics. Using this device, the user controls the time at which various reagents are

introduced which eliminates the need for built-in timing mechanisms. This device is especially

useful for assays in which the volumes dispensed are systematically varied, such as titrations. This

device architecture is compatible with many common fabrication techniques currently used to

make channel-based microfluidics at low cost.149 As with other microfluidic technologies, the

replacement of PDMS with more cost-effective alternatives (e.g. plastics, glasses) might also

accompany the transition to larger scale manufacturing.149 For our device this would have the

added benefit of enabling us to choose materials whose surface chemistry is more stable and

suitable for long-term storage.

Along with these advantages, the Digit Chip does have some weaknesses. The nature of the

capillary valve requires that metered fluid be displaced and replaced with air before the valve is

re-established, meaning that the chambers need to be flushed with air between subsequent

66

dispensings. Furthermore, valve dimensions need to be precise to allow for correct behaviour.

However, these drawbacks are relatively minimal given the simplicity and accuracy of the device.

We applied the Digit Chip to evaluating antimicrobial resistance, which is an important target

application to enable in resource-limited settings. The features of the Digit Chip would make it

straightforward to carry out the type of drug titration that is typically used to assess whether a

frontline antibiotic will be efficacious. This approach is also applicable to any of the multitude of

tests in biology and chemistry that require titration. For example, the testing of dissociation

constants and the activity of enzymes could be tested using minute quantities of reagents, provided

that a measurable color change occurs.

The adoption of this technology in a wider variety of assays will also include its functioning with

complex fluids, such as whole blood, that have a higher viscosity and contain suspended particles.

Although the core principles of our device are independent of viscosity, whose only independent

effect on our chips was to change the speed with which the chambers filled, the presence of

suspended particles (e.g. blood cells) would place size restrictions on our chambers, with the

smaller geometries having the capacity to become easily clogged.

Increasing the complexity of manipulations and involving more reagents in the Digit Chip would

require more inlets, and a 2D configuration of these buttons and valves. Valving mechanisms must

be added at each node to control the directionality of pressure applied, just as was illustrated here

to make the concentration gradient. One natural extension of the existing device would be to have

two parallel Digit Chip motifs emptying into a common chamber, with one having chambers 10×

larger than the other. This “ones and tens” configuration would allow any integer value between 1

and 100 to be metered and dispensed with only a small increase in complexity.

Just as has been shown with digital microfluidics,130–132 a vast modular library of manipulations

and different assays could be built from the Digit Chip using a few stock 1D and 2D configurations

of wells, valves and buttons. These device configurations would need to be optimized for the

maximum number and type of fluid manipulations possible, but not with a need for any reagents

to be loaded beforehand. Therefore a single chip design could be used for many assays (e.g. any

titration involving a given number of reagents) or rapid experiments performed in the field.

However, this modularity comes at the cost of easy assay automation. A general limitation of user-

67

programmable digital fluidic platforms is that more sophisticated assays may require more buttons

and thus, greater user involvement and more complicated instructions

4.4 Conclusion

In summary, we introduce a device for power-free and user programmable manipulation of small

sample volumes. A sequence of microfluidic chambers connected by capillary valves connected

to pressure-regulated buttons allow for discretized dispensing of sub-nanoliter volumes. This

device architecture enabled power-free dispensing of volumes in increments as low as 21 pL and

near 97% accuracy. The volume dispensed could be fine-tuned in 10 increments across one order-

of-magnitude. The number of volume increments could be increased in future iterations of the

device. A series of devices is used to generate a concentration gradient with 10 discrete

concentrations in sub-microliter volumes, and to determine the susceptibility of bacteria to various

concentrations of antibiotic. This device could help bring a new class of assays, which require

sophisticated manipulations of small volumes, to low-resource settings.

4.5 Methods

4.5.1 Digit chip fabrication

Using standard photolithography, we patterned a 50 μm tall SU-8 3050 (Microchem, MA) layer

on a silicon wafer (University Wafer, MA). PDMS (Dow Chemical, MI) was dispensed onto the

wafer and cured for 1 hour at 67 °C. After curing, the PDMS was removed from the silicon wafer

and holes were punched to form the inlet and outlet.

4.5.2 Fabrication of user-friendly interface

The mold was printed using a μPrint 3D printer. PDMS was dispensed onto the mold and cured at

67 °C for 1 hour. The PDMS was removed and holes were punched to create outlets and a small

vent in the button chamber. The PDMS was plasma treated and bonded to a glass slide. Silicone

tubing was used to attach the button to the Digit Chip.

4.5.3 Contact angle measurements

PDMS was treated with an oxygen plasma for a variety of exposure times. Advancing contact

angles were measured using ImageJ to analyse images of the droplets acquired using a camera.

68

4.5.4 Bursting pressure measurements

Valve bursting pressures were measured using a syringe pump connected to the chip. After

introducing PBS (pH 7.4) into the Digit Chip, the syringe pump was connected to the chip with

9.5 cm of silicone tubing (0.76 mm inner diameter). The initial volume of gas in the 1 mL syringe

and tubing was measured. The syringe was slowly compressed at a rate of 20–50 μL min-1 while

monitoring the capillary valve under a microscope. The pump was stopped and the pressure was

relieved as soon as the valve burst. The volume change of gas contained in the syringe and tubing

was recorded. The induced pressure was calculated using the initial and final volumes using the

ideal gas law. The changes in pressure were measured at least 5 times per chip.

4.5.5 Measurements of applied pressure using the elastomeric button

The outlet of the user-interface was connected to silicone tubing (0.76 mm inner diameter) filled

with a plug of PBS buffer (pH 7.4). After compressing the button, we recorded the change in

displacement of the plug. This change in displacement of the plug was used to calculate the volume

of gas displaced while pressing the button. This volume change was converted into a pressure

change using Boyle's law.

4.5.6 Chamber filling percentage measurements

The filling percentage of chambers was measured by acquiring optical images under a microscope

(Nikon) and analyzing the images using ImageJ.

4.5.7 Generation of a discretized concentration gradient

The device was fabricated from PDMS as stated above. Auxiliary air outlets were connected to

each fluid inlet to enable fluid to be pumped in from both sides. Screw valves were fabricated on

each air outlet by 3D printing a chuck to suspend the screws above the channel. The valves were

closed or opened by turning the screw clockwise or counter-clockwise. Before dispensing the dyes,

the air channels were blocked by closing the valves to ensure that liquid did not enter the air

channels. Both samples were introduced and dispensed for the desired number of chambers. The

sum of the number chambers filled with blue and yellow dyes was held constant at 10 for all

concentrations. Samples used were ddH2O with food coloring at 2 drops per mL. Devices were not

plasma treated or bonded to allow them to be reusable so 20% EtOH was added to the solutions to

lower the contact angle. To prevent leakage, the channels were held against the substrate under

69

light pressure. After dispensing, the valve was removed. Using a 1 mL syringe, air was injected by

hand to move the droplets to the middle chamber. Mixing was aided by pressing on the middle

chamber 3 times. Using an optical microscope, a picture of the mixed sample was recorded. The

ratio of the dye was measured using ImageJ by recording the intensity of the dye and comparing

it to the intensity of a bulk solution of dye mixed using standard laboratory pipettes and injected

into the on-chip mixing chamber.

4.5.8 Antibiotic susceptibility testing

The device was fabricated in PDMS as stated above. Ampicillin in 1× PBS (pH 7.4), either 100 μg

mL-1 or 1000 μg mL-1, was introduced into the antibiotic inlet. 1, 2, 5 or 10 chambers were filled

in the standard manner. Air was then used to push the antibiotic into the growth chamber. The

bacteria metering chamber was then filled with the bacterial solution (12.5 × 106 cfu mL-1 E. coli

in LB broth with 50 mM TBS pH 8.5 and 1 mM resazurin). Air was used to push this plug of fluid

into the growth chamber where it diffusively mixed with the antibiotic. The chip was then

incubated in a 37 °C water bath (to prevent evaporation) for 8 h.

4.5.9 Fluorescence image acquisition and analysis

A Luxeon Rebel Color LED (Green, LEDSupply) was used as a light source. A collimating lens

was used to focus the light into a beam. This beam was passed through a plastic green filter

(Roscolux) to remove any extraneous wavelengths and directed onto the growth chamber of the

Digit Chip. Emitted light was passed through a 589 nm bandwidth filter (Edmund Optics) and

imaged using a LG G3 cell phone camera equipped with an adhered PDMS lens.150 Images were

captured using the free Open Camera app and image analysis was performed using ImageJ. The

red channel of the RGB image was extracted and the average intensity of the pixels in the growth

chamber was measured. Each concentration of antibiotic was performed in triplicate and values

were normalized to maximum and minimum pixel values, with error bars showing standard error.

70

Conclusions and Future Outlook

5.1 Thesis Findings

In this thesis we explored a variety of topics in device design and application towards the probing

of biology, with the ultimate purpose of shedding light on the disease state and informing potential

treatments. Multiple aspects of devices were investigated with an eye towards application at the

POC, including microfluidic innovations, improvements in electrode design towards enhanced

sensing, simplification of platforms for greater accessibility, and testing using genuine biological

samples

In chapter 2, we challenged a microfluidic device to reveal the evolution of cancer in an animal

model. This technique allowed us to probe the progression of the disease by using the blood as a

liquid biopsy, thereby identifying changes in circulating tumour cells over time. The unique nature

of the device allowed both the number of CTCs and their invasive potential to be measured

multiple times a week, giving a high-resolution glimpse into the disease course. It was revealed

that both the total number of CTCs and the fraction of low-epithelial character CTCs increased

monotonically, in step with the growth of the primary tumour. This approach also allowed us to

observe the decrease in CTC load upon removal of the primary tumor, as the source for the majority

of CTCs was resected. Perhaps most importantly, we saw a marked rebound in CTC levels at a

later time, implying the presence of metastatic disease. The distinctive capacity of the device to

discriminate between CTC subpopulations had further utility in highlighting the biological changes

enacted by tumor growth and seeding.

One of the most effective means of biosensing is the use of microelectrodes. In chapter 3, we

pursued a fundamental understanding of the mechanisms underlying gold electrodeposition,

towards the development of improved sensing performance. We discovered significant and

separate effects for each of applied voltage, concentration of gold precursor, and modulation of

71

solution viscosity. A specific combination of conditions was established which reliably produced

a high surface area electrode with an intriguing geometry. Deeper probing of the underlying

mechanisms showed how this structure was rooted in nucleation and crystal twinning events early

in electrode growth. Finely tuned electrodes showed improved performance in the electrochemical

detection of nucleic acids, indicating possible application in biosensing applications.

A further necessity in the widespread implementation of point-of-care devices is the ability to

manipulate microscopic quantities of fluid in a reliable manner while not requiring the use of an

external instrument. As a step towards this end we examined the potential of capillary valves for

the metering and dispensing of minute quantities of fluid. Following a parametric search to

elucidate the dimensions best suited for reproducible behaviour, we devised a repeated valve motif

for discrete volume measurement. The actuation was made intuitive and power-free by pairing this

motif with a small external button, each press of which triggered the bursting of a single valve.

This architecture was first used to create a concentration gradient, demonstrating a functionality

with broad applications across chemistry and biology. Next, an application for the determination

of bacterial resistance and requisite antibiotic concentration was illustrated.

5.2 Future Outlook

The desire to create new and more informative devices for probing human disease is pushing

research forward on a large number of fronts. The work performed here aims to help advance this

goal. The work in chapter 2 served as a good initial assessment of the technology for tracking

cancer progression and identifying changes in cancer biology. However, this investigation was

confined to a single animal model of disease. A natural next step is investigating a variety of cancer

types, ideally in human patients. In fact, some studies to this effect have been performed in our

lab, and reveal the utility of the technology across cancers of varied origin and invasiveness.

Moreover, there is great potential in further probing the biology of the captured cells, since the

device retains cellular viability. Together with improvements in the sensitivity and reliably of

single cell “omics” techniques, devices such as this should soon be able to give unprecedented

insight into the abnormal cellular processes at work in cancerous cells. With this information in

hand, the ability to create more targeted and effective therapies will be significantly enhanced,

paving the way for a host of new cancer drugs with reduced side effects and reduced chances of

recurrence.

72

Similarly, the development of new electrodes with new properties is important not only for its own

sake, but for the advances in POC performance that it enables. Improved understanding of the

processes dictating electrode growth during electrodeposition allows for the custom tuning of

electrode properties. By controlling surface area and crystallinity of electrodes, improved ability

to capture biomolecular targets can be achieved without requiring increases in electrode footprint.

This in turn permits the incorporation of electrodes into more compact devices, a necessity if we

are to produce surgically implantable sensors such as those required for an artificial pancreas.

Apart from POC applications, there is evidence that gold electrodes with controlled surface

morphology have the ability to efficiently catalyze the reduction of greenhouse gases into useful

fuel stocks. Further refining the existing electrode architecture towards this end would be an

exciting avenue to pursue.

The primary idea behind the Digit Chip is allowing laboratory-style operations to be performed

using small samples volumes and in an intuitive manner, so as to allow for use by individuals with

minimal training. This simplification of testing is paramount for widespread adoption in areas with

limited resources. One envisioned extension of this technology is a small handheld cartridge which

features a small number of buttons facilitating basic laboratory functions (mixing, metering,

moving, etc.). By incorporating a simple optical or electrochemical sensor into such a device, a

variety of typical biochemical tests could be performed on a single platform. Such a device should

remain self-contained and power-free, with the possible exception of a simple battery for powering

the sensor element.

In addition to the specific outlooks of each project individually, there is potential for combining

the elements of each to provide additional functionality. Microelectrodes have the potential to

probe cell metabolic activity, which could add an extra dimension of characterization to the CTC

capture device. By placing electrodes in each of the chambers it may be possible to measure the

concentrations of key metabolites and see if these correlate to the degree of epithelial character.

Similarly, incorporation of electrodes into the antibiotic susceptibility model of the Digit Chip

would allow for resazurin to be detected electrochemically. This alteration would increase

sensitivity while simultaneously obviating the need for an external smart phone. Finally,

modification of the Digit Chip would allow for the discrete movement of cells between

compartments. By releasing cells from the CTC capture device and introducing them into this cell

manipulation Digit Chip, each CTC could potentially be individually cultured and characterized.

73

Taken together, the results of this thesis represent a step forward in the production of advanced

devices for probing the disease state. Their future application will improve disease

characterization, allowing for more accurate prognoses and better informed treatment.

74

References

1. Yager, P. et al. Microfluidic diagnostic technologies for global public health. Nature 442,

412–418 (2006).

2. Kelley, S. O. New technologies for rapid bacterial identification and antibiotic resistance

profiling. SLAS Technol. Transl. Life Sci. Innov. 22, 113–121 (2017).

3. Kuhn, L. et al. Human papillomavirus DNA testing for cervical cancer screening in low-

resource settings. J. Natl. Cancer Inst. 92, 818–825 (2000).

4. McDermott, U. & Settleman, J. Personalized cancer therapy with selective kinase inhibitors:

an emerging paradigm in medical oncology. J. Clin. Oncol. 27, 5650–5659 (2009).

5. Bissonnette, L. & Bergeron, M. G. Infectious disease management through point-of-care

personalized medicine molecular diagnostic technologies. J. Pers. Med. 2, 50–70 (2012).

6. Kissinger, P. T. Biosensors—a perspective. Biosens. Bioelectron. 20, 2512–2516 (2005).

7. DeLisa, M. P. et al. Evanescent wave long-period fiber Bragg grating as an immobilized

antibody biosensor. Anal. Chem. 72, 2895–2900 (2000).

8. Yemini, M., Reches, M., Gazit, E. & Rishpon, J. Peptide nanotube-modified electrodes for

enzyme−biosensor applications. Anal. Chem. 77, 5155–5159 (2005).

9. Sassolas, A., Leca-Bouvier, B. D. & Blum, L. J. DNA biosensors and microarrays. Chem.

Rev. 108, 109–139 (2008).

10. El-Sagheer, A. H. & Brown, T. Click chemistry with DNA. Chem. Soc. Rev. 39, 1388

(2010).

11. Cunningham, B. et al. A plastic colorimetric resonant optical biosensor for multiparallel

detection of label-free biochemical interactions. Sensors Actuators B Chem. 85, 219–226

(2002).

12. Arlett, J. L., Myers, E. B. & Roukes, M. L. Comparative advantages of mechanical

biosensors. Nat. Nanotechnol. 6, 203–215 (2011).

13. Wang, J. et al. DNA electrochemical biosensor for the detection of short DNA sequences

related to the human immunodeficiency virus. Anal. Chem. 68, 2629–2634 (1996).

14. Ehrenkranz, J. R. L. Home and point-of-care pregnancy tests: a review of the technology.

Epidemiology 13, S15–S18 (2002).

15. Oliver, N. S., Toumazou, C., Cass, A. E. G. & Johnston, D. G. Glucose sensors: a review

of current and emerging technology. Diabet. Med. 26, 197–210 (2009).

16. Osteryoung, J. G. & Osteryoung, R. A. Square wave voltammetry. Anal. Chem. 57, 101A–

110A (1985).

17. Kunduraci, M., Al-Sharab, J. F. & Amatucci, G. G. High-power nanostructured LiMn2 -

xNixO4 high-voltage lithium-ion battery electrode materials: electrochemical impact of

electronic conductivity and morphology. Chem. Mater. 18, 3585–3592 (2006).

18. Pang, S.-C., Anderson, M. A. & Chapman, T. W. Novel electrode materials for thin-film

ultracapacitors: comparison of electrochemical properties of sol-gel-derived and

electrodeposited manganese dioxide. J. Electrochem. Soc. 147, 444 (2000).

75

19. Xu, S., Luo, Y. & Zhong, W. Investigation of catalytic activity of glassy carbon with

controlled crystallinity for counter electrode in dye-sensitized solar cells. Sol. Energy 85,

2826–2832 (2011).

20. Heinze, J. Ultramicroelectrodes in electrochemistry. Angew. Chemie Int. Ed. English 32,

1268–1288 (1993).

21. Li, F., Han, X. & Liu, S. Development of an electrochemical DNA biosensor with a high

sensitivity of fm by dendritic gold nanostructure modified electrode. Biosens. Bioelectron.

26, 2619–2625 (2011).

22. Li, C. W., Ciston, J. & Kanan, M. W. Electroreduction of carbon monoxide to liquid fuel

on oxide-derived nanocrystalline copper. Nature 508, 504–507 (2014).

23. Lucarelli, F., Marrazza, G., Turner, A. P. F. & Mascini, M. Carbon and gold electrodes as

electrochemical transducers for DNA hybridisation sensors. Biosens. Bioelectron. 19, 515–

530 (2004).

24. Schouten, K. J. P., Qin, Z., Pérez Gallent, E. & Koper, M. T. M. Two pathways for the

formation of ethylene in CO reduction on single-crystal copper electrodes. J. Am. Chem.

Soc. 134, 9864–9867 (2012).

25. Purcell, E. M. Life at low Reynolds number. Am. J. Phys. 45, 3–11 (1977).

26. Kamholz, A. E. & Yager, P. Theoretical analysis of molecular diffusion in pressure-driven

laminar flow in microfluidic channels. Biophys. J. 80, 155–160 (2001).

27. Inglis, D. W. Efficient microfluidic particle separation arrays. Appl. Phys. Lett. 94, 013510

(2009).

28. Lee, J. & Kim, C.-J. Surface-tension-driven microactuation based on continuous

electrowetting. J. Microelectromechanical Syst. 9, 171–180 (2000).

29. Safavieh, R. & Juncker, D. Capillarics: pre-programmed, self-powered microfluidic circuits

built from capillary elements. Lab Chip 13, 4180–4189 (2013).

30. Liu, K. et al. Simple fabrication scheme for sub-10 nm electrode gaps using electron-beam

lithography. Appl. Phys. Lett. 80, 865–867 (2002).

31. Vong, T. et al. Site-specific immobilization of DNA in glass microchannels via

photolithography. Langmuir 25, 13952–13958 (2009).

32. Xia, Y. & Whitesides, G. M. Soft lithography. Annu. Rev. Mater. Sci. 28, 153–184 (1998).

33. Bélanger, M.-C. & Marois, Y. Hemocompatibility, biocompatibility, inflammatory and in

vivo studies of primary reference materials low-density polyethylene and

polydimethylsiloxane: a review. J. Biomed. Mater. Res. 58, 467–477 (2001).

34. Bhagat, A. A. S. et al. Microfluidics for cell separation. Med. Biol. Eng. Comput. 48, 999–

1014 (2010).

35. Kang, J. H. et al. A combined micromagnetic-microfluidic device for rapid capture and

culture of rare circulating tumor cells. Lab Chip 12, 2175–2181 (2012).

36. Pratt, E. D., Huang, C., Hawkins, B. G., Gleghorn, J. P. & Kirby, B. J. Rare cell capture in

microfluidic devices. Chem. Eng. Sci. 66, 1508–1522 (2011).

37. El-Ali, J., Sorger, P. K. & Jensen, K. F. Cells on chips. Nature 442, 403–411 (2006).

76

38. Copley, M. R., Beer, P. A. & Eaves, C. J. Hematopoietic stem cell heterogeneity takes center

stage. Cell Stem Cell 10, 690–697 (2012).

39. Raj, A., Rifkin, S. A., Andersen, E. & van Oudenaarden, A. Variability in gene expression

underlies incomplete penetrance. Nature 463, 913–918 (2010).

40. Elowitz, M. B., Levine, A. J., Siggia, E. D. & Swain, P. S. Stochastic gene expression in a

single cell. Science 297, 1183–6 (2002).

41. Navin, N. et al. Tumour evolution inferred by single-cell sequencing. Nature 472, 90–94

(2011).

42. Jordan, N. V. et al. HER2 expression identifies dynamic functional states within circulating

breast cancer cells. Nature 537, 102–106 (2016).

43. Quirke, P. et al. The future of the TNM staging system in colorectal cancer: time for a

debate? Lancet Oncol. 8, 651–657 (2007).

44. Baxter, N. N. & Tuttle, T. M. Inadequacy of lymph node staging in gastric cancer patients:

a population-based study. Ann. Surg. Oncol. 12, 981–987 (2005).

45. Epstein, F. H. & Schnipper, L. E. Clinical implications of tumor-cell heterogeneity. N. Engl.

J. Med. 314, 1423–1431 (1986).

46. Junttila, M. R. & de Sauvage, F. J. Influence of tumour micro-environment heterogeneity

on therapeutic response. Nature 501, 346–354 (2013).

47. Aceto, N. et al. Circulating tumor cell clusters are oligoclonal precursors of breast cancer

metastasis. Cell 158, 1110–1122 (2014).

48. Thiery, J. P. Epithelial–mesenchymal transitions in tumour progression. Nat. Rev. Cancer

2, 442–454 (2002).

49. Heitzer, E. et al. Complex tumor genomes inferred from single circulating tumor cells by

array-CGH and next-generation sequencing. Cancer Res. 73, 2965–2975 (2013).

50. Zong, C., Lu, S., Chapman, A. R. & Xie, X. S. Genome-wide detection of single-nucleotide

and copy-number variations of a single human cell. Science 338, 1622–6 (2012).

51. Cohen, S. J. et al. Relationship of circulating tumor cells to tumor response, progression-

free survival, and overall survival in patients with metastatic colorectal cancer. J. Clin.

Oncol. 26, 3213–3221 (2008).

52. Stopeck, A. et al. Circulating tumor cells, disease progression, and survival in metastatic

breast cancer. 781–791 (2004).

53. Bidard, F.-C. et al. Single circulating tumor cell detection and overall survival in

nonmetastatic breast cancer. Ann. Oncol. 21, 729–733 (2010).

54. Chaffer, C. L. & Weinberg, R. A. A perspective on cancer cell metastasis. Science 331,

1559–64 (2011).

55. Yu, M. et al. Circulating breast tumor cells exhibit dynamic changes in epithelial and

mesenchymal composition. Science 339, 580–584 (2013).

56. Gorges, T. M. et al. Circulating tumour cells escape from epcam-based detection due to

epithelial-to-mesenchymal transition. BMC Cancer 12, 178 (2012).

77

57. Tsukamoto, H. et al. Irradiation-induced epithelial–mesenchymal transition (EMT) related

to invasive potential in endometrial carcinoma cells. Gynecol. Oncol. 107, 500–504 (2007).

58. Nagrath, S. et al. Isolation of rare circulating tumour cells in cancer patients by microchip

technology. Nature 450, 1235–1239 (2007).

59. Nichols, A. C. et al. Detection of circulating tumor cells in advanced head and neck cancer

using the Cellsearch system. Head Neck 34, 1440–1444 (2012).

60. Rao, C. et al. Circulating melanoma cells and survival in metastatic melanoma. Int. J. Oncol.

38, 755–760 (2011).

61. Bidard, F. C. et al. Disseminated and circulating tumor cells in gastrointestinal oncology.

Crit. Rev. Oncol. Hematol. 82, 103–115 (2012).

62. Naoe, M. et al. Detection of circulating urothelial cancer cells in the blood using the

Cellsearch system. Cancer 109, 1439–1445 (2007).

63. Mohamadi, R. M. et al. Nanoparticle-mediated binning and profiling of heterogeneous

circulating tumor cell subpopulations. Angew. Chemie Int. Ed. 54, 139–143 (2015).

64. Besant, J. D. et al. Velocity valleys enable efficient capture and spatial sorting of

nanoparticle-bound cancer cells. Nanoscale 7, 6278–6285 (2015).

65. Cristofanilli, M. et al. Circulating tumor cells in metastatic breast cancer: biologic staging

beyond tumor burden. Clin. Breast Cancer 7, 34–42 (2007).

66. Alix-Panabieres, C. & Pantel, K. Circulating tumor cells: liquid biopsy of cancer. Clin.

Chem. 59, 110–118 (2013).

67. EBCTCG (Early Breast Cancer Trialists’ Collaborative Group) et al. Effect of radiotherapy

after mastectomy and axillary surgery on 10-year recurrence and 20-year breast cancer

mortality: meta-analysis of individual patient data for 8135 women in 22 randomised trials.

Lancet (London, England) 383, 2127–35 (2014).

68. Baum, B., Settleman, J. & Quinlan, M. P. Transitions between epithelial and mesenchymal

states in development and disease. Semin. Cell Dev. Biol. 19, 294–308 (2008).

69. Hiltermann, T. J. N. et al. Circulating tumor cells in small-cell lung cancer: a predictive and

prognostic factor. Ann. Oncol. 23, 2937–2942 (2012).

70. Steinert, G. et al. Immune escape and survival mechanisms in circulating tumor cells of

colorectal cancer. Cancer Res. 74, 1694–1704 (2014).

71. Wang, D. & Bodovitz, S. Single cell analysis: the new frontier in ‘omics’. Trends

Biotechnol. 28, 281–290 (2010).

72. Joosse, S. A. & Pantel, K. Biologic challenges in the detection of circulating tumor cells.

Cancer Res. 73, 8–11 (2013).

73. Valiev, R. Materials science: nanomaterial advantage. Nature 419, 887–889 (2002).

74. Wang, J. Nanomaterial-based electrochemical biosensors. Analyst 130, 421 (2005).

75. Rose, M. et al. Hierarchical micro- and mesoporous carbide-derived carbon as a high-

performance electrode material in supercapacitors. Small 7, 1108–1117 (2011).

76. Gittens, R. A. et al. The roles of titanium surface micro/nanotopography and wettability on

78

the differential response of human osteoblast lineage cells. Acta Biomater. 9, 6268–6277

(2013).

77. Potyrailo, R. & Naik, R. R. Bionanomaterials and bioinspired nanostructures for selective

vapor sensing. Annu. Rev. Mater. Res. 43, 307–334 (2013).

78. Noorduin, W. L., Grinthal, A., Mahadevan, L. & Aizenberg, J. Rationally designed

complex, hierarchical microarchitectures. Science 340, 832–7 (2013).

79. Grinthal, A. et al. Steering nanofibers: an integrative approach to bio-inspired fiber

fabrication and assembly. Nano Today 7, 35–52 (2012).

80. Phillips, K. R. et al. Tunable anisotropy in inverse opals and emerging optical properties.

Chem. Mater. 26, 1622–1628 (2014).

81. Akdim, B., Pachter, R. & Naik, R. R. Self-assembled peptide nanotubes as electronic

materials: an evaluation from first-principles calculations. Appl. Phys. Lett. 106, 183707

(2015).

82. Pacholski, C., Kornowski, A. & Weller, H. Self-assembly of ZnO: from nanodots to

nanorods. Angew. Chemie - Int. Ed. 41, 1188–1191 (2002).

83. Whitesides, G. M. & Grzybowski, B. Self-assembly at all scales. Science 295, 2418–21

(2002).

84. Jasuja, K. & Berry, V. Implantation and growth of dendritic gold nanostructures on

graphene derivatives: electrical property tailoring and raman enhancement. ACS Nano 3,

2358–2366 (2009).

85. Grzelczak, M., Pérez-Juste, J., Mulvaney, P. & Liz-Marzán, L. M. Shape control in gold

nanoparticle synthesis. Chem. Soc. Rev. 37, 1783–1791 (2008).

86. Woo, S. M., Gabardo, C. M. & Soleymani, L. Prototyping of wrinkled nano-

/microstructured electrodes for electrochemical DNA detection. Anal. Chem. 86, 12341–

12347 (2014).

87. Hashmi, A. S. K. & Hutchings, G. J. Gold catalysis. Angew. Chemie Int. Ed. 45, 7896–7936

(2006).

88. Liu, M. et al. Enhanced electrocatalytic CO2 reduction via field-induced reagent

concentration. Nature 537, 382–386 (2016).

89. Campbell, C. T. The active site in nanoparticle gold catalysis. Science 306, 234–235 (2004).

90. Ming, T. et al. Growth of tetrahexahedral gold nanocrystals with high-index facets. J. Am.

Chem. Soc. 131, 16350–16351 (2009).

91. Golan, Y., Margulis, L. & Rubinstein, I. Vacuum-deposited gold films. Surf. Sci. 264, 312–

326 (1992).

92. Lin, Y., Zou, Y., Mo, Y., Guo, J. & Lindquist, R. G. E-beam patterned gold nanodot arrays

on optical fiber tips for localized surface plasmon resonance biochemical sensing. Sensors

10, 9397–9406 (2010).

93. Zhao, P., Li, N. & Astruc, D. State of the art in gold nanoparticle synthesis. Coord. Chem.

Rev. 257, 638–665 (2013).

94. Cherevko, S. & Chung, C.-H. Direct electrodeposition of nanoporous gold with controlled

79

multimodal pore size distribution. Electrochem. commun. 13, 16–19 (2011).

95. Soleymani, L., Fang, Z., Sargent, E. H. & Kelley, S. O. Programming the detection limits

of biosensors through controlled nanostructuring. Nat. Nanotechnol. 4, 844–848 (2009).

96. Soleymani, L. et al. Hierarchical nanotextured microelectrodes overcome the molecular

transport barrier to achieve rapid, direct bacterial detection. ACS Nano 5, 3360–3366

(2011).

97. Soleymani, L. et al. Nanostructuring of patterned microelectrodes to enhance the sensitivity

of electrochemical nucleic acids detection. Angew. Chemie Int. Ed. 48, 8457–8460 (2009).

98. Bin, X., Sargent, E. H. & Kelley, S. O. Nanostructuring of sensors determines the efficiency

of biomolecular capture. Anal. Chem. 82, 5928–5931 (2010).

99. Fang, Z. et al. Direct profiling of cancer biomarkers in tumor tissue using a multiplexed

nanostructured microelectrode integrated circuit. ACS Nano 3, 3207–3213 (2009).

100. Sage, A. T. et al. Fractal circuit sensors enable rapid quantification of biomarkers for donor

lung assessment for transplantation. Sci. Adv. 1, e1500417 (2015).

101. Scharifker, B. & Hills, G. Theoretical and experimental studies of multiple nucleation.

Electrochem. Acta 28, 879–889 (1983).

102. Scharifker, B. R. On the theory of the potentiostatic current transient for diffusion-

controlled three-dimensional electrocrystallization processes. J. Electrochem. Soc. 146,

1005–1012 (1999).

103. Hyde, M. E. & Compton, R. G. A review of the analysis of multiple nucleation with

diffusion controlled growth. J. Electroanal. Chem. 549, 1–12 (2003).

104. Johnson, C. J., Dujardin, E., Davis, S. A., Murphy, C. J. & Mann, S. Growth and form of

gold nanorods prepared by seed-mediated, surfactant-directed synthesis. J. Mater. Chem.

12, 1765–1770 (2002).

105. Ghahremaninezhad, A. & Dolati, A. Diffusion-controlled growth model for

electrodeposited cobalt nanowires in highly ordered aluminum oxide membrane. in ECS

Transactions 28, 13–25 (The Electrochemical Society, 2010).

106. Xia, Y., Xiong, Y., Lim, B. & Skrabalak, S. E. Shape-controlled synthesis of metal

nanocrystals: simple chemistry meets complex physics? Angew. Chemie Int. Ed. 48, 60–

103 (2009).

107. Das, J. & Kelley, S. O. Protein detection using arrayed microsensor chips: tuning sensor

footprint to achieve ultrasensitive readout of CA-125 in serum and whole blood. Anal.

Chem. 83, 1167–1172 (2011).

108. Mahshid, S. S., Camiré, S., Ricci, F. & Vallée-Bélisle, A. A highly selective electrochemical

DNA-based sensor that employs steric hindrance effects to detect proteins directly in whole

blood. J. Am. Chem. Soc. 137, 15596–15599 (2015).

109. Peterson, A. W. The effect of surface probe density on DNA hybridization. Nucleic Acids

Res. 29, 5163–5168 (2001).

110. Nakano, S., Miyoshi, D. & Sugimoto, N. Effects of molecular crowding on the structures,

interactions, and functions of nucleic acids. Chem. Rev. 114, 2733–2758 (2014).

80

111. Yager, P. et al. Microfluidic diagnostic technologies for global public health. Nature 442,

412–418 (2006).

112. Kelley, S. O. et al. Advancing the speed, sensitivity and accuracy of biomolecular detection

using multi-length-scale engineering. Nat. Nanotechnol. 9, 969–980 (2014).

113. Lutz, B. et al. Dissolvable fluidic time delays for programming multi-step assays in

instrument-free paper diagnostics. Lab Chip 13, 2840–2847 (2013).

114. Renault, C., Koehne, J., Ricco, A. J. & Crooks, R. M. Three-dimensional wax patterning of

paper fluidic devices. Langmuir 30, 7030–7036 (2014).

115. Liu, H. & Crooks, R. M. Three-dimensional paper microfluidic devices assembled using the

principles of origami. J. Am. Chem. Soc. 133, 17564–17566 (2011).

116. Laksanasopin, T. et al. A smartphone dongle for diagnosis of infectious diseases at the point

of care. Sci. Transl. Med. 7, 273re1 (2015).

117. Dimov, I. K. et al. Stand-alone self-powered integrated microfluidic blood analysis system

(simbas). Lab Chip 11, 845–850 (2011).

118. Gong, M. M., MacDonald, B. D., Vu Nguyen, T. & Sinton, D. Hand-powered microfluidics:

a membrane pump with a patient-to-chip syringe interface. Biomicrofluidics 6, 044102

(2012).

119. Begolo, S., Zhukov, D. V, Selck, D. a, Li, L. & Ismagilov, R. F. The pumping lid:

investigating multi-material 3d printing for equipment-free, programmable generation of

positive and negative pressures for microfluidic applications. Lab Chip 14, 4616–4628

(2014).

120. Safavieh, R., Zhou, G. Z. & Juncker, D. Microfluidics made of yarns and knots: from

fundamental properties to simple networks and operations. Lab Chip 11, 2618–2624 (2011).

121. Safavieh, R., Tamayol, A. & Juncker, D. Serpentine and leading-edge capillary pumps for

microfluidic capillary systems. Microfluid. Nanofluidics 18, 357–366 (2015).

122. Yetisen, A. K., Akram, M. S. & Lowe, C. R. Paper-based microfluidic point-of-care

diagnostic devices. Lab Chip 13, 2210–2251 (2013).

123. Thom, N. K., Lewis, G. G., Yeung, K. & Phillips, S. T. Quantitative fluorescence assays

using a self-powered paper-based microfluidic device and a camera-equipped cellular

phone. RSC Adv. 4, 1334–1340 (2014).

124. Funes-Huacca, M. et al. Portable self-contained cultures for phage and bacteria made of

paper and tape. Lab Chip 12, 4269–4278 (2012).

125. Deiss, F., Funes-Huacca, M. E., Bal, J., Tjhung, K. F. & Derda, R. Antimicrobial

susceptibility assays in paper-based portable culture devices. Lab Chip 14, 167–171 (2014).

126. Luo, L., Li, X. & Crooks, R. M. Low-voltage origami-paper-based electrophoretic device

for rapid protein separation. Anal. Chem. 86, 12390–12397 (2014).

127. Guo, M. T., Rotem, A., Heyman, J. A. & Weitz, D. A. Droplet microfluidics for high-

throughput biological assays. Lab Chip 12, 2146–2155 (2012).

128. Rosenfeld, L., Lin, T., Derda, R. & Tang, S. K. Y. Review and analysis of performance

metrics of droplet microfluidics systems. Microfluid. Nanofluidics 16, 921–939 (2014).

81

129. Kim, M. et al. Optofluidic ultrahigh-throughput detection of fluorescent drops. Lab Chip

15, 1417–1423 (2015).

130. Ng, A. H. C., Chamberlain, M. D., Situ, H., Lee, V. & Wheeler, A. R. Digital microfluidic

immunocytochemistry in single cells. Nat. Commun. 6, 7513 (2015).

131. Eydelnant, I. A., Betty Li, B. & Wheeler, A. R. Microgels on-demand. Nat. Commun. 5,

3355 (2014).

132. Witters, D., Knez, K., Ceyssens, F., Puers, R. & Lammertyn, J. Digital microfluidics-

enabled single-molecule detection by printing and sealing single magnetic beads in

femtoliter droplets. Lab Chip 13, 2047–2054 (2013).

133. Puntambekar, A., Choi, J.-W., Ahn, C. H., Kim, S. & Makhijani, V. Fixed-volume metering

microdispenser module. Lab Chip 2, 213–218 (2002).

134. Hitzbleck, M. et al. Capillary soft valves for microfluidics. Lab Chip 12, 1972–1978 (2012).

135. Iwai, K. et al. Finger-powered microfluidic systems using multilayer soft lithography and

injection molding processes. Lab Chip 14, 3790–3799 (2014).

136. Shen, F. et al. Multiplexed quantification of nucleic acids with large dynamic range using

multivolume digital RT-PCR on a rotational slipchip tested with hiv and hepatitis C viral

load. J. Am. Chem. Soc. 133, 17705–17712 (2011).

137. Shen, F. et al. Nanoliter multiplex pcr arrays on a slipchip. Anal. Chem. 82, 4606–4612

(2010).

138. Pompano, R. R., Platt, C. E., Karymov, M. A. & Ismagilov, R. F. Control of initiation, rate,

and routing of spontaneous capillary-driven flow of liquid droplets through microfluidic

channels on slipchip. Langmuir 28, 1931–1941 (2012).

139. Liu, H., Li, X. & Crooks, R. M. Paper-based slippad for high-throughput chemical sensing.

Anal. Chem. 85, 4263–4267 (2013).

140. Li, W. et al. Squeeze-chip: a finger-controlled microfluidic flow network device and its

application to biochemical assays. Lab Chip 12, 1587–1590 (2012).

141. Xu, K., Begley, M. R. & Landers, J. P. Simultaneous metering and dispensing of multiple

reagents on a passively controlled microdevice solely by finger pressing. Lab Chip 15, 867–

876 (2015).

142. Leng, J. & Salmon, J.-B. Microfluidic crystallization. Lab Chip 9, 24–34 (2009).

143. Li Jeon, N. et al. Neutrophil chemotaxis in linear and complex gradients of interleukin-8

formed in a microfabricated device. Nat. Biotechnol. 20, 826 (2002).

144. Cho, H., Kim, H.-Y., Kang, J. Y. & Kim, T. S. How the capillary burst microvalve works.

J. Colloid Interface Sci. 306, 379–385 (2007).

145. Elizabeth Hulme, S., Shevkoplyas, S. S. & Whitesides, G. M. Incorporation of prefabricated

screw, pneumatic, and solenoid valves into microfluidic devices. Lab Chip 9, 79–86 (2009).

146. Cira, N. J., Ho, J. Y., Dueck, M. E. & Weibel, D. B. A self-loading microfluidic device for

determining the minimum inhibitory concentration of antibiotics. Lab Chip 12, 1052–1059

(2012).

147. Greenwood, D. & O’Grady, F. A comparison of the effects of ampicillin on escherichia coli

82

and proteus mirabilis. J. Med. Microbiol. 2, 435–441 (1969).

148. Kitagawa, M. et al. Complete set of ORF clones of escherichia coli ASKA library (a

complete set of e. coli k-12 ORF archive): unique resources for biological research. DNA

Res. 12, 291–299 (2006).

149. Chin, C. D., Linder, V. & Sia, S. K. Commercialization of microfluidic point-of-care

diagnostic devices. Lab Chip 12, 2118–2134 (2012).

150. Sung, Y.-L., Jeang, J., Lee, C.-H. & Shih, W.-C. Fabricating optical lenses by inkjet printing

and heat-assisted in situ curing of polydimethylsiloxane for smartphone microscopy. 20,

47005–47006 (2015).

151. Papalexi, E. & Satija, R. Single-cell RNA sequencing to explore immune cell heterogeneity.

Nat. Rev. Immunol. 18, 35–45 (2017).

152. Narsinh, K. H. et al. Single cell transcriptional profiling reveals heterogeneity of human

induced pluripotent stem cells. J. Clin. Invest. 121, 1217–1221 (2011).

153. Chattopadhyay, P. K., Gierahn, T. M., Roederer, M. & Love, J. C. Single-cell technologies

for monitoring immune systems. Nat. Immunol. 15, 128–135 (2014).

154. Roeder, I., Loeffler, M. & Glauche, I. Towards a quantitative understanding of stem cell–

niche interaction: experiments, models, and technologies. Blood Cells, Mol. Dis. 46, 308–

317 (2011).

155. Jordan, N. V. et al. HER2 expression identifies dynamic functional states within circulating

breast cancer cells. Nature 537, 102–106 (2016).

156. Brouzes, E. et al. Droplet microfluidic technology for single-cell high-throughput

screening. Proc. Natl. Acad. Sci. 106, 14195–14200 (2009).

157. Lagus, T. P. & Edd, J. F. A review of the theory, methods and recent applications of high-

throughput single-cell droplet microfluidics. J. Phys. D. Appl. Phys. 46, 114005 (2013).

158. Chung, J., Kim, Y.-J. & Yoon, E. Highly-efficient single-cell capture in microfluidic array

chips using differential hydrodynamic guiding structures. Appl. Phys. Lett. 98, 123701

(2011).

159. Lin, C.-H. et al. A microfluidic dual-well device for high-throughput single-cell capture and

culture. Lab Chip 15, 2928–2938 (2015).

160. Miyanishi, K. et al. Effects of hydrostatic pressure and transforming growth factor- β 3 on

adult human mesenchymal stem cell chondrogenesis in vitro. Tissue Eng. 12, 1419–1428

(2006).

161. Pertoft, H. & Laurent, T. C. Isopycnic separation of cells and cell organelles by

centrifugation in modified colloidal silica gradients. in Methods of Cell Separation 25–65

(Springer US, 1977). doi:10.1007/978-1-4684-0820-1_2

162. Yamada, M. & Seki, M. Hydrodynamic filtration for on-chip particle concentration and

classification utilizing microfluidics. Lab Chip 5, 1233–1239 (2005).

163. Yamada, M. & Seki, M. Microfluidic particle sorter employing flow splitting and

recombining. Anal. Chem. 78, 1357–1362 (2006).

83

Single Cell Capture Device

A.1 Background

It was historically the practice in cell biology to treat all of the cells from a particular tissue or

culture as being essentially identical. Part of this paradigm was rooted in the poor sensitivity of

early techniques, which often required that large quantities of cells be pooled and lysed in order to

detect biomolecules of interest. In this approach, the differences in genotype or phenotype within

these populations were ignored or treated as noise obscuring the desired information. However,

more recent investigations have shed light on the fundamental importance of this heterogeneity in

diverse biological processes.151–153 These variations stem from a number of different sources. One

of the most fundamental is small changes in the genotypes of cells due to unique mutations.41 Other

differences are a consequence of the history of cell, including the particular niche it inhabits and

the stimuli it has been exposed to.154 More transitory discrepancies arise as a consequence of the

fundamentally stochastic nature of molecular interactions, especially in the case of regulatory

proteins whose binding prompt a cascade of effects.40 In general, these differences are of greatest

import in scenarios where a small number of cells can have a disproportionately large impact. This

is the case during embryonic development, stem cell development and tumor biology.39,155

The desire to more fully characterize this cellular heterogeneity has motivated the development of

assays which allow for the analysis of protein, mRNA, DNA, and various metabolites at a single

cell level. Consequently, there exists a need for devices that allow for the precise manipulation of

single cells: specifically, the ability to isolated, treat, incubate and release cells in an individually

addressable manner. These devices can roughly be divided into single-phase and multi-phase

types. Multi-phase devices employ droplets of one medium immersed in a second medium.156,157

By trapping individual cells in these droplets, cells can then be manipulated using the full suite of

droplet fluidic techniques. Single-phase devices don’t isolate cells in different media but instead

84

use physical obstacle or force fields to keep cells apart.158,159 An ideal device would be as simple

as possible while still supporting all of the desired functions; this suggest the use of single phase

device which employs primarily passive elements.

A.2 Device Design

The first stage of any cell manipulation device must accept a stream of cells and individually

position them into isolated traps. There are two methods by which this can be done: the stochastic

“shotgun” approach and the deterministic approach. In the stochastic approach, a large number of

traps are used and the design depends on the probability that a cell will encounter a trap at some

point in its path through the device. In this case, which trap the cell encounters is largely random.

By contrast, the deterministic approach involves a series of traps such that the first cell to enter the

device is directed towards the first trap, the next towards the subsequent trap, and so on. Each

method has its advantages. As compared with the deterministic approach, shotgun approach is

typically capable of considerably higher throughput and needs less stringent design. However, this

approach may allow cells to escape and does not preserve the identity of the cell. The identity of

the cell is important if the incoming cells are being delivered from an upstream measurement, e.g.

magnetic or fluorescent sorting; maintaining cell identity will allow the measured properties to

remain associated with the correct cell. Moreover, such designs typically require more traps than

the number of cells, which increases the difficulty of addressing particular captive cells. Both of

these issues are accentuated in situations where the number of cells being analyzed is small, such

as in rare cell applications.38,39

To preserve the ability of the device to work with diverse cell populations and circumvent the need

for cell tagging/external force fields, the decision was made to capture cells purely on the basis of

physical obstruction. The structure which allows for this physical capture is the microfluidic weir

(Figure A-1).

85

Figure A-1 Structure of a microfluidic weir. The channel abruptly narrows to a diameter too

small for cell passage.

A weir is a simply an abrupt narrowing of the microfluidic channel that does not allow the passage

of particles above a certain diameter. Once a cell encounters the weir it is lodged in place, held

trapped by the pressure difference between the regions upstream and downstream of the

obstruction. It is important that the pressure the cell is exposed to be kept to a minimum both in

magnitude and duration, since exposure to high pressures can modify cell biology or damage cell

integrity.160 In order to capture multiple cells, a series of weir structures is required. However, the

structures cannot simply be chained together, since the filling of the first weir arrests flow and

subsequent weirs would remain unfilled. Instead, a shunt structure is included in parallel to each

weir, which allows later cells to circumvent filled traps (Figure A-2).

86

Figure A-2 A microfluidic weir with a parallel shunt channel. The lower channel includes a

weir structure, whereas the upper shunt is a simple wide channel.

The inclusion of a shunt, however, introduces a further complication. Even when approaching an

empty weir, there is a possibility that the cell will enter the shunt channel rather than the weir

channel. The path that a cell follows is determined by the streamline that the cell occupies, as well

as the relative resistances of the two channels. Specifically, the fraction of the flow that enters each

fork is inversely proportional to the resistance of that fork (Figure A-3).

87

Figure A-3 Equivalent circuit diagram for a sequence of weir-shunt pairs. At each fork the

cell will be guided to either the weir or the shunt depending on flow resistance. The lower the

relative resistance of each channel, the greater portion of the flow that enters it.

If the centre of mass of a cell occupies one of the streamlines that enters a given channel, the cell

will do so as well. Thus, to maximize the chance that a cell enters an empty weir instead of

diverting around it, the resistance of the weir relative to the shunt must be minimized. Since the

weir must be small enough to occlude the cell, it must be smaller than the size of the cell to be

captured (on the order of 10 μm). Thus, the resistance of the weir is bounded by a minimum value

and is not freely adjustable. Consequently, the resistance of the shunt must be increased in order

to encourage cell flow into the weir (Figure A-4).

Figure A-4 Encouraging flow to enter the weir. A distribution of cells across the channel is

shown, with each cell in a different colour. The resistance of the shunt must be much greater than

RWEIR

RSHUNT

RWEIR

RSHUNT

RWEIR

RSHUNT

RWEIR

RSHUNT

RW

RS >> RW

88

that of weir to direct the majority of the cells into the capture area. RS is shunt resistance, RW is

weir resistance.

Specifically, the portion of flow that enters the weir channel, QW, is:

𝑄𝑊 =

𝑅𝑠

𝑅𝑠 + 𝑅𝑊 , (A.1)

Where RS is the shunt resistance and RW is the weir resistance. A shunt resistance 10-fold that of

the weir ensures that approximately 90% of fluid approaching an empty weir will enter it rather

than diverting around it. Assuming that cells are roughly evenly distributed within the flow, a

matching 90% will enter the trap. Increasing the resistance of the shunt is not without consequence,

however. Once a weir is occupied, the pressure difference that a cell is exposed to is the product

of the volumetric flow and the shunt resistance. In order to keep the pressure on the cell low, the

volumetric flow would thus need be kept minimal, which severely limits device throughput. The

only way to solve this problem is to adjust the cell distribution across the streamlines such that it

is no longer random. Specifically, by moving the cells as close to the wall proximal the weir as

possible, a large fraction of the cells can enter the weir despite only a small fraction of flow doing

so (Figure A-5).

Figure A-5 Reducing shunt resistance by translating cells across streamlines. By driving the

cells to one lateral edge, only a small amount of flow is required to enter the weir, allowing a

RW

RS << RW

89

dramatic reduction in shunt resistance. Shunt resistance is now much smaller than weir

resistance, instead of the converse.

This allows for greatly improved performance, and is limited only by the size of the cell.

Specifically, the cell cannot lie on any streamline which is closer to the wall than the cell radius.

By integrating all of the flow which lies between the wall and this streamline of nearest approach,

the fraction of flow which must enter the weir to ensure complete capture (assuming cells have

been fully aligned to the wall) can be determined. This quantity is further reduced by the parabolic

nature of flow in a channel. Since flow is slowest near the wall, the fraction of flow which lies

between the wall and a distance x from the wall is even less than the ratio of x to the entire width

of the channel. Employing a channel width of 200 μm and a cell diameter of 15 μm, this means

that only 1-2% of the flow is required to enter the weir to ensure complete capture. Consequently,

the resistance of the shunt can be decreased to less than 1/10 the resistance of the weir, a more than

two order of magnitude reduction in the pressure that captured cells experience. This modification

is crucial for a sequential series of traps to capture cells in a deterministic fashion while also

retaining high throughput.

Moving cells across streamlines is especially difficult in the microfluidic regime, where fluid

lamina do not mix and tend to maintain their relative positioning. Nonetheless, it is possible, and

can be performed either by an externally generated force field or by passive structures within the

device. Candidate external forces include electrophoretic, dielectrophoretic, magnetic, and

gravitational forces. Of these, gravity is the simplest, since it is omnipresent and does not need to

be generated manually. The displacement of cells due to gravity is known as cell settling. Cell

settling is a consequence of the mismatch in densities between a cell and a typical aqueous solution.

In general, a cell is a few percent more dense than water, due to the various biomolecules contained

within.161 Consequentially, the gravitational force on a cell is slightly greater than the buoyancy

force, and cells will tend to settle downwards. The velocity with which a cell settles is determined

by the interaction between this net gravitational force and the drag force, and is calculated here:

𝑣𝑠𝑒𝑡𝑡𝑙𝑖𝑛𝑔 =

𝑔𝑑2(𝜌𝑐 − 𝜌)

18𝜇, (A.2)

where g is the acceleration due to gravity, ρC is the density of the cell, ρ is the density of the fluid,

and μ is the dynamic viscosity of the fluid. For a typical cell, this velocity is on the order of 1-10

90

μm/s. Thus, for a cell to traverse a 200 μm wide channel, it must remain in this channel for a time

on the order of 1 minute. The time that a cell spends in this alignment channel is a function of three

parameters: the depth of the channel, the length of the channel, and the volumetric flow rate. More

specifically, the residence time is proportional to both the length and depth of the alignment

channel, and inversely proportional to the flow rate. In order to maintain as high a volumetric flow

rate as possible to enable good throughput, the depth and length of the channel should be

maximized. The height of the channel is constrained to some degree by the height and aspect ratio

of the SU8, with 200 μm a suitable dimension. In order to maintain a relatively rapid 10 μL/min

flowrate with a height and depth of 200 μm and a residence time of one minute, the length of the

channel needs to be 25 cm! This is prohibitively long for a single, straight channel. By instead

using a serpentine channel, the effective length of the alignment channel can be increased without

a significant change in device footprint. This produces an additional issue; the settling in each

return channel is equal and opposite that in each forward channel, and as a consequence there is

no net cell displacement. This can be fixed by reducing the depth of the return channels. By

reducing the height of these channels to 50 μm, the time spent therein is one quarter the time spent

in the forward channels, and the net cell displacement remains in the downward direction. Using

8 of these switchback channels, each 50 mm in length, was shown to be sufficient to move all cells

to the intended wall, and allow for capture using the sequential weir traps (Figure A-6).

Figure A-6 Design of the gravity driven cell alignment and capture device, which is

operated in the vertical orientation. The green channel has a depth of 200 μm and the red

channel has a depth of 50 μm to allow for net displacement due to gravity. Aligned cells then

encounter a series of 8 weir-shunt pairs.

This approach allows for very simple and robust fabrication of the alignment portion of the device.

However, it does present a number of drawbacks. Firstly, the device must be oriented in the vertical

direction during cell capture, which complicates imaging. Secondly, and more importantly, the

91

cells spend a large amount of time near the wall. This is especially true of cells which happened to

enter the device proximal to this wall. This can cause cells to slow down significantly and requires

that the wall be pretreated with surfactants to discourage sticking.

An alternative approach that uses passive fluidic design in place of gravity was implemented to

mitigate these issues. In general, channel design in the laminar regime is incapable of diverting

cells due to the tendency of streamlines to retain their relative position. Although this is true, it

omits an important fact which can be exploited to great effect; namely, the finite size of cells. Since

cells have a certain radius (on the order of 5 – 10 μm), they cannot approach any wall more closely

than this limiting distance. This is not true of the fluid flow as a whole, which consists of essentially

infinitesimal water molecules. One design motif that takes advantage of this is the “bleeding flow”

element (Figure A-7).162

Figure A-7 Function of the bleeding flow motif. (Top) The fluid located between the dashed

line and the wall exits through the side channel, with the remaining fluid remaining in the main

channel and migrating outwardly to replace the removed volume. As long as the centre of a

particle lies medial to the dashed line, it will remain in the main channel. (Bottom) Resistances

need to be determined such that the shaded portion of flow exits at each branch point. Adapted

with permission from 162. Copyright 2005 The Royal Society of Chemistry.

92

This consists of a large primary channel through which particles are flowing, and from which small

quantities of fluid are siphoned off by a series of smaller “bleeder” channels. By judicious selection

of the resistance of these channels, small volumes of fluid are diverted at each junction. As long

as these quantities are small enough, only streamlines which lie between the wall and a distance

of one cell radius will be diverted. This means that cells which are not yet touching the wall are

moved closer to it, whereas cell which are already touching the wall simply remain touching it.

With enough of these channels, most of the cells in a channel will eventually come in contact with

the wall. The asymptotic nature of this approach means, however, that cells which begin at the far

side of the channel will tend not to migrate the entire distance. This can be rectified by using two

of these motifs in series, with a slight modification (Figure A-8).163

Figure A-8 Two stage cell aligner. The device consists of two sequential elements each

employing the “bleeding flow” motif. Each coloured region is expanded in a separate figure.

93

Figure A-9 First “bleeding flow” element. Each side channel draws a small amount of fluid,

causing the cell to migrate closer to the wall.

Figure A-10 End of the first “bleeding flow” element. The cell has approached the right wall,

so it remains near the bottom of the wider channel.

94

Figure A-11 Reintroduction of removed fluid. The fluid siphoned off in the first element is

returned. As long as at least half of the total flow was siphoned, this creates a buffer (shown by

blue dotted line) which ensures the cell will be leftward of the centre line when entering the

second “bleeding flow” element.

95

Figure A-12 Second “bleeding flow” element. Since the cell begins on the left half of the

channel, it quickly migrates the remaining distance to the left wall.

In this case, the flow siphoned off during the first element is reintroduced between the first and

second elements. As long as at least half of the flow had been siphoned, this reintroduced fluid

pushes all cells to one side of the channel centerline.

From this point, a second motif can be used to sequester cells along one wall. The design of such

a device is made possible by modelling the channels as resistors and simulating flow in analogy

with an electric circuit. The device was designed using a custom Python script which determined

the requisite dimensions of the channels while keeping in accordance with manually defined

parameters for SU8 tolerances (max/min height, max/min height, max aspect ratio). Testing of this

device demonstrated that cells are regularly moved to the intended wall and remain there. Upon

interfacing of this alignment module with the shunt and weir capture module (Figure A-13), it was

determined that cells are indeed moved to the intended streamlines and enter the weirs rather than

bypassing them via the shunts (Figure A-14).

96

Figure A-13 The cell alignment module interfaced with the weir and shunt capture module.

Figure A-14 Single cell captured in weir trap.

The device as outlined is capable of accepting a stream of cells and directing them, in single file,

to a series of traps. Additional features are required in order to allow for targeted introduction of

chemicals, isolated incubation, and addressable release. These features all require the introduction

of non-passive elements into the design. To this end, membrane deflection valves were

incorporated (Figure A-15, Figure A-16)

97

Figure A-15 Photograph of integrated devce including membrane deflection valves. Scale

bar is 1 cm.

Figure A-16 Complete design of integrated devce. The lower region performs the cell

alignment feature while the upper region performs cell capture and addressing. White channels

are 25 μm tall with rectangular cross section for flow, green channels are 25 μm tall with

rounded cross section for valves, and blue channels are control lines for opening and closing

valves.

98

These valves add a second channel which lies parallel to, and vertically offset from, the flow

channels, a so-called “control” channel. The two channels are separated by a thin PDMS

membrane. Upon application of pressure to the control channel this membrane deflects vertically

and blocks passage through the flow channel. A binary-tree channel, hereafter called the

addressing channel, was introduced to allow individual access to specified traps. This addressing

channel is controlled by 6 valves that are arranged to allow binary addressing. Specifically, these

channels comprise a multiplexer which accepts 6 inputs and outputs one of the eight traps. These

inputs are not independent; rather, they are arranged in complementary pairs (A and ~A, B and ~B

and C and ~C), such that the pressure state of any valve should be opposite its complement. This

means that the states of A, B and C uniquely determine which trap is addressed, as shown in Table

A-1.

Table A-1 Truth table for the multiplexer.

A B C Selected Trap #

0 0 0 1

0 0 1 2

0 1 0 3

0 1 1 4

1 0 0 5

1 0 1 6

1 1 0 7

1 1 1 8

These control lines were actuated using a pressure manifold outfitted with three separate 4-way

valves. These valves each supplied a pair of complementary lines, and ensured that the pair

remained in opposite pressurization (Figure A-17). These lines pressurize water reservoirs which

in turn actuate the on-chip valves (Figure A-18).

99

Figure A-17 Pressure manifold with three 4-way valves (left) and two 3-way valves (right).

The 4-way valves control the addressing channel, the 3-way valves control the device mode.

Figure A-18 Pressure manifold feeding pressurized water reservoirs.

100

Two additional valve lines were required to allow for different device modes (capture, treatment,

incubation) to be accessed. One valve blocks vertical flow and must be in the closed position

during capture and incubation, and in the open position during addressable release. The other valve

blocks horizontal flow and must be closed during addressable release and incubation, and open

during capture. These valves increase the total number of valves to 8, and are also actuated by the

external pressure manifold (Figure A-19).

Figure A-19 Cell capture region of the devce including all 8 valves.. The addressable channel

at the bottom branches in a binary-tree manner into each of the eight weir capture sites above.

The lower six blue channels form the multiplexer, with the two upper blue channels controlling

the mode of operation.

The possible device modes are illustrated below, along with the states of the valves required to

enable each.

101

Figure A-20 A single trap in capture mode. Vertical flow is blocked (blue superposed on

green) whereas horizontal into the weir is unblocked (Green superposed on blue). Cells enter the

weir channel and are captured.

102

Figure A-21 A single trap in incubation mode. Flow in both the vertical and horizontal

directions is blocked (blue superposed on green). The cell is contained in a small, isolated

region.

103

Figure A-22 A single trap in release mode. Horizontal flow is blocked (blue superposed on

green) but vertical flow is unblocked. By addressing the multiplexer to a single trap, the cell

contained therein is selectively released.

Together, the alignment module, weir and shunt pairs, and membrane deflection valves comprise

a device which is able to deterministically allocate an incoming series of cells into sequential traps,

isolate them spatially for incubation, and release a selected cell from the device in an addressable

manner. By placing this device upstream of a single-cell analysis module, such as a device for

single cell sequencing, rare cells can be triaged for downstream analysis and released individually

for measurement.

104

Supporting Information

Supporting Information for Chapter 2

Characterization of magnetic nanoparticles: In order to determine if the nanoparticle were suitable

to this application a characterization of their size and morphology was conducted. This was

performed using two different techniques. First, the hydrodynamic particle diameter was measured

using dynamic light scattering (DLS). This revealed the average hydrodynamic radius to be 70-80

nm (Figure B-1, Top) This is considerably higher than the nominal value of 50 nm provided by

the manufacturer. However, since larger particles produce larger signals in DLS, this technique

often overestimates particle size. As such, the discrepancy may be quite minor. This small diameter

should facilitate close packing of particles and allow the number of beads to closely reflect the

surface expression of EpCAM. In order to further characterize the beads, SEM images were taken.

There images show the beads to have a consistent round morphology, which should again engender

close packing (Figure B-1, Bottom)

105

Figure B-1 Nanoparticle characterization. (Top) DLS measurement of nanoparticle size

distribution. (Bottom) TEM image showing nanoparticle size and morphology.

In order to confirm that SK-BR-3 and MDA-MB-231 both expressed EpCAM and that there was

a large difference in expression between the two, flow cytometry was performed. An anti- EpCAM

antibody conjugated directly to Alexa Fluor 488 was used (Figure B-2). Notably, both cells stain

positive for EpCAM. The majority of MDA-MB-231 cells showed an intensity below 103 whereas

the majority of SKBR3 cells appear between 103 and 104. However, there are still a considerable

number of SKBR3 cells below 103, which explains why some SK-BR-3 cells tend to be captured

in the later zones. In addition, In order to confirm that VX2 cells had sufficient EpCAM expression

for capture using the device, flow cytometry was performed. Cells were stained with APC anti-

EpCAM antibodies (Figure B-3). As shown below, VX2 cells showed a considerable expression

of EpCAM.

106

Figure B-2 Cell line flow cytometry.

Figure B-3 VX2 EpCAM flow cytometry.

107

Supporting Information for Chapter 4

Figure B-4 The critical advancing contact angle of PBS on PDMS as a function of oxygen

plasma exposure time. Error bars represent standard error.

Figure B-5 Accuracy of filling as a function of chamber size.

108

Figure B-6 A replication of Figure 4-2 F-H using buttom pressess rather than using a

hydraulic pump, showing the same ability to reliably meter discrete quantities of liquid.