86
Molecular and Nanoscale Strategies for Enhanced Biosensing by Wendi Zhou A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of Electrical and Computer Engineering University of Toronto © Copyright by Wendi Zhou 2018

Molecular and Nanoscale Strategies for Enhanced Biosensing · 2018-11-19 · Molecular and Nanoscale Strategies for Enhanced Biosensing Wendi Zhou Doctor of Philosophy ... readout

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

  • Molecular and Nanoscale Strategies for Enhanced Biosensing

    by

    Wendi Zhou

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

    Department of Electrical and Computer Engineering University of Toronto

    © Copyright by Wendi Zhou 2018

  • ii

    Molecular and Nanoscale Strategies for Enhanced Biosensing

    Wendi Zhou

    Doctor of Philosophy

    Department of Electrical and Computer Engineering

    University of Toronto

    2018

    Abstract

    As personalized medicine advances, the need continues to grow for new sensors in applications

    ranging from clinical diagnostics to therapeutic development. Biosensors are powerful tools and

    their evolution will play an important role in healthcare by enabling the detection of specific

    analytes that are clinically significant.

    The central aim of this thesis is to formulate new sensing strategies to detect clinically relevant

    biomolecules in a sensitive and specific manner. Many target analytes are present at extremely

    low concentrations while remaining relevant. The utility of a biosensor is limited by its

    sensitivity; but also, as we investigate herein, its ability to maintain its function in a

    heterogeneous environment. Here, we develop a novel sensing method for low concentration

    analysis of soluble signaling factors by coupling DNA hybridization engineering and antibody-

    capturing chemistry with an electrochemical-based reporting method. This technique – when

    combined with a chip-based approach – allows for both sensitive and rapid detection.

    Through the use of nanostructured electrodes, we showcase our technologies in applications that

    include protein quantification and low-level light detection. We design a device that analyzes

    multiple small signaling proteins in stem cell cultures with 10 pg·mL-1 sensitivity and validate

    our results by comparing them to the gold standard enzyme-linked immunosorbent assay, with

  • iii

    more suitable integration capability and in-line monitoring along with smaller sample sizes and

    decreased process times. This platform is improved by further developing a reusable,

    multiplexed one-step assay and sensor, simplifying and speeding up the steps required to obtain

    readout. Finally, we report on nanogap-separated fractal electrodes as a method for the formation

    of nanoscale features, exhibited by the fabrication of quantum dot photodetectors with nearly a

    one hundredfold improvement in performance over that of conventional devices.

    Overall, these strategies present important steps toward the development of more sensitive and

    faster sensors. By demonstrating low concentration analysis in complex environments and

    unprecedented performance with novel methods, we forge a pathway to more cost-effective and

    powerful sensing.

  • iv

    Acknowledgments

    I would like to thank Professor Edward H. Sargent and Professor Shana O. Kelley for the

    opportunity to work in their labs. Their support and guidance were invaluable over the course of

    this work and they have continuously challenged me to think creatively, encouraged me to find

    new solutions, and pushed me to be a better engineer.

    I would like to extend a thank you to Professor Stewart Aitchison as my committee member for

    his helpful suggestions.

    This work would not be possible without all the members past and present of the Sargent and

    Kelley Labs. I would like to particularly thank Dr. Valerio Adinolfi, Dr. Justin Besant, Dr. Alex

    Ip, Dr. André Labelle, Dr. Brian Lam, Dr. Mahla Poudineh, Dr. Andrew Sage, and Dr. Brandon

    Sutherland for introducing me to and educating me on many aspects of the lab. Thank you to

    Barbara Alexander, Jeannie Ing, Damir Kopilovic, Elenita Palmiano, Dr. Mark Pereira, Dr. Ali

    Seifitokaldani, Remigiusz Wolowiec, and Alexander Zaragoza for help in and outside of the lab.

    I would also like to thank Dr. Jagotamoy Das, Dr. Mahmoud Labib, Dr. Reza Mohamadi, and

    Dr. Sahar Mahshid for sharing their knowledge and expertise. I am grateful to Dr. Sharif Ahmed,

    Dr. Ian Burgess, Bill Duong, Dr. Fengjia Fan, Surath Gomis, Dr. Brenda Green, Dr. Sae Rin

    Jean, Andrew Johnston, Dr. Leyla Kermanshah, Dr. Laili Mahmoudian, Dr. Sara Mahshid, Mona

    Mukhopadhyay, David Philpott, Dr. Tina Saberi Safaei, Dr. Ying Wan, Daniel Wang, Dr.

    Guangli Wang, Dr. Simon Wisnovsky, Fan Xia, Xiaolong Yang, and Dr. Libing Zhang, and Dr.

    Yige Zhou for their help and support. I would especially like to thank Peter Aldridge, Jenise

    Chen, Dr. Eric Lei, Wenhan Liu, Adam Mepham, Carine Nemr, Tanja Sack, and Dr. Sarah Smith

    for their helpful insights and advice, particularly in teaching me about biology and chemistry

    and, most importantly, in making lab life fun.

    I am thankful for my friends, with special appreciation for my NDRC and Iron Dragons families

    to whom I owe a whole host of memorable experiences. A big thank you goes to Dr. Chris Wong

    for his support throughout the years. Last but not least, I would most like to thank my parents,

    Dr. Guanghan Wang and Dr. Bai Zhou, for their continued support in all things I do.

  • v

    Table of Contents

    Acknowledgments.......................................................................................................................... iv

    Table of Contents .............................................................................................................................v

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

    List of Abbreviations ................................................................................................................... xiii

    Chapter 1 ..........................................................................................................................................1

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

    1.1 The Development of Advanced Biosensors .........................................................................1

    1.2 Electrochemical Biosensors .................................................................................................2

    1.3 Scope of Thesis ....................................................................................................................4

    Chapter 2 ..........................................................................................................................................6

    2 Background .................................................................................................................................6

    2.1 Electrochemistry Principles and Techniques .......................................................................6

    2.2 Protein Detection ...............................................................................................................11

    2.3 Nanostructured Electrodes and Nanogaps .........................................................................16

    2.3.1 Nanostructured Electrodes .....................................................................................16

    2.3.2 Nanogaps................................................................................................................18

    Chapter 3 ........................................................................................................................................21

    3 Steric Hindrance Assay for Secreted Factors in Stem Cell Culture ..........................................21

    3.1 Introduction ........................................................................................................................21

    3.2 Background ........................................................................................................................22

    3.3 Results and Discussion ......................................................................................................25

    3.3.1 Assay and Sensor Chip ..........................................................................................25

    3.3.2 DNA-Protein Conjugates and Chip Preparation ....................................................25

    3.3.3 Sensor Characterization .........................................................................................27

  • vi

    3.3.4 Determining Dynamic Range and Sensitivity........................................................27

    3.3.5 Electrochemical Detection in Cell Culture Media Samples ..................................28

    3.4 Methods..............................................................................................................................29

    3.4.1 Materials ................................................................................................................29

    3.4.2 Chip Fabrication.....................................................................................................30

    3.4.3 Sensor Fabrication .................................................................................................30

    3.4.4 Culture Media ........................................................................................................30

    3.4.5 Conjugation of Protein with DNA .........................................................................30

    3.4.6 Preparation of Capture Probe-Modified Chip ........................................................31

    3.4.7 Detection in Buffer and Media ..............................................................................31

    3.4.8 Detection in Stem Cell Culture Samples ...............................................................31

    3.4.9 Electrochemical Measurements .............................................................................31

    3.4.10 Binding Curves ......................................................................................................32

    3.5 Conclusions ........................................................................................................................33

    Chapter 4 ........................................................................................................................................35

    4 A Reusable One-Step Electrochemical DNA-Based Sensor for the Quantitative Detection

    of Soluble Signaling Proteins ....................................................................................................35

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

    4.2 Background ........................................................................................................................36

    4.3 Results and Discussion ......................................................................................................37

    4.3.1 Assay and Sensor Chip ..........................................................................................37

    4.3.2 Sensor Characterization and Sensitivity ................................................................39

    4.3.3 Multiplexing ...........................................................................................................40

    4.3.4 Sensor Regeneration ..............................................................................................41

    4.3.5 On-Chip Resistive Heater ......................................................................................43

    4.4 Methods..............................................................................................................................45

  • vii

    4.4.1 Materials ................................................................................................................45

    4.4.2 Chip Fabrication.....................................................................................................46

    4.4.3 Sensor Fabrication .................................................................................................46

    4.4.4 Culture Media ........................................................................................................46

    4.4.5 Conjugation of Antibodies with DNA ...................................................................46

    4.4.6 Preparation of Capture Complex ...........................................................................46

    4.4.7 Detection in Media .................................................................................................47

    4.4.8 Electrochemical Measurements .............................................................................47

    4.4.9 Sensor Regeneration ..............................................................................................47

    4.5 Conclusions ........................................................................................................................47

    Chapter 5 ........................................................................................................................................49

    5 Programmable Definition of Nanogap Electronic Devices Using Self-Inhibited Reagent

    Depletion ...................................................................................................................................49

    5.1 Introduction ........................................................................................................................49

    5.2 Background ........................................................................................................................50

    5.3 Results and Discussion ......................................................................................................51

    5.3.1 SIRD Method Overview ........................................................................................51

    5.3.2 Solution-Processed Photoconductors .....................................................................53

    5.4 Methods..............................................................................................................................56

    5.4.1 SIRD Device Fabrication .......................................................................................56

    5.4.2 SIRD Electrodeposition .........................................................................................56

    5.4.3 Fabrication of CQD Photoconductor Devices .......................................................57

    5.4.4 EQE Measurements ...............................................................................................57

    5.4.5 Responsivity and Irradiance Measurements ..........................................................57

    5.4.6 Noise Current and Detectivity Measurements .......................................................58

    5.5 Conclusions ........................................................................................................................58

  • viii

    Chapter 6 ........................................................................................................................................60

    6 Conclusions and Future Work ...................................................................................................60

    6.1 Summary ............................................................................................................................60

    6.2 Future Work .......................................................................................................................61

    References ......................................................................................................................................64

  • ix

    List of Figures

    Figure 1.1. Schematic showing the basic components of a biosensor. ........................................... 2

    Figure 2.1. Schematic for a typical three-electrode electrochemical system consisting of the

    working electrode (WE), reference electrode (RE), and counter electrode (CE). .......................... 7

    Figure 2.2. Cyclic voltammetry. (A) Sample curve depicting linear potential ramp. (B) Sample

    curve depicting current with redox reaction peaks. ........................................................................ 9

    Figure 2.3. Differential pulse voltammetry. (A) Sample curve depicting potential ramp with a

    superimposed pulse. (B) Sample curve depicting current versus the potential. ............................. 9

    Figure 2.4. Sample curve depicting the potential step for chronoamperometry. .......................... 10

    Figure 2.5. Square wave voltammetry. (A) Sample curve depicting potential staircase ramp with

    superimposed square wave pulse. (B) Sample curve depicting current versus the potential. ...... 11

    Figure 2.6. Schematic for tandem mass spectrometry. A sample is injected, ionized, and

    accelerated. The ions are separated according to mass and charge through electromagnetic

    deflection. They are analyzed first by mass analyzer 1 (MS1), selectively fragmented, and then

    analyzed by mass analyzer 2 (MS2), which generates the spectra. .............................................. 12

    Figure 2.7. Schematic for surface plasmon resonance. Incident light on the metal film is

    reflected, collected, and analyzed. At a specific incident angle, the plasmons resonate with light,

    thus resulting in light absorption at that angle resulting in a dark line in the reflected light beam.

    The angular position of the dark line moves as a binding event or molecular conformational

    change occurs. ............................................................................................................................... 13

    Figure 2.8. Schematic demonstrating the change in efficiency at which the redox reporter bound

    to the reporting protein reaches the electrode surface upon target binding[47]

    . Reprinted (adapted)

    with permission from [47]

    . Copyright 2017 American Chemical Society. .................................... 14

    Figure 2.9. Schematic illustrating the bio-barcode assay. (Upper) Barcode DNA-functionalized

    gold nanoparticles are conjugated to antibodies. (Lower) The magnetic antibody-functionalized

  • x

    microparticles and the antibody-DNA-gold nanoparticles form a sandwich around the target

    antigen and the barcode DNA is detected using a scanometric assay[48]

    . ..................................... 15

    Figure 2.10. Scanning electron microscope image of a nanostructured microelectrode. ............. 17

    Figure 2.11. Schematic illustrating the blocking assay, where protein is captured using antibody-

    functionalized NMEs, preventing [Fe(CN)6]3-/4-

    from reaching the electrode surface, reducing the

    measured current[12]

    . Reprinted (adapted) with permission from [12]

    . Copyright 2011American

    Chemical Society. ......................................................................................................................... 18

    Figure 2.12. (A) Schematic and (B) Scanning electron micrograph of a nanogap sensor[62]

    ©

    2016 IEEE. .................................................................................................................................... 19

    Figure 3.1. Schematic showing the influence of secreted factors in stem cell culture and an

    overview of the chip-based electrochemical detection scheme. (A) Simplified schematic of the

    interactions between soluble factors and cell subpopulations. As the concentrations of signaling

    proteins increase, mature cells (pink) accumulate as HSCs (blue) tend toward differentiation. As

    the concentration of secreted factors is decreased through media dilution, their impact is reduced,

    promoting the proliferation and self-renewal of HSCs. (B) Chip layout. Contacts are formed

    from circular apertures in a layer of SU-8 covering a gold pattern on chip surface. (C) Schematic

    representation of ASHHA on NMEs. Samples containing the target protein are preincubated with

    the blocking antibody. The samples are then mixed with signaling DNA strands labeled with

    both the electrochemical reporter and the recognition element before on-chip incubation. ......... 24

    Figure 3.2. (A) Signal changes upon conjugation of the RANTES, MDC, and LAP recognition

    complexes and their respective antibodies to the DNA signaling strands. (B) Electrochemical

    signals obtained as a function of time and concentration for the RANTES assay. ...................... 26

    Figure 3.3. Square wave voltammetry-derived currents detected at electrodes with varying

    concentrations of RANTES in (A) buffer and (B) cell culture media. ......................................... 28

    Figure 3.4. Electrochemical detection directly in cell culture media samples. (A) At-line protein

    monitoring schematic. Measured concentrations obtained for specific detection of (B) RANTES,

    (C) LAP, and (D) MDC in culture samples compared against ELISA measurements from the

  • xi

    same samples. Currents obtained from electrochemical measurements for each factor were

    normalized according to calibration curves and converted to concentrations. ............................. 29

    Figure 3.5. Signaling reporter validation. Electrochemical signals obtained from the capture

    strand-functionalized electrodes after 30 min incubation, (1) in the absence of DNA signaling

    strands, and in the presence of DNA signaling strands tagged with (2) biotin, (3) the recognition

    complex, and (4) the recognition complex bound to the blocking antibody. ............................... 32

    Figure 3.6. Binding curves as measured by the sensor for (A) MDC, (B) LAP, and (C) RANTES

    in buffer. (D) Representative binding curve as measured by a standard ELISA for RANTES in

    buffer. ............................................................................................................................................ 33

    Figure 4.1. Schematic illustrating chip-based detection scheme. (A) Chip layout. Gold contacts

    are formed through the apertures in the SU-8 layer and are electroplated to create NMEs. Chip is

    divided into individual sections for multiplexing, separated by lines drawn by a hydrophobic pen.

    (B) Chip layout with individually addressed heaters surrounding each electrode. (C) Schematic

    representation of the one-step proximity principle assay. Two thiolated DNA strands (P1 and P4)

    are tethered to the surface while two complementary strands (P2 and P3, respectively) act as both

    the capture complex and the signaling mechanism. Upon addition of protein, a decrease in

    measured electrochemical current is seen. .................................................................................... 38

    Figure 4.2. Square wave voltammetry-derived currents detected at electrodes with varying

    concentrations of RANTES in cell culture media......................................................................... 39

    Figure 4.3. Electrochemical currents detected at the electrodes for (A) Sample positive for

    RANTES, (B) Sample positive for MDC, and (C) Sample positive for both targets. .................. 41

    Figure 4.4. Currents detected at electrodes before heating, after heating, and after DNA

    rehybridization for the capture complex. ...................................................................................... 43

    Figure 4.5. Relationship between on-chip heater temperature and applied bias. ......................... 44

    Figure 4.6. Electrochemical currents detected at electrodes before and after sensor regeneration

    over five cycles using the on-chip heater. ..................................................................................... 45

  • xii

    Figure 5.1. Self-inhibiting reagent depletion electroplating platform. (A) Lithographic template

    fabrication. Electrical leads with a 2 μm separation distance are passivated with a layer of SU-8

    photoresist. (B) 10-μm apertures (denoted with an arrow) are imaged into the SU-8 passivation

    layer above the 2 μm separation. (C) Imaging after SIRD electroplating electrodes in parallel

    with 2 mM HAuCl4 + 50 mM HCl at 200 mV for 3 min. (D, E) SEM top-down imaging after

    SIRD electroplating. (F) Focused ion beam cross-sectional imaging after SIRD electroplating.

    (Scale bar, 1 μm). .......................................................................................................................... 52

    Figure 5.2. Electroplating SIRD electrodes showing overlapping electrodes that remain

    electrically isolated. (Scale bar = 1 µm). ...................................................................................... 53

    Figure 5.3. SIRD solution-processed photoconductors. PbS colloidal quantum dots are dip coated

    on surface of SIRD devices and infiltrate narrow region creating nanoscale photoconductor

    junctions. (A) Image of SIRD photoconductors with corresponding SEM images. (B, C) SEM

    images. (Scale bar, 1 μm).............................................................................................................. 54

    Figure 5.4. SIRD photoconductor characterization. (A) Spectral external quantum efficiency. (B)

    Responsivity with respect to applied bias. (C) Noise current with respect to applied bias. (D)

    Detectivity with respect to applied bias. ....................................................................................... 56

  • xiii

    List of Abbreviations

    DNA: deoxyribonucleic acid

    RNA: ribonucleic acid

    WE: working electrode

    RE: reference electrode

    CE: counter electrode

    ELISA: enzyme-linked immunosorbent assay

    NME: nanostructure microelectrode

    PNA: peptide nucleic acid

    HSC: hematopoietic stem cell

    ASHHA: amplified steric hindrance hybridization assay

    RANTES: regulated on activation, normal T cell expressed and secreted

    MDC: macrophage-derived chemokine

    TGF-β1: transforming growth factor-β1

    LAP latency-associated peptide

    MCH: 6-mercaptohexanol

    PBS: phosphate buffered saline

    TPO: thrombopoietin

    SIRD: self-inhibited reagent depletion

    SEM: scanning electron microscopy

  • xiv

    CQD: colloidal quantum dots

    NEP: noise-equivalent power

    EQE: external quantum efficiency

    LED: light-emitting diode

  • 1

    Chapter 1

    1 Introduction

    1.1 The Development of Advanced Biosensors

    Biosensors are used for a wide variety of applications, including medical diagnostics, therapeutic

    development, and general healthcare monitoring. The first enzyme electrodes were designed by

    Leland C. Clark and colleagues in 1962[1]

    , and research has since enabled the development of

    sophisticated and advanced devices to detect a variety of biological analytes. With applications

    ranging from glucose monitors to food analysis to cancer diagnosis, biosensors form powerful

    tools that enable the sensitive and rapid detection of important analytes. Recent advances in

    biosensing technology have improved the sensitivity and speed of these devices, allowing for the

    detection of a multitude of targets such as proteins, cells, and nucleic acids.

    In its simplest form, a biosensor is an analytical device designed to identify or quantify a

    biological molecule of interest and create a measurable signal output. As one might expect, its

    name implies that a biosensor is a combination of a biological element and a sensor element. A

    typical device consists specifically of a bio-recognition site, a biotransducer component, and a

    system for readout. The bio-recognition site interacts with a biological sample such as blood, cell

    cultures, or environmental samples to detect the target of interest using designed biological

    receptor elements such as antibodies, proteins, or nucleic acids. The biotransduction component

    converts this interaction into a measurable signal generally through optical, gravimetric, or

    electrochemical means, which can be detected and quantified by the readout system [2]

    . The basic

    components of a biosensor are illustrated in Figure 1.1.

    Biosensors have been designed using a variety of readout techniques. A common example of a

    biosensor is a home pregnancy test, which is based on lateral flow and the detection of human

    chorionic gonadotropin in urine[3]

    . In this case, a visual readout provides an immediate diagnosis.

    Optical sensors[4]

    provide direct visual readout, but often have poor sensitivity. Other methods,

    such as electronic[5]

    , optical[4]

    , and electrochemical[6]

    , calorimetric biosensors[7]

    , and

  • 2

    piezoelectric biosensors[8]

    are more sensitive and more easily designed to detect multiple

    analytes at once, but can be more expensive and require more complex equipment.

    Figure 1.1. Schematic showing the basic components of a biosensor.

    The most commercially successful biosensor is the glucose monitor, which measures the blood

    glucose level through the electrochemical detection of an enzymatic reaction[9]

    . Glucose oxidase

    converts glucose to glucono delta-lactone and hydrogen peroxide, which is then measured by

    amperometric sensing[9]

    . These sensors are relatively inexpensive and can be used by patients

    with minimal training, enabling regular at-home monitoring of blood glucose levels. In this

    work, we place particular emphasis on electrochemical biosensors due to their potential for

    excellent sensitivity, low cost, and ease of miniaturization.

    1.2 Electrochemical Biosensors

    Electrochemical biosensors have several benefits and are of special interest for use in

    applications where high levels of sensitivity are required[10]

    . Electrochemical sensors function

    through the detection of oxidation or reduction reactions at or near the surface of an electrode. A

    potentiostat is used to detect an electronic signal and can take the form of a benchtop or portable,

    handheld device. In contrast, though optical sensors have a long history of use, instrumentation

    needed for sensitive and quantitative detection, such as fluorescence microscopes, is often bulky

    and expensive, limiting their in-line or at-line monitoring applications.

    Electrochemical biosensors have been designed for the detection of many different biological

    molecules, including proteins[11-13]

    , DNA[14-15]

    , and small molecules[16-18]

    , with recognition

  • 3

    elements such as antibodies, nucleic acids, enzymes, and aptamers[19]

    . In many cases, the

    receptor component is designed to be selective for the target and sample preparation is often

    minimal. In some instances, the specific binding of an analyte to a receptor or an enzymatic

    reaction in the presence of a small molecule can directly be detected electrochemically.

    However, there is often no direct electrochemical signal and a system is designed to convert a

    biological recognition event into an electrochemical output. Various reporter methods include

    using covalent[10]

    and non-covalent reporters[20]

    , in addition to enzymatic labels[21]

    and

    conformational changes to DNA structures[22]

    .

    Electrochemical biosensors are, by comparison, less extensively characterized. While

    electrochemical sensors have great potential for high sensitivity due to their ability to measure

    small changes in charge, they must continue to be developed to analyze analytes for more

    exacting applications as detection limits are challenged further and further. Other disadvantages

    of electrochemical biosensors include multiple and complex assay steps and the necessary

    addition of several reagents. These issues need to be addressed for more convenient and

    widespread use, such as improving reaction speed or time to obtain readout through the

    development of one-step assays.

    Novel materials and device and system architectures could provide avenues to overcome these

    limitations. Nanomaterials and nanostructures in particular have found utility in improving the

    performance of electrochemical biosensors and great strides have been made with regards to

    sensitivity and speed from their contributions to the field[23]

    . Electrode materials and structures

    play an important role in sensing applications, and nanomaterials have been shown to act as

    highly effective immobilization matrices[24]

    . Efficient immobilization of molecules onto

    electrode surfaces has been demonstrated with a variety of nanostructured materials with

    different shapes, sizes, and compositions, such as graphite nanoparticles[25]

    , palladium

    nanoparticles[26]

    , electrodes made from nanomaterials (NiFe2O4[27]

    , Tm2O3[28]

    , and Cu2O[29]

    ), and

    nanostructured silica-phytic acid materials with diverse morphologies[30]

    . New sensors need to be

    developed, taking advantage of such electrode materials and structures. As a new generation of

    nano-inspired biosensors emerges, robust characterization and investigation into how to best

    exploit nanotechnology remains a challenge.

  • 4

    In addition to new electrode materials and fabrication techniques, new measurement assays for

    specific analyte detection must be developed in order to fully embrace the advantages of

    electrochemical sensors. Emerging sensors are challenged to breach the current limits of

    detection and function in complex environments. Furthermore, lengthy assay times and multiple

    reagent addition steps necessitate the development of more rapid and user-friendly procedures.

    In this work, we endeavour to confront the shortcomings of currently available sensors and to

    improve on existing methods with new solutions. We aim to tackle issues of sensitivity for

    challenging molecules and increase assay speed while decreasing complexity of use. We display

    enhanced efficacies through the development of novel assays and harness nanostructuring to

    lower detection limits and increase efficiency. To exhibit our findings, we apply our sensors to

    the detection of soluble signaling proteins in stem cell culture samples. Finally, we demonstrate

    improved performance for a solution-processed quantum dot photodetector through the use of

    nanoscale features.

    1.3 Scope of Thesis

    The objective of this thesis is to develop creative approaches to quantitative biosensing. We will

    explore methods for overcoming limitations in sensitivity, speed, and specificity by developing

    novel assays and combining them with a chip-based platform for electrochemical sensing. We

    will also investigate new nanofabrication techniques and their impact on sensor performance.

    Our methods are showcased with a variety of sensors such as photodetectors and with clinical

    applications such as monitoring signaling proteins in stem cell culture.

    In Chapter 2, we present an overview of state-of-art of biosensing as related to protein detection

    as well as a summary of the electrochemical techniques used in this thesis. We also describe

    nanostructured electrodes and fabrication techniques used to create nanoscale features.

    Our goal in Chapter 3 was to explore methods into improving sensitivity for the quantification of

    small proteins. As previous sensors were unable to achieve the desired sensitivity for targets of

    this size in complex environments, we hypothesized that the addition of a large blocking

    antibody could mitigate the challenge of small protein detection. We thus investigated steric

    hindrance assays and techniques to amplify the effects to obtain low detection limits. We

    describe here the results of a combination of DNA hybridization engineering with antibody-

  • 5

    capturing chemistry in an amplified steric hindrance hybridization assay on nanostructured

    microelectrodes, representing the first universal nonsandwich technique that allows for pg·mL-1

    quantification of small proteins in complex media without signal loss.

    While we were able to achieve good levels of sensitivity, we then had a new goal of developing

    simple and rapid sensing to reduce the number of steps, time to readout, and complexity of the

    assay. We investigated one-step assays, particularly proximity-based assays, which have the

    benefit having the potential for a built-in capture surface without the need for additional

    reagents. We further probed methods for sensor regeneration to increase the utility of our device.

    While chemical regeneration is the most common technique, we sought to implement a system

    that did not require additional reagents and came upon the solution of heating, or DNA melting.

    Thus, in Chapter 4, we demonstrate a one-step assay for rapid multiplexed protein detection on a

    reusable sensor. Using a proximity-based assay along with a toehold latch and competitive

    binding between an antibody-protein pair and DNA hybridization, quantification is achieved

    quickly and with the use of an on-chip resistive heater, the sensor can be regenerated over

    multiple cycles.

    In Chapter 5, we investigate nanostructuring and nanoscale features as related to sensors. As

    nanoscale features can enhance sensitivity and other sensor characteristics, we sought to explore

    low-cost fabrication techniques that can achieve the same or better performance and

    controllability compared to established methods. Low costs and simple, scalable techniques are

    highly attractive for multiple applications. We describe here a novel technique for the

    programmable definition of nanogaps separating fractal electrodes, showcased through quantum

    dot photodetectors with low voltage responsivities a hundred times higher than previously

    reported devices.

    Finally, Chapter 6 provides a summary of this work and offers insight into potential future

    directions for biosensing and improvements still to be made.

  • 6

    Chapter 2

    2 Background

    2.1 Electrochemistry Principles and Techniques

    Electrochemistry is a branch of physical chemistry whose early studies included Luigi Galvani’s

    first establishing a link between chemical reactions and electricity in 1791[31]

    . Galvani performed

    experiments on frogs and in one instance, accidentally touched an exposed sciatic nerve with a

    charged metal scalpel, which resulted in sparks and the dead frog’s leg kicking. This observation

    led Galvani to make the connection between electricity and life.

    Following Galvani’s discoveries, Alessandro Volta conducted his own studies, which contributed

    to the eventual invention of the voltaic pile, an early version of the battery capable of producing

    a continuous electric current[32]

    . This invention was the first in a series of other discoveries such

    as electrolysis, and was one of the most important innovations leading to the development of the

    field of electrochemistry. Today, the field has broadened and extended to many other

    applications and devices, such as biosensors.

    Electrochemistry is the study of typically heterogeneous chemical reactions in which electrons

    are transferred between species, whereby one substance loses an electron and is oxidized, while a

    second substance gains an electron and is reduced. Electrochemical systems generally consist of

    three electrodes: a working electrode (WE), a reference electrode (RE), and a counter electrode

    (CE), seen in Figure 2.1. The working electrode is where the electrochemical reaction takes place

    and acts as the charge transfer interface. This electrode is often the site for biological binding

    events. The reference electrode provides a stable reference point by which to measure the

    potential of the working electrode. The most commonly used type of reference electrode is

    silver/silver chloride, as it is inexpensive and non-toxic. The counter electrode, also known as the

    auxiliary electrode, is used to measure the current through the working electrode. It completes

    the circuit along with the working electrode, thus ensuring a stable potential for the reference

    electrode through which current does not pass. The counter electrode generally has a larger area

    than the working electrode so as not to limit the current passing through. Typical materials for

  • 7

    counter electrodes are inert conductors such as gold, platinum, or carbon. In some cases, a two-

    electrode system is used, where a known current or potential is applied between the working

    electrode and a second electrode that replaces both the reference and counter electrodes.

    However, this method is less precise due to the ohmic drop in solution and is generally used with

    energy storage or conversion devices such as batteries or photovoltaic panels. A two-electrode

    system may also be used when the ohmic drop and current are small, such as in the case of

    microelectrodes.

    Figure 2.1. Schematic for a typical three-electrode electrochemical system consisting of the

    working electrode (WE), reference electrode (RE), and counter electrode (CE).

    A potentiostat is an electronic instrument used to control the electrochemical system and

    measure signals between electrodes. It measures and controls the potential difference between

    the working and reference electrodes using a control amplifier in addition to measuring the

    current flow between the working and counter electrodes via the ohmic drop across a series

  • 8

    resistor. Various electrochemical techniques are classified according to the type of measurement

    that is being made and include amperometric/voltammetric, impedimetric, conductometric and

    capacitive, and field-effect transistor-based sensors.

    Voltammetric and amperometric sensors measure a change in current due to an electrochemical

    oxidation or reduction while applying a potential to the working electrode versus the reference

    electrode. These sensors allow for good selectivity as the oxidation or reduction potential used is

    usually specific to the target. We offer a brief overview here of a few of the most common

    techniques, namely cyclic voltammetry, differential pulse voltammetry, chronoamperometry, and

    square wave voltammetry.

    Cyclic voltammetry is a frequently used method that measures current in an electrochemical cell

    while the potential changes, typically sweeping from an initial voltage and increasing (or

    decreasing) linearly with time until it reaches a defined endpoint before decreasing (or

    increasing) at the same rate back to the initial voltage. A redox reaction will occur as the voltage

    approaches the redox potential and the rate of this reaction will increase as the potential is

    increased, which will in turn result in an increase in the current. Once the electroactive species is

    consumed and the rate of the reaction exceeds that of the electroactive species diffusive

    replenishment, the current will then decrease. In the reverse scan, the opposite reaction will

    occur with a similar peak in the opposite direction. An example of sample curves depicting the

    varying potential and resulting current with time are shown in Figure 2.2. Cyclic voltammetry is

    a useful technique that offers information on analyte concentration, redox potential, and reaction

    rates through analysis of the peak current and width. Information can also be provided regarding

    the electrode surface area and electron transport efficiency as the surface adsorbents go through a

    redox reaction. This information can be obtained by calculating the area under the current peak,

    adjusting the scan rate, and observing the peak shape.

  • 9

    Figure 2.2. Cyclic voltammetry. (A) Sample curve depicting linear potential ramp. (B) Sample

    curve depicting current with redox reaction peaks.

    A second technique, differential pulse voltammetry, is similar to cyclic voltammetry, where the

    potential is linearly increased (or decreased), with the addition of a small superimposed pulse.

    The current is measured immediately before the pulse and immediately before the end of the

    pulse, with the difference between them plotted against the varying potential. This technique

    offers the advantage of low background currents and high sensitivity due to the background

    current being removed by the subtraction of the two pre- and post-pulse current values, resulting

    in a differential current being reported. Sample curves illustrating the voltage pulse and resulting

    current peak are shown in Figure 2.3.

    Figure 2.3. Differential pulse voltammetry. (A) Sample curve depicting potential ramp with a

    superimposed pulse. (B) Sample curve depicting current versus the potential.

    A B

    A B

  • 10

    Chronoamperometry is an electrochemical technique where the potential is maintained at a

    constant value and the current is measured as a function of time. The potential may be modeled

    after a step function (seen in Figure 2.4), where it is initially at a constant value where no redox

    reaction occurs, and is then suddenly stepped to a value at which a redox reaction may occur. For

    short time periods after charging is complete, the current can be modeled by the Cottrell equation

    (eq. 2.1),

    𝑖(𝑡) = 𝑛𝐹𝐴𝐶√𝐷

    √𝜋𝑡 (eq. 2.1)

    where n is the number of electrons, F is Faraday's constant, A is the area of the (planar)

    electrode, C is the concentration, D is the diffusion coefficient, and t is time. This technique is

    useful for measuring the electroactive area of the electrode or the diffusion coefficient for the

    electroactive species in question, using the Cottrell equation. It is also used for real-time

    monitoring, especially once a system reaches steady-state such as for spherical or

    microelectrodes, as a change in electroactive moiety would result in a subsequent change in

    current. An additional application is in electroplating electrodes.

    Figure 2.4. Sample curve depicting the potential step for chronoamperometry.

    In this work, square wave voltammetry is the main technique used for electrochemical

    measurements due to its high level of sensitivity and speed. It is a pulse method similar to

    differential pulse voltammetry and consists of a linear potential sweep in the form of a square

    wave pulse superimposed onto a staircase potential. The current is obtained by measuring the

    difference between the forward and reverse potential pulses before the potential direction is

    reversed. This sampling technique results in minimal capacitive current. This technique is a

  • 11

    preferred choice as very low detection limits may be obtained, making it an excellent option for

    the detection of analytes with low concentrations. A sample waveform of the potential and

    current for this method are shown in Figure 2.5.

    Figure 2.5. Square wave voltammetry. (A) Sample curve depicting potential staircase ramp

    with superimposed square wave pulse. (B) Sample curve depicting current versus the potential.

    2.2 Protein Detection

    Due to their low limits of detection, electrochemical biosensors have been used to detect a

    multitude of analytes. In this thesis, we will highlight our work on developing novel sensors to

    detect proteins, which are biologically relevant to many applications, including tumor markers,

    clinical diagnostic testing, drug dosing, and as signaling factors. We focus on an application to

    monitor soluble signaling proteins in stem cell cultures, which are heterogeneous and complex

    environments. Though present in low quantities (around the pg·mL-1

    level), these proteins play

    an influential role in cell expansion and differentiation[33-35]

    and thus control over these factors

    through frequent or continuous monitoring would allow for improved cell expansion. Ideally, a

    biosensor for protein detection would provide a rapid response and be highly sensitive and

    specific for the protein of interest with the ability to perform integrated or automated monitoring,

    making this a difficult challenge.

    Currently, the gold standard method for protein quantification is the enzyme linked

    immunosorbent assay (ELISA). In this method, proteins are immobilized and specifically labeled

    via antibodies in sandwich-type reactions and are subsequently quantified through

    spectrophotometry[36]

    . Commercial ELISA kits are widely available and are sensitive, with low

    A B

  • 12

    detection limits in the pg·mL-1 range. However, ELISAs usually require several hours for

    operation due to multiple steps including many washes, as well as large sample volumes and

    auxiliary equipment, and so are more suitable for off-line measurements and require trained

    personnel.

    There are currently a few techniques that are capable of directly performing protein analysis

    without the need for additional reagents. Mass spectrometry combined with immunoassay

    techniques, along with selected reaction monitoring, is a highly sensitive protein analysis

    technique. In mass spectrometry, chemical species are ionized and these ions are separated

    according to their mass to charge ratio and then quantified with a mechanism capable of

    detecting ions. Selected reaction monitoring is used in tandem mass spectrometry, in which

    multiple steps of mass spectrometry are performed, seen in Figure 2.6. An ion with a specific

    mass is selected in the first stage of tandem mass spectrometry and in the second stage, an ion

    product from the precursor fragmentation reaction is selected. Unfortunately, these techniques

    require complex equipment and thus as with ELISAs, they have limited utility for in-line

    measurements[36-39]

    . Another direct technique, Raman spectroscopy, may also be used to identify

    proteins via their spectral signatures using lasers. A sample is irradiated with a laser, resulting in

    a small amount of Raman scattering, which is then detected as a Raman spectrum. However, this

    technique has not yet been used for quantification[40]

    .

    Figure 2.6. Schematic for tandem mass spectrometry. A sample is injected, ionized, and

    accelerated. The ions are separated according to mass and charge through electromagnetic

    deflection. They are analyzed first by mass analyzer 1 (MS1), selectively fragmented, and then

    analyzed by mass analyzer 2 (MS2), which generates the spectra.

  • 13

    Another method used for the multiplexed detection of many proteins is the usage of microarrays,

    such as in surface plasmon resonance[41]

    (Figure 2.7). In this technique, proteins are captured

    onto a surface via a ligand, such as an antibody, which has been immobilized on to a substrate

    coated in a metal film, known as the sensor chip. The presence of the proteins generates an

    optical response, which is then measured by a detector. Though on-line measurements are able to

    be made quickly, detection limits are around the ng·mL-1 level and are therefore unsuitable for

    applications with lower protein concentrations.

    Figure 2.7. Schematic for surface plasmon resonance. Incident light on the metal film is

    reflected, collected, and analyzed. At a specific incident angle, the plasmons resonate with light,

    thus resulting in light absorption at that angle resulting in a dark line in the reflected light beam.

    The angular position of the dark line moves as a binding event or molecular conformational

    change occurs.

    Aptamers have been developed to bind tightly to specific proteins, and have been used to

    generate an electrical or optical response when bound to the protein of interest. Aptamer binding

    has faster reaction kinetics compared to antibody binding, which reduces the time needed for

    traditional immunoassays[42]

    . However, the widespread availability of aptamer-based protein

  • 14

    detection kits is limited due to the extensive and expensive development required to generate

    aptamers for individual proteins.

    Electrochemical sensors are an attractive option for direct protein detection due to their

    sensitivity and potential for miniaturization[43]

    . Many such sensors have already been

    demonstrated with a variety of assays and detection platforms. Several studies have made use of

    the immobilization of proteins on gold electrodes functionalized with target-specific antibodies

    combined with electrochemical reporters, but demonstrated long assay times[44-46]

    . Another group

    described a reagentless one-step assay based on a redox-reporter-modified protein anchored to an

    electrode[47]

    . Upon specifically binding to the target antigen, the efficiency at which the reporter

    reaches the electrode surface is altered, resulting in a change in the measured redox current

    (Figure 2.8). However, the sensitivity of this assay only reached the ng·mL-1

    level.

    Figure 2.8. Schematic demonstrating the change in efficiency at which the redox reporter

    bound to the reporting protein reaches the electrode surface upon target binding[47]

    .

    Reprinted (adapted) with permission from [47]

    . Copyright 2017 American Chemical Society.

    Nanomaterials and nanostructures have been garnering especial interest in recent years for

    sensing applications, particularly for their benefits in improving sensitivity[12,23]

    . One example is

    a bio-barcode approach, where 330 fg·mL-1

    detection of prostate specific antigen was achieved

    using a specific DNA sequence as a reporter[48]

    . In this assay, shown in Figure 2.9, magnetic

    microparticle probes functionalized with antibodies are mixed with the target antigen, washed

    free of excess serum, and combined with gold nanoparticles functionalized with antibody-

    conjugated DNA barcode probes in a sandwich assay. After magnetic separation and wash steps,

  • 15

    the DNA barcodes are detected using the scanometric assay. While this technique is extremely

    sensitive, it involves the use of multiple wash steps and significant lead time.

    Figure 2.9. Schematic illustrating the bio-barcode assay. (Upper) Barcode DNA-

    functionalized gold nanoparticles are conjugated to antibodies. (Lower) The magnetic antibody-

    functionalized microparticles and the antibody-DNA-gold nanoparticles form a sandwich around

    the target antigen and the barcode DNA is detected using a scanometric assay[48]

    .

    Silicon nanowire based field-effect transistor sensors have been described for the detection (92

    pg·mL-1

    ) of cardiac troponin I using immobilized antibodies on the silicon nanowire surfaces, but

    this method was only validated in buffered solutions[49]

    . Another technique for electrochemical

    sensing is electrode modification, such as with gold nanoparticle-coated glassy carbon electrodes

    for the detection of prostate specific antigen in the range of 2 pg·mL-1

    to 10 ng·mL-1

    using a

    sandwich assay and antibodies labeled with the nanocomposite of ferrocene monocarboxylic acid

  • 16

    hybridized graphene oxide[50]

    . While sensitive, its multiple steps and wash cycles and thus time

    required resemble those of an ELISA.

    2.3 Nanostructured Electrodes and Nanogaps

    2.3.1 Nanostructured Electrodes

    For the detection of pg·mL-1

    levels of analytes, ultrasensitive sensors must be developed. There

    have been many advances, notably nano-inspired approaches using nanomaterials and

    nanostructured surfaces. With a combination of micro- and nanomaterials, these sensing schemes

    have made significant improvements in both sensitivity and speed[12]

    . In this work, we focus on

    nanostructured microelectrodes and nanogap-based sensors.

    As electrodes shrink in size, the diffusional regime changes from linear to radial[51]

    . A higher

    flux of analytes is able to reach the electrode, which in turn allows for increased signal to noise

    ratios and lower limits of detection[52]

    . There is, however, a limit to the miniaturization of the

    working electrode, as the interactions between the analyte and the electrode surface may not

    occur within a reasonable timeframe with too small of a surface. This issue has been addressed

    by the Kelley Laboratory, who worked to increase sensitivity by increasing the surface area of

    their microelectrodes through nanostructuring[53]

    .

    Whereas classical electrodes are gold disk electrodes, the Kelley group developed non-planar

    metal electrodes electroplated with metal alloys or pure noble metals such as gold or platinum[54]

    .

    Under different electrodeposition conditions, various sizes and morphologies could be obtained

    to form nanostructured microelectrodes (NMEs)[55]

    . With these microelectrodes, the metal

    surface becomes textured at the nanoscale with the structures extended three-dimensionally

    outward into the solution. Using chips pre-patterned with 5 µm gold apertures with the rest of the

    electrode covered by an insulating layer, three-dimensional NMEs are fabricated through

    electrodeposition using a metal salt, seen in Figure 2.10. Other methods such as thermal

    wrinkling[56]

    , in-situ formation of the nanostructured materials[57-58]

    , and hierarchical growth[55]

    have also been used to form nanostructures. This nanostructuring has the effect of increasing the

    surface area of the working electrode and consequently the signal intensity and sensitivity, as

    well as speed of detection by increasing the probability of interaction between an analyte and a

  • 17

    receptor at the electrode surface. The use of NMEs has resulted in detection of biomolecules at

    concentrations as low as attomolar[54]

    .

    Figure 2.10. Scanning electron microscope image of a nanostructured microelectrode.

    Once nanostructuring is complete, electrodes can be functionalized using thiolated probe

    molecules. Due to the rough electrode surface, probes such as thiolated DNA capture molecules

    may have an increased range of immobilization deflection angles compared to planar surfaces,

    increasing their accessibility and hybridization efficiency[59]

    .

    Several chip-based platforms utilizing nanostructured electrodes have been reported that can

    achieve multiplexed sensing of analytes with improved sensitivity compared to that of planar

    surfaces[60]

    . Combined with a simple electrochemical readout, these sensor chips can detect

    DNA, RNA, protein, and small molecules qualitatively within 30 minutes with high specificity in

    diverse solutions, from human serum to whole blood[12]

    . These chips can be further integrated

    into microfluidic devices for automated sample processing and detection.

    A variety of assays have been developed using this platform and various electrochemical

    reporters have been used. One such assay specifically detects cancer antigen 125, a large 200

    kDa protein, with a detection limit of 100 pg·mL-1

    using [Fe(CN)6]3-/4-

    as the electrochemical

    reporter (Figure 2.11)[12]

    . A capture antibody was assembled on the surface of the electrode and

    the presence of the target analyte results in the inhibition of the interfacial electron transfer, as

    [Fe(CN)6]3-/4-

    can no longer interact with the electrode surface. However, this assay is unable to

  • 18

    detect small proteins as their small size is insufficient to block the electrochemical reporter from

    reaching the electrode surface.

    Figure 2.11. Schematic illustrating the blocking assay, where protein is captured using

    antibody-functionalized NMEs, preventing [Fe(CN)6]3-/4-

    from reaching the electrode

    surface, reducing the measured current[12]

    . Reprinted (adapted) with permission from [12]

    .

    Copyright 2011American Chemical Society.

    Another assay developed to detect a number of analytes including small molecules employs a

    capture aptamer nucleic acid probe attached to the surface of an electrode, bound to a

    neutralizing complementary PNA molecule[61]

    . The aptamer preferentially binds to the target as

    the neutralizer contains base-pair mismatches, after which a change in the charge at the sensor

    surface is detected due to the dissociation of the aptamer-neutralizer complex. This assay can

    detect small molecules, but the detection limit of 1 μg·mL-1

    is quite high for our application.

    Nanostructured microelectrodes have significant potential for ultrasensitive biosensors and,

    combined with an electrochemical readout and a creative assay, present an appealing choice for

    the purposes of this work.

    2.3.2 Nanogaps

    Nanogap sensors offer another attractive option for sensing. A nanogap is formed when two

    electrodes are separated by submicron distances. Benefits of nanogaps include reducing the

    interdiffusion time, low power consumption and reagent volumes, trapping biomolecules, and the

    ability to create large electric fields, and in the case of an electrochemical sensor, enabling

    analytes to quickly cross the gap and switch between the oxidized and reduced states. The

    amperometric response is amplified, thereby increasing the signal.

  • 19

    Several sensors have been reported that take advantage of nanogaps to enhance sensitivity. Using

    coplanar nanogap electrodes (Figure 2.12) and a conductive linker, 10 pg·mL-1

    detection of

    cardiac troponin T was achieved[62]

    . Conductive linkers were immobilized onto the nanogap

    surface between the electrodes and crosslinked with antibodies. As the target antigen was bound

    to the antibodies, the measured conductance decreased according to the concentration of antigen

    in solution.

    Figure 2.12. (A) Schematic and (B) Scanning electron micrograph of a nanogap sensor[62]

    © 2016 IEEE.

    In another case, nanogaps were used to trap nanoparticles for analysis using dielectrophoresis[63]

    .

    By shrinking the distance between electrodes to sub-10 nm, strong trapping forces were created

    at low biases due to the strong electric field gradients and without the typical challenges

    associated with high voltages such as heat generation, bubble formation, and unwanted surface

    electrochemical reactions.

    A third group fabricated electrochemical nanogap devices based on signal amplification by redox

    cycling[64]

    . In this work, a current amplification factor of 2.5 was achieved in redox cycling dual

    mode compared to single mode, and detection of 5 μM of Fc(MeOH)2 was performed with a

    volume of 20 aL, demonstrating extremely low sample volumes.

    One type of nanogap is sharp pointed electrodes with interelectrode spacing to match that of the

    target analyte. Sensors with this format can result in fringing effects and poor redox

    amplification. The second type is better employed for sensing, where there exists both submicron

    spacing between electrodes and large electrode surface area[65]

    . While most early nanogap

  • 20

    devices form coplanar horizontal point-like junctions, many current fabrication techniques also

    allow for the formation of larger electrode surface areas, which, combined with the narrow gaps

    lead to improved device performance[65]

    .

    Nanogaps can be formed by a variety of different methods. Conventional nanofabrication

    includes techniques such as atomic layer deposition [66-67]

    , dip-pen nanolithography[68]

    , electron

    beam lithography[69-70]

    , molecular lithography[71]

    , focused ion beam milling[72]

    , molecular beam

    epitaxy[73]

    , nanosphere lithography[74]

    , interference lithography[75-76]

    , block copolymer

    lithography[77-79]

    , and galvanic displacement[80-81]

    . Unfortunately, the high costs, time

    consumption, or lack of control associated with fabrication for these methods act as deterrents

    against more widespread use. Thus, new simpler and cheaper fabrication methods are in demand

    for this technology to have more applicability.

  • 21

    Chapter 3

    3 Steric Hindrance Assay for Secreted Factors in Stem Cell Culture

    3.1 Introduction

    Biosensors often display great sensitivity in controlled conditions, but can suffer from loss or

    drift in signal in complex environments, such as in media or solutions from real samples.

    Additionally, many targets are small in size and difficult to detect, particularly at low

    concentrations, using size-based techniques such as blocking assays. In this chapter, we

    investigate methods into improving sensitivity for the detection of small proteins. We discuss

    here the development of an electrochemical biosensor based on a novel amplified steric

    hindrance assay using nanostructured microelectrodes for the purpose of monitoring soluble

    signaling proteins in stem cell culture. As stem cell culture media is heterogeneous and complex,

    specific and sensitive quantification is challenging.

    We made use of a steric hindrance assay with amplified effects to obtain the necessary

    sensitivity. This amplification is performed via careful size-based DNA hybridization

    engineering and the use of nanostructured microelectrodes developed by the Kelley Laboratory.

    In order to maintain specificity, we used an antibody-based competition scheme to capture the

    targets of interest.

    This work was an effort to accomplish low-level detection of small signaling proteins influential

    in the expansion and differentiation of hematopoietic stem cells (HSCs) grown in culture. We

    endeavoured to find a sensing strategy that would allow for possible future integration with a

    bioreactor while also minimizing sample volumes and achieving the necessary sensitivity. We

    were able to attain 10 pg·mL-1

    quantification comparable to that of the current gold standard

    while improving on assay time and requiring very little sample volume.

    This chapter contains materials from the manuscript:

  • 22

    Reprinted with permission from Zhou, W., Mahshid, S.S., Wang, W., Vallée-Bélisle, A.,

    Zandstra, P.W., Sargent, E.H., Kelley, S.O., “Steric hindrance assay for secreted factors in stem

    cell culture.” ACS Sensors, 2017, 2, 4, 495-500. Copyright 2017 American Chemical Society.

    Disclosure of work within this manuscript: W.Z., S.S.M, E.H.S., and S.O.K. designed the

    experiments. A. V-B. provided assistance with experimental design. W.Z. performed all

    experiments unless otherwise specified and interpreted the results with assistance from E.H.S.,

    and S.O.K. W.W. and P.W.Z. designed the experiments regarding the fed-batch bioreactor

    culture system and ELISA validation. W.W. provided the cell culture media and performed the

    fed-batch bioreactor culture system experiment and ELISA validation. W.Z., S.S.M., E.H.S., and

    S.O.K. wrote the manuscript.

    3.2 Background

    HSC transplantation is used as a clinical therapy for hematological pathologies including blood

    cancers and immune system disorders[82-83]

    . Umbilical cord blood is an appealing source of

    HSCSs[82,84]

    , but its clinical use is limited by the low cell numbers available[85]

    , prompting the

    need for ex vivo expansion. Expansion is made especially difficult by the accumulation of

    endogenously produced signaling protein factors secreted from off-target cell populations, which

    promote unwanted differentiation[33-35]

    . Sensitive and specific detection of the various secreted

    proteins that regulate HSC expansion would enable control over their concentrations, improving

    ex vivo HSC growth.

    Strategies to promote HSC expansion include attempts to minimize the influence of mature

    cells[35,86-88]

    , through the supplementation of additional factors[89-92]

    and regular media exchange

    to slow the accumulation of secreted proteins. Even at low concentrations, these factors have a

    strong impact on cell fate decisions[93-95]

    , with signals from mature blood cells leading to a net

    negative effect on HSC expansion (Figure 3.1)[33,35]

    . The quantification of signaling factors

    would allow for the development of process control strategies to monitor and regulate the

    concentrations of these proteins. An integrated sensor to provide sensitive, real-time feedback on

    secreted proteins would thus be highly attractive as it would reduce the impact of secreted factors

    on HSC differentiation and enhancing HSC expansion.

  • 23

    The enzyme-linked immunosorbent assay (ELISA) is the current gold standard method for

    protein quantification, but the long process times involved, along with the labels and equipment

    required, make this method less convenient for in-line monitoring applications. Several other

    techniques exist for protein analysis, such as selected reaction monitoring[37]

    , Raman

    spectroscopy[96]

    , and surface plasmon resonance[97]

    , but these methods either do not lend

    themselves easily to integration due to the complex equipment required or do not have

    sufficiently low limits of detection. Other sensors are based on improving reaction kinetics to

    reduce process times and make use of assays employing aptamers[42]

    or microbeads[98]

    , but are

    limited in their widespread use due to the considerable development involved, high costs, and

    variability. Specific pg·mL-1

    protein quantification in complex media using techniques amenable

    to automation is difficult to achieve and poses a significant challenge.

    Electrochemical sensors are an attractive option for protein monitoring due to their versatility,

    integration capability, and excellent sensitivies/low limits of detection[13,18,54,99-111]

    . In particular,

    chip-based platforms that make use of modified surfaces such as nanostructured microelectrodes

    (NMEs) can achieve multiplexed immunosensing of several factors and have improved

    sensitivity compared to that of planar surfaces. While several blocking assays or sandwich assays

    have been developed for the analysis of proteins using antibody-modified sensors[12,61,112-113]

    , the

    detection of low molecular weight proteins at low concentrations has remained difficult. Given

    that most secreted factors involved in stem cell differentiation in culture are small proteins with

    100 amino acids or fewer, new assay configurations are needed to target the important

    application of stem cell culture engineering.

    Here, we describe a novel method for the quantification of signaling proteins in primary stem

    cell cultures using a sensitive on-chip detection strategy. Drawing inspiration from the design of

    a recently developed assay that uses steric hindrance effects to detect large proteins, namely,

    antibodies and streptavidin[13]

    , we report on a powerful approach to the analysis of small secreted

    proteins. Only by combining the use of size-controlled DNA hybridization engineering on three-

    dimensional gold NMEs with an alternative competitive antibody attachment scheme, we were

    able to improve on the original assay to enhance steric hindrance effects in a new amplified steric

    hindrance hybridization assay (ASHHA) for the detection of small proteins. We present a highly

    specific protein capturing system and engineer a wide dynamic range from 10 pg·mL-1

    to 10

    ng·mL-1

    for a number of targets that are important for stem cell expansion.

  • 24

    Figure 3.1. Schematic showing the influence of secreted factors in stem cell culture and an

    overview of the chip-based electrochemical detection scheme. (A) Simplified schematic of the

    interactions between soluble factors and cell subpopulations. As the concentrations of signaling

    proteins increase, mature cells (pink) accumulate as HSCs (blue) tend toward differentiation. As

    the concentration of secreted factors is decreased through media dilution, their impact is reduced,

    promoting the proliferation and self-renewal of HSCs. (B) Chip layout. Contacts are formed

    from circular apertures in a layer of SU-8 covering a gold pattern on chip surface. (C) Schematic

    representation of ASHHA on NMEs. Samples containing the target protein are preincubated with

    the blocking antibody. The samples are then mixed with signaling DNA strands labeled with

    both the electrochemical reporter and the recognition element before on-chip incubation.

  • 25

    3.3 Results and Discussion

    3.3.1 Assay and Sensor Chip

    The glass microchips used as the sensor platform and the protein detection strategy are illustrated

    in Figure 3.1. Gold contacts covered in SU-8 with 5 μm apertures formed the templates for the

    100 μm electroplated NMEs. The surface of the NMEs is functionalized with immobilized

    thiolated capture DNA strands to form a high density DNA monolayer.

    As shown in Figure 1, the detection of small secreted proteins is accomplished by monitoring the

    competitive binding of a blocking antibody. The target protein is attached to a strand of DNA,

    which is also labeled with the redox-active reporter methylene blue; this is referred to as the

    signaling DNA. If the target protein is present in solution at a high concentration, the blocking

    antibody will not bind to the conjugated target on the signaling DNA, and this molecular species

    is therefore free to bind to the electrode surface, producing a significant level of electrochemical

    current. If the target protein is not present, the blocking antibody binds to the signaling DNA,

    suppresses the hybridization of the DNA at the sensor surface because of steric hindrance, and

    decreases the level of electrochemical current observed. For intermediate levels of protein, the

    amount of current would then be proportional to the concentration of target protein. This

    approach links increases in current with increased concentrations of protein in solution

    irrespective of the size of the target. We take advantage of the competition chemistry in the

    ASHHA approach and, through the incorporation of the NME platform and the inclusion of a

    large antibody, enable the sensitive detection of small secreted proteins at low concentrations.

    3.3.2 DNA-Protein Conjugates and Chip Preparation

    We selected three analytes to demonstrate the effectiveness of this approach: (1) regulated on

    activation, normal T cell expressed and secreted (RANTES); (2) macrophage-derived chemokine

    (MDC); and (3) transforming growth factor-β1 (TGF-β1). All three are secreted factors

    deleterious to HSC expansion[114-115]

    . These factors are significant modulators in stem cell

    culture and therefore monitoring and controlling the concentrations of these proteins is

    important. As TGF-β1 is only present in culture in latent form and cannot be measured through

    antibody detection without an activation step, we used the latency-associated peptide (LAP) as a

    surrogate. LAP binds to TGF-β1 to form the Small Latent Complex[116-117]

    and its concentration

  • 26

    correlates very well to that of TGF-β1[98]

    . We generated target-DNA complexes corresponding to

    all three proteins and monitored the steric effects resulting from the binding of the blocking

    antibody to the DNA signaling reporter by measuring the differences among electrochemical

    signals obtained upon hybridization (Figure 3.2A). The DNA-protein conjugates were incubated

    on-chip at a concentration of 30 nM and the signal was measured after 40 min. Square wave

    voltammetry was used to scan the sensors for the detection of methylene blue, with the reduction

    potential peak located at -0.25 V versus Ag/AgCl.

    Figure 3.2. (A) Signal changes upon conjugation of the RANTES, MDC, and LAP

    recognition complexes and their respective antibodies to the DNA signaling strands. (B)

    Electrochemical signals obtained as a function of time and concentration for the RANTES

    assay.

  • 27

    3.3.3 Sensor Characterization

    As shown in Figure 3.2, the presence of antibodies against RANTES, MDC, or LAP all caused

    large changes in the electrochemical current in assays featuring corresponding DNA-protein

    signaling conjugates. Addition of the blocking antibodies caused signal changes of 58%, 59%,

    and 75% in terms of gain reduction for RANTES, MDC, and LAP, respectively, confirming the

    ability of the sensor to detect each of the three target proteins (Figure 3.2A). The largest protein,

    LAP produced the largest change in signal, but the smaller proteins also produced measurable

    signal changes. We also investigated the time and concentration dependence of the assay for

    RANTES (Figure 3.2B), and observed that discernible signal changes could be detected as early

    as 10 min.

    3.3.4 Determining Dynamic Range and Sensitivity

    To determine the dynamic range of this sensor for the detection of signaling proteins, a

    concentration series was performed using RANTES and a human RANTES antibody (Figure

    3.3). A sample of each concentration of RANTES was mixed with the blocking RANTES

    antibody and then incubated with the RANTES-bound signaling strands. The solution was

    pipetted onto chips containing NMEs functionalized with capture strands and each electrode was

    scanned after 30 min. In buffer solution, the sensors were able to detect concentrations ranging

    from 10 pg·mL-1

    to 10 ng·mL-1

    (Figure 3.3A). This range of detection sensitivity has practical

    significance as these proteins are typically present in culture at concentrations between 10

    pg·mL-1

    to 1 ng·mL-1[118]

    . The change in current was measured for each concentration of protein,

    and the greatest reduction in current was observed for the lowest protein concentrations.

    In order to demonstrate the utility of ASHHA for in-line monitoring of signaling proteins,

    detection was also performed in cell culture media, with RANTES spiked into media (Figure

    3.3B). The assay is shown to be sensitive even in media, with no signal drift or change in current

    compared to results obtained from using buffer. Furthermore, it is specific enough to function

    selectively in a heterogeneous solution.

  • 28

    Figure 3.3. Square wave voltammetry-derived currents detected at electrodes with varying

    concentrations of RANTES in (A) buffer and (B) cell culture media.

    3.3.5 Electrochemical Detection in Cell Culture Media Samples

    Electrochemical sensors have the advantage of being versatile, and can be easily integrated into a

    culture system with the potential for multiplexing. We highlighted our approach by using these

    sensors to analyze samples drawn from a HSC fed-batch bioreactor culture system that was made

    to dilute soluble factors (Figure 3.4A). Fresh media was added to the culture each day to dilute

    the solution, lowering the concentration of the secreted factors and thus minimizing their impact

    on HSC expansion. Samples were taken from the culture every 4 days over the course of 12

    days, with the sample from day 0 contained no secreted factors and treated as the baseline signal

    control sample. The electrochemical current levels of each of the signaling proteins of interest

    (RANTES, MDC, and TGF-β1) were measured from these samples, compared to measured

    calibration curves, and converted to concentrations, the results of which were validated through

  • 29

    comparison with measurements from an ELISA (Figure 3.4)[118]

    . The results obtained using the

    electrochemical strategy compared very favorably with the ELISA method, indicating that this

    new approach displays comparable accuracy relative to the gold standard.

    Figure 3.4. Electrochemical detection directly in cell culture media samples. (A) At-line

    protein monitoring schematic. Measured concentrations obtained for specific detection of (B)

    RANTES, (C) LAP, and (D) MDC in culture samples compared against ELISA measurements

    from the same samples. Currents obtained from electrochemical measurements for each factor

    were normalized according to calibration curves and converted to concentrations.

    3.4 Methods

    3.4.1 Materials

    The DNA probes were purchased from Biosearch Technologies Inc., Novato, CA. The sequence

    of the capture strand used in this work is GGA ATG AAG TCG ATG GAC CTT ACC TGC

    CTT GT, with a thiolated 5' terminal. The sequence of the signaling strand is ACA AGG CAG

    GTA AGG TCC ATC GAC TTC ATT CC with methylene blue at the 3' terminal and biotin at

  • 30

    the 5' terminal. RANTES, MDC, TGF-β1, and their respective antibodies were purchased from

    R&D Systems (Minneaopolis, USA). Unless otherwise specified, all other chemicals were

    purchased from Sigma-Aldrich and used without any further purification. 1x phosphate buffered

    saline (PBS) with a pH of 7.4 was used throughout the experiments.

    3.4.2 Chip Fabrication

    Chips were fabricated using glass substrates from Telic Company (Valencia, USA) that were

    precoated with 5 nm Cr, 50 nm Au, and AZ1600 positive photoresist. Twenty electrodes were

    patterned using standard constant lithography while the Cr and Au were etched using their

    respective etchants, after which the positive photoresist was removed. A layer of SU-8 2002 was

    spin-coated at 5000 rpm, 30 s and patterned using contact lithography to create 5 μm apertures.

    Chips were then cleaned with acetone, rinsed with isopropyl alcohol and deionized water, and

    dried with a flow of air.

    3.4.3 Sensor Fabrication

    NMEs were fabricated by electroplating at room temperature in a solution of 20 mM HAuCl4

    and 0.5 M HCl at a constant potential of -400 mV for 60 s. Electrodeposition was performed

    using an Epsilon potentiostat using a standard three-electrode system with an Ag/AgCl reference

    electrode and a Pt counter electrode.

    3.4.4 Culture Media

    The stem cell culture media used in this work consisted of serum-free IMDM media from Gibco

    (Rockville, USA), 20% BIT serum substitute from StemCell Technologies (Vancouver, Canada),

    and 1% Glutamax (Gibco). Media was further supplemented with 100 ng·mL-1

    stem cell factor,

    100 ng·mL-1

    F1t3 ligand, 50 ng·mL-1

    thrombopoietin, and 1 μg·mL-1

    low-density lipoproteins.

    3.4.5 Conjugation of Protein with DNA

    The conjugation of the proteins (RANTES, MDC, and LAP) to streptavidin was performed using

    the Lighting-Link Streptavidin Antibody Labeling Kit from Novus Biologicals (Oakville,

    Ontario). These protein-streptavidin complexes were then incubated with the biotinylated

    signaling probes for minimum 30 minutes at a 1:3 molar ratio of protein to DNA to form as tock

    solution.

  • 31

    3.4.6 Preparation of Capture Probe-Modified Chip

    A 50 μL immobilization solution containing the 100 nM capture probe solution was incubated on

    the chip overnight at 4°C. After the removal of excess solution and washing the chip with PBS,

    50 μL of 1.37 mM MCH solution was added to the chip as a backfilling step as well as to block

    potential remaining active sites from nonspecific adsorption. Excess solution was further

    removed and the chip washed with PBS after 3 h, after which the chip was ready for use and

    either kept at room temperature in a humid environment for short-term (less than a day) or at 4°C

    for long-term use.

    3.4.7 Detection in Buffer and Media

    Varying concentrations of RANTES from 10 pg·mL-1

    to 10 ng·mL-1

    spiked in PBS were

    incubated with 5 nM specific antibody for one hour after which the solution was added to a 30

    nM signaling probe solution that had previously been conjugated with RANTES. After an hour,

    100 μL of this solution was added to the chip for electrochemical detection. For detection in

    media, PBS was replaced with stem cell culture media.

    3.4.8 Detection in Stem Cell Culture Samples

    A solution of 5 nM specific antibody for RANTES, MDC, or LAP was incubated with separate

    samples drawn from the culture (from days 0, 4, 8, and 12) for an hour. The samples were then

    incubated for an hour with 30 nM signaling probe solution that had been conjugated with the

    specific target to be measured. Finally, 100 μL of the sample was added to the chip for

    electrochemical detection.

    3.4.9 Electrochemical Measurements

    All electrochemical measurements were performed using an EmStatMUX potentiostat

    multiplexer (Palmsens Instruments, Netherlands) and a standard three-electrode setup with an

    Ag/AgCl reference electrode and a Pt counter electrode. For protein detection, the chip was

    scanned using square wave voltammetry