132
A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION SUBMITTED TO THE DEPARTMENT OF MATERIALS SCIENCE AND ENGINEERING AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Sebastian Jeremias Osterfeld December 2009

A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

  • Upload
    others

  • View
    9

  • Download
    1

Embed Size (px)

Citation preview

Page 1: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

A MICROFLUIDIC BIOCHIP

BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES

A DISSERTATION

SUBMITTED TO THE DEPARTMENT OF MATERIALS SCIENCE AND ENGINEERING

AND THE COMMITTEE ON GRADUATE STUDIES

OF STANFORD UNIVERSITY

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Sebastian Jeremias Osterfeld

December 2009

Page 2: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

http://creativecommons.org/licenses/by-nc-sa/3.0/us/

This dissertation is online at: http://purl.stanford.edu/px491tp4561

Includes supplemental files:

1. This file is an open-access publication of some of the magnetic biochip assay results. (SJ

Osterfeld 2008 PNAS Publication.PDF)

2. This file is an open-access publication supplement which shows photos, e.g., of the magnetic

biochip readout hardware. (SJ Osterfeld 2008 PNAS Publication Supplement.PDF)

3. This file is a conference poster detailing the magnetic biochip research progresss in 2006. (SJ

Osterfeld 2006 Conference Poster.pdf)

4. This file is a copy-and-pastable code for Wolfram Mathematica, which calculates the

nanoparticle-sensor interaction a... (SJ Osterfeld 2009 Wolfram Mathematica (R) Code

Example.txt)

© 2010 by Sebastian Jeremias Osterfeld. All Rights Reserved.

Re-distributed by Stanford University under license with the author.

This work is licensed under a Creative Commons Attribution-Noncommercial-Share Alike 3.0 United States License.

ii

Page 3: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Shan Wang, Primary Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Nicholas Melosh

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Robert White

Approved for the Stanford University Committee on Graduate Studies.

Patricia J. Gumport, Vice Provost Graduate Education

This signature page was generated electronically upon submission of this dissertation in electronic format. An original signed hard copy of the signature page is on file inUniversity Archives.

iii

Page 4: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

iv

ABSTRACT

The detection of magnetic nanoparticle (MNP) labels is a promising alternative

to optical detection of fluorescent labels in biomolecular assays, in part because MNPs

are not susceptible to pH, bleaching, or autofluorescence, but especially because

microscopic quantities of MNPs can be detected with simple and inexpensive

magnetoresistive sensors such as spin valves. The goal of this dissertation was to

develop and demonstrate a biochip based on this detection principle.

The particular novelty of this work is the extensive demonstration of magnetic

biochips in real assays, the establishment of a compatible microfluidic fabrication

process, and the development of a simple mathematical model which explains the

experimentally observed signal scaling trends.

Process challenges included finding a sufficiently durable ultra-thin biosensor

passivation and developing a fabrication process that is compatible with the delicate

nature of spin valve sensors, which cannot withstand high temperatures or corrosive

reagents. For the fluidics, a 30 micron layer of silicone elastomer was affixed to a rigid

glass wafer, thereby combining the advantages of soft lithography microfluidics, such

as low-temperature bonding and conformity, with the high alignment accuracy,

mechanical rigidity, and wafer-level integration that traditionally could only be

achieved with anodically bonded microfluidics.

The resulting open-well and multi-channel fluidic biochips have been validated

in several protein and DNA detection assays. Without employing molecular

amplification, protein detection sensitivities of approximately 1 pg/mL or 5 fM

concentration levels can be easily achieved. Even better performance is anticipated in

the near future as there are many avenues towards additional improvements of the base

technology.

Page 5: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

v

ACKNOWLEDGMENTS

I would like to thank my advisor, Professor Shan X. Wang, whose outstanding

character, intellect, vision, and resourcefulness make him one of the best academic

leaders a student could hope for. The tools, infrastructure, and scientific guidance that

he provided allowed me to be productive and creative in my work, for which I am

truly grateful.

I would also like to thank Professor Robert L. White, whose advice and

enthusiasm have been a source of inspiration for me throughout the years.

I also would like to thank Professor Nick Melosh, who has taken an interest in

my work and kindly agreed to serve on my thesis defense and reading committee. I

also thank Professor Mike McGehee and Professor Joseph Liao for chairing my

dissertation defense.

My friend and co-worker Dr. Heng Yu has my sincere thanks and respect for

working tirelessly on the challenging aspects of the assay biochemistry. My thanks

also go out to Professor Nader Pourmand, who has supported this work with his

experience in assay development, and who was instrumental in getting some of the

assay results published in the Proceedings of the National Academy of Sciences.

At Hitachi’s Global Storage Technology division I would like to thank Dr.

Robert Fontana, Dr. Thomas Boone, Stefan Maat, and Jordan Katine for their interest

and for a great scientific collaboration which resulted in some very important data

presented in Chapter 5, Optimization and Characterization.

I also very much would like to thank my fellow students and coworkers in

Professor Wang’s group for much kindness, scientific collaboration, and many great

discussions and ideas: Drew Hall, Richard Gaster, Mingliang Zhang, Donkoun Lee,

Chris Earhart, Dok Won Lee, Liang Xu, Shu-Jen Han, LiangLiang Li, Wei Hu,

Guanxiong Li, Dong-Woon Shin, Seung-Young Bae, Aihua Fu, and Ai Leen Koh.

Page 6: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

vi

Special thanks also go out to Dr. Robert Wilson, who first suggested to try MACS

nanoparticles, which turned out to work very well.

The Stanford Nanofabrication Facility was also essential to this work, because

the entire biochip fabrication process had been developed in the SNF cleanroom in

long hours. The people who supported me there were Mahnaz Mansourpour, Mary

Tang, and Uli Thumser.

The Materials Science Department and the Geballe Laboratory for Advanced

Materials were my scientific home, and I would like to thank Christina Konjevich, Fi

Verplanke, Jane Edwards, Professor Robert Sinclair, Professor Bruce Clemens,

Professor Dauskardt, and many more, for welcoming me there.

I also would like to thank The Whitaker Foundation, Leonard Shustek through

the Stanford Graduate Fellowship program, the ARCS Foundation, DARPA, and the

National Institutes of Health for generous funding and support.

Stanford University in general has been a wonderful place, and I have a great

amount of respect for all the kind, inspiring and accomplished people here. I am

thankful to be a part of this truly unique place.

And of course, I would like to thank my parents, Dr. Karina Krasomil-

Osterfeld and Dr. Karl-Hermann Osterfeld, and my dear wife Sheryl Lin, very much

for their trust and support throughout this endeavor.

Page 7: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

vii

TABLE OF CONTENTS

List of Figures......................................................................................................................x

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

1.1. Research Aim .........................................................................................................1

1.2. Bioassays ................................................................................................................1

1.2.1. DNA Assays ..................................................................................................3

1.2.2. Protein Assays ...............................................................................................3

1.2.3. Label-Free Bioassay ......................................................................................4

1.2.4. Label-Based Bioassay ...................................................................................5

1.2.5. Homogeneous vs. Heterogeneous Bioassay ..................................................5

1.2.6. Multiplex Bioassay........................................................................................6

1.3. Magnetic Biochips..................................................................................................7

1.3.1. Principle of Operation – Magnetoresistive Sensors ......................................7

1.3.2. Principle of Operation – Nanoparticle Detection ........................................10

1.3.3. Benefits of Magnetic Labeling ....................................................................13

1.3.4. Prior Developments in the Field of Magnetic Biosensors...........................15

Chapter 2. Magnetic Biochip Development ......................................................................19

2.1. Biochip Fabrication Process .................................................................................19

2.1.1. Sensor Passivation .......................................................................................21

2.1.2. Sensor Geometry Development...................................................................23

2.2. Magnetic Nanotags...............................................................................................28

Chapter 3. Fluidic Biochip Development ..........................................................................32

3.1. Rationale for Microfluidics ..................................................................................33

3.2. The Need for a New Microfluidic Fabrication Technology.................................34

3.3. Thin PDMS on a Rigid Support ...........................................................................36

3.4. Dry-Etching and Bonding of PDMS ....................................................................39

3.5. Alignment Tolerant Two-Layer Fluidics..............................................................42

3.6. Packaging and Fluidic Connections .....................................................................44

Page 8: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

viii

3.6.1. Snap-Off Edge Fluidic Connections............................................................46

3.6.2. Backside Port Fluidic Connections .............................................................47

3.7. Microfluidic Measurements..................................................................................51

3.8. Microfluidics Conclusion and Suggested Future Work .......................................54

Chapter 4. Assay Results...................................................................................................57

4.1. Direct-Binding Assay for Interferon-Gamma ......................................................60

4.2. Sandwich Assay for Interferon-Gamma...............................................................62

4.3. Sandwich Assay for Interferon-Gamma in 50% Serum.......................................66

4.4. Standard Curve for hCG in 50% Serum...............................................................67

4.5. hCG Assay Signal Scaling and Dynamic Range..................................................70

4.6. Magnetic Biochip Assay Conclusion ...................................................................72

Chapter 5. Optimization And Characterization .................................................................73

5.1. Development of a Simple 64-Sensor Signal Preamplifier....................................73

5.2. Sensor-to-Sensor Reproducibility ........................................................................77

5.3. Chip-to-Chip Reproducibility...............................................................................79

5.4. Reducing the Impact of Sensor Drift....................................................................80

5.5. Signal Dependence on Nanoparticle Distance .....................................................83

5.6. Signal Dependence on Tickling and Bias Fields..................................................86

5.7. Signal Dependence on Sensor Segment Width ....................................................89

Chapter 6. Mathematical Modeling...................................................................................92

6.1. The Resistance of a Spin-Valve Biosensor ..........................................................93

6.2. The Magnetization of Superparamagnetic Nanoparticles ....................................95

6.3. Effect of Nanoparticle on Sensor Resistance .......................................................96

6.4. Definition of Assay Signal ...................................................................................98

6.5. Model and Experiment – Optimal Tickling Field at Zero Bias............................99

6.6. Model and Experiment – Tickling and Bias Field Dependence.........................100

6.7. Model and Experiment – Sensor Segment Width Dependence..........................101

6.8. Insight Derived from Mathematical Modeling...................................................102

6.9. Mathematical Modeling Conclusion ..................................................................104

Page 9: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

ix

Appendix A – Biochip Fabrication Process at SNF ........................................................105

Appendix B – Temperature Correction ...........................................................................111

Appendix C – Mathematica Code ...................................................................................114

Bibliography ....................................................................................................................115

Page 10: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

x

LIST OF FIGURES

Number Page

Figure 1: Magnetoresistive sensor principle of operation. ...............................................................8

Figure 2: Nanoparticle detection principle of operation.................................................................10

Figure 3: Confocal-like label detection in magnetic assays. ..........................................................13

Figure 4: List of representative research groups developing magnetic biochips. ..........................18

Figure 5: Schematic outline of the biochip fabrication process. ....................................................20

Figure 6: Effectiveness of global-, lead-, and sidewall-passivation. ..............................................23

Figure 7: Sensor geometry evolution. ............................................................................................25

Figure 8: Comparison of water response of 2 kΩ and 40 kΩ biochips. .........................................26

Figure 9: Two actual spin valve sensors with different degrees of segmentation. .........................27

Figure 10: Illustration of three classes of magnetic nanotags. .......................................................28

Figure 11: Comparison of Miltenyi MACS and Immunicon magnetic nanoparticles. ..................30

Figure 12: Three generations of spin-valve sensor fluidic biochips...............................................33

Figure 13: Channel collapse in a PDMS section under compression.............................................36

Figure 14: Thickness dependence of spin-cast PDMS on solvent addition. ..................................37

Figure 15: Example of dry-etched PDMS pattern fidelity. ............................................................40

Figure 16: Microfluidic fabrication procedure. ..............................................................................41

Figure 17: Two-layer fluidics.........................................................................................................43

Figure 18: Three generations of fluidic interconnect technology...................................................44

Figure 19: Photo of wafer-level PDMS microfluidics fabrication. ................................................45

Figure 20: Schematic illustration of “snap-off edge” fluidic interconnects. ..................................46

Figure 21: Schematic illustration of backside port fluidic interconnects. ......................................47

Page 11: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

xi

Figure 22: Fluidic spin-valve sensor biochips with backside port connections. ............................49

Figure 23: Fluidic layout of the 8-fluidic-channel biochip. ...........................................................50

Figure 24: Face-down microfluidic biochip with backside ports during measurement. ................51

Figure 25: First microfluidic measurements. .................................................................................52

Figure 26: Photo of open-well biochip and signal generation schematic.......................................58

Figure 27: Nanoparticle coverage image from scanning electron microscope...............................59

Figure 28: Direct binding interferon-gamma assay........................................................................60

Figure 29: Schematic illustration of magnetic label sandwich immunoassay................................62

Figure 30: Example of real-time data from IFN-γ sandwich assay quantification.........................64

Figure 31: Analyte concentration determines the nanoparticle binding curves. ............................65

Figure 32: IFN-γ sandwich assay in PBS buffer and in 50% serum compared (June 2006)..........66

Figure 33: Signal as a function of hCG concentration in 50% serum. ...........................................67

Figure 34: Offline sensor quantification example. .........................................................................68

Figure 35: Example of nanoparticle amplification.........................................................................69

Figure 36: Effect of nanoparticle amplification on standard curve. ...............................................70

Figure 37: Standard curve for hCG in 50% serum. ........................................................................71

Figure 38: New 64-channel signal preamplifier architecture from late 2007.................................74

Figure 39: Example of dynamic range and channel separation......................................................76

Figure 40: Example of sensor-to-sensor signal reproducibility in multiplex assay. ......................77

Figure 41: Example of chip-to-chip assay reproducibility. ............................................................79

Figure 42: Nanoparticle adsorption followed by nanoparticle release. ..........................................80

Figure 43: Quantification from nanoparticle adsorption vs. nanoparticle release..........................81

Figure 44: Signal vs. sensor-to-nanoparticle distance....................................................................83

Figure 45: Continuous measurement of the average nanoparticle distance. ..................................84

Figure 46: Determination of the optimal tickling and bias field for 1.5 µm sensors......................87

Page 12: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

xii

Figure 47: Schematic illustration of sensor segment width evaluation. .........................................89

Figure 48: Signal and noise dependence on spin valve sensor segment width. .............................90

Figure 49: Experimental observations were explained with a mathematical model. .....................92

Figure 50: The resistance of a spin valve sensor segment..............................................................93

Figure 51: Example of calculated spin valve sensor MR transfer curves. .....................................94

Figure 52: Measured magnetization curve and model for MACS nanoparticles. ..........................95

Figure 53: Mathematical description of the sensor-nanoparticle interaction. ................................96

Figure 54: Model and experiment of two different types of sensors at zero bias field. .................99

Figure 55: Model and experiment of signal dependence on fields...............................................100

Figure 56: Model and experiment of signal dependence on sensor segment width. ....................101

Figure 57: Example of temperature-induced drift in the magnetoresistive sideband signal ........111

Figure 58: The centertone (sense current) drift can indeed be used to correct the signal ............112

Page 13: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

1

CHAPTER 1. INTRODUCTION

1.1. RESEARCH AIM

The research efforts described in this thesis aim to significantly expand upon

the results of earlier students, which demonstrated the technical feasibility and

theoretical bioassay potential of magnetic biochips1,2. Specifically, this research work

attempts to analyze the shortcomings that existed before, and to develop appropriate

solutions that improve the ruggedness, manufacturability, performance, and ease of

use of magnetic biochips, to a point where actual analytic bioassays can be carried

out, reproducibly and under realistic conditions, with ease, routine protocol,

acceptable costs, and excellent assay results. Another important aim of this thesis is

the development of a microfluidic magnetic biochip, which is robust and suitable for

mass-production. The effort to develop such a microfluidic magnetic biochip is in

many ways inseparable from the overall optimization effort. Manufacturability and

practicality were important guiding principles at all stages of this work, even if it

meant eschewing solutions which permit good results with a lot of manual work in the

lab, but which are difficult to scale up to mass production.

1.2. BIOASSAYS

Molecular bioassays, which are used to quantify the concentration of specific

biological molecules in a sample, are an important analytical tool in many fields, such

as basic research, medicine, pharmacology, and forensics. While there are plenty of

simple bioassays which measure the concentration of small organic molecules such as

glucose, urea, and creatinine, the term “bioassay” more typically implies measuring

the concentration of complex biological macromolecules such as proteins or particular

segments of DNA. Such advanced bioassays are challenging for several reasons:

These macromolecular analytes in question are often present at very low

concentrations, and furthermore usually not distinguishable from similar molecules by

Page 14: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

2

macroscopic physical measurements like mass spectral analysis or spectroscopy. The

reason, of course, is that the same large number of atoms and molecular bonds in a

biological macromolecule can give rise to many different conformations. Protein

folding is such an example: The final protein obtains its function not primarily from its

constituent atoms and molecular bonds, but from its overall shape and functional

domains.

This means that the goal of a bioassay is to identify and quantify, with a very

high degree of specificity, biological macromolecules on the basis of their shape and

structure. In theory, this could be accomplished with a sufficiently high resolution

microscopic technique, such as electron microscopy. However, aside from the fact that

these techniques would probably denature proteins before they could be identified,

such microscopic techniques suffer from their low throughput: The rate at which

macromolecules could be identified with today’s technology would be so low that it

would be extremely difficult to achieve a representative count at acceptable cost.

For these reasons, the vast majority of today’s bioassays hand off the task of

identifying the macromolecular analyte in question (the target) to other, highly

specialized complementary macromolecules (the probe). A particular probe binds to a

particular target analyte, typically with a very high degree of specificity and affinity.

As a rule of thumb, the specificity and affinity of the target-probe interaction increases

with the size of both macromolecules, and as a result the probes used in bioassays are

usually rather large – ranging from a few tens to more than a hundred kilodaltons, and

usually around 10 – 20 nanometers in size.

The two most common molecular probes are short segments of single-stranded

DNA and a class of proteins called immunoglobulins. While these are very different

types of molecules with very different properties – for example, DNA probes tend to

be very robust and durable, while immunoglobulins tend to be perishable in the open

air – they both serve a single purpose: To recognize a very specific complementary

macromolecule.

Page 15: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

3

1.2.1. DNA ASSAYS

DNA probes, naturally, will reliably bind to complementary DNA strands.

DNA probes of commonly used lengths (25 – 100 base pairs) are also relatively easy

to synthesize and handle, and are predictable in the sense that the ideal probe for a

particular DNA target can be readily inferred, and that the affinity and specificity of

the probe can be reasonably well calculated in advance.

DNA bioassays are commonly used to test if a particular DNA sequence is

present in a sample or not, and occasionally also at what concentration. Because the

genome of an organism is relatively static, gene tests are primarily used for

identification and classification purposes of individuals, species, bacteria, and viruses.

Gene tests can also be used to assess someone’s risk for certain diseases such as breast

cancer3. However, exactly because of the static nature of the genome of most living

tissues not including tumors, DNA assays can usually not determine the current state

of health or disease progression/regression of an organism.

1.2.2. PROTEIN ASSAYS

Protein assays can be more challenging to set up and reproduce than DNA

assays, in part because there can be a multitude of different immunoglobulins (also

called antibodies), all of which can bind to a particular antigen, i.e., a specific

polypeptide, protein, or glycolipid4, with various affinities and specificities. A mixture

of such immunoglobulins for one particular antigen is called a “polyclonal antibody”,

which is what one typically obtains from batch fabrications with variable success. It is

possible to isolate and clone a particular immunoglobulin to obtain a “monoclonal

antibody” which raises costs but which provides a more reproducible performance.

The concentration of certain proteins, for example the blood level of interferon,

can change significantly and quickly with someone’s state of health. When a particular

protein has been positively linked to a certain condition, it is called a disease

biomarker and assumed to have significant diagnostic value. Testing for known

Page 16: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

4

disease markers can be a useful tool in early diagnosis and medical treatment

monitoring, especially when done repeatedly over a period of time. Such proteomic

profiling of perhaps 4-20 biomarkers is expected to be a key to improving the

survival rate of patients with complex diseases such as cancers, autoimmune

disorders, infectious diseases, and cardiovascular diseases5,6,7.

1.2.3. LABEL-FREE BIOASSAY

One concept for utilizing these macromolecular probes (DNA or antibodies) in

an assay is as follows: First, locate the probes and measure a suitable, macroscopically

accessible property of the probes, such as their mass, conductivity, or refractive index.

Second, let the target molecules bind, and re-measure said property of the probes.

Now the mass of the probes should have increased, and their optical and electric

properties should also be different.

When such intrinsic properties of the probe are measured, one typically speaks

of a label-free assay. The challenge lies in the fact that these intrinsic property changes

are difficult to pick up because they are “diluted” by the surroundings of the probes –

the macroscopically measurable parameter is often mostly determined by the support

structure, the surrounding container, liquid, and other molecules, and only to a very

small percentage by the probes themselves. This means, first of all, that a large

number of potential phantom signal sources exist, and secondly that an extremely

sensitive method is needed to detect, for example, the change in mass of the probe –

nevertheless, this has been demonstrated to work, for example with the use of

microscopic tuning forks, or microcantilevers, onto which the probes are

immobilized8. Another technique of label-free detection in bioassays is based on

surface plasmon resonance9, in which the binding of the target to the probe cause a

change in the optical properties of the surface onto which the probes are immobilized.

The two significant advantages of label-free detection are the fact that the

analyte remains unaltered, and that the signal is generated in real-time as the analyte

Page 17: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

5

binds. This allows one to measure the kinetics of the probe-target interaction, such as

the rates of binding and release10.

1.2.4. LABEL-BASED BIOASSAY

Instead of looking for changes in the intrinsic properties of the molecular

probes, as is done in label-free assays, it is also possible to attach a reporter molecule

(also called tag or label) to the target analyte of interest. The benefit of this method is

that the signal obtained from the reporter molecule can often be stronger and more

clearly distinguishable from the background than the intrinsic changes in the probe.

This is especially the case when the reporter molecule has some macroscopically

measurable property that is not typically found in the sample, such as radioactivity.

Using a radioactive reporter molecule would be called radiolabeling. More commonly,

however, an optical dye or fluorescent reporter molecule is attached to the target in

optical labeling. The change in emission or adsorption spectra can then be quantified

with optical systems. A third method, on which the work in this thesis is based, is

magnetic labeling, in which a tiny quantity of magnetic material is attached to the

analyte, which can then be detected with magnetic field sensors.

1.2.5. HOMOGENEOUS VS. HETEROGENEOUS BIOASSAY

One disadvantage of label-based bioassays is the need to distinguish between

labels that are bound to the target, and excess labels that are just floating around. Two

readily apparent solutions to this problem of excess labels are: washing, in which the

bound labels are held in place via the probes, while the excess labels are rinsed away;

or appropriately restricting the volume of observation to just the area of interest, i.e.,

looking only for labels at the surface onto which the probes are immobilized. The

former method requires multiple assay steps and rinsing away of the excess reagents,

which is called a heterogeneous assay. The latter method of restricting the observation

volume to the probes can allow measurements without removing excess reagents,

which is called a homogeneous assay.

Page 18: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

6

A homogeneous assay is potentially much simpler to carry out, since in theory

reagents can just be added one by one. However, the larger number of concurrent

reagents present in label-based homogeneous assays can increase the chance of

unexpected cross-reactions, which could lower the overall specificity of the assay.

1.2.6. MULTIPLEX BIOASSAY

A “singleplex” bioassay employs only one type of molecular probe, which can

specifically recognize just one particular target molecule. If one wants to test for

several different targets, then several singleplex bioassays would need to be carried

out, each isolated in its own reaction volume and needing its own supply of reagents.

This is commonly done in microtiter well plates for example in Enzyme Linked

Immunoassays (ELISA), where each well has one probe.

Multiplex assays, on the other hand, have multiple probes in the same reaction

volume. The probes are usually in different locations, but all are in contact with the

same sample, and sharing all reagents. This results in a tremendous reduction of

reagent consumption and work effort, since a single 20-probe bioassay can

theoretically provide the same data as twenty individual singleplex assays.

This works particularly well with DNA assays, where up to 10,000 different

probes can be used in a single test tube11. In protein assays, the consensus is that a

high level of multiplexity is much more difficult to achieve12, because unexpected

cross-reactions tend to appear in protein assays which degrade the assay results. To

avoid such cross-reactions, multiplex protein assays need to be thoroughly tested,

theoretically with every conceivable analyte combination, which soon becomes

impractical. In general, there seems to be a tradeoff between the multiplexity and

specificity of protein assays, which as a result are usually limited to around a hundred

different probes13,14.

Page 19: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

7

1.3. MAGNETIC BIOCHIPS

Magnetic biochips use magnetic sensors to measure the concentration of

specific analytes in bioassays quickly and inexpensively. The analytes, which often are

DNA segments or proteins, become visible to the magnetic sensors after they have

been tagged with small magnetic labels.

While magnetic biochips are in several ways similar to fluorescent label-based

bioassay chips, the use of magnetic labels can lead to many distinct advantages, such

as better background rejection, no signal fading due to bleaching, simpler and less

expensive hardware, higher sensitivity, real time signal monitoring, and seamless

integration with magnetic separation techniques.

1.3.1. PRINCIPLE OF OPERATION – MAGNETORESISTIVE SENSORS

At the heart of the magnetoresistive (MR) sensor technology stands an

elaborate multilayer thin film. This MR film is deposited layer-by-layer with utmost

care onto a non-conducting substrate wafer, and later divided into individual sensors

by photolithography and ion beam etching. The MR response of these sensors to

magnetic fields is very fast (nanoseconds or less) and is a static function of the

magnetic field strength and orientation. This is an important distinction from inductive

sensors such as pick-up coils, which respond only to changing magnetic fields.

The three types of magnetoresistive elements commonly used in magnetic

biochips are giant magnetoresistive (GMR) multilayer stacks, spin valves (SV), and

magnetic tunnel junctions (MTJ), all of which are examples of spintronic

(magnetoelectronic) sensors, meaning that spin interactions are used to modulate the

electronic properties of the structure15.

Page 20: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

8

R1

Parallel

a.) Low Resistance

H

eR2

Perpendicular

b.) Intermediate Resistance

H

e Reference Layer

Free Layer

Resistive LayerR3

Antiparallel

c.) High Resistance

H

e

Figure 1: Magnetoresistive sensor principle of operation. The overall resistance of a magnetoresistive sensor varies with the degree of alignment of the two magnetic layers that sandwich a nonmagnetic layer. While the magnetization of the reference layer is fixed, the free layer will easily rotate and align itself with the applied magnetic field H. The actual current path depends on the electrical contact points and relative resistance of each layer.

There are many good books available which describe magnetoresistance in

great detail16, so it should suffice to use the spin valve as an example to illustrate the

general concept as shown in Figure 1. As electrons travel through a magnetized

material, they tend to align their spin with the magnetization of the material

surrounding them. If such spin polarized electrons cross an interface and enter a

differently magnetized region, they tend to be scattered, which causes an increase in

the apparent electrical resistance of the overall structure. In Figure 1 electrons emerge

from a magnetic reference layer with a fixed (pinned) magnetization, cross a non-

magnetic layer, and enter a soft magnetic layer with variable magnetization. The

magnetization of this so-called free layer closely follows the direction and magnitude

of the surrounding magnetic field H, while the magnetization of the pinned layer is

largely independent of H. The resistance of the magnetoresistive sensor therefore

depends on the orientation of the applied field, as illustrated in Figure 1.

The nonmagnetic layer helps to decouple the free layer from the pinned layer,

and it is also typically the primary determining factor of the base resistance of a spin

valve device. In spin valves the decoupling layer is usually a noble metal such as

copper or gold, and it transports the bulk of the electrons. As a result, SV films have

low sheet resistances on the order of 20 Ohms per square, which makes them suitable

for in-plane current transport, such as along a simple linear segment. Multiple SV

segments can be connected end-to-end in series to cover a large sensing area. A single

Page 21: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

9

spin valve sensor can thus easily cover an area of about 100 µm in diameter, a size that

is comparable to a typical spot in DNA or protein arrays.

In contrast, MTJs utilize spin-dependent tunneling across a very thin insulating

oxide barrier, and accordingly have a much higher resistivities. As a result, they need

to be patterned into sensor elements with much larger electrical cross-sections, and the

measuring current is run perpendicular to the plane of the film, while spin valves can

be operated either current-in-plane (CIP) or current-perpendicular-to-plane (CPP).

Creating an MTJ sensor that covers a large area is challenging because a single small

defect in the thin but highly resistive tunneling layer can create a pinhole short, which

disables the entire MTJ segment, and the probability of such a defect increases with

the total sensor area. Electrostatic discharge is also a greater risk for MTJs than it is

for SV sensors, where a small defect would have minimal consequences for the

performance of the final device. Additionally, the relatively thick top lead on MTJs,

which is needed to minimize current crowding and the resultant highly localized

tunneling, is a potential complication which might decrease the effective sensitivity of

an MTJ in nanoparticle-sensing experiments, but with a careful design of the top

electrode shape it is possible to detect 10 nm sized particles 17.

For the work in this thesis, the first concern was to have an MR sensor that is

able to cover a large measuring area, because the resulting larger number of sampled

sites will reduce the stochastic noise in low concentration measurements, where

binding events are widely scattered and sporadic. Furthermore, for development work

it is desirable to select an MR sensor which is defect-tolerant, has low noise, and

which exhibits good baseline signal stability over time, which is important for

quantitative analytic assays which typically take several minutes. Considering these

requirements for large area coverage, defect tolerance, low noise, and signal stability,

a spin valve sensor with synthetic antiferromagnetic pinning was chosen for the

magnetic biochip on which this thesis is based.

Page 22: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

10

Externally Applied Magnetic Field Ht

Particle Stray Field

Particle Magnetization M

SN

SN

SN

Pinned Layer

Free Layer

Ht

Hb

SN SN

A B

Figure 2: Nanoparticle detection principle of operation. To generate a detectable magnetic moment, superparamagnetic nanoparticles require the application of an external magnetic field Ht (A). On actual sensors, up to two magnetic fields are externally applied, a time-varying tickling field Ht and a static free layer stabilizing bias field Hb. For details see Chapter 6.

1.3.2. PRINCIPLE OF OPERATION – NANOPARTICLE DETECTION

The MR thin film on which the work in this thesis is based is a spin valve (SV)

structure which at room temperature can achieve a magnetoresistance of ∆R/Rmin =

12%. A simple linear stripe of this SV thin film can be used as the actual sensing

element on a magnetic biochip. In a very simplistic thought experiment, a miniature

permanent magnet could be used to label a biological molecule of interest. If this

molecule then attaches to the sensor, for example due to a specific binding reaction, a

small change in sensor resistance could be registered.

In reality, using miniature permanent magnets as labels would not be feasible,

since the labels would tend to steadily attract each other, just like real magnets would,

until they eventually would cluster and precipitate, largely canceling out each other’s

field in the process. Stabilizing surfactants would probably be insufficient to prevent

such magnetic aggregation of permanently magnetized labels.

To prevent aggregation of the magnetic nanoparticles, the magnetic labels

which are actually used are so small that their individual magnetization is weak and

continuously randomized by the thermal energy at room temperature. The resulting

time-average of zero net magnetic moment is called superparamagnetism. Materials

Page 23: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

11

which are ferromagnetic in bulk are generally superparamagnetic in particle form with

diameters below ca. 10 nanometers.

Unlike a permanent magnet, superparamagnetic nanoparticles would be very

difficult if not impossible to detect directly with a spin valve sensor due to their lack

of a discernible magnetic field. So to generate a detectable magnetic signal from a

collection of superparamagnetic labels, a magnetic polarizing field, or “tickling field”

Ht, is externally applied to the nanoparticles as shown in Figure 2a. The tickling field

stabilizes the magnetic moment of the superparamagnetic labels, making them act like

permanently magnetized labels, which then are easy to detect. The magnetic tickling

field Ht can be alternated at a particular frequency ωHt, which also allows for the

possibility of frequency-based detection schemes such as narrowband detection.

To distinguish induced currents (electromotive forces, EMF) from the sensor’s

MR signal, it is furthermore possible to use an alternating sense current at a frequency

ωisense, which is amplitude modulated by the MR sensor at frequency ωHt. The

resulting modulated sense current contains two AM sidebands at frequencies ωisense ±

ωHt with amplitudes which are primarily a function of the MR effect and sense

current, but not of the EMF. In a typical setup (see also Figure 26d), the sense current

would have a frequency of ωisense = 500 Hz, while the tickling field would have an

amplitude of 80 Oe (rms) and a frequency of ωHt = 208 Hz. This would mean that the

EMF signal is contained in the 208 Hz band, while the actual sensor signal can be

found at 292 Hz and 708 Hz. This makes high signal to noise ratios possible in this

magnetic nanoparticle detection scheme18. On the other hand, if a DC sense current

had been used, i.e., ωisense = 0 Hz, then both the EMF and the sensor signal would be

found at 150 Hz, and the measurement would be less precise.

As shown in Figure 2b, an external magnetic bias field Hb is also applied to

the sensor and nanoparticles. Hb is a static field of ca. 50 Oe, and its purpose is to

reduce the sensor noise to acceptable levels by providing a default orientation for the

sensor’s free layer when Ht transitions through zero.

Page 24: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

12

One complication in this scheme is that the AC tickling Ht field creates a very

strong signal in the MR sensors, which can be regarded as the signal baseline. The

tickling field Ht is minimally altered in the immediate vicinity of a magnetic particle,

which on binding to the MR sensor induces a small deviation from the MR sensor’s

signal baseline. It is this small deviation from the signal baseline which constitutes the

magnetic label signal. In relative terms, the nanoparticle signal in actual experiments

is equivalent to a sensor resistance change of a few tens to a few hundred parts per

million, while the baseline signal is roughly equivalent to a 5% - 10% resistance

change.

With a reasonable degree of circuit complexity (see Chapter 5), an array of

such sensors can be read out by successively polling individual sensors for the relevant

frequency components at ωisense ± ωHt. A typical sensor polling durations is 1 second,

which results in a frequency resolution bandwidth of 1 Hz. An 8-channel ADC card

(NI PCI-6281) makes it possible to have an aggregate polling rate of 8 sensors per

second. Additionally, ωisense could be different for various sensors (frequency

multiplexing), which permits simultaneous measurement of multiple sensors with one

acquisition channel. This increases the polling rate further. For example, towards the

end of this work, a 2-frequency, 8-channel data acquisition system had been

established with a cumulative polling rate of 16 sensors per second and 1 Hz

bandwidth.

Page 25: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

13

Magnetoresistive Sensor

dNS

dNoise Floor

Observation

Volume

Observation Volume

Out of Range

Signal

d min d max

(1/d)3

Figure 3: Confocal-like label detection in magnetic assays. The signal from a magnetic label drops off significantly as the separation d between the label and sensor increases beyond a few hundred nanometers. This results in an observation volume which encompasses primarily surface-bound magnetic labels, while the background signal from distant magnetic labels beyond is relatively small. In this work, the closest practically attainable separation dmin is ca. 100 nm, and labels beyond dmax of ca. 500 nm tend to be indistinguishable from the noise floor.

1.3.3. BENEFITS OF MAGNETIC LABELING

Magnetic biochips are expected to have several technological advantages when

compared to more traditional fluorescence-based biochips. One of the most important

advantages may be the very small background signal, which stems from the fact that

ordinary assay ingredients and biological samples have usually no magnetic signal

sources. Another important advantage is the extreme simplicity of the hardware and

signal transduction pathway: With just a small, simple, inexpensive stripe of spin-

valve film, the surface concentration of magnetic labels is directly translated into a

linearly proportional19 electrical signal. The signal transduction pathway is also

immune from other common sources of measurement error, such as signal fading due

to label degradation (photobleaching in optical systems), chemical changes such as pH

or osmolarity, or changes in opacity and level of autofluorescence. The effect of

temperature drift is also easily accounted for, as shown in Appendix B – Temperature

Correction.

The required instrumentation (chip reader) is also simple, inexpensive, and

very suitable for miniaturization – reducing it to the size of a USB memory stick

Page 26: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

14

seems entirely feasible. Furthermore, magnetically labeled molecules can also be

manipulated and extracted with magnetic fields, for example to pre-concentrate certain

analytes, which might work particularly well in combination with microfluidics.

Another important benefit of the magnetic labeling scheme is the dependence

of the signal on distance. To a first order approximation, the signal induced in an MR

sensor by a properly oriented magnetic label would be approximately proportional to

1/d3 (where d is the sensor-to-label distance), i.e., the signal attenuation with distance

would be that of a simple dipole field. Theoretical calculations20 predict that in some

circumstances there is an optimal sensor-to-label distance of ca. 70nm, however in

actual bioassays in this work (see Chapter 5, Optimization and Characterization), the

closest attainable distances were 120 nm or more, so that this prediction could not be

tested.

The rapid signal attenuation with distance leads to a very limited observation

volume around the magnetic sensors, as shown in Figure 3. Because of the finite

observation volume, properly designed MR sensors are ideal for detecting surface-

bound labels. Unbound magnetic labels, if they are adequately stable in suspension,

will remain largely outside the observation volume, and will therefore not interfere

with the detection of surface-bound labels, which are very close to the sensor. Simply

put, the rejection of the background signal from excess labels is very high – so high

that excess labels may not need to be removed, which means that homogeneous assays

can be performed. This is an important advantage of magnetic labeling over optical

labeling methods, where much more complex equipment would be needed to achieve a

similar effect, for example with confocal microscopy.

Page 27: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

15

1.3.4. PRIOR DEVELOPMENTS IN THE FIELD OF MAGNETIC BIOSENSORS

It appears that the use of magnetic nanoparticles as labels in immunoassays

was first reported in 1997 by Kötitz et al. who used a superconducting quantum

interference device (SQUID) to detect the binding of antibodies21. While their

experiment was successful, it was performed in a magnetically shielded room, and the

SQUID magnetometer required cooling with liquid helium.

At around the same time, giant magnetoresistive (GMR) stacks22, and spin

valves (SV), which had been introduced in hard disk drives as read head sensors23 in

1995, were reaching sufficiently high performance levels at room temperature to

become suitable for magnetic biochips. Modern spin valve read heads are sensitive

and stable enough to detect magnetic data bits from a hard disk at temperatures up to

about 100 °C. Each magnetic bit typically contains a few hundred cobalt alloy

magnetic nanoparticles, but the spin valve sensors in hard disk drives operate at very

high frequencies (up to ~500 MHz) and benefit from the high signal modulation rate

which is beyond the 1/f noise range of the detection process. This advantage is absent

in biological detection assays, where the magnetic fluctuations that need to be detected

occur much more slowly. On one hand, slow changes permit longer sampling times

and correspondingly a better resolution of the absolute signal level, but on the other

hand, this also means that the requirements with respect to 1/f noise, interference,

drift, and long-term measurement stability are much more stringent when GMR and

spin valve sensors are used on biochips.

One of the earliest papers on biomagnetic detection assays using GMR sensors

was published in 1998 by Baselt et al. with a research group at the US Naval Research

Laboratory (NRL). Their bead array counter (BARC) chip was able to detect a single

2.8 µm diameter polystyrene bead containing dispersed maghemite24, albeit in a dry

state. Their data showed that the signal to noise ratio improved significantly as the

sensor width was decreased from 20 µm to 5 µm. Due to its potential for

miniaturization, Edelstein et al. later proposed the BARC sensor for use in a portable

Page 28: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

16

detector for biological warfare agents25. In this paper, the NRL group also

demonstrated the application of a magnetic force to manipulate the magnetic beads

and improve the assay outcome. In 2003 the same group, using a multi-segment GMR

sensor, measured a signal change resulting from biologically bound 2.8 µm beads in

an aqueous solution. However, the binding event could not be recorded in real-time,

apparently because the application of the tickling field that magnetizes the beads

would also lead to clustering of the particles and hence obscure the natural binding

process26.

Graham et al. and a group based in Lisbon, Portugal, were probably the first to

publish real-time magnetic label capture curves, using a short single-segment SV

sensor, and a magnetic gradient to concentrate the particles in the vicinity of the

sensor. The biological signal was obtained by comparing the GMR signals before and

after washing off the nonspecifically bound magnetic particles. They also reported

particle clustering problems with 400 nm high magnetic content particles, which were

however resolved through the use of 2 µm lower magnetic content microspheres27.

A direct performance comparison of magnetic biochips with a fluorescent

detection method for DNA hybridization was first carried out by Schotter et al. with a

research group in Bielefeld, Germany, who defined the relative sensitivity of each

assay as the signal ratio between positive probes and negative probes, the latter of

which generate only the signal from nonspecific adsorption. The conclusion of this

group was that the performance of the magnetic detection method was superior to the

fluorescent method, primarily because at low concentrations the fluorescent method

had a higher background signal level28, which may stem from autofluorescence of the

negative probes.

Our research group at Stanford University, California, is one of the first to

focus on truly nanometer-sized magnetic labels. Unlike other groups which mostly

used particles that ranged from 200 nm to 3 µm, at Stanford the original aim had been

to develop a biochip based on high-moment monodisperse 11 nm diameter Co

nanoparticles29 and 16 nm diameter Fe3O4 nanoparticles19. To advance this approach

Page 29: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

17

of using very small nanoparticles, the feasibility of using very thin passivation layers

was first evaluated using 4 nm of tantalum oxide29. Another distinction is the early

adaptation of spin valve sensors with line widths below two micrometers. In an earlier

implementation, such sensors with widths of 0.2 µm have already been shown to

detect a few tens of said 16 nm particles in a dry-environment before-and-after capture

experiment30.

GMR spin valve sensors have remained the dominant read head technology in

hard disk drives until roughly 2005, when magnetic tunnel junctions (MTJ) began

replacing GMR spin valve sensors in hard disk drives. However, whether MTJs will

also become the primary sensors in magnetic biochips remains to be seen – after all,

the different requirements in biological applications such as the need for low drift and

data collection over a large binding surface may well favor GMR spin valve sensors

over MTJs, which are better suited for highly localized measurements. On the other

hand, MTJs have significantly larger magnetoresistances and may have better

corrosion resistances than the all-metallic spin valves. In an early example, MTJs were

being used for real-time detection of 2.8 µm beads in an aqueous solution, albeit

without biological binding events31.

Several originally academic research efforts in magnetic biochips have

attracted commercial interests. The NRL group joined forces with NVE Corporation

and more recently Seahawk Biosystems Corporation to advance the development of

the BARC sensor. The IST group in Portugal collaborates with Micro Magnetics Inc.

Similarly, the good results of the research group at Stanford has led to the

formation of MagArray Inc., which pursues commercialization of magnetic biochips

for medical and research uses. On the side of corporate research, Philips Research in

the Netherlands has published research articles about their development of magnetic

biochips for use in point-of-care diagnostic medical devices.

Page 30: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

18

Institutionand Site

Principal

Investigators

Magnetic

Particles

Sensor

Technology

Sensor

Passivation

NRL, Washington

NVE, Eden Prairie

Whitman, LJ

Tondra, M

Dynal M280

2.8 µm

GMR, Multi-Segment

1.6 x 8000 µm, 42 kΩ

Si3N4

250 nm

IST,

Lisbon, Portugal

Ferreira, HA

Freitas, PP

Nanomag-D

250 nm

SV, Single-Segment

2.5 x 100 µm, 1 kΩ

Al2O3/SiO2

100/200 nm

University of Bielefeld, Germany

Reiss, G Brueckl, H

Bangs CM01N 350 nm

GMR, Spiral 1 x 1800 µm, 12 kΩ

SiO2

100 nm

Stanford University,Stanford

Wang, SX

Pourmand, N

Miltenyi MACS

40 nm

SV, Multi-Segment

1.5 x 2800 µm, 45 kΩ

SiO2/Si3N4/SiO2

20/20/20 nm

Brown University,Providence

Xiao, G

Dynal M280 2.8 µm

MTJ, Ellipse Patch 2 x 6 µm, 142 Ω

Au/SiO2 200/200 nm

Philips Research,

Netherlands

Prins, M Ademtech

300 nm

GMR, Gradiometer

3 x 100 µm, 250 Ω est.

Unknown

>1000 nm est.

GMR

SV

MTJ

= GMR Stack

= Spin Valve

= Magnetic Tunnel Junction

Figure 4: List of representative research groups developing magnetic biochips. Basic design parameters are listed which can be used to estimate the theoretical performance limit of each platform. However, in practice the performance of a magnetic biochip may depend on additional factors such as the choice of binding chemistry, modulation, and signal processing.

Some representative research groups which are actively developing magnetic

biochips are listed in Figure 4. This list is necessarily not complete, but it can serve as

a starting point for a further literature search. Some of the basic parameters of each

group’s platform are also included, but it should be noted that each group typically

evaluates several different designs at any given time. The great variety of sensors and

nanoparticles under investigation is a reflection of the ongoing active development in

the field of magnetic biochips, in which definite design guidelines had not yet been

established.

Page 31: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

19

CHAPTER 2. MAGNETIC BIOCHIP DEVELOPMENT

At the beginning of this work, the magnetic biochip development at Stanford

had not yet produced actual data from real bioassays. The sensors had to be measured

before and after the magnetic nanoparticles had been captured, which made it more

difficult to identify and eliminate data from malfunctioning sensors. Therefore, the

first task was to achieve a reproducible functionality of the magnetic biochip in actual

assays.

2.1. BIOCHIP FABRICATION PROCESS

As part of this thesis, the magnetic biochip fabrication process was refined

multiple times, and the final process, dimensions, and material choices will be

described and reasoned in this chapter. For better communication, the general process

is outlined immediately in Figure 5. Starting with a magnetoresistive spin valve film

composed of nanometer-scale metallic layers (e.g., Ta 5 / seed layer 2 / IrMn 8 / CoFe

2 / Ru 0.9 / CoFe 2 / Cu 2.3 / CoFe 1.5 / Ta 3, all thicknesses in nm) on a non-

conductive or passivated silicon support wafer (1), large portions of the MR film are

removed by ion beam milling, leaving only the patterned sensors remaining (2). A

typical sensor consists of several linear MR segments, each of which is ca. 0.75 x 100

micrometers in size. The removed portions of the MR film are then replaced with an

insulating oxide. This backfill step (3) planarizes the wafer surface and passivates the

easily corroded sides of the MR sensor. Corrosion-resistant leads (e.g., Ta 5 / Au 300 /

Ta 5, all thicknesses in nm), which electrically connect the sensor with the external

world, are then put in place (4). A thin global oxide passivation (e.g., SiO2 15 / Si3N4

15 / SiO2 15, all thicknesses in nm) is then applied to the entire wafer (5), but even

though it is globally applied, its primary purpose is the encapsulation of the sensor,

which it protects from the corrosive assay reagents.

Page 32: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

20

Insulating Substrate

1. Substrate 2. Pattern sensor from film

3. Backfill / sidewall passivation 4. Apply conductive leads

5. Apply thin global passivation 6. Apply thick lead passivation

Magnetoresistive Film

critical dimension ca. 750 nm

ActiveArea

Figure 5: Schematic outline of the biochip fabrication process. Proportions are not to scale. Note especially the three different types of passivation: Sidewall passivation, global passivation, and lead passivation.

In a final step (6), a significantly thicker oxide passivation (e.g., SiO2 100 /

Si3N4 100 / SiO2 100, all thicknesses in nm) is applied over the conductive leads while

sparing the area of the sensor. This defines the active area of the sensor, i.e., the

roughly 100 x 100 micrometer central area of the sensor which is only protected by the

thin global passivation. The remaining area of the biochip is non-sensing and protected

by a much thicker and hence more durable passivation.

A detailed description of the fabrication procedure is given in Appendix A –

Biochip Fabrication Process at SNF. However, it should be noted that the exact spin

valve structure and fabrication procedure vary significantly according to what

fabrication equipment and capabilities are available at a given time. For example,

much of the fabrication work in this thesis was carried out at the Stanford

Page 33: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

21

Nanofabrication Facility, which at this time had a certain range of processing

capabilities aimed at 100 mm wafers. As a result, the fabrication process as described

has been developed around the equipment at hand at the SNF, and will probably not be

directly transferable to another wafer foundry.

2.1.1. SENSOR PASSIVATION

In a bioassay it is highly desirable to be able to monitor the sensor signal

continuously. Such continuous real-time measurements can show the dynamics of

label and/or analyte adsorption, from which analyte binding kinetics can be inferred.

More importantly, however, a real-time signal is a very useful troubleshooting tool,

because it can help distinguish between relevant, assay-induced signals, and unwanted

extraneous signal changes from temperature drift, sensor malfunction, and various

other unexpected signal sources. This is not trivial: the real-time signal monitoring

ability revealed that the earliest magnetic biochips generated signal errors in response

to vibration (bad contacts), water exposure (dielectric losses), corrosion, temperature

changes, magnetic pre-conditioning (removal of irregular magnetic domains). Even

light sensitivity (unexplained but limited to a particular wafer) was observed at one

point. These and many other signal errors were identified and eventually removed with

the help of the real-time signal monitoring ability.

However, to achieve real-time measurements it is necessary to passivate the

sensors adequately to withstand the conditions of electrolytic corrosion that are created

when measurement currents and assay liquids are applied at the same time. At the

beginning of this work, the risk of electrolytic corrosion was considered to be

significant enough that the spin valve sensors were only measured before and after the

assay.

Magnetic biochips also face a particular challenge with regards to the thickness

of the sensor passivation. On one hand, the magnetoresistive sensor passivation needs

to be durable enough to minimize leakage currents and prevent sensor corrosion, but

on the other hand, the passivation needs to be as thin as possible. The thinnest possible

Page 34: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

22

passivation will allow the closest possible proximity of the magnetic nanoparticles to

the sensor’s free layer, which will maximize the sensitivity of the finished chip. The

importance of the passivation thickness for chip sensitivity is illustrated schematically

in Chapter 1, Figure 3, where it is shown that the theoretical signal falls off with the

sensor-to-nanoparticle separation cubed.

For that reason, several experiments were conducted to determine how the spin

valve biochips needed to be passivated to permit real-time measurements. Different

passivation materials, passivation thicknesses, and passivation architectures were

tested.

Eventually, three types of passivation (sidewall, global, and leads) were

applied to the spin valve sensors as shown in Figure 5. To maximize the sensitivity,

the initial approach was to use the sensors without the thin global passivation, and to

instead rely on the tantalum capping layer of the spin valve film to provide sensor

topside passivation via the native tantalum oxide. To test the feasibility of this

approach, three wafers with test devices were fabricated, each with a different

combination of passivations: Wafer RA01 had only the sidewall passivation, wafer

RA06 had sidewall and lead passivations, while wafer RA04 further added a global

passivation of 30 nm SiO2.

To simulate the conditions of a simple bioassay, chips from each of these three

wafers were exposed to 1x phosphate buffered saline for an extended period of time,

and the change in resistance recorded as shown in Figure 6. Wafer RA01 fared worst,

showing resistance increases of more than 100%. The addition of the lead passivation

on RA06 reduced the corrosion to only 10% resistance increase, which is a hint that

unpassivated gold leads are probably forming a corrosion-accelerating galvanic couple

with the sensor in the PBS solution (this was later corroborated). Wafer RA04, which

finally adds the global passivation, fared best in this static PBS exposure experiment.

This showed that the global passivation could not be omitted.

Page 35: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

23

% Increase in DC Resistance of Various MagArrayIV S ensorsAfter Immersion in Buffer (1xPBS + 0.01% Tween, pH 7.4)

0.1

1

10

100

1000

0 20 40 60 80 100 120

Sensor Number

Incr

ease

, %

of O

rigin

al V

alue

After 1hr

After 3hrs

Wafer RA01+ Sidewall Pass.- No Lead Pass.- No Global Pass.

Wafer RA06+ Sidewall Pass.+ Lead Pass.- No Global Pass.

Wafer RA04+ Sidewall Pass.+ Lead Pass.+ Global Pass.

Figure 6: Effectiveness of global-, lead-, and sidewall-passivation. All three types of passivation are required for adequate corrosion protection of the magnetic biosensor during prolonged exposure to phosphate buffered saline.

Subsequently, an additional improvement to the corrosion resistance of the

biochip was made by adopting a tri-layer silicon oxide-nitride-oxide (ONO) film

instead of plain SiO2 of the same thickness to passivate the sensors. This ONO film

was reported as a promising passivation structure in other people’s work32,33. In

general, it was found that the spin valve sensors are adequately protected against

corrosion with the addition of 30-50 nanometers this ONO film, meaning that this

passivation permitted continuous (live, real-time) readout of the sensors throughout

the assay, at a sense voltage of 0.5 V, with individual sensor failure rates of roughly

1% in simple actual assays, which is low enough to be inconsequential if redundant

sensors are used in conjunction with continuous sensor quality monitoring

implemented in software.

2.1.2. SENSOR GEOMETRY DEVELOPMENT

When designing an MR sensor for magnetic biochips, the first consideration

should be to find a shape which facilitates orderly magnetic domain formation. Edges

Page 36: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

24

create a local demagnetizing field which favors alignment of the magnetization

parallel to any edges that the sensor has, therefore the default magnetic free layer

orientation would tend to fall along the axis of a linear segment. This shape anisotropy

effectively is a magnetization bias, which stabilizes the free layer and lowers the noise

of the sensor (see also Chapter 6 – Mathematical Modeling). On the other hand, rough

sensor edges, sharp corners, curved sensor segments, or a lack of shape anisotropy can

result in complex magnetic domain structures which reduce the linearity and

reproducibility of the sensor. For that reason, simple straight segments with a high

length/width ratio and smooth edges (good photolithography) are the basis of the

magnetic biosensor design.

At the beginning of this thesis, the shape of the magnetic sensor was chosen to

be similar to the designs of earlier students30, which had worked with sensors

consisting of a single linear spin valve segment (e.g., 1 x 2 µm, and 0.3 x 4 µm). To

eliminate the need for Electron-Beam Lithography and to permit the use of the much

faster contact mask optical photolithography equipment at the Stanford

Nanofabrication Facility (resolution limited to around 1 µm), the initial sensor

geometry was scaled up to a single line, 3 x 14 µm in size, as shown in Figure 7a. This

type of sensor also featured a small “gold patch” in its center, which was meant to

provide a means of anchoring capture probes via a special thiol-based binding

chemistry exclusively to the middle of the sensor, where theory predicts the highest

sensitivity17. However, in practice it turned out to be challenging to achieve a reliable

and exclusive localization of the capture probes only on the gold patch. Nanoparticle

binding occurred wherever the capture probe solution had contacted the chip, with

little distinction between the gold patch and the passivation. Furthermore, the gold

patch needed to be even narrower than the sensor, and well aligned, which meant that

the sensor itself needed to be wider than it could have been without a gold patch.

Another problem was the rather low resistance of this type of sensor: The sensor had

100 Ω, while the leads on the chip had ca. 20 Ω, which meant that the effective

magnetoresistance was noticeably reduced in practice.

Page 37: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

25

Figure 7: Sensor geometry evolution. The blue areas indicate the spin valve film, the brown areas indicate the leads, which make electrical connections to the sensor. Sensor design A features a thin gold stripe in its center, intended to be a preferred anchoring site for the capture probes.

Without the ability to chemically localize the capture probe only on the center

of the sensor, the alternative was to spot the entire sensor and its surroundings with a

capture probe. In this scenario, the small size of the first sensor didn’t match well with

the spot size that DNA array printers could readily achieve, which was ca. 100 µm in

diameter (~7854 µm^2). Most of the spotted capture probe would have been wasted,

possibly even depleting the samples of low-abundance analytes, which would only

have a small (42 µm^2 / 7854 µm^2 = 0.5%) probability of binding on the small

sensor’s area. For that reason, another sensor was soon designed, which had a total

footprint of ca. 100 x 100 µm, as shown in Figure 7b. This sensor was better matched

to the size of a DNA spot, and it was also thought that such a sensor would provide

more reproducible data due to observing and averaging a larger number of capture

events. Indeed, this sensor became the first to result in actual magnetic bioassay data,

and it greatly facilitated systematic technical optimizations and bioassay development.

However, two flaws of the design shown in Figure 7b soon became apparent:

First, the sensor’s high resistance of ca. 40 kΩ, combined with the sensor’s passivation

failure threshold of 2V when immersed in water, meant that the sense current was

necessarily very small. This meant that even very small leakage currents across the

passivation and into the water became noticeable, and signal shifts would occur when

the water was applied and removed from the sensor.

Page 38: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

26

IFN-γ Detection Assay - 300ng/mL2kΩ Chip SJO7-WD1-9-4, May-18-2007 - Raw Data Minus In itial Value

-5

0

5

10

15

20

0 5 10 15 20 25 30 35 40

Time, Minutes

Sig

nal A

mpl

itude

, µV

Anti-IFN-γ

Anti-IFN-γ

BSA 10%

BSA 10%

Signal w/o Wash - Start

Signal w/o Wash - End

Signal with Wash - Start

Signal with Wash - End

VA

C MACS Nanoparticles in PBS

H2O

H2O

PB

S

PB

S

PB

S

PB

S

IFN-γ Detection Assay - 300ng/mL40kΩ Chip RB2-7-6, May-18-2007 - Raw Data Minus Initial Value

-5

0

5

10

15

20

0 5 10 15 20 25 30 35 40

Time, Minutes

Sig

nal A

mpl

itude

, µV

Anti-IFN-γ

Anti-IFN-γ

BSA 10%

BSA 10%

Signal w/o Wash - Start

Signal w/o Wash - End

Signal with Wash - Start

Signal with Wash - End

VA

C MACS Nanoparticles in PBS

H2O

H2O

PB

S

PB

S

PB

S

PB

S

Figure 8: Comparison of water response of 2 kΩ and 40 kΩ biochips. The 40 kΩ biochip, due to its very high impedance and sensitivity to leakage currents, picks up a lot of signal swing when water is applied and removed (left). The revised design with a lower resistance sensor is insensitive to the application of water (right).

These “water signals” were usually on the order of ca. 10 µVrms (implying a

total passivation impedance of ca. 70 MΩ at 500 Hz, see footnote1), comparable to the

actual assay signals. Another lesser design flaw was that the patches of lead material,

which shunt out the irregular magnetic domains where spin valve segments join, were

rather small and closely spaced (see Figure 7b), and difficult to manufacture reliably.

Both of these issues were resolved with the final sensor geometry revision,

which reduces the number of SV segment joints to just five, resulting in larger shunts

and a significantly simpler lead layer design shown in Figure 7c. More importantly, by

connecting the SV segments in parallel and in series, a total sensor resistance of ca.

2.4 kΩ was realized. The lower resistance of latest sensor design allows a higher sense

current, which in turn lessens the impact of current leakage through the passivation,

which takes place when the sensor is immersed. As a result, by changing the sensor

resistance from 40 kΩ to 2 kΩ, the “water signals” are reduced from formerly 10

µVrms to theoretically 0.6 µVrms – in practice, the water signals are no longer

noticeable, as can be seen in actual data in Figure 8.

1 Assume the sensor is operated at 7 %MR, sense voltage (centertone) is 1 Vrms at

500 Hz, and magnetic field frequency is 200 Hz. The sideband at 700 Hz (our signal) is then 1*0.07/4 = 17.5 mVrms. To cause a change in the sideband of 10 µVrms, the centertone needs to change by 10/0.07*4 = 571 µVrms. This is 571 PPM of 1 Vrms. A bypass impedance of 70 MΩ changes a 40 kΩ sensor by 571 PPM.

Page 39: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

27

Figure 9: Two actual spin valve sensors with different degrees of segmentation. The latest sensor design allows changing of the spin valve segment width without altering any other parameters of the sensor – for example, the total area and electrical resistance of the two sensors shown are identical.

Another intended feature of the latest sensor design is that the spin valve

segment width can be easily changed, without changing the overall area, resistance, or

sense current density in the spin valve layer, and without having to re-design the

electrical leads. For example, two spin valve segments of 3 µm in width can be

replaced by four spin valve segments which each are 1.5 µm wide but occupy

essentially the same footprint. This is illustrated with an actual chip in Figure 9, where

the left sensor has 12 segments, each 3.0 µm wide, while the right sensor has 24

segments, each 1.5 µm wide. This flexibility in sensor design can be exploited to put

comparable sensors, which differ only in their segment width, on the same biochip.

This makes it possible to vary and study the spin valve segment width (see also

Chapter 5, Optimization and Characterization). By placing different but comparable

sensors in close proximity, one can be certain that the effects of assay chip-to-chip

variability, magnetic field non-uniformity, etc., will be minimized.

An unresolved question in the sensor design is what the impact of the gold

patch shown in Figure 7a would be, if such a fine degree of selective surface

functionalization could be reliably achieved in real assays. Would the sensors be more

sensitive if nanoparticles would bind just to the center of the spin valve segments, as

theoretical considerations suggest? Additional work is needed before this question can

be tested in actual experiments.

Page 40: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

28

Organic polymer (dextran) Magnetic material (Fe3O4) Non-magnetic material (Ru)

A B C

50 nm 16 nm 100 nm

40 nm

14 nm

Figure 10: Illustration of three classes of magnetic nanotags. In this work, the most frequently used magnetic nanotags consisted of an irregular-shaped organic matrix with several small clusters of magnetic material embedded within (A). Monolithic spherical magnetic nanoparticles with an organic surface coating (B) and disc-shaped, lithographically produced multilayer antiferromagnetic nanoparticles (C) may offer better magnetic performance.

2.2. MAGNETIC NANOTAGS

As mentioned earlier, superparamagnetic nanoparticles (also called nanotags)

are used to avoid clustering and precipitation. Furthermore, using nanotags with a

small diameter enhances their rate of diffusion and helps limit the sensor’s observation

volume to surface-bound nanotags only. While nanotags should also have as high a

magnetic moment as possible, their most important performance-determining

characteristics are probably their surface functionalization and colloidal stability.

Particle precipitation must not occur at any rate, because it would lead to a continuous

rise in the signal baseline that could obscure the equilibration of the binding reactions,

especially at low concentrations. Similarly, it is important that the surface

functionalization of the particles leads to highly selective and strong binding reactions

so that the molecules of interest are labeled exclusively and irreversibly, while other

areas need to remain particle-free.

In practice, it is surprisingly difficult to create nanotags which simultaneously

fulfill the requirements of small size, high magnetic moment, suspension stability, and

excellent binding selectivity. Prior to this work, magnetic biosensors used large

magnetic particles for proof of concept which ranged in size from ~250 nm to

3 µm24,28,34,35.

Page 41: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

29

These micron-sized labels are not optimal for biomolecular assays – they

diffuse slowly, are prone to magnetic interaction and subsequent precipitation, and are

very bulky compared to the analyte molecules. Truly nanometer-sized magnetic labels

where therefore evaluated soon after, for example 16 nm nanotags, which were nearly

completely metallic to ensure a strong magnetic signal17. Synthetic antiferromagnetic

nanotags36,37, which were produced by imprint lithography and which offer a tightly

controlled size and custom-tailored magnetic properties, were also tested. An

illustrative overview of the different classes of nanotags investigated in this thesis is

given in Figure 10.

One general observation was that magnetic nanoparticles, despite being stable

in suspension in storage, would often start to precipitate when used on the biochip.

The probable reason for this precipitation is that the nanoparticles, once exposed to the

magnetic tickling field, attract each other much more strongly and form clusters which

are no longer stable in suspension. The telltale signature of such nanoparticle

precipitation can only be observed if the sensor can be measured continuously (which

is one of the achievements of this thesis) while the reagents are applied: It is a steady,

linear increase in the signal, which starts and stops with the nanoparticle incubation,

and which is independent of the biochemistry. A second often observed shortcoming

of various nanoparticles was a lack of binding selectivity. This could be tested by

functionalizing a biochip with two surface functionalizations in close proximity: one

which should bind the nanoparticles via specific molecular interactions – such as

streptavidin on the nanoparticles and biotin on the sensors – while other nearby

sensors functionalized with a suitable non-particle-binding protein, such as bovine

serum albumin (BSA), would serve as a control. A high particle binding selectivity

would result in a large difference in particle coverage between the two types of

functionalizations.

Page 42: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

30

Immunicon vs. MACSTested on 600 pM HCG Assay, Sept-15-2007

0

5

10

15

20

25

30

35

40

MACS MACS Control Immunicon Immunicon Control

Sig

nal,

µV

Median: 22.7 Median: 0.8 Median: 23.5 Median: 2.5

MACS SNR: 29 dB Immunicon SNR: 19 dB

Figure 11: Comparison of Miltenyi MACS and Immunicon magnetic nanoparticles. Both nanoparticles generated similar signal levels, but Immunicon particles had more non-specific binding to the biochip, as revealed by the control sensors.

In Figure 11, two types of nanoparticles are compared in such a binding

selectivity experiment. Two biochips were identically prepared, each with 16 hCG-

specific sensors and 16 control sensors blocked with BSA solution. An actual bioassay

(as described in Chapter 4 – Assay Results) was then carried out, using a 600 pM hCG

solution as the analyte. To quantify the signal, one chip was exposed to 45 nm

streptavidin-functionalized MACS nanoparticles from Miltenyi, which are irregular-

shaped and have a structure as illustrated in Figure 10a. The other chip was exposed to

70 nm streptavidin-functionalized nanoparticles from Immunicon, which have a

structure as illustrated in Figure 10b. The signal levels after five minutes of

nanoparticle exposure are shown.

Both nanoparticles worked well, generating comparable signal levels on the

hCG-specific sensors. However, the control signal, which ideally should be zero, was

significantly higher with the Immunicon nanoparticles, probably due to a higher level

of non-specific binding. As a result, the MACS particles resulted in a ca. 28:1 signal to

background ratio (29 dB), while Immunicon particles only achieved a 9.4:1 signal to

background ratio (19 dB).

Page 43: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

31

At the beginning of this work, MACS nanoparticles, which are estimated to

have only a small amount of magnetic material embedded in a largely organic matrix,

were thought to be too weak for use as magnetic nanotags. Eventually it was found

that these particles did actually generate useful signals on the magnetic biochip, in part

probably due to the very thin sensor passivation used in this thesis. Furthermore,

MACS nanoparticles are small – 35 to 50 nm in overall diameter depending on which

method is used to measure their diameter – which is a reasonable size for labeling

biological molecules. Due to their reliable suspension stability in the presence of an

applied magnetic field, and due to their consistent binding selectivity, MACS particles

eventually became the nanotags of choice for the work in this thesis, despite having a

somewhat lower magnetic signal than several other available magnetic nanotags38.

An important but unexpected side benefit of using small and magnetically

weak nanotags is that the spin valve sensors will only detect them when they are very

close – such as when they are attached the sensor surface. Due to their small size and

good suspension stability, the signal contribution from unbound nanotags is negligible.

Washing steps, which are typically required to remove unbound but strongly signal-

generating labels, can be omitted.

From actual experiments it seems that the sensor’s detection range for MACS

nanoparticles is indeed small enough to encompass only surface-bound nanotags (the

sensor’s detection range will be estimated more precisely in Chapter 5, Optimization

and Modeling). In practice, this suppression of unbound labels means that the true

amount of currently bound nanotags can be observed in real-time, and that negative

control sensors experience no signal changes during nanotag application and removal.

Simple one-step homogeneous assays with no washing steps are thus a possibility.

Page 44: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

32

CHAPTER 3. FLUIDIC BIOCHIP DEVELOPMENT

A simple question was raised in the early stages of the magnetic biochip

development: How can one ensure that the sample, which one wants to analyze, is

maximally interacting with the sensor elements? For example in the case of a limited

number of available analyte molecules, any adsorption of analyte molecules to the

reaction chamber walls would reduce the number of analyte molecules that can still

interact with the sensor elements. This loss of analyte would manifests itself as a loss

of analytical sensitivity. So the question is, if analyte depletion is a concern, how can

one minimize the loss of analyte to non-sensing surface areas?

One immediate idea would be to alter the surface chemistry of the non-sensing

areas so as to adsorb as little of the analyte molecules as possible, e.g., by coating

these areas with anti-biofouling molecules such as poly(ethylene glycol) chains39 or a

blocking protein like bovine serum albumin (BSA). However, in that case one would

have to ensure that only the non-sensing areas are modified to prevent analyte

adsorption, and that the actual sensor elements maintain their affinity. Achieving this

kind of locally differential functionalization is not trivial and often works less reliably

than expected, especially if the non-sensing area is large, and the sensing area

comparatively small.

Another approach might be to confine the sample physically to the location

where it is needed, i.e., right above the sensor. This could be achieved with a channel

that flows the sample across the sensor, which would have the additional benefit of

providing a means of forced sample agitation, which should further help to maximize

the sensor-sample interaction and hence the analytical sensitivity. Such “microfluidic”

sample handling systems are already widely established in other academic projects,

often even with significant complexity40, and a simple microfluidic sample handling

system might be very useful on a magnetic biochip.

Page 45: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

33

Figure 12: Three generations of spin-valve sensor fluidic biochips. Left: The earliest fluidic biochips contained a single fluidic channel, accessible from the chip edge, and short-segment spin valve sensors. Center: The second generation fluidic biochips contained a single fluidic channel, also accessible from the chip edge, which branched out to cover 64 large multi-segment spin valve sensors. Right: The latest generation fluidic biochip contained eight separate fluidic channels, accessible from twelve backside ports, and 64 large multi-segment spin valve sensors. In all three chip designs, the bond pads to make electrical connections are on the left and right sides of the chip.

3.1. RATIONALE FOR MICROFLUIDICS

The original motivation behind the development of microfluidics for the

magnetic biochip was the notion that a microfluidic assay should have better

sensitivity than a well-based assay, for reasons of better sample confinement and

better sample agitation, both of which would reduce the analyte diffusion distances

and maximize the sensor-sample interaction.

However, there are several additional potential benefits of moving from a well-

based assay to a microfluidic assay. The laminar flow of reagents in a microfludic

channel is a potential benefit, since it might result in more predictable, reproducible

interactions between the liquid reagents and the solid surfaces. Washing steps, for

Page 46: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

34

example, could be carried out with a prescribed and uniform shear force at the channel

walls, which should be an improvement over the turbulent and non-uniform manual

washing procedure in well-based assays. A fluidic biochip would also have the

additional advantage of permitting the sample and reagent handling to be automated

with relatively modest hardware. Automation of the bioassay should further increase

the reproducibility of the results over manually pipetted assays. The closed reaction

compartment of a microfluidic biochip would also better prevent the contamination or

uncontrolled evaporation of reagents in response to room humidity fluctuations, which

can be a source of variability in open-well chips. Lastly, microfluidics can also be

used to compartmentalize a biochip into several isolated reaction volumes, while still

retaining a very small chip size. Reaction compartmentalization is often an important

strategy in highly multiplexed protein bioassays, where cross-reactivity between the

various antibodies can severely degrade the sensitivity and specificity of an assay.

3.2. THE NEED FOR A NEW MICROFLUIDIC FABRICATION TECHNOLOGY

Several existing methods of fabricating microfluidic channels were evaluated

and deemed unsuitable for use on a magnetic biochip for various reasons. The most

durable and chemically inert microfluidic channels would probably consist of silicon

or silicon dioxide, and indeed such channels can be fabricated on a separate wafer and

fused to the sensor wafer with one of several wafer-to-wafer bonding methods41.

However, even so-called low-temperature bonding methods42 employ temperatures of

at least 200°C, and most wafer bonding methods use temperatures far higher than that.

The temperature requirements alone ruled out the majority of these methods, because

they would almost certainly destroy the spin valve sensors in the process.

Additionally, several wafer bonding methods require that the mating surfaces be

perfectly clean and flat, which would be very difficult to achieve on the magnetic

biosensor chip.

An interesting alternative was therefore the so-called “soft lithography”, where

channels are fabricated by casting a slab of polydimethyl siloxane (PDMS) on a mold

Page 47: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

35

template43. PDMS has two distinct advantages, one being that it is an elastomer, which

will conform to surface irregularities without reflowing. The other advantage is that

the PDMS surface can be chemically activated with a brief oxygen plasma exposure,

after which it adheres spontaneously to clean oxide surfaces such as the spin valve

sensor wafer, even forming permanent bonds under the right conditions.

However, the traditional PDMS-based microfluidic fabrication method has

several drawbacks, the most serious one probably being that the PDMS slab is sloppy

– at larger sizes it does not maintain its dimensions very well, and it can not be

handled like a regular, rigid wafer. This means that an accurate registration of the

entire PDMS slab to the entire sensor wafer is nearly impossible. To get around this

flaw, the PDMS slab is typically made almost 1000x thicker than the channels that it

provides, and then cut up into many individual pieces, which are aligned one-by-one

to the finished biochips as needed. The obvious downside is that this approach is very

manual and is most likely not suitable for large-scale production of biochips in an

industrial setting.

Since PDMS microfluidics with their low-temperature elastic sealing abilities

seemed like a great fit for the spin valve biochip in principle, the plan was to try to

develop an improved type of PDMS microfluidics which would also permit wafer-like

handling and alignment accuracies. Not only would this eliminate the chip-by-chip

PDMS attachment, but it would open up the possibility for wafer-based, possibly

large-scale industrial PDMS microfluidic fabrication.

Page 48: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

36

The PDMS (blue) deforms when compressed between two rigid wafers (grey). Qualitatively shown are Van Mises stresses, which indicate the overall shear force concentration. Blue indicates zero, and red indicates a large positive Van Mises stress.

255

0

Figure 13: Channel collapse in a PDMS section under compression. A thick section of PDMS (left) is more likely to buckle and collapse a fluidic channel than a thin section of PDMS (right).

3.3. THIN PDMS ON A RIGID SUPPORT

To make a wafer-like handling and alignment accuracy possible for the very

flaccid PDMS material, the first idea was to add a rigid support layer to the PDMS.

Instead of just curing the PDMS material on the channel mold template, the PDMS

was covered with a glass wafer, to which it remained attached after curing and

separation from the mold. This approach was initially pursued with Dow Corning

Sylgard 184 PDMS polymer, 10:1 volume mixing ratio of base to curing agent.

However, two problems became apparent as this approach was pursued. First, the

separation of the PDMS from the mold was rather difficult, since the rigid-backed

PDMS could no longer be “peeled” off the mold, as it was done in traditional PDMS-

based fluidic fabrication. Second, channels fabricated this way were observed to

collapse easily with the slightest application of pressure from above. This was

especially a problem with very shallow channels which were supposed to be roughly

half a micrometer in height and roughly five micrometers wide – such shallow, wide

channels would often spontaneously collapse, with the PDMS sticking to the substrate

where the channel was supposed to be.

To remedy this problem and to make even very shallow channels possible, we

set a goal to make the PDMS layer thinner, comparable in thickness to the height of

the intended fluidic channel – a channel collapse should then be impossible. A finite

element simulation, shown in Figure 13, supported this notion.

Page 49: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

37

Sylgard 184 Spin Casting Thickness

at 2500 rpm, 60 seconds

0.1

1

10

100

0.1 1 10

Parts Solvent per Part PDMS (vol)

Thic

knes

s, µ

m

Xylene or IPA Chloroform

Figure 14: Thickness dependence of spin-cast PDMS on solvent addition.

Unfortunately this further complicated the issue of separating the PDMS from

the mold: The thinner the PDMS layer was cast in the mold, the more difficult it

became separate the cured and rigidly supported PDMS from the mold afterwards.

Additionally, the high viscosity of the uncured PDMS made it impossible to squeeze

the PDMS thinner than ca. 100 µm between the mold and supporting cover wafer.

Since it seemed impossible to pattern very thin PDMS layers by molding,

another method of PDMS patterning was sought. A literature search revealed that

other researchers had successfully patterned cured PDMS films with dry etching

methods44,45. However, initial attempts to implement the reported method weren’t

satisfactory: The spin-cast PDMS was still relatively thick, and attainable etch rates

did not realistically permit full-depth dry etching of PDMS above 50 µm in thickness.

To obtain thinner PDMS films by spin-casting, the uncured PDMS was thinned

with organic solvents which would later evaporate during the curing of the PDMS on a

hot plate. Several PDMS spin-casting tests with various solvent loading were

conducted with the results summarized in Figure 14. Chloroform evaporated too

quickly to be an effective thinning solvent during the spin-casting. Xylene worked

Page 50: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

38

very well as a PDMS thinning agent, resulting in 20 µm thick PDMS films at ca. 10

vol-% solvent loading, and permitting thicknesses down to 300 nm with high solvent

loading. Eventually, xylene was replaced by isopropanol alcohol, which was an

equally effective PDMS thinner at the thicknesses of interest while being less

hazardous.

Another reason for the use of a solvent thinner was that spin-cast PDMS films

would often have several defects in the form of raised protrusions. This was a

significant problem in later attempts to bond the PDMS wafer to the sensor wafer,

because these PDMS protrusions would prevent the wafers from making full and

uniform contact during the bonding step. However, with the addition of the thinning

solvent, the PDMS viscosity was lowered just enough to permit filtration of the mixed

but uncured PDMS prior to spin-casting. The filtration step was very effective in

reducing the PDMS film irregularities and yielded visually perfect PDMS films most

of the time. Due to the still significant viscosity of the uncured thinned PDMS

polymer, most 1 µm pore size syringe filters would clog almost immediately. 10 µm

and larger pore size syringe filters would often not resolve or even worsen the PDMS

film defect issue. Eventually, the optimum filter appeared to be a 5 µm pore size,

32 mm diameter syringe filter with a Supor® membrane, which was strong enough to

permit filtration of ca. 5 mL per filter before clogging, which was the amount needed

to coat one 4” wafer. Centrifugation of the PDMS prior to spin-casting removed

trapped air bubbles and further enhanced the PDMS film quality.

Page 51: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

39

3.4. DRY-ETCHING AND BONDING OF PDMS

Starting with low-defect 30 µm thick films of Sylgard 184 PDMS spin-cast and

cured on 4” Pyrex wafers, the patterning of PDMS by dry etching was optimized next.

A 150 nm aluminum film was sputtered onto the PDMS as a hard mask. The hard

mask itself was patterned with photolithography and ion beam milling, which yielded

a much improved pattern fidelity compared to wet etching of the hard mask.

Complications in this step were the fact that the soft PDMS film would yield if the

photoresist was baked with the normal parameters, which created defects in the hard

mask that would later be transferred to the PDMS. The solution was to bake the

photoresist only at low temperatures before exposure, and not at all after development.

This way, defect-free patterned hard masks with the required etch selectivity could be

obtained on PDMS. Frequently the aluminum hard mask would also develop

characteristic surface wrinkles, which however did not interfere with the PDMS

etching and patterning resolution.

With the hard mask in place, extended dry etching times which exceeded the

durability of the photoresist were then tested in Drytek 1 at the SNF. The originally

reported dry etching recipe with a mixed O2-CF4 plasma did not result in a satisfactory

etch profile, possibly due to equipment-specific issues. Instead, an SF6-CHClF2 plasma

etch gave much better anisotropy with near etch vertical profiles. After several

variations, it was determined that in Drytek 1 the best results were obtained with a 105

W plasma at 200 mTorr, using 100 sccm of SF6 as a source of fluorine and 150 sccm

of CHClF2 (Freon® 22) to passivate the sidewalls and obtain better anisotropy. The

PDMS etch rate was rather low at approximately 100 nm/min, but it was sufficient that

the 30 µm PDMS film could be etched cleanly all the way down to the supporting

glass wafer in a few hours. This optimized recipe resulted in very good pattern fidelity

and near vertical channel sidewalls, as shown in Figure 15.

Page 52: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

40

Figure 15: Example of dry-etched PDMS pattern fidelity. The optimized and highly anisotropic PDMS etching recipe is used to create a fluidic filtration grate with pillars that are 2 µm in diameter and ca. 10 µm tall (left). The PDMS is etched all the way down to the supporting glass wafer, and is later covered with another glass wafer to create a 10 µm high fluidic channel (bottom right).

After the completion of the dry etch, the photoresist was usually no longer

present, but the 150 nm aluminum hard mask was still in place. It was conveniently

removed by exposing the PDMS wafer to TMAH-based developer for about two

minutes. The PDMS was then thoroughly cleaned with acetone, methanol,

isopropanol, and extended sonication in distilled water. Following that, the PDMS

surface was first baked to remove moisture and solvent traces and then activated by

exposure to an O2 plasma in Drytek 1 (500 mTorr, 100 sccm O2, 85 W, 30 seconds).

In this condition, the PDMS surface was hydrophilic and ready for bonding to

another oxide surface which had been similarly cleaned and activated with an O2

plasma. Accurate alignment of the PDMS wafer to the spin-valve wafer was then

achieved on the Electronic Visions 620 Aligner using the wafer-to-wafer alignment

chuck and viewing the mating surface through the transparent PDMS and support

wafer. Final bond strength was achieved by compression and, optionally, by heating of

the wafer sandwich to 100°C for 2 hours. The entire fluidic wafer fabrication

procedure is illustrated in Figure 16.

Page 53: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

41

1.) Spin Cast PDMS onto Wafer

• Use Dow Corning Sylgard® 184 PDMS (Polydimethyl Siloxane)

• Add isopropanol to reduce the viscosity for filtering and spin casting

• Use 20:2:3 of Base : Curing Agent : IPAby volume, mix thoroughly

• De-bubble in centrifuge, filter through 5 µm Supor®32 mm syringe filter

• Spin cast at 2000 – 3000 rpm, 45 seconds, thickness ca. 20 – 30 µm

• Heat cure at 150°C for 30 minutes (longer is also o.k., 120-160°C also o.k.)

Glass (Pyrex, Quartz) Wafer, 6” diameter

PDMS, 30 µm

3.) Coat Photoresist over Hard Mask & Develop

• Standard Photoresist process can be used

• Expose features with sizes of ca. 10 µm critical dimension or larger

• Develop as usual, but do not use any post-develop bake

• A post-develop bake will cause cracks in the resist and hard mask!

Glass (Pyrex, Quartz) Wafer, 6” diameter

PDMS 30 µm

Aluminum 150 nm

Resist Resist

5.) Transfer Pattern into Silicone Elastomer

• Dry Etch: 100 sccm SF6, 150 sccm CHClF2 (Freon 22), 200 mTorr, 105 Watts

• Photoresist will usually disintegrate after ca. 1 hour (hence use hard mask)

• Etch rate ca. 100 nm / min (5 hours), sidewall angle ca. 90°

Glass (Pyrex, Quartz) Wafer, 6” diameter

PDMS

Aluminum Aluminum

PDMS

SF6 Plasma

7.) Clean and Activate PDMS

• Solvent clean with Acetone, Methanol, IPA

• Sonicate in DI water for extended time

• Finally, activate PDMS with 30-second 100W RF O2 plasma

Glass (Pyrex, Quartz) Wafer, 6” diameter

PDMS PDMS

O2 plasma O2 plasma

2.) Sputter Al Hard Mask onto PDMS

• Ca. 150 nm of Al recommended

• Any sputtering method is acceptable

• Temperature is not a problem up to 300°C• Interesting effect: Aluminum film will

often form surface wrinkles on PDMS

• Wrinkles as described in J. Mech. Phys. Solids, ZY Huang, W Hong, Z Suo, 2005

Glass (Pyrex, Quartz) Wafer, 6” diameter

PDMS, 30 µm

Aluminum 150 nm

4.) Transfer Pattern into Hard Mask

• Ideally done by Ion Beam Milling for best pattern fidelity

• Alternative: Wet Etch with Aluminum Etchant or TMAH developer

Glass (Pyrex, Quartz) Wafer, 6” diameter

PDMS 30 µm

Aluminum

Resist Resist

Aluminum

Aluminum Etch

6.) Remove Hard Mask

• Immerse wafer in Aluminum Etchant to strip Hard Mask

• Alternatively, TMAH-based developer will also strip the Al Hard Mask

Glass (Pyrex, Quartz) Wafer, 6” diameter

PDMS PDMS

Aluminum Aluminum

8.) Align and Bond PDMS to Sensor Wafer

• Align PDMS features to sensor wafer

• Lower PDMS wafer into contact with sensor wafer

• Gently apply pressure to seal wafers to each other

• Heat to 100°C for 2 hours to enhance bond strength

Glass (Pyrex, Quartz) Wafer, 6” diameter

PDMS PDMS

6” Silicon Wafer with Sensors

S

. O

ster

feld

, S

tan

ford

Uni

vers

ity,

200

7

Figure 16: Microfluidic fabrication procedure.

Page 54: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

42

The microfluidic channels created in this way are very durable. Since the

microfluidic channels are no longer just a surface feature on a much thicker PDMS

slab, but instead cut across the entire thickness of the 30 µm PDMS film, they can no

longer experience “roof collapse” if external forces were applied to the final

microfluidic device. Similarly, they are much less likely to leak than traditional soft

lithography microfluidics when the channels were pressurized. The transparent glass

wafer on which the PDMS layer is supported becomes a glass top window on the final

biochip. This means that the exterior of the biochip is both transparent and also nearly

as impervious as all-solid fusion-bonded glass microfluidics would be. Yet at the same

time one has all the conveniences associated with traditional soft lithography

microfluidics: The cold-sealing ability, and the compliance of PDMS, which can

accommodate non-planar substrate topographies, and which minimizes the impact of

slight particle contamination on the mating surfaces.

In theory it is even feasible to encapsulate biological molecules in the channels

by applying them first to a properly prepared substrate wafer and then cold-sealing the

PDMS without heat application over this functionalization chemistry. Of course, some

experimentation would be needed to ensure that the substrate wafer chemistry is both

suitable for surface functionalization and PDMS adhesion at the same time.

3.5. ALIGNMENT TOLERANT TWO-LAYER FLUIDICS

A further refinement of the thin PDMS fluidics is the creation of a two-layer

fluidic structure as shown in Figure 17. The PDMS is patterned on the support wafer

as usual, but in addition, a layer of ca. 200 nm thick SiO2 is fabricated on the sensor

wafer by lift-off patterning. This SiO2 layer contains short trenches, for example at the

location of the spin valve sensor, which are later covered by a section of the PDMS

layer. If designed properly, this effectively forms a localized fluidic channel

constriction above the sensor. This is useful if the flow of the analyte needs to be

constricted above the sensor for sensitivity reasons.

Page 55: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

43

Sensor Wafer (silicon)

PDMS Support Wafer (glass)

PDMS PDMS PDMS

SiO2 SiO2Sensor

Large Channelin PDMS

Small Channel in SiO2

Figure 17: Two-layer fluidics. Two large microfluidic channels are patterned in the 20 µm thick PDMS layer. They are joined through a much smaller channel which is patterned in a 300 nm thick SiO2 layer on the substrate wafer. This way a channel constriction is created above the sensor.

Such a structure failed to work with thick PDMS layers of traditional soft

lithography, which would sag and close the constriction, but with the rigid-supported

dry-etched thin PDMS it worked as designed. The benefit of fabricating only a small

local constriction with a 2-layer fluidic architecture is that the overall flow impedance

of the fluidic circuit is much lower, and the overall flow rate and clogging resistance

greater, than if the fluidic circuit had been implemented exclusively in the 200 nm

SiO2 layer.

Another benefit of the two-layer microfluidic structure is that the trenches in

the PDMS and SiO2 layers can be designed to have generous overlap in those areas

where the fluid switches from one layer to the other, thereby making the accurate

alignment of the PDMS to the SiO2 layer less critical. Lastly, the finest, most

alignment-critical channels can be defined exclusively in the SiO2 layer, which as a

standard process will have even better alignment to the substrate than the PDMS layer.

A provisional patent application for this method was filed on Dec. 10, 2004,

eventually resulting in the issuance of US Patent 11388223.

Page 56: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

44

Figure 18: Three generations of fluidic interconnect technology. Left: A prototype fluidic-only single-channel chip with edge connections that interface with an aluminum manifold. Center: An early single-channel fluidic biochip with edge connections that interface with cast epoxy manifolds. Right: The latest design, an 8-fluidic-channel biochip which has fluidic inlet/outlet ports on the backside.

3.6. PACKAGING AND FLUIDIC CONNECTIONS

One of the most challenging aspects of the microfluidic chip development was

the fabrication of the fluidic connections, through which the microfluidic channels can

be connected to the macroscopic world. For a durable, rugged chip which could

potentially be mass-produced, such connections had to withstand rough handling, high

fluid pressures, high temperatures, and a variety of chemicals. This alone is a

challenging task, but the greatest challenge lies in making such a fluidic connection to

the chip in a very small form factor. After all, silicon real estate is expensive, and as

little as possible should be wasted on interconnects, be they electrical of fluidic in

nature. None of these important concerns are usually addressed in traditional soft

lithography microfluidics.

Page 57: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

45

Figure 19: Photo of wafer-level PDMS microfluidics fabrication. The last step of the whole-wafer microfluidic fabrication procedure consists of accurately aligning the completed spin valve sensor wafer (left) to the PDMS-coated glass wafer. This can be done in a regular wafer alignment machine. The aligned wafers are then brought in contact to form a wafer sandwich (right) which then needs to be sawed into multiple individual devices.

Drilling holes in the PDMS support wafer, which would have been equivalent

to the traditional soft lithography approach where holes are simply punched into the

PDMS, was not possible. Such holes would have to be drilled prior to PDMS spin

casting, which would have rendered a good spin casting film nearly impossible.

Alternatively, drilling such holes into the support wafer after spin casting would most

likely ruin and contaminate the fine PDMS structures. A different approach had to be

found.

A further complication was the fact that the fluidic channels were completed

before the wafer had been diced into individual chips. As shown in Figure 19, the last

step of the fluidic wafer fabrication is the alignment and bonding of the sensor

substrate wafer to the PDMS supporting wafer. After this step, the wafer sandwich

needs to be diced into individual chips on a wafer saw, which is a dirty process that

could easily contaminate the chip and render it useless if the cooling water and

sawdust should accidentally enter the fluidic channels. For this reason the wafersaw

must not cut into or open any of the fluidic channels.

Page 58: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

46

1.) 2.) 3.) 4.)

PDMS

PDMS Support Wafer (glass)

Sensor Wafer (silicon)

SiO2

Partial cut

PDMS

PDMS Support Wafer (glass)

Sensor Wafer (silicon)

SiO2

Insert blade and twist edge off

PDMS

PDMS Support Wafer (glass)

Sensor Wafer (silicon)

SiO2

Microfluidic opening

PDMS

PDMS Support Wafer (glass)

Sensor Wafer (silicon)

SiO2

Epoxy

Inlet / Outlet

Microfluidic Manifold

Figure 20: Schematic illustration of “snap-off edge” fluidic interconnects. This was the earliest method of making strong connections to the glass-topped PDMS fluidics. It only requires that a partial cut be made in the PDMS support wafer at the right location. This can be conveniently done prior to actual wafersawing.

3.6.1. SNAP-OFF EDGE FLUIDIC CONNECTIONS

On earliest generation fluidic biochips, the channels terminated near the edge

of the chip under a section of the PDMS support wafer which could later be removed

to reveal the channel entrance. This entirely eliminated the need for drilling or etching

holes into the wafers. To make a section of the PDMS support wafer removable, a

partial cut was made across the biochip with the wafersaw, deep enough to cut most of

the way through the PDMS support wafer, but not deep enough to cut into the sensor

wafer. This formed a designed breakpoint which could be opened when access to the

fluidic channel was needed. To make a permanent fluidic connection to the channel,

these partially removable chip edges were opened in a clean environment, and then

sealed to a matching manifold with a carefully metered amount of epoxy resin.

The “snap-off edge” fluidic connection concept is illustrated in Figure 20 and

shown in actual use in Figure 12 and Figure 18. Several fluidic chips were made with

such connections, which were sturdy, worked as intended, and held up well to high

pressures.

Page 59: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

47

A

Chip Carrier

In Out

Contact Contact

Epoxy

PDMS Fluidics Rigid Cover (Glass)PDMS

Biochip

B

PDMS

Biochip

Chip Carrier

In Out

Contact Contact

PDMS Fluidics Rigid Cover (Glass)

Wirebond

Figure 21: Schematic illustration of backside port fluidic interconnects. This was the later method of making fluidic connections to the biochip. It requires a special chip carrier with drilled holes that line up with the etched backside holes in the biochip. The chip is glued to the carrier with adhesive epoxy. To make sure that the epoxy doesn’t enter the fluidic channels, it is usually best to apply the epoxy indirectly, i.e., by placing the chip in its correct position and letting the epoxy wick (by capillary forces) under the chip from the side.

It could be argued that such fluidic connections at the chip edge are potentially

mass-producible: With the appropriate amount of epoxy resin applied around the

microfluidic chip opening, assembly simply consists of placing the chip into the

microfluidic manifold. However, the most significant drawback to this method is that

despite the relatively small footprint – estimated at ca. 12 mm2 per fluidic connection

– there is only enough room for two of these connections on the 120 mm2 size biochip.

Unless the chip size was increased, this packaging method would yield just a single

microfluidic channel.

3.6.2. BACKSIDE PORT FLUIDIC CONNECTIONS

Another fluidic connection approach was therefore soon developed to achieve

a larger number of fluidic ports. It was based on backside fluidic connections etched

into the biochip, and required a chip carrier with pre-drilled holes that would line up

with the holes in the chip. The concept is relatively simple and illustrated in Figure 21.

This method requires that access holes are etched into the sensor wafer backside at the

end of its manufacturing cycle with deep reactive ion etching (see Figure 19 for a

wafer with such holes).

Page 60: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

48

Care must be taken that the spin valve sensors are not exposed to the etching

gases in this step, however this is usually helped by the native oxide of the spin valve

wafer which, when properly timed, can serve as an etch stop just before the hole

breaks through to the sensitive frontside of the wafer. The remaining oxide film which

formed the etch stop is so thin that it is easily blown out with compressed air or a brief

sonication after the etch. Alternatively, the frontside of the sensor wafer can be

temporarily protected with another wafer during the port etch step.

During wafer sawing, the backside ports of the chips are sealed shut by the

wafer sawing tape. This prevents contamination with sawing debris. To assemble such

a chip, the wafer sawing tape is removed, and the chip is aligned to a chip carrier with

matching pre-drilled holes. The chip is then glued to this carrier, preferably by letting

the epoxy glue wick under the chip, i.e., letting it fill a tiny air gap between the chip

and the carrier by capillary forces, so that only the required amount of epoxy is drawn

in. If done properly, this technique results in no epoxy entering the fluidic pathways.

This worked actually quite well with a gold-coated ceramic chip carrier and fluidic

holes that were ca. 0.8 mm in diameter, and it is also feasible that this step could be

automated with precise epoxy metering.

To complete the biochip, stainless steel tubing segments (needle gauge 21)

were press-fitted into the holes in the ceramic chip carrier and potted with a generous

amount of epoxy resin. After this procedure, the chip can be wirebonded as usual, with

the wirebonds being encapsulated in epoxy afterwards (see Figure 21 and Figure 22).

The result is an extremely well-packaged fluidic biochip which can withstand a lot of

mechanical, thermal, chemical, and hydraulic abuse. This is a very significant

improvement over traditional soft lithography PDMS microfluidics, which often are

used with minimal or no mechanical reinforcement, and which therefore are extremely

sensitive to mechanical disturbances or slight fluidic overpressure, which in traditional

soft lithography fluidics quickly causes leaks or de-bonding of the PDMS from the

sensor chip surface.

Page 61: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

49

Figure 22: Fluidic spin-valve sensor biochips with backside port connections. On the backside each chip has twelve stainless steel tubes, needle gauge 21, supported in a socket of clear epoxy, which connect to the fluidic channels of the biochip. These fluidic ports are extremely robust and yet occupy only minimal real estate on the chip. The biochip itself is also very robust, with the microfluidic channels being protected by a strong 0.5 mm thick glass cover (formerly the PDMS support wafer).

5 mm

30 mm

Page 62: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

50

Air Flow Carries Fluid to DrainAir Flow Carries Fluid to DrainATM VACATM VACATM VAC

3 mm 500 µm

Figure 23: Fluidic layout of the 8-fluidic-channel biochip. Each of the eight channels has an individual inlet, but the outlet ends in a shared drain channel which is open to the atmosphere at one end and attached to a vacuum at the other end. The constant stream of air in this drain channel carries the reagents to a waste carboy.

In addition to their great mechanical strength, the fluidic backside ports are

also much more densely packed. Compared to the snap-of edge fluidic connections,

the backside ports have a 4x smaller footprint, and require ca. 3 mm2 of chip real

estate each. While it would have been possible to accommodate 16 fluidic backside

ports on the existing biochip, which would have been enough for eight completely

isolated fluidic channels, the actual design used only 12 ports as shown in Figure 23.

By using a shared drain line, the waste collection from the chip is simplified and the

required number of fluidic ports is slightly reduced. This design could also have

operated in reverse, flowing a reagent from one port to several of the fluidic channels.

This would have made it possible to functionalize (or program) each fluidic channel

with an individual chemistry first, after which an analyte could have been delivered to

all channels at once to run multiple analysis tasks in parallel.

A compartmentalization of separate analysis tasks makes sense if the analysis

chemistries would interfere with each other if they were in direct contact. This actually

is frequently the case with protein detection chemistries if multiple analytes are to be

detected simultaneously. Beyond a certain level of multiplexity, protein detection

assays will most likely benefit from compartmentalization.

Page 63: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

51

200 µm

30 mm

Figure 24: Face-down microfluidic biochip with backside ports during measurement. Close-up of microfluidic channels made by dry etching PDMS, as seen through clear glass top of the spin valve biochip (right). The channels are approximately 120 µm wide and 30 µm high.

3.7. MICROFLUIDIC MEASUREMENTS

Several measurements were carried out with the microfluidic chips. In a typical

setup, the biochip was inserted face down into the measurement socket inside a

Helmholtz coil as shown in Figure 24. Just two tubes were connected to the chip: One

tube from the chip’s waste drain channel (see Figure 23) to a vacuum pump with a

waste carboy, and the other tube from a particular fluidic channel inlet to a syringe

pump which had been loaded with a few microliters of the desired reagent(s). This of

course meant that only one of the eight fluidic channels on the chip would be utilized,

while the other seven fluidic channels would remain empty until they were utilized in

a later experiment. This one-by-one measurement of fluidic channels was both a

demonstration of the chip compartmentalization – now eight independent experiments

could be run on one chip, even at different times – and it was also necessitated by the

fact that in 2007 the electronics were only able to record four (freely selectable)

sensors of the 64 sensor array.

To avoid switching valves and dead volumes, the reagents were sequentially

loaded into the syringe pump, so that a sequential flow assay could be performed.

Long air bubbles were used to separate the reagents and reduce the intermixing of

reagents as they traveled through the tubing.

Page 64: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

52

First Microfluidic Signal from Magnetic Particles - Real Time DataCa. 5 microliters of Ferridex Iron Particles, May-5 -2007

-40

-20

0

20

40

60

80

100

120

10 15 20 25 30

Time, Minutes

Sig

nal A

mpl

itude

, µV

Sensor 1

Sensor 2

Sensor 3

Sensor 4

Ferridex FlowAir Air

BiotinBSA Test Assay - Real Time Data2kΩ Fluidic Chip SJO7-WD3-8-2, July-02-2007

-10

0

10

20

30

40

50

0 5 10 15 20 25 30 35 40

Time, Minutes

Sig

nal A

mpl

itude

, µV

Sensor 1

Sensor 2

Sensor 3

Sensor 4

Initial Value

Final Value - MACS

Final Value - Water

Flowing MACS Nanoparticles Flowing DI WaterAir

Figure 25: First microfluidic measurements. This experiment verified that a highly concentrated ferrofluid could be detected as it was pumped through the fluidic channel (left). Later experiments demonstrated specific binding of streptavidin-coated nanoparticles to a biotinylated channel (right).

The first fluidic measurements were designed to simply deliver a uniform, non-

specific stimulus to the spin valve sensors via the fluidic channels, which should

ideally elicit an identical response from all sensors that doesn’t depend on chemistry.

For this test, Ferridex® magnetic nanoparticles, which are highly concentrated iron

oxide particles and usually marketed as a non-gadolinium MRI contrast agent, were

pumped at a rate of 1 µL per second through a channel of the biochip. Four sensors,

which was all that the electronics could monitor concurrently at that time, were

recorded. An approximate step function in the signal level to 85 µV was observed,

coinciding with the Ferridex® nanoparticles entering and leaving the sensor area, as

shown in Figure 25a. The step function shape was an indication that no accumulation

was occurring, i.e., there was no settling and continuous addition of particles to the

sensor, instead the signal was mostly just caused by bulk of the Ferridex® particles.

This is in contrast to MACS particles, which are dilute enough that the bulk MACS

fluid does not cause a signal, rather, only the adsorbing and accumulating MACS

particles generate a signal. Also, as the Ferridex® fluid was pushed out of the fluidic

channel by air, a thin layer of the magnetic fluid remained on the channel walls, as can

be seen by the fact that the signal settles down at roughly 16 µV instead of returning to

its initial value (zero) after the bulk of the fluid has passed.

Page 65: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

53

In Figure 25b, a specific nanoparticle binding experiment is shown. Prior to the

measurement, the fluidic channel had been functionalized by first flowing a 2%

polyethyleneamine through the channel, followed by a solution of biotinylated BSA,

which will adsorb onto the entire channel surface and the spin valve sensors.

Streptavidin-coated MACS nanoparticles are entering the fluidic channel at t = 7

minutes and start adsorbing and accumulating on the sensor and channel surfaces due

to the strong biotin-streptavidin affinity. The resulting signal curve shows a very

typical adsorption behavior, with a rapid signal rise in the beginning, which eventually

slows down and levels off as the number of available binding sites approaches an

equilibrium value. Such binding curves appear in many surface adsorption processes

and can be described as a Langmuir adsorption isotherm, or more precisely, as a two-

compartment adsorption process46,47,48.

The specificity of the binding reaction in Figure 25b is corroborated by the

high binding strength, which results in a stable signal steady even as the bulk of the

MACS magnetic nanoparticles is flushed out by air at t = 28 minutes, and furthermore

as the sensor is washed with distilled water beginning at t = 31 minutes. Also

encouraging is the relative similarity of the signals reported by the four recorded

sensors. At the time of this experiment, open-well measurements of multiple sensors

did not usually achieve the same level of uniformity as this microfluidic measurement,

which suggested that the use of microfluidics could indeed make measurements more

reproducible.

Page 66: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

54

3.8. MICROFLUIDICS CONCLUSION AND SUGGESTED FUTURE WORK

The microfluidic fabrication method presented in this work has reached a level

of maturity which permits routine fabrication of very durable microfluidic structures

even on thermally and chemically sensitive devices such as magnetic spin-valve

biochips. The microfluidic fabrication method is entirely implemented on a wafer

level and could be scaled up for mass production. Sturdy high-density fluidic

interconnects, compatible in principle with automated assembly tools, were also

developed. The overall result is a highly compatible manufacturing route for fully

integrated microfluidics, which could be automated to mass-produce microfluidic

biochips which are robust enough to withstand rough handling e.g., outdoors in a

portable device.

However, it is interesting to note that the basic spin-valve biochip, which had

been developed and improved concurrently with the microfluidics, took on a much

larger role in actual assay experiments than the microfluidic version of the same chip.

In fact, assaying experiments carried out on an open-well version of the same basic

chip became so successful (see Chapter 4, Assay Results) that the microfluidic

development eventually slowed down. Why was this the case? The most important

reason might be ease of use and flexibility in handling. For example, several open-well

biochips chips can be quickly sealed with caps and placed in a refrigerator for several

hours. With a microfluidic chip, one would have to seal at least two connections

without accidentally purging the tiny fluidic channels, or one would have to place the

entire fluidic pump and the chip into the refrigerator. Similarly, every reagent

exchange on a microfluidic chip requires loading and programming the fluidic pump

and connecting the chip, while on the open-well chip the same step can be done

directly with a pipette. Lastly, the open-well chip permitted direct access to the

individual sensors. This is a great benefit for individual-sensor biofunctionalization

(e.g., with robotic contactless spotting) and sensor inspection by optical and electron

microscopy.

Page 67: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

55

It appears that the success of the open-well chip is also partially due to the

demands of the ongoing development work, where many parameters are frequently

changed. In such an environment, direct manipulation and flexible handling are more

important than a high degree of automation. Automation requires a great deal of time

to be set up properly, and hence only becomes viable when all design parameters have

been thoroughly optimized and locked in. Additionally, while the microfluidic spin-

valve biochip was potentially suitable for automatic assaying, in actuality it was still

operated in a rather manual fashion: All reagents were manually and sequentially

loaded into a syringe pump – which clearly was far from an automatic process – which

was then connected to the biochip and slowly emptied. This way, the microfluidic

biochip resulted in added complexity without actually reducing the researcher’s

workload – yet.

To better utilize the assay automation potential of a microfluidic biochip, one

would need to develop a sufficiently capable reagent delivery system, which might,

for example, incorporate several multi-position valves to select and deliver various

reagents to the biochip(s) on demand. Similarly, it would be desirable to develop a

means of automatically making and breaking the fluidic connections between such a

reagent delivery system and various fluidic biochips under test. An alternative would

be to develop an on-chip miniaturized reagent handling system, which consisted of

microfluidic valves and on-chip pumps – no small task indeed, especially if true

autonomy from external machinery is the goal.

Furthermore, it seems highly desirable to combine individual-sensor

biofunctionalization with microfluidics. This is not easy: If the microfluidics are

completed first, then one no longer has direct physical access to individual sensor

surfaces. If, on the other hand, the sensors are coated with individual probes first and

then covered with microfluidics, then the additional limitations placed on

manufacturing and assembly are extremely restrictive. In such a case, mechanically

sealed macrofluidics, such as a small flow chamber sealed with an O-ring, are

probably more suitable than bonded mirofluidics. The downside is that mechanically

Page 68: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

56

sealed macrofluidics would most likely be bulkier and less precisely aligned than

prefabricated microfluidics, so that high-density fluidic compartmentalization might

no longer be possible.

Lastly, it also appeared that a slow corrosion process was limiting the useful

shelf life of the fluidic biochips after fabrication, apparently due to a liquid which was

slowly precipitating inside the microfluidic channels. This liquid was probably either

formed or already present in the PDMS at the end of manufacturing and slowly

released over the course of several weeks. Possible candidates are traces of TMAH

developer, uncured PDMS precursors, water, or sulfur-containing residues formed in

the SF6-based dry etching process, which occasionally produced PDMS with a faintly

noticeable hydrogen sulfide smell. To resolve this issue, the microfluidic fabrication

method should be varied to evaluate a purely CF4-based dry etching chemistry. Other

possible solutions may be the elimination of direct contact between TMAH and

PDMS, and using other more fully cross-linked PDMS polymers such as h-PDMS49,50.

Page 69: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

57

CHAPTER 4. ASSAY RESULTS

In the development of the magnetic biochip, an essential challenge was to

optimize the chip and the assay chemistry concurrently: Without a working chip the

assay chemistry can’t be tested, but without a proven chemistry the chip is difficult to

evaluate. To facilitate this interdependent optimization task without the added

complexity of microfluidics, many assay experiments were carried out on an open-

well magnetic biochip of the type shown in Figure 26. A low viscosity two-component

epoxy (EP5340, Eager Plastics, Chicago, IL) was chosen to encapsulate the wirebonds

and attach the reagent well (Tygon® tubing, 1/4” ID x 3/8” OD, 6 mm long) to the

chip.

The chip itself was supported on a ceramic 84-pin chip carrier (LCC08423,

Spectrum Semiconductor Materials, San Jose, CA). A ~0.5 mm layer of the same

epoxy was used to mask some of the sensors. The masked sensors, no longer able to

detect nanotag binding, would serve as electrical signal references. The electronic

signal generation method based on an AC current source was implemented by Shu-Jen

Han18 and capable of recording only four differential signals (sensors) in actual

experiments – a further complication when unproven assays need to be tested on

initially unreliable chips.

To accelerate the early assay development, simple specific binding

experiments were carried out on dummy chips, which were then inspected in a

scanning electron microscope as shown in Figure 27. The question was whether a

reliable correlation could be observed between varying the concentration of a dummy

analyte, such as a biotinylated antibody, and the final density of streptavidin-coated

nanoparticles on the chip surface. The MACS nanoparticles were always applied at the

same (stock) concentration and should, ideally, bind to the surface only as facilitated

by the analyte.

Page 70: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

58

AC Current Source

500 Hz

MR

Sensor

208 Hz

AC Field

80 Oe

AM Modulated

Voltage

Ht

CoilRef.

Differential

Pre-Amplifier

FFT

208 292 708500Hz

µV

SB SB

CT

Figure 26: Photo of open-well biochip and signal generation schematic. The chip has a 200 µL reaction well and is supported by an 84-pin ceramic base (A). At the bottom of the reaction well are 64 sensors in an 8x8 array, some of which are covered with epoxy to provide a reference signal (B). Each sensor has an active area of roughly 90x90 µm2 and consists of 32 linear spin valve segments, each 1.5 µm wide, which are connected in series (C). The signal generation method is shown for a single sensor in (D), where the sensor modulates an alternating 500 Hz current at 208 Hz. A nanoparticle binding event will increase the AM sidebands at 292 Hz and 708 Hz, which are recorded with 1 Hz bandwidths. A reference sensor signal subtraction at the differential amplifier helps reduce drift artifacts.

Page 71: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

59

1000 ng / mL 100 ng / mL

10 ng / mL 1 ng / mL

Figure 27: Nanoparticle coverage image from scanning electron microscope. Before the chip and the assay chemistry were working reliably, assays were sometimes evaluated from the magnetic nanoparticle coverage, which can be assessed in a scanning electron microscope (SEM). In this series of images, varying the concentration of biotinylated antibody did change the final nanoparticle coverage accordingly, which demonstrates that the nanoparticle binding is specific and mediated by the adsorbed analyte. However, due to the small field of view and the irregular, SEM-setting dependent appearance the nanoparticles, this method is only qualitative.

Page 72: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

60

Probe Control Probe Control

c MNT-based Analyte Quantification

IFN-γ (Probe) Biotinylated Anti-IFN-γ (Analyte)IL6-sR (Control) MACS Nanotags

b Analyte Incubation

IFN-γ IL6-sR

Probe Control

a Sensor Functionalization

d Select Real-Time Data from 3 Chips Combined e Results

GMR Sensor

Concentration of Anti-IFN-γ in 1x PBS Buffer

67 pM 670 pM 6.7 nM

Probe Sensor

Control Sensor

0

10

20

30

Time, minutes

∆ S

ign

al, µ

Vrm

s

∆ S

ign

al, µ

Vrm

s

0 5 10 15 20-5

Magnetic Nanotags in PBSAirPBS

0

10

30

20

6.7 nM

670 pM

67 pM

6.7 nM control

10

5.4

25

0.4

10

5.4

25

0.4

Figure 28: Direct binding interferon-gamma assay. This was one of the earliest successful protein assays. Three biochips were used to generate this data, with four sensors being recorded on each chip, for a total of twelve datapoints as shown in the bar graph (May 2006).

4.1. DIRECT-BINDING ASSAY FOR INTERFERON-GAMMA

One of the earliest successful protein assays used a relatively simple assay

chemistry, where the analyte is specially prepared, biotinylated anti-IFN-γ, which is

captured by IFN-γ functionalized sensors and quantified with streptavidin-coated

magnetic nanotags as shown in Figure 28. Interferon gamma is an inflammatory

marker, which for example is released by lymphocytes during an immune system

response. This permits ex-vivo immune response tests, e.g., to reveal past exposure to

tuberculosis51.

Three spin valve biochips were first thoroughly cleaned with acetone,

methanol, and isopropanol, and then further cleaned with a brief exposure to oxygen

plasma. To establish an anchoring layer for the biological surface functionalization,

the chips were exposed to a 2% polyethyleneimine (PEI) solution in water for two

minutes, followed by rinsing with deionized (DI) water, blow-drying with compressed

Page 73: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

61

nitrogen, and baking on a 100°C hotplate for ten minutes. On all three chips, the

exposed (not epoxy-coated) sensors were divided into two groups, a “probe” and

“control” group. The probe sensors were identically functionalized with a 1 µL droplet

of IFN-γ, 100 µg/mL in PBS buffer. Control sensors were functionalized with a 1 µL

droplet of IL6-sR, 100 µg/mL in PBS buffer (Figure 28a). After incubation for 30

minutes at 4 °C the chips were washed with a 1% BSA in 1x PBS buffer solution to

block any remaining non-specific adsorption sites.

The three chips were then incubated for 1.5 hours at 30°C with 100 µL of

different concentrations (0.067, 0.67, and 6.7 nM, i.e., 10, 100, and 1,000 ng/mL) of

analyte, consisting of biotinylated anti-IFN-γ in PBS buffer (Figure 28b). The chips

were then rinsed with 0.1% BSA in TPBS and transferred to the measuring station for

subsequent analyte quantification. To quantify the amount of analyte captured on the

spin valve sensors of a particular chip, the chip well was filled with 100 µL of

Miltenyi MACS nanoparticle stock solution, and the developing nanoparticle binding

signal was recorded (Figure 28c and Figure 28d). The nanoparticle binding signals

after 20 minutes were taken as a measure of analyte concentration (Figure 28e).

The probe sensors, functionalized with IFN-γ, developed strong signals, which

approximately doubled with every 10x increase in analyte concentration. Yet at the

same time, the control sensors had near zero signal, which was a strong indication that

both the binding of the analyte to the probe, and the binding of the nanoparticles to the

analyte, was highly selective. Achieving the required high cumulative specificity of

these two dependent adsorption steps together with proper chip functionality were

significant milestones.

The small error bars in Figure 28e indicate the electrical noise of the

measurement, which is much smaller than the variance of signals from identically

functionalized sensors in this experiment. The sensor-to-sensor variance, which can

stem from local irregularities in the surface chemistry and from physical sensor

variations, is examined in more detail in Chapter 5 - Optimization And

Characterization.

Page 74: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

62

Probe Control

c Linker incubation

Probe Control

d Nanotag-based quantification

Analyte Biotinylated AntibodyBSA

Streptavidin-coated Magnetic Nanotag

Capture Antibody

Probe Control

b Analyte incubation

Capture Antibody BSA

Probe Control

a Sensor functionalization

Magnetoresistive Sensor

Figure 29: Schematic illustration of magnetic label sandwich immunoassay. This figure outlines the use of a sandwich immunoassay for highly sensitive protein quantification with magnetic nanoparticles. Probe sensors are functionalized with capture antibodies specific to the desired analyte, while control sensors are functionalized with a non-matching antibody or protein such as 1 wt% BSA solution (A). During analyte incubation, the probe sensors capture a fraction of the analyte molecules (B). A biotinylated linker antibody is subsequently incubated which binds to the captured analyte (C), and which provides binding sites for the streptavidin-coated magnetic nanoparticles. Streptavidin-coated magnetic nanoparticles are then incubated (D), and the binding signal, which quickly saturates at an analyte concentration-dependent level, is used to quantify the analyte concentration. Image source: Original work by S. Osterfeld. First published in Osterfeld, S.J. et al. “Multiplex protein assays based on real-time magnetic nanotag sensing.” Proceedings of the National Academy of Sciences 105, 20637-20640 (2008).

4.2. SANDWICH ASSAY FOR INTERFERON-GAMMA

One disadvantage of the direct-binding assay (described earlier) is that it

required an analyte which is already biotinylated. This is avoided in a sandwich assay

as shown in Figure 29 where an unmodified analyte is sandwiched between two

specific antibodies, one which anchors the analyte to the sensors, and the other of

which provides a biotin site for magnetic labeling. In addition to being able to quantify

analytes in their natural state, sandwich assays also usually outperform direct-binding

Page 75: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

63

assays in terms of sensitivity and specificity52. This is probably due to the fact that the

analyte incubation, which due to the low concentration of analyte is often the

throughput-limiting step of the assay, can occur faster and more specifically with the

higher diffusion speed of a small, unlabeled analyte. For example, the IFN-γ analyte in

the sandwich assay (Figure 29b) has a molecular weight of ca. 17 kDa, whereas the

biotinylated antibody analyte in the direct assay (Figure 28b) has a molecular weight

of ca. 150 kDa.

After cleaning and PEI coating, several chips were prepared with two

functionalizations: 500 µg/mL Anti-IFN-γ for the probe sensors, and 1 wt% BSA

solution for the control sensors (Figure 29a). Different concentrations of analyte (0.64,

1.9, 6.4, and 64 nM IFN-γ in PBS) were incubated on different chips at room

temperature for one hour (Figure 29b). Following a rinse with 0.1 wt% BSA in PBS,

the chips were then incubated for one hour at room temperature with 100 µL of a

linker antibody solution (Figure 29c), consisting of biotinylated anti-IFN-γ, 2 µg/mL

in PBS. The chips were then rinsed again with 0.1 wt% BSA in PBS, and transferred

to the measuring station for quantification with magnetic nanoparticles (Figure 29d).

Analyte quantification with magnetic nanoparticles was then carried out

according to a precisely timed protocol as shown in Figure 30. After the chip has been

primed a few times with PBS buffer and the stability of the signal baseline has been

confirmed, at time t = 15 minutes, 100 µL of undiluted nanoparticle solution (Miltenyi

MACS, part # 130-048-102) are delivered to the reaction well of the chip and

incubated for 20 minutes at room temperature. At the end of the 20 minute

nanoparticle incubation period, excess particles are removed, and the well is twice

washed with deionized water for 1 minute each. This rinsing step with DI water

removes any salt residues and thereby enables a later assessment of the nanoparticle

coverage in the scanning electron microscope (SEM).

Page 76: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

64

3.6939.2345.1449.19∆ Signal with Wash

4.1040.7542.6552.33∆ Signal w/o Wash

Sensor 4: BSA 1%Sensor 3: Anti-IFN-γSensor 2: Anti-IFN-γSensor 1: Anti-IFN-γ

3.6939.2345.1449.19∆ Signal with Wash

4.1040.7542.6552.33∆ Signal w/o Wash

Sensor 4: BSA 1%Sensor 3: Anti-IFN-γSensor 2: Anti-IFN-γSensor 1: Anti-IFN-γ

IFN-γ Detection Assay - 1 µg/mL (59 nM)Chip RB3-7-3, June-09-2006 - Raw Data Minus Initial Value

-10

0

10

20

30

40

50

0 5 10 15 20 25 30 35 40

Time, Minutes

Sig

nal A

mpl

itude

, µV

Anti-IFN-γ

Anti-IFN-γ

Anti-IFN-γ

BSA 1%

Signal w/o Wash - Start

Signal w/o Wash - End

Signal with Wash - Start

Signal with Wash - End

VA

C MACS Nanoparticles in PBS

H2O

H2O

PB

S

PB

S

PB

S

PB

S

Figure 30: Example of real-time data from IFN-γ sandwich assay quantification. In minutes 4 to 15 some fluid cycling is performed to wash the chip and verify that it doesn’t respond excessively to wet/dry cycling alone. Magnetic nanoparticles were incubated from t = 15 to 35 minutes.

As shown in Figure 30, the final signal can be assessed in two ways, either at t

= 35 minutes, i.e., before the nanoparticle removal and washing procedure, or at t = 42

minutes, i.e., after washing. Since the probe-to-control signal ratio with washing was

not consistently better than the signal without washing, and for matters of

convenience, most assays were soon quantified without a final washing step.

However, the signal levels after the final rinses with DI water are more appropriate for

correlation with nanoparticle coverage densities in SEM images, because such images

can only be obtained on DI-cleaned chips.

Overall the sandwich assay worked as expected, with signals that scaled

reasonably well with the analyte concentration as shown in Figure 31. Occasionally,

however, inexplicable variations were observed, such as for example in Figure 30,

where only two out of three identically functionalized sensors agree well, while a third

sensor reports a 52 / 42 = 24% higher signal.

Page 77: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

65

Time, minutes

∆ S

ign

al, µ

Vrm

s0

0 5 10 15 20-5

50

10

20

30

40

Magnetic Nanoparticle Labels in PBSAirPBS

59 nM Control

(adjacent to probe)

0.59 nM

1.8 nM

5.9 nM

59 nM

Figure 31: Analyte concentration determines the nanoparticle binding curves. This figure shows real-time magnetic nanoparticle binding curves from different IFN-γ sandwich assay chips combined into one figure. The rate of nanoparticle binding is initially rapid, but slows down after a few minutes as the available binding sites are being saturated. Binding site saturation occurs fastest when few binding sites exist to begin with, e.g., when the analyte was only present in a low concentration. Roughly 90% of the specific nanoparticle binding occurs in the first five minutes (June 2006).

Such variations could not always be explained, but were occasionally found to

be the result of experimental error, such as when accidental merging and cross-

contamination of the functionalizations occurred during their incubation step (Figure

29a). Another likely source of variability is uneven drying of these small

functionalization droplets. The resulting protein deposition may be uneven due to

evaporation and convection inside the droplet – this is akin to fluorescent arrays where

DNA spots often take the form of rings, rather than evenly filled circles.

The IFN-γ sandwich assays were the first successful detection of a native (non-

biotinylated) protein analyte on the spin-valve biochips, which is a significant

milestone in actual utility. The sandwich assay chemistry increased the complexity of

the assay over the direct-binding assay by adding another instance of specific

adsorption, for a total of three specific binding events which all need to preserve

specificity in order to obtain good results: First, the specific binding of the analyte

only to the correct capture antibody (Figure 29b). Secondly, the specific binding of the

biotinylated linker only to the captured analyte (Figure 29c). Lastly, the specific

binding of the magnetic nanoparticles only to the biotinylated linker (Figure 29d).

Page 78: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

66

MagArray4 Sandwich Assay - Signal vs. IFN- γ ConcentrationHomogeneous Assay - Datapoints Taken Before Washing

20.3

24.8

15.8

9.6

15.8

14.7 15

.6

12.9 13.6

12.9

0.2

3.7

0.4 1.

7

0.41.

0 2.4

1.3 2.

5

1.3

0

5

10

15

20

25

30

RB3-6-4, 30ng/mL,Buffer

RB3-6-5, 30ng/mL,Buffer

RB3-6-6, 30ng/mL,Serum

RB2-2-6, 30ng/mL,Serum

RB3-6-7, 30ng/mL,Serum

Chip & IFN- γ Concentration

Sig

nal,

uV

Anti-IFN-γAnti-IFN-γAnti-IFN-γBSA 10%

1.2 ± 0.3 µV

1.8 ± 0.8 µV

Negative, Average ± Std.Err.

~10.8 (~20.7 dB)13.4 ± 0.9 µVSERUM

~10.2 (~20.2 dB)18.9 ± 2.3 µVBUFFER

Pos. Sig. / Neg. Sig. (dB)Positive, Average ± Std.Err.

1.2 ± 0.3 µV

1.8 ± 0.8 µV

Negative, Average ± Std.Err.

~10.8 (~20.7 dB)13.4 ± 0.9 µVSERUM

~10.2 (~20.2 dB)18.9 ± 2.3 µVBUFFER

Pos. Sig. / Neg. Sig. (dB)Positive, Average ± Std.Err.

BUFFER SERUM

Figure 32: IFN-γ sandwich assay in PBS buffer and in 50% serum compared (June 2006).

4.3. SANDWICH ASSAY FOR INTERFERON-GAMMA IN 50% SERUM

One concern that was raised in results to the successful sandwich assay was

that the results had been obtained with a single analyte in plain PBS buffer. While

such an experiment is a good proof of concept, real assays would need to quantify

analytes in blood serum, which contains vast numbers of other proteins, many of them

in very high concentrations. A reasonable expectation is that these proteins interfere

with the sensor or the detection chemistry in a way that makes measurements in serum

perform much worse than measurements in PBS buffer.

The IFN-γ detecting sandwich assay was therefore duplicated in both PBS and

50% human blood serum (balance PBS). In all cases IFN-γ was added to the samples

to establish an analyte concentration of 1.8 nM (30 ng/mL). Two measurements were

carried out with PBS, and three measurements with 50% serum. The results are shown

in Figure 32. The serum-based assays, on average, resulted in IFN-γ signals which

were ca. 29% lower than in PBS buffer. However, the signal from the control sensors

was also lowered by 33%, so that the overall probe to control signal ratio was similar.

Page 79: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

67

MagArray Standard Curve - 16 Sensor Median ± 1 StDe vHCG in Serum; Diluted to 1:1 with PBS buffer for me asurement

y = 11.453x0.1981

R2 = 0.9918

10

100

1 10 100 1000 10000 100000

HCG Concentration (in serum, before dilution), mIU/ mL

Sig

nal,

mic

rovo

ltsMACS + 2xAmp

Nonlinear Regression

S17

S18

Figure 33: Signal as a function of hCG concentration in 50% serum. The standard curve was obtained from samples provided by the National Cancer Institute. Two unknown samples S17 and S18 were also measured and interpreted as 8-12 and 500-700 mIU/mL, respectively. Note the exponent of the nonlinear regression (0.1981), which indicates that a ca. 10^5 increase in analyte concentration would result in a ca. 10x larger signal.

4.4. STANDARD CURVE FOR HCG IN 50% SERUM

One of the most carefully planned and executed magnetic biochip assays

performed in the scope of this thesis was the determination of a standard curve for

human chorionic gonadotropin (hCG) with samples which were provided by the

National Institute of Cancer.

It is relatively easy to find high quality antibodies and calibration standards for

hCG, which is typically quantified in international units (IU) per mL. While the

definition of one IU is occasionally readjusted, it appears that 1 mIU/mL is roughly

equal to an hCG concentration of 1.9 pM according to Sturgeon et al53,54. Five

calibration samples were individually measured on magnetic biochips functionalized

with anti-hCG to obtain the standard curve shown in Figure 33. Two unknown

samples S17 and S18 were also quantified and interpreted. This assay series resulted

in positive feedback from the NCI program managers.

Page 80: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

68

Supplementing 4-Channel Real-Time Data with Before & After Off-Line Measurements2kΩ MagArray Chip WD1-3-2, Sept-15-2007

22.7

0

5

10

15

20

25

30

35

0 5 Minutes 15 20 Real-Time Offline Median

Sig

nal,

µV

MACS nanoparticle incubation - 20 min

Figure 34: Offline sensor quantification example. To enhance the accuracy of the hCG measurements, off-line evaluation of additional sensors and nanoparticle amplification were employed. In addition to four real-time sensors, 12 hCG sensors were quantified with before-and-after measurements. This way a 16-sensor median signal value could be obtained, even though the electronics had only 4 independent data channels.

The hCG assay series is noteworthy for several new techniques that were

employed to get better data, and for the insights gained from the assay results. First of

all, the measurements were performed at a time when the readout electronics were still

limited to recording only four sensors at a time. However, since the choice of the

sensors could be manually changed during a live experiment without significantly

disturbing the measurement integrity, additional sensors could be evaluated with some

effort by recording their signals off-line, i.e., before and after nanoparticle incubation.

This of course required a stable measurement setup and a fair amount of certainty that

no unexpected signal artifacts would go unnoticed in the off-line sensors. However,

the before-and-after data obtained this way was indeed roughly comparable in mean

value and variance to the real-time data, as shown in Figure 34. For that reason it was

deemed safe to combine off-line and real-time measurement data into an overall

median signal value and an overall error estimation.

Page 81: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

69

HCG Assay 12.5 mIU Oct-08-2007 - Example of Nanopar ticle Amplification Protocol

-5

0

5

10

15

0 10 20 30 40 50 60

Minutes

Sig

nal,

µV

MACS - 20 min MACS - 20 minLinker Wash Offline Wash

Figure 35: Example of nanoparticle amplification. To enhance the accuracy of the hCG measurements, after the initial nanoparticle adsorption , the incubation of a biotinylated nanoparticle-nanoparticle linker molecule allowed the chip to adsorb a second layer of nanoparticles onto existing nanoparticles, which roughly doubled the assay signal level.

Additionally, as shown in Figure 35, the hCG assays systematically applied a

technique for nanoparticle amplification which was introduced by Heng Yu: After the

initial 20-minute nanoparticle incubation, the chips were rinsed with clean PBS buffer

and then incubated for five minutes with a multiply biotinylated nanoparticle-

nanoparticle linker molecule. This linker molecule would attach to any nanoparticles

on the chip surface and, colloquially speaking, make them “sticky” again towards

additional streptavidin-coated nanoparticles. After the incubation of the biotinylated

linker molecules, a second round of magnetic nanoparticles could be adsorbed onto the

sensors. The resulting secondary binding curves saturated at a level that was ca. 1.5x

to 3.5x higher than the earlier signal levels. This method of amplification was applied

twice on each chip. The result is that the hCG standard curve was obtained in three

forms: One from the original (traditional, unamplified) nanoparticle adsorption, one

from the first round of amplification, and one from the second round of amplification,

as shown in Figure 35. An interesting trend is that the amplification is strongest for the

lowest signal levels, and weakest for the highest signal levels. This means that overall

slope of the signal vs. concentration curve is flattened with each amplification. This

enhances the dynamic range of the nanoparticle quantification, but may increase the

measurement error.

Page 82: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

70

MagArray Standard CurveHCG in Serum; Diluted to 1:1 with PBS buffer for me asurement

y = 1.0641x0.3364

y = 4.4422x0.2554

y = 11.453x 0.1981

1.00

10.00

100.00

1 10 100 1000 10000 100000

HCG Concentration (in serum, before dilution), mIU/ mL

Sig

nal,

mic

rovo

lts

MACSMACS + 1xAmpMACS + 2xAmp#REF!

Figure 36: Effect of nanoparticle amplification on standard curve. The hCG quantifications systematically employed two rounds of nanoparticle amplification, which each increased the signal ca. 1.5x to 3.5x over the previous signal level.

4.5. HCG ASSAY SIGNAL SCALING AND DYNAMIC RANGE

The systematic hCG assay series was the first rigorous experiment to explore

the signal scaling trend and sensitivity limits of the spin-valve biochip. The generally

noticed signal scaling trend of “twice the signal for every 10x increase in analyte

concentration” is more precisely quantified by a non-linear regression fit in Figure 36.

There, the data from the original nanoparticle application fits a power law relationship

of Signal = [conc]^0.3364 which means that for every 10x increase in concentration

the signal increases, on average, by a factor of 10^0.3364 = 2.17x. The benefit of this

signal scaling trend is that a concentration increase by one million results in only a

10^(6*0.3364) = 104x larger electrical signal and nanoparticle coverage (the linear

dependence of the signal on nanoparticle coverage was already demonstrated by

former students, e.g., thesis G. Li30). This means that the magnetic biosensor has a

very large dynamic range of at least six orders of magnitude in theory. Good signal

scaling over four orders of magnitude is already shown in Figure 36.

Page 83: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

71

∆ S

ign

al, µ

Vrm

s

hCG concentration in 50% serum

Standard curve for hCG in 50% serum

1

10

100

1 nM 1 µM1 pM1 fM

Control Signal Range in PBS: 2.5 - 3.5 µVrms

Figure 37: Standard curve for hCG in 50% serum. The detection threshold for the hCG assay was estimated from PBS control assays to be a few fM. However, in actual 50% serum samples, the lowest readings were consistent with a metabolic baseline concentration of 1 pM hCG, which could have been the result of using real serum in the preparation of the samples.

However, the Signal = [conc]^0.3364 scaling relationship is altered if

nanoparticle amplification is used. From Figure 36 we see that if one and two rounds

of nanoparticle amplification are added, then for each 10x increase in analyte

concentration the signal level increases only by 10^0.2554 = 1.80x and 10^0.1981 =

1.58x, respectively. This would, theoretically, increase the sensor’s dynamic range

even more, but at the cost of accuracy. For example, a 10% measurement uncertainty

would translate into a roughly 10^(Log(1.1) / 0.3364) = 33%, 45%, and 62% analyte

concentration uncertainty for zero, one, and two rounds of nanoparticle amplification.

Another interesting aspect of the hCG assay series was the evaluation of the

detection limits. On the upper end of the scale, no signal saturation was detected, and

the standard curve was in good agreement with the power law fit up to the highest

concentration hCG sample (ca. 50 nM). Going to lower concentrations, the signal

scaling was well behaved across four decades, down to a few pM, where the signal

levels seemed to stabilize. This was the case even for serum samples which had no

additional hCG added, and which were labeled as “zero hCG”. Yet when a zero-hCG

assay was performed with plain PBS buffer, a significantly lower background signal

level of only 3 µV (after two rounds of amplification) was obtained. As shown in

Figure 37, this signal background in PBS would suggest that the lower detection

Page 84: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

72

threshold for the hCG assay should be in the femtomolar concentrations, rather than in

the observed pM concentrations. One possible reason for this discrepancy between the

serum and PBS control assays might have been that the “zero hCG” serum sample was

not actually devoid of hCG. In fact, the natural metabolic baseline level for hCG in

real serum is around 1 pM for non-pregnant women according to Korhonen et al55. It

is therefore feasible that in this assay series the lower detection threshold was limited

by the natural metabolic baseline concentration of hCG in human serum.

4.6. MAGNETIC BIOCHIP ASSAY CONCLUSION

The results shown in this chapter are some the earliest true assay experiments

performed on the spin-valve biochip. Previously unobtainable, these results were made

possible by countless hours of sensor and biochip fabrication development at the

Stanford Nanofabrication Facility, the right choice of nanoparticles, lots of electrical

troubleshooting, and by the ingenuity of Heng Yu and Nader Pourmand, who were

developing the biochemistry for these chips. Chip fabrication and assembly was done

by myself, while Heng Yu would usually carry out most of the assay steps on the chip,

with the exception of the nanoparticle-based analyte quantification measurement,

which also was done by myself. Many additional assaying experiments were carried

out successfully later on, with increasingly more people obtaining results on the spin-

valve chip independently. To date, the open-well chips developed in this thesis were

used for assaying experiments by Heng Yu, Richard Gaster, Andy Mak, Liang Xu,

Shu-Jen Han, Drew Hall, and Dokyoon Kim. Their work often pursues much more

advanced assays, which increase the number of concurrent analytes in multiplex

protein detection, which push the limits of sensitivity, and which outperform existing

assaying systems such as ELISA and hence produce truly new scientific biochemistry

data. As a result, several high-profile publications based on this open-well biochip

have been published (in PNAS56,, Nature Medicine57, Biosensors and Bioelectronics58)

or are about to be published59 by S. J. Osterfeld, H. Yu, R. S. Gaster, A. Mak, N.

Pourmand, Drew A. Hall, L. Xu, and thesis advisor S. X. Wang. Additional

publications are likely to follow.

Page 85: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

73

CHAPTER 5. OPTIMIZATION AND CHARACTERIZATION

This chapter will present several additional original developments and

experiments which are significant for the performance and understanding of the spin

valve biochip, but which also occasionally needed time and testing to be fully

understood and adopted.

5.1. DEVELOPMENT OF A SIMPLE 64-SENSOR SIGNAL PREAMPLIFIER

The spin valve biochip, as designed, had 64 individual sensors. Yet the signal

preamplifier was limited to recording only four sensors concurrently, which however

were freely selectable from the array. This was enough to obtain many very

encouraging assay results as shown in Chapter 4, Assay Results, but it left much to be

desired. What was the sensor-to-sensor signal variation across the entire chip? How

many analytes could be detected simultaneously? Was the magnetic tickling field

uniform enough across the chip? And what types of optimizations could be performed

with the statistical power of 64 sensors combined?

As shown in Figure 38, scaling up the old signal generation architecture to 64

channels would have required the use of at least 128 operational amplifiers, one in the

AC current source and in the buffer of each signal pathway. This seemed excessive for

a discrete electronics design. The idea of connecting the 64 sensors in an 8x8 matrix,

with just one external connection per row and column, had been floated around before.

However, it had not been entirely clear (to me) how sufficient signal isolation between

the sensors could be achieved in such a setup. A common node shared among several

sensors would need to be a perfect current sink, such as a connection to ground, to

prevent the sensors from interacting. But how could one measure a signal at a ground

node? The crucial insight came in a December 2007 discussion where Drew Hall

explained the concept of a virtual ground, i.e., an operational amplifier set up in such a

way that it maintained a ground potential at its signal input.

Page 86: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

74

Old signal preamplifier: Applies an AC sense current, and measures the voltage across the spin valve sensors. Requires ca. 128 opamps and at least 65 connections to a discrete 64 sensor biochip.

New signal preamplifier: Applies an AC sense voltage and measures the current in the spin valve sensor. Requires ca. 16 opamps and a total of 16 connections to a discrete 64 sensor biochip.

Demonstration of 64 Sensor Magnetic Biochip Measure ment

53 Sensors with Biotin-BSA, 8 Reference Sensors, 3 N/CS. Osterfeld and H. Yu, July-14-2008

-20

30

80

130

180

1 3 5 7 9 11 13 15Minutes

Sig

nal,

mic

rovo

lts (r

ms)

53 Biotinylated Sensors

8 Reference Sensors

One current source and preamplifier for each sensor

500 Hz1mApp

SV SV SV

500 Hz1mApp

500 Hz1mApp

Vsig 1 Vsig 2 Vsig 3

SV

Etc.

500 Hz1mApp

Vsig 4

Biochip with N sensors and N+1 pins

One current source and preamplifier for each sensor

500 Hz1mApp

SV SV SV

500 Hz1mApp

500 Hz1mApp

Vsig 1 Vsig 2 Vsig 3

SV

Etc.

500 Hz1mApp

Vsig 4

Biochip with N sensors and N+1 pins

One voltage source per row

Virtual GND:0V

SVSV

540 Hz1Vpp

i1 i2i3+i4

Rf

570 Hz1Vpp

~

~

SVSVi3 i4

Vsig 3+4

Biochip with N sensors and 2*N0.5 pins

One transimpedance amplifier per colum

-+i1+i2

Rf

Vsig 1+2-+

One voltage source per row

Virtual GND:0V

SVSV

540 Hz1Vpp

i1 i2i3+i4

Rf

570 Hz1Vpp

~

~

SVSVi3 i4

Vsig 3+4

Biochip with N sensors and 2*N0.5 pins

One transimpedance amplifier per colum

-+-+i1+i2

Rf

Vsig 1+2-+-+

Figure 38: New 64-channel signal preamplifier architecture from late 2007. A new preamplifier design was developed to permit a complete readout of all 64 sensors on the biochip with a realistic part count and significantly fewer connections per chip. This resulted in the first ever 64-sensor magnetic biochip measurement being performed in July of 2008.

Page 87: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

75

Such an amplifier could perform what seemed previously inconceivable: It

would extract useful information from a “ground” node shared by multiple sensors,

effectively by performing a linear addition of the individual signals. Despite being

summed at this common node, the individual signals would remain recoverable if each

of the summed sensor signals had a unique frequency. This would amount to

frequency domain multiplexing, as shown in Figure 38 with two AC voltage sources at

540 Hz and 570 Hz. Alternatively or in conjunction, these voltage sources could be

turned on and off sequentially, which would amount to time domain multiplexing.

This suddenly permitted that all 64 sensors be connected into a passive array

with just 16 external connections. This would permit smaller, more economic

biochips, and it would make building a 64-sensor preamplifier simple enough that it

could be done without too much effort. Additionally, the new signal preamplifier

could be built with much fewer discrete components, in part because the passive

sensor array could now be driven with AC voltage sources instead of the more

complex AC current sources needed in the old design.

The new preamplifier circuit architecture was immediately promoted for actual

implementation, but nevertheless, it took some time to gain the attention it deserved. It

was not until July of 2008 that such an amplifier had been built (by myself), which

allowed the first 64-sensor magnetic biochip measurement (Figure 38) to be

performed. The 64-sensor signal preamplifier made many important experiments

possible, such as the detection of eight analytes at 1 pg/mL (ca. 50 fM) concentrations

in the presence of a much larger signal from adjacent biotinylated sensors. This is

shown in Figure 39, where the signal separation was well maintained in the new

amplifier design, as signal ratios of up to 100:1 were easily recorded without any

evidence of electronic cross-talk. Due to the signal scaling relationship (Sig. =

[conc.]^0.3364, see Chapter 4), the 100:1 signal ratio shows that analytes with

concentration differences of ca. one million to one could be measured on the same

chip.

Page 88: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

76

Detection of Multiple Analytes, Each at 1 pg/mL in PBSS. Osterfeld and H. Yu, October-2-2008

000E+0

50E-6

100E-6

150E-6

200E-6

250E-6

300E-6

350E-6

400E-6

0 5 10 15 20 25 30 35 40

Minutes

Sig

nal,

Vp

100:1

75:1

Biotinylated Sensors

1 pg/mL Analytes

Photo of Individually Functionalized Sensors:

Biotin / BSA

GC

SF

Biotin

Epoxy

TN

F-α

IL-1α

Eotaxin

INF

LFVE

GF

CE

A

Figure 39: Example of dynamic range and channel separation. The signals from individual sensors remain well separated despite being very different in magnitude. The good similarity of signals from identically functionalized sensors is also apparent. The onset of particle binding results in a very sudden deflection of the data, indicating that no temporal averaging filter was used to smooth the data. The three signal saturation plateaus stem from three successive magnetic nanoparticle adsorptions, i.e., nanoparticle amplification as previously explained in Figure 35: Example of nanoparticle amplification.

Page 89: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

77

Sensor-to-Sensor Signal ReproducibilityMultiplex Sandwich Assay, 7 Analytes at 100 pg/mLS. Osterfeld and H. Yu, Chip 408-4-3, Nov-19-2008

-5

0

5

10

15

20

25

30

35

40

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

Minutes

Sig

nal.

mic

rovo

lts1.08%

4.04%

98% (flawed probe)

BSA Control Range: 0.72 ± 0.42 µV

Epoxy Reference

5.48%

9.45%

4.57%

5.83%

Figure 40: Example of sensor-to-sensor signal reproducibility in multiplex assay. Data from a single chip with multiple analytes in an actual protein sandwich assay. Seven analytes were measured, each with eight identically functionalized sensors. The median sensor-to-sensor coefficient of variation is roughly 5%. One functionalization resulted in a 98% CV, but this was an unusual probe consisting of a mix of several antibodies.

5.2. SENSOR-TO-SENSOR REPRODUCIBILITY

How many sensors should be allocated to a single analyte to ensure good

reproducibility? Would the assay results be better if the data from multiple identically

functionalized sensors was averaged? To answer these questions, it is important to

have an estimate of the reproducibility of signals obtained from identically

functionalized sensors in real assays.

The data shown in Figure 40 was obtained in a regular multiplex sandwich

assay with seven protein analytes at 100 pg/mL concentrations (ca. 5 pM in most

cases). The resulting CV for each group of eight sensors was ca. 1% to 10%, with a

median CV of approximately 5%. One particular analyte could not be reliably

detected, with very poor agreement across the eight corresponding sensors resulting in

a 98% CV. However, this was an experimental probe, which consisted of a mix of

several antibodies. The rest of the better performing probes were all monoclonal

antibodies.

Page 90: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

78

The initial question, i.e., how many sensors need to be allocated to a single

analyte, is not fully answerable with this data alone. However, a lower bound can be

recommended: It seems like a good idea to use at least three sensors for each analyte.

This redundancy would help to detect problems, such as the malfunctioning probe

with 98% CV, from the resulting sensor disagreement.

Additionally, one should expect that the CV increases at lower analyte

concentrations. Therefore, low concentration analytes can most likely be more reliably

quantified with a larger number of redundant sensors. On the other hand, if one only

needs a better average without an estimate of the sensor-to-sensor CV, then it would

be more economical to simply increase the sensor size to average more analyte

binding events. Doing so should have the same effect as averaging the data from

multiple sensors, without allocating too many of the limited data channels to a single

analyte.

In Figure 40, the BSA control sensors are expected to have an average signal

close to zero, and hence the BSA signal range is indicated rather than the CV.

Similarly, the average signal of the epoxy reference sensors is used to define the zero

signal reference electronically, which is why they fall almost exactly on the horizontal

axis.

Assays results with redundant sensors, such as shown in Figure 40, are often

simplified by first removing data from sensors which are electrically malfunctioning

(e.g., unstable baseline resistance, out of spec initial resistance, etc.) and then

estimating an average signal for each analyte.

Page 91: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

79

Chip-to-Chip Signal ReproducibilityMultiplex Sandwich Assay, 7 Analytes at 100 pg/mL

S. Osterfeld and H. Yu, Nov-19-2008

-5

0

5

10

15

20

25

30

35

40

Analyte 1 Analyte 2 Analyte 3 Analyte 4 Analyte 5 Analyt e 6 Flawed Probe BSA

Functionalization

Sig

nal,

mic

rovo

ltsChip 408-2-6

Chip 408-3-3

Chip 408-4-3

Chip 408-4-4

Chip 408-7-3

CV = 1.5%

CV = 11.3%

CV = 10.1%CV = 6.3%

CV = 13.3%

CV = 11.1%

CV = 29.8%

BSA Control0.15 ± 0.54 µV

Figure 41: Example of chip-to-chip assay reproducibility. A multiplex sandwich assay with analytes at concentrations of 100 pg/mL was repeated five times on separate chips. If the flawed probe is excluded, the average chip-to-chip CV in this test was 8.9%.

5.3. CHIP-TO-CHIP REPRODUCIBILITY

To estimate the chip-to-chip reproducibility of assay results, the measurement

shown in Figure 40 was repeated five times on separate chips. On each of the five

identically functionalized chips, mean analyte signals were determined by averaging

of the redundant sensors. Ideally, all five chips should return identical analyte signals,

and the actual signals are shown in Figure 41.

The average chip-to-chip coefficient of variation was 11.9% in this round of

measurements if the flawed probe with 29.8% chip-to-chip CV is included. Without

the flawed probe, the chip-to-chip CV is 8.9% on average. This degree of chip-to-chip

assay reproducibility is quite good, and in practice means that one can usually

compare results from different chips without the inclusion of on-chip calibration

(a.k.a. housekeeping) probes. Still, on-chip control sensors which provide a chemistry-

dependent reference signal for assay calibration and quality assurance could be very

helpful for identifying general chemistry problems on individual chips.

Page 92: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

80

Recycled Chip WD1-2-3, Mar-16-2008Only HPV-18 analyte is present (1nM)

-5

0

5

10

15

20

0 5 10 15 20 25 30 35

Time, minutes

Sig

nal,

mic

rovo

lts

HPV-39

EPOXY

EPOXY

HPV-39

HPV-39

EPOXY

HPV-39

HPV-39

EMPTY

EMPTY

HPV-18

HPV-31

HPV-18

HPV-18

HPV-31Nanoparticles NanoparticlesLinker DI Water

Initial Signal AmplifiedSignal

RinseWith

DIWater

Figure 42: Nanoparticle adsorption followed by nanoparticle release. Adsorbing nanoparticles generates a signal within a few minutes. Releasing the adsorbed nanoparticles by denaturing the hybridized DNA with DI water generates a reverse signal almost instantaneously. Using the magnitude of the near instantaneous nanoparticle dissociation signal for assay quantification, rather than the slower nanoparticle association signal, can reduce the impact of sensor drift on the results.

5.4. REDUCING THE IMPACT OF SENSOR DRIFT

In the early days of the spin-valve biochip experiments it used to be standard

procedure to rinse the chip with deionized water at the end of the measurement to

remove surface salt residues, so that the nanoparticles, which otherwise would have

been obscured, could be inspected in the scanning electron microscope. In protein

sandwich assays the rinse with DI water did not change the signal levels very much.

However, several experiments were also performed where the capture probe and

analyte consisted of complementary DNA strands. In these DNA detection

experiments, the final DI water rinsing step would cause the previously established

nanoparticle binding signal to disappear almost instantaneously and completely, due to

the DNA dissociating and the nanoparticles being released, as shown in Figure 42.

Page 93: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

81

Adsorption Measurement vs. Release MeasurementHPV DNA Detection Experiment

S. Osterfeld / Heng Yu March 19th, 2008

-10

-5

0

5

10

15

20

25

30

Known PositiveProbes

Known NegativeProbes

Known PositiveProbes

Known NegativeProbes

Sig

nal,

mic

rovo

lts

Nanoparticle Adsorption Data Nanoparticle Release Data

3 σ LOD

10 σ LOQ

3 σ LOD

10 σ LOQ

Figure 43: Quantification from nanoparticle adsorption vs. nanoparticle release. The nanoparticle release data is obtained faster and hence less affected by sensor drift, which improves the negative probe signals in particular. This in turn results in a lower limit of detection (LOD) and a lower limit of quantitation (LOQ), which are both determined by the standard deviation of the negative probes (σ).

The feasibility of using the nanoparticle release data for a more accurate assay

quantification is shown in Figure 43. This set of data was obtained in a DNA detection

experiment performed on a chip which exhibited a slightly elevated amount of sensor

drift, which could have been either of magnetic origin (domain noise) or due to slow

sensor corrosion during the measurement. The sensor drift was random from sensor to

sensor, rather than uniform, resulting in a spreading of signals over time which could

not be corrected by reference subtraction or similar means.

In Figure 43 the “nanoparticle adsorption data” corresponds to the regular

method of magnetic assay quantification, where the signal gain during the nanoparticle

incubation period is recorded. On the other hand, the “nanoparticle release data” is the

near instantaneous drop in signal (multiplied by -1 for better comparison) which

happened upon rinsing with DI water.

In Figure 43 the most obvious difference between the adsorption and the

release data can be found in the set of DNA probes which are known to be negative,

Page 94: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

82

i.e., which should ideally have resulted in a zero signal. The nanoparticle release data

from the negative probes is closer to the ideal value and has a much smaller standard

deviation. This is significant, because the standard deviation of the known negative

probes, designated as σ in Figure 43, determines the limits of detection (LOD) and

quantitation (LOQ). 3 σ is a common definition for the LOD. Any signal above the

LOD is a positive result with a 99% confidence. Similarly, 10 σ is often assumed to be

a reasonable LOQ, which is the minimum signal required to obtain a reasonably

accurate measure of concentration, rather than a simple positive/negative result.

In Figure 43, only 63% of the positive probes are above the LOD in the

nanoparticle adsorption dataset. On the other hand, 88% to 100% of positive probes

are above the LOD in the nanoparticle release dataset. This improvement in detection

is solely the result of the shorter observation period needed to record the nanoparticle

release signal, which hence is less affected by sensor drift.

Some degree of sensor drift is inevitable in high sensitivity measurements. In

fact, in many cases slow and random sensor drift is one of the sensitivity-limiting

factors on the spin valve biochip, second in importance only to functionalization

uniformity. It therefore would be desirable to find a similar method of sudden

nanoparticle release for the protein assays. Maybe a suitable antibody dissociation

reagent could be found. A better solution might be to anchor the antibody probes to

the sensor surface with a double stranded DNA segment, which could be dissociated

with benign DI water on demand at the end of the sandwich assay. Another solution

might be the use of a protein in which a sizeable conformation change (contraction or

expansion) could be induced on demand. This also would also create an observable

and sudden signal, as will be shown in the next section, Signal Dependence on Height.

Page 95: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

83

Sensor 1 Sensor 2 Sensor 3

20nm SiO220nm SiO2

Signal: 241 µV Signal: 162 µV Signal: 114 µV

(241/162)^(1/3) = 1.14

Sensor 2is 1.14x farther away

Z0 + 20 nm = 1.14 Z0

Z0 = 20/0.14 = 142 nm

(241/114)^(1/3) = 1.28

Sensor 3is 1.28x farther away

Z0 + 40 nm = 1.28 Z0

Z0 = 40/0.28 = 142 nm(confirmation)

Sensor 1unknown distance

Z0

Signal vs. Sensor-to-Nanoparticle DistanceS. Osterfeld, Nov-13-2008

y = 7E+08x -3.0016

R2 = 0.9991

0

50

100

150

200

250

140 180 220 260 300

Calculated Distance, nanometers

Obs

erve

d S

igna

l, m

icro

volts

A B

Figure 44: Signal vs. sensor-to-nanoparticle distance. By creating a staircase structure of additional oxide on a spin valve biochip, the actual distance of surface-bound nanoparticles Z0 could be measured.

5.5. SIGNAL DEPENDENCE ON NANOPARTICLE DISTANCE

One fundamental question that needed to be confirmed was the signal

dependence on sensor-to-nanoparticle distance. In general, the field of a dipole magnet

diminishes with the third power of the distance. However, former students had also

carried out calculations which suggested that this r3 relationship might not be valid

very close to the sensor, and that for a specific sensor geometry and particle size, there

might be an optimum distance and strongest signal at ca. 80 nanometers20.

The availability of the 64-sensor amplifier made it possible to investigate this

scaling relationship carefully in an actual experiment. A regular 64-sensor spin valve

chip was modified by coating groups of eight sensors with successive depositions of

SiO2 in an ion beam sputter system, resulting in a staircase of additional oxide on the

chip as shown in Figure 44a. In total, eight groups of sensors with increasingly thicker

SiO2 layers ranging from +20 nm to +160 nm were created. To obtain a signal, this

chip was uniformly biotinylated, and MACS nanoparticles from stock solution were

adsorbed for roughly eight minutes, at which point the signal was stabilized by

replacing the MACS solution with PBS. At this point the signal was averaged for each

group of eight identical sensors for better data accuracy. The unknown sensor-to-

MACS distance on the unmodified sensors Z0 was then calculated as shown in Figure

44a while assuming an r3 scaling relationship.

Page 96: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

84

MACS Nanoparticle to Sensor Distance - Continuous M easurement S. Osterfeld, Nov-13-2008

-50

0

50

100

150

200

250

300

350

0 2 4 6 8 10 12 14 16 18

Minutes

Sig

nal,

mic

rovo

lts

120

130

140

150

160

170

180

MA

CS

Dis

tanc

e, n

anom

eter

s

Original Sensors

+20 nm SiO2

+40 nm SiO2

+60 nm SiO2

+100 nm SiO2

+120 nm SiO2

+140 nm SiO2

+160 nm SiO2

Distance Calc.MACS Nanoparticle Incubation PBS Air Dried ChipDI

142 nm

129 nm

Figure 45: Continuous measurement of the average nanoparticle distance. The organic surface chemistry layer contracts by 13 nanometers as the chip surface is dried.

The sensor-to-MACS distances on the modified sensors was then estimated by

adding the known thickness of the deposited oxide layers to the initial distance Z0. The

resulting signal vs. distance data is shown in Figure 44b, and fitted with a power law

by nonlinear regression. The resulting fit is excellent and no deviation from a simple r3

scaling relationship is evident even if slightly different values for Z0 are tested.

Of greater interest than the scaling law confirmation is the fact that this method

can yield a seven-fold independently calculated estimate of the sensor-to-nanoparticle

distance, which is averaged over several million nanoparticles. The so calculated

distance should be a rather precise measurement of Z0 = 142 nm.

The most interesting result is that the spin-valve biochip can provide a real-

time measurement of the average nanoparticle distance. By performing the Z0

calculation for every data interval, and not just at the end of the assay, a continuous

distance measurement can be obtained as shown in Figure 45. This capability could

potentially be used to measure conformation changes in surface-bound proteins in

real-time, for example in response to a signaling molecule which is added to the

solution.

Page 97: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

85

To demonstrate a measurement of a simple protein conformation change, the

biotin assay in Figure 45 was finished with a brief rinse with deionized water,

followed by air drying of the sensors. This caused the biotin surface chemistry, and

probably also the streptavidin and/or dextran matrix of the MACS nanoparticles, to

contract by a combined amount of 13 nanometers due to dehydration. This resulted in

a corresponding sudden signal increase on all sensors, with the largest signal increase

of +38% occurring, as one would expect, on the unmodified sensors.

In conclusion, Z0 = 142 nm is somewhat larger than expected. The spin valve’s

free layer is ca. 10 nm below the sensor’s surface, and the sensor passivation accounts

for another 40 nm, for an approximate total of 50 nm. This would suggest that the

relatively simple biotin surface functionalization chemistry, plus one MACS particle

radius, must have accounted for roughly 90 nm. Possible explanations include that

hydrated MACS particles may be significantly larger than they appear in the electron

microscope, and that the r3 scaling relationship diminishes in the near-field20, which

would tend to inflate Z0 calculations done according to Figure 44a.

These results also highlight the importance of minimizing and controlling the

thickness not just of the sensor passivation, but also of the surface functionalization

chemistry, which may have an even greater thickness in real sandwich assays – this

should be measured in a future experiment. So far, the data suggests that even just a

13 nm change in functionalization thickness can result in a 38% difference in

sensitivity and signal levels. This result also reinforces the notion that the surface

functionalization needs to be applied in a way that prevents uncontrolled evaporation

and uneven drying, which almost certainly would result in functionalization thickness

variations, and binding site density variations as well. This means, for example, that

the initial PEI layer in sandwich assays should not be slowly air dried, but instead

quickly and uniformly dried with compressed nitrogen. Lastly, the hypothesized

optimum nanoparticle distance of roughly 80 nm could be neither verified nor refuted,

since it lies outside the experimentally accessible range of distances, but a diminished

r3 scaling relationship might be observable in a meticulous follow-up experiment.

Page 98: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

86

5.6. SIGNAL DEPENDENCE ON TICKLING AND BIAS FIELDS

As described in Chapter 1.3.2 – Principle of Operation, the superparamagnetic

nanoparticles need to be polarized with an external magnetic tickling field Ht, before

they can be readily detected by the spin valve sensors. A magnetic bias field Hb is also

applied to bias the sensor and lower the measurement noise. Hb is a static field applied

along the spin valve segments’ geometric axis, while Ht is an alternating field applied

perpendicular to the spin valve sensors’ geometric axis, with a frequency of 200 Hz to

facilitate narrowband detection. Details of this signal generation scheme have been

previously described in the works of S. X. Wang, G. X. Li, and SJ. Han.

Without a tickling field, there is no signal to measure, and without a bias field

the sensor does not respond linearly and is too noisy to be useful. On the other hand, if

either of these fields is too strong, then the sensor is saturated and can no longer detect

the small field changes due to the nanoparticles. However, the appropriate strength of

these externally applied tickling and bias fields had never been systematically

evaluated. Instead, default settings of 50 Oersteds for the bias field and 80 Oersteds

(rms) for the tickling field were routinely used, which had been roughly estimated

from the spin valve sensor’s MR transfer curve, which in a 50 Oe bias field appears to

be pretty linear to roughly ± 100 Oersteds before gradually saturating.

The obvious method for testing the bias and tickling field dependence would

have been to run several identical measurements with different Ht and Hb settings.

However, this would have been exceedingly laborious, in part because at the time the

signals varied somewhat from chip to chip even if nothing is changed. Thankfully, it

was determined that the sensor response is highly reproducible when alterations and

restorations of Ht and Hb are performed while the nanoparticle coverage is held

constant. This made it possible to determine the signal dependence on Ht and Hb on a

single biotinylated chip, by measuring the initial signal baseline level for various Ht

and Hb settings before the experiment, and again after the nanoparticle binding had

been positively stopped with a wash, rinse, and drying step, as shown in Figure 46a.

Page 99: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

87

Magnetic Biochip Signal vs. Bias Field (Hb) and Tic kling Field (Ht)First Experimental Data, July-22-2007, S. Osterfeld

-250

0

250

0 10 20 30 40 50 60 70 80 90 100Minutes

Sig

nal,

mic

rovo

lts

25 Alternate Settings for Ht, Hb Recreate 25 Alternate Settings for Ht, Hb Particle Adsorption,

Default Settings

Example of Alternate Setting(Hb 132, Ht 25 Oe), Signal = 44 µV

Default Settings(Hb 54, Ht 78 Oe)Signal = 125 µV

A

Figure 46: Determination of the optimal tickling and bias field for 1.5 µm sensors.

Page 100: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

88

Five different settings for the tickling field (25, 38, 51, 63, and 78 Oe rms) and

five different settings for the bias field (1, 14, 27, 50, and 54 Oe) were initially

evaluated in this way, resulting in twenty-five combinations of settings that had to be

dialed in and held for ca. one minute each, both before and after the actual

nanoparticle binding. The before-and-after difference in signal level was determined

for each setting, together with the sensor noise at each setting, which was determined

from the amount of signal variation during the one-minute steady hold for each

setting. Such data was obtained from a total of four sensors (the limit at the time of

this experiment), but only one such sensor is shown in Figure 46a for visual clarity.

The results revealed a surprising trend of increasing signals with decreasing

fields, so much in fact that both the first experiment (Figure 46b) and a second

experiment (Figure 46c) failed to reveal a clear signal maximum, which was

apparently closer to zero than the range of fields investigated. Similarly surprising was

the strength of the scaling dependence, as signals between 473 µV (Ht = 25, Hb = 1)

and 41 µV (Ht = 131, Hb = 27) could be obtained, depending on the choice of fields.

The underlying reasons for this signal scaling dependence remained somewhat of a

mystery until a mathematical model of the sensor and nanoparticle magnetization was

developed almost a year later, which showed that this signal scaling relationship was,

in fact, in good agreement with theory (see Chapter 6, Mathematical Modeling).

On the other hand, the noise dependence on fields was as expected, with the

sensor noise increasing in the absence of a bias field as shown in Figure 46d. The

noise measurements were combined with the signal levels to obtain the signal to noise

ratio at each field setting as shown in Figure 46e. The results suggested an optimum

SNR at ca. Ht = 38 Oe rms and Hb = 27 Oe. Fields were both ca. 50% lower than the

former default settings, resulting in 2x larger signals and ca. 6 dB better SNR. These

results were used for any subsequent assays on this type of chip. However, as later

results and mathematical modeling revealed (see next sections), these specific settings

were only optimal for sensors with 1.5 µm wide, 90 µm long spin valve segments.

Page 101: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

89

Magnetic Nanoparticle Labels in PBSAirPBS

0

Time, minutes

∆ S

ign

al,

µV

rms

0 10 20 30

100

20

40

60

80

3 µm sensors

1.5 µm sensors

a Wide sensor vs. narrow sensor geometry b Signal development vs. sensor size

i i

Three segments,

each 3.0 µm wide

Six segments,

each 1.5 µm wide

Figure 47: Schematic illustration of sensor segment width evaluation. This figure illustrates how the effect of sensor segment width can be evaluated while keeping the current density and total resistance constant (A). Initial tests revealed that varying the sensor width can indeed affect sensitivity (B).

5.7. SIGNAL DEPENDENCE ON SENSOR SEGMENT WIDTH

Several models exist which predict that a spin valve sensor will be more

sensitive in nanoparticle detection assays if its aspect ratio (length/width) is

increased60,61. To investigate this prediction experimentally, several biochips chips

were made which featured multiple sensors in close proximity which were differently

patterned in such a way that only the sensor segment width was varied, while all other

operating parameters (current density, overall resistance, total area, etc.) were kept

constant. This is shown schematically in Figure 47a. Initially, such experiments were

limited by the electronics and the use of the available optical photolithography to

measuring only four sensors and segment widths of more than 1 µm respectively. Such

initial results revealed a promising trend which seemed to agree with predictions of

better sensitivity for more finely segmented sensors, as shown in Figure 47b. An

interesting and yet unexplained aspect of these early results is that the more finely

patterned sensors did not just simply generate a stronger signal, but also a signal with

a disproportionally faster initial response. In other words, while the signal of the

1.5 µm sensors is ca. 1.4x larger than that of the 3.0 µm sensors at t > 15 minutes, it is

up to 2.6x larger in the first 15 – 45 seconds of nanoparticle application.

Page 102: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

90

Figure 48: Signal and noise dependence on spin valve sensor segment width. The optimal width for 90 µm long spin valve segments seems to be between 600 to 800 nanometers if noise is considered too. In this experiment no bias field was applied.

Later, with the availability of the 64-sensor amplifier it became feasible to

measure a single biochip which had a much larger number of different sensor on the

same chip. Additionally, due to a fortuitous collaboration with Hitachi Global Storage

researchers Robert Fontana, Thomas Boone, Stefan Maat, and Jordan Katine, it

became possible to fabricate a spin valve biochip with test sensors patterned by

electron beam photolithography. This easily permitted sensor segment widths of down

to 300 nanometers.

A biochip was fabricated with eight different spin valve sensor types, each

having segments that were 90 µm long, but of various widths. This chip was uniformly

biotinylated and the nanoparticle adsorption signal was measured for all eight sensor

types under a range of different tickling fields. No bias field was applied. The results

are shown in Figure 48a, which reveal that starting with a 1.5 µm sensor, increasingly

finer sensor segmentation will initially result in increased sensitivity. However, after

reaching a maximum sensitivity at ca. 0.75 µm segment width, reducing the sensor

width further actually decreases the sensitivity. This is contrary to most theoretical

models, which predict no limit to the benefits of decreasing the sensor width.

However, the diminishing returns of increasingly finer sensor segmentation

were partially anticipated. Spin valve sensors typically have a “dead zone” or “inverse

Page 103: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

91

zone” at their edges where the magnetoresistance effect is suppressed or inverted. This

edge effect is thought to have a fixed width extending maybe 50 – 100 nm from the

edge into the segment. For increasingly narrower segments, this would leave an

increasingly smaller intact magnetoresistive region available for nanoparticle

detection, which would explain the diminishing returns of narrower sensors.

Another interesting result of this experiment is the sensor noise, which is

shown in Figure 48b. Unfortunately, the chip which was investigated had a few

defects, and as such the noise data is not deemed to be accurate enough for a reliable

signal to noise ratio calculation. However, the general trend of increasing noise with

increasing sensor width is accurate and has been confirmed in several other

experiments. In fact, it was eventually found that an external bias field was no longer

beneficial for sensors with segment widths of less than 750 nm, because such sensors

already had low noise due to their internal magnetic bias, or shape bias field, which

stems from their high aspect ratio of more than 120 (length/width). The resulting

simplification of the external magnetic field setup – a reduction of the Helmholtz coil

from two to one axis – is another engineering accomplishment of this thesis work.

Taking into account the experimentally observed noise and signal trends of the

spin valve sensors, it was determined that the optimal spin valve sensor segment width

lies between 600 nm and 800 nm, with corresponding optimal tickling fields of 40 to

20 Oe (rms), respectively. Narrower segments would give lower signals without

significantly improving noise, while wider segments would produce lower signals and

increasingly more noise.

It is anticipated that these results should be more general in terms of sensor

segment aspect ratio – i.e., optimal sensors probably have a segment aspect ratio

ranging from 110 to 150 (length/width), regardless of actual size. Additional

experiments would be needed to clarify these results and the exact noise scaling.

Page 104: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

92

CHAPTER 6. MATHEMATICAL MODELING

In Chapter 5 the dependence of the nanoparticle signal on various parameters,

such as tickling field strength, bias field strength, and sensor width was determined

experimentally. The signal scaling with the applied fields was particularly puzzling at

first, in part because significantly higher signals were obtained for increasingly

smaller tickling and bias fields Ht and Hb, as shown in Figure 49a.

Almost a year after these results were obtained, it seemed like a good idea to

investigate whether these results were actually in agreement with theory or not. Simple

equations for the resistance of a spin valve sensor and for the magnetization of

superparamagnetic nanoparticles existed, but they had not yet been combined in an

intuitive fashion, and hence a new and simple modeling approach was pursued. In the

end, summing up the magnetic field x- and y-components acting on the sensor, and

then defining what an assay signal is in this context, yielded a surprisingly

straightforward model which is entirely analytical and which showed that the results

were actually in almost perfect agreement with theory, as is qualitatively highlighted

in Figure 49b.

Figure 49: Experimental observations were explained with a mathematical model.

Page 105: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

93

( )θSinR

RR Avg *2max

.

∆+=

Sensor θ

Hy

Hx

x

y

x

y

HkHb

Ht

++∆+=

22

max.

)(*

2 HkHbHt

HtRRR Avg

Figure 50: The resistance of a spin valve sensor segment is determined by the externally applied tickling field Ht, the externally applied bias field Hb, and the segment’s geometric aspect ratio, which manifests itself as an apparent anisotropy field Hk.

6.1. THE RESISTANCE OF A SPIN-VALVE BIOSENSOR

In Figure 50, according to Li et al., the resistance of a spin valve sensor which

has its reference (pinned) layer oriented along the y-direction can be described as

(1)

where θ is the angle between the free layer and the x-axis. Rather than using

the hyperbolic tangent approximation by Li et al., it is assumed that the angle θ can be

calculated exactly from Hy (the y-component of all magnetic forces acting upon the

free) and Hx (the x-component of all magnetic forces acting upon the free) as

(2)

Central to the upcoming model is the idea that Hy and Hx are similar to, but

not exactly the same as the applied tickling and bias fields Ht and Hb. For example,

the sensor’s shape introduces an apparent anisotropy field Hk, which supplements the

external bias field in terms of free layer stabilization. Thus, for the sensor’s free layer,

(3)

( )θSinR

RR Avg *2max

.

∆+=

( )22 HxHy

Hy

Hx

HyArcTanSinSin

+=

HkHbHx +=

Page 106: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

94

Example of Calculated MR Transfer Curves in Zero Bi as FieldFor Spin Valve Sensors with 10% MR and 100 µm Long Segments

1880

1920

1960

2000

2040

2080

2120

-200 -150 -100 -50 0 50 100 150 200

Applied Transverse (Tickling) Field, Oe

Res

ista

nce,

Ohm

s

3.0 µm Sensor, Hk ~ 17 Oe

1.5 µm Sensor, Hk ~ 33 Oe

0.75 µm Sensor, Hk ~ 67 Oe

0.5 µm Sensor, Hk ~ 100 Oe

Figure 51: Example of calculated spin valve sensor MR transfer curves.

This leads to the final equation of the sensor’s resistance

(4)

In equation (4), Ht and Hb are controlled with external electromagnets, while

RAvg. and ∆Rmax are material- and geometry-dependent properties. For example, a

typical spin valve film has a sheet resistance of ca. 20 Ω/sq and maybe 10%

magnetoresistance, meaning that a single sensor segment of 100 µm length and 1 µm

width would have RAvg. = 2,000 Ω and ∆Rmax ~ 200 Ω.

The apparent shape anisotropy field Hk is determined by the sensor’s

geometric aspect ratio (width/height), and can be calculated from first principles if

desired. Very approximately, from actual experiments, it seems that

(in Oersteds) (5)

However, in practice Hk should be measured directly from the finished

sensor’s magnetoresistance transfer curve to guard against manufacturing variations.

++∆+=

22

max.

)(*

2 HkHbHt

HtRRR Avg

width

lengthHk *4.0≈

Page 107: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

95

MACS Nanoparticle Magnetization CurveCourtesy of Wei Hu

-100%

-50%

0%

50%

100%

-200 -150 -100 -50 0 50 100 150 200

Applied Field, Oe

Mag

netiz

atio

n, P

erce

nt

AGM Measurement

Langevin Fit, α = 0.025 / Oe

Figure 52: Measured magnetization curve and model for MACS nanoparticles.

6.2. THE MAGNETIZATION OF SUPERPARAMAGNETIC NANOPARTICLES

Superparamagnetic nanoparticles develop a magnetization M in the presence of

an externally applied magnetic field H. The degree of nanoparticle magnetization,

normalized by the saturation magnetization MSat, is commonly described by the

Langevin Function as

(6)

where α is a proportionality constant which depends on the temperature, the

material, and the diameter of the nanoparticles. In practice, α is often experimentally

determined from a magnetization measurement of specific particles. For the MACS

nanoparticles, α ~ 0.025 * Oe-1 is a reasonably accurate value at room temperature.

However, an assembly of nanoparticles with a range of sizes will not follow equation

(6) exactly. Such a size distribution is the likely reason why the Langevin Function fit

in Figure 52 is not perfect. In fact, sometimes the underlying nanoparticle size

distribution can be inferred by weighing an assembly of Langevin Functions with

different values for α until an optimal fit with the measured data has been achieved.

This method has been termed Langevin Granulometry62.

( )H

HCothM

M

Sat *

1*

αα −=

Page 108: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

96

( )βCosMpcHpx **3=

( )βSinMpcHpy **3=β

P

Nanoparticle

MP

Sensor θ

Hy

Hx

x

y

x

y

HbHk

Ht

Ht

HbHpx

Hpy

( )θSinR

RR Avg *2max

.

∆+=

Figure 53: Mathematical description of the sensor-nanoparticle interaction. The proportionality factor c3 depends on the size and location of the nanoparticle relative to the sensor.

6.3. EFFECT OF NANOPARTICLE ON SENSOR RESISTANCE

The essence of the mathematical model is captured in Figure 53. The

nanoparticle magnetization Mp is determined solely by the externally applied fields

Hb and Ht. Specifically, we can now rewrite equation (6) and specify the angle β

(7)

(8)

The nanoparticle magnetization Mp generates additional field components Hpx

and Hpy, which act on the sensor’s free layer. How large these x- and y-components

are depends on several factors, such as the distance and position of the nanoparticle

relative to the sensor, as well as the nanoparticle size. These factors could be

accurately calculated for an individual nanoparticle.

However, on a biosensor which is interacting with millions of nanoparticles, it

is assumed that these factors can be folded into a single, constant, proportionality

factor c3, which represents the average sensor-nanoparticle interaction coefficient for

surface-bound nanoparticles.

( )22

22

*

1*

HtHbHtHbCothMp

+−+=

αα

( )22 HtHb

HtSin

+=β ( )

22 HtHb

HbCos

+=β

Page 109: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

97

With these assumptions it is now possible to rewrite equation (4) for a single

nanoparticle as

(9)

To account for multiple nanoparticles, we introduce the variable n, which

simply is the number of nanoparticles that act on the sensor. Adding this variable to

equation (9) gives

(10)

where Mp and β are given by equations (7) and (8).

Equation (10) is already a complete description of the biosensor’s

instantaneous response to fields and magnetic nanoparticles. It could be further

modified to reflect the time-varying tickling field by changing the term Ht

accordingly. However, it is easier to consider Ht to be the effective (rms) value.

Also, it is not the absolute resistance R which is important, rather the change

∆R in response to a change in the number of nanoparticles ∆n is important. The goal

therefore is to maximize the sensor’s signal ∆R/∆n. Rather than calculating discrete

differences, the derivative δR/δn is therefore defined on the following page.

( )( )( ) ( )

++++

+∆+=22

max.

)**3(**3

**3*

2 HkCosMpcHbSinMpcHt

SinMpcHtRRR Avg

βββ

( )( )( ) ( )

++++

+∆+=22

max.

)**3*(**3*

**3**

2 HkCosMpcnHbSinMpcnHt

SinMpcnHtRRR Avg

βββ

Page 110: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

98

6.4. DEFINITION OF ASSAY SIGNAL

As stated, the goal is to maximize the sensor’s signal ∆R/∆n in equation (10).

For this reason it is convenient to look at the derivative δR/δn. Furthermore, since the

goal is to detect the smallest possible number of nanoparticles n, it is sufficient to look

only at the initial response, i.e., the derivative in the limit of n 0, which simplifies

the resulting derivative further:

(11)

Calculating (11) in Mathematica (see Appendix C) yields the relatively

compact result

(12)

Note that Mp has been factored out but is still given by equation (7).

With equation (12) it is now possible to calculate the magnetic biochips’

response to nanoparticle adsorption as a function of tickling field Ht and bias field Hb.

Additionally, equation (12) makes it possible to calculate the effect of varying the

sensor width or aspect ratio, which will be reflected in a corresponding change of the

sensor shape bias field Hk. The results may need to be scaled by a constant multiplier

due to uncertainty about c3 and n. However, the experimentally observed signal

scaling trends and local maxima, which are very important for sensor optimization, are

independent of n and c3 and strongly corroborated by this simple theoretical model.

∂∂≡

→ n

RLimitnSignal

n 0*

( )( )

+++

+∆=2

32222

max *)(***3*

2*

HtHkHbHtHb

HtHkHbHkMpc

RnSignal

Page 111: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

99

Sensitivity vs. Tickling Fields: Model and Experime ntS. Osterfeld, Apr-14-2008

0

50

100

150

200

250

300

0 10 20 30 40 50 60 70 80 90 100

Tickling Field, Oe (rms)

Sig

nal,

mic

rovo

lts

500nm Data500nm Model750nm Data750nm Model

Figure 54: Model and experiment of two different types of sensors at zero bias field.

6.5. MODEL AND EXPERIMENT – OPTIMAL TICKLING FIELD AT ZERO BIAS

As mentioned before, spin valve sensors with an aspect ratio (length/width) of

more than 100:1 have sufficiently low noise even without a bias field Hb. Two such

sensor types were evaluated for optimal signal generation in a biotin-BSA dummy

assay in a range of different tickling fields Ht. The results of these experiments,

together with the model according to equation (12), are shown in Figure 54. The

following four parameters were used to obtain a good fit:

500 nm Sensor 750 nm Sensor

∆Rmax 200 Ω

n*c3 1 mV * Oe / Ω 0.67 mV * Oe / Ω

α 0.025 / Oe

Hk 91 Oe 45 Oe

∆Rmax and n*c3 are simply vertical scaling coefficients. Parameter Hk is the

primary determinant of the optimal tickling field strength, which was found to be ca.

27 Oe (rms) for the 750 nm sensors and ca. 55 Oe (rms) for the 500 nm sensors.

Page 112: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

100

Signal Dependence on Fields – Model S. Osterfeld, Apr -14-2008

Signal Dependence on Fields – Experiment S. Osterfeld, July -22-2007

Figure 55: Model and experiment of signal dependence on fields. See also Figure 46c for the original experiment.

6.6. MODEL AND EXPERIMENT – TICKLING AND BIAS FIELD DEPENDENCE

One of the first applications of equation (12) was to check if the dependence of

the signal on bias and tickling field could be explained mathematically. Indeed, the

mathematical model was in very good agreement with the experimentally observed

signal scaling trends. This is shown in a 3D graph in Figure 49 for larger fields, and

again for lower fields in a 2D surface contour plot in Figure 55.

The sensor on which the experimental data in Figure 55 was obtained was

designed to have 1.5 µm wide segments, 100 µm long. It was modeled with equation

(12) using the following parameters:

1.5 µm Sensor

∆Rmax 200 Ω

n*c3 2.62 mV * Oe / Ω

α 0.025 / Oe

Hk 15 Oe

Page 113: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

101

Signal Dependence on Sensor Width – Model S. Osterfeld, September 2009

Signal Dependence on Sensor Width – Experiment Osterfeld, Fontana, Boone, et. al, Oct-21-2008

Figure 56: Model and experiment of signal dependence on sensor segment width.

6.7. MODEL AND EXPERIMENT – SENSOR SEGMENT WIDTH DEPENDENCE

The model given by equation (12) can also be used to help interpret the signal

vs. sensor width results which were obtained in collaboration with Hitachi researchers

(see Chapter 5). However, in this application the agreement between model and

experiment is not as good as in the field dependence tests. The reason may be that

different sensors are combined into one experiment, which will introduce fabrication

accuracy and sensor-to-sensor biochemistry uniformity as sources of error.

Nevertheless, the model and experiment agree qualitatively as shown in Figure 56. In

these calculations, the shape bias Hk was allowed to deviate somewhat from the

scaling relationship given in equation (5) to reflect likely deviations of the actual

width from the designed sensor width. Also, the existence of a ca. 100 nm “dead zone”

at each sensor edge (see Chapter 5) was corroborated by the model, and was simulated

by decreasing the ∆Rmax parameter accordingly for successively smaller sensors.

Width, nm 1500 1000 750 600 500 400 375 300

∆Rmax, Ω 87 80 73 67 60 50 47 33

n*c3 0.65 mV * Oe / Ω

α 0.025 / Oe

Hk, Oe 18 26 34 50 70 102 109 136

Page 114: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

102

6.8. INSIGHT DERIVED FROM MATHEMATICAL MODELING

Developing the mathematical model was a very helpful exercise, because it

deepened the understanding of the spin valve biosensor – nanoparticle interaction. For

example, initially the expectation existed that a more strongly magnetized nanoparticle

would generate a stronger signal. However, the mathematical set-up shown in Figure

53, by virtue of its leading to a mathematical model which agrees so well with the

experiments, makes it clear that it is not the maximum nanoparticle magnetization that

we seek, but instead the maximum magnetic field rotation in the vicinity of the

particle. In essence, the mathematical model reminds us that the spin valve sensor is

primarily a magnetic field orientation sensor, rather than a field strength sensor. It is

only by appropriately setting up the spin valve sensor that we can measure a magnetic

field strength, i.e., if the external field provides a tangential component (a torque, so to

say), which disturbs the spin valve’s equilibrium orientation somewhat proportionally,

then it becomes possible to infer the magnetic field strength indirectly from its effect

on the free layer orientation.

By recognizing that the spin valve sensor is primarily a field orientation sensor,

it becomes clear that the optimal strategy is to induce a magnetization in the

nanoparticles which is oriented differently from the magnetization of the sensor. This

way, the nanoparticle can exert a torque on the spin valve’s free layer, and the

resulting free layer rotation is the actual source of the signal.

The logical conclusion is that there needs to be a magnetic force component

which acts on the sensor but not on the nanoparticle – if the free layer and particle

were to experience identical forces, they would both be magnetized along the same

direction, and no signal would occur.

The only such magnetic force component which acts on the sensor’s free layer,

but apparently not on the magnetic nanoparticle, is the sensor’s shape anisotropy,

which is modeled by the fictional field Hk in Figure 53. Whether this field is actually

fictional or simply locally constrained so that doesn’t significantly affect the

Page 115: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

103

nanoparticle may need to be elaborated in more detail. However, it is clear that the

assumption that Hk affects only the free layer, and never the nanoparticle, leads to a

good mathematical model.

The utility of this conclusion is that Hk can be custom-tailored by choosing the

spin valve segment width appropriately during the sensor design. With increasing Hk,

the spin valve sensor initially becomes more sensitive to nanoparticles due to

increasingly larger differences in the orientation of the magnetizations of the free layer

and the nanoparticle. Eventually, Hk begins to limit the sensitivity of the spin valve

due to excessive free layer stabilization – an excessively stabilized free layer will not

rotate optimally despite the presence of the magnetic torque from nanoparticle.

In summary, the mathematical model helped clarify that it is not the

nanoparticle magnetization which needs to be maximized, but rather the free layer

rotation induced by the nanoparticle is what needs to be maximized. The free layer

rotation was recognized to be a function of the rotational torque exerted by the

nanoparticle (proportional to Hk) times the rotational sensitivity of the free layer

(inversely proportional to Hk). This insight made the experimentally observed signal

scaling trends much more intuitive.

Page 116: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

104

6.9. MATHEMATICAL MODELING CONCLUSION

The mathematical model given in equation (12) is simple in the sense that it

only needs two physically measurable parameters Hk and α to describe the sensor-

nanoparticle interaction in relative terms, i.e., to identify optimal field settings. With

the addition of two more parameters c3 and ∆Rmax, the model can furthermore

achieve quantitative accuracy and calculate actual signal levels.

The model has shown that the experimentally obtained signal scaling results

presented in Chapter 5 might have been anticipated from theoretical considerations,

and at a minimum are not in contradiction with the existing models of spin valve

resistance and nanoparticle magnetization.

In fact, the mathematical model has shown very good agreement with

experimental data in several cases, to the extent that the model can itself become a tool

for experimentation. For example, the model shows that even an ideal spin-valve

sensor (without a dead zone at the edge) would not benefit from decreasing the sensor

width indefinitely, because at some point the shape anisotropy field Hk becomes so

large that the sensor’s response is being limited.

Another theoretical experiment that can be performed to some extent with

equation (12) is magnetic noise analysis. One can, for example, derive how strongly

the nanoparticle signal depends on noise, or fluctuations, in the applied and shape bias

fields Ht, Hb, and Hk. More precisely, the optimal choice for Ht, Hb, and Hk is

probably that which maximizes δSignal/δn, and which also achieves a near-zero

magnitude for the terms δSignal/δHt¸ δSignal/δHb¸ and δSignal/δHk. In the signal

vs. parameter landscape, such a location would correspond to a hilltop which is high,

yet relatively flat, and which doesn’t have steep cliffs nearby. With such

considerations it might be possible to define a theoretical magnetic signal to noise

ratio from the various derivatives of equation (12), which could then be used to

identify the optimal sensor parameters in terms of magnetic SNR.

Page 117: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

105

APPENDIX A – BIOCHIP FABRICATION PROCESS AT SNF

• Refer to Figure 5 for visualizing the instructions given in this appendix.

• Starting Substrate: The starting substrate is a silicon wafer with 100 – 500 nm

thermal oxide (for electrical insulation) on which the ca. 35 nm thick GMR film

has been deposited. The GMR film should have at least 10% magnetoresistance, a

symmetric MR loop (centered around zero applied field), and the dynamic MR

loop (measured at ca. 100 – 300 Hz) should not differ too much from the static

(measured at ca. 0.1 Hz) MR loop. Additional details about the substrate structure

and procurement are given in several publications by former students (e.g.,

doctoral thesis and papers by Guanxiong Li).

• General Precautions: The metallic GMR film should never be exposed to dirt

(dust), greasy substances (fingerprints), salts (tap water), any temperature above

200° Celsius, and never to corrosive chemicals such as halogen gases or halogen

radicals, acids, or bases.

• Substrate Cleaning: Any protective photoresist should be removed with acetone,

followed by methanol and isopropanol. If dust contamination is suspected, it might

be possible to improve the wafer with a wash on the ultrasonic spin washer on the

EVBonder station. Finally, the wafer should be cleaned with DI water in one of the

automatic wafer washer/spin dryers.

Page 118: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

106

• Mask M1 (Sensor): This is the highest resolution layer of the chip. The wafer

should be prepared with a surface silanization in the HMDS YES oven. A suitable

single-layer positive photolithography process (e.g., 1 µm 3612 resist, 2 second

exposure) should then be tested and optimized on several dummy wafers before

being applied to the real wafer. Actual recipes vary with equipment, but for M1 the

lithography should be easy (single layer resist, no alignment needed). If the correct

photoresist pattern has been achieved (verify with microscope), the resist pattern

can then be transferred into the GMR film with ion beam etching, e.g.,

o Veeco Etcher: Primary etch 200 mA, 500 V, 75° Angle, 90 seconds,

followed by edge clean-up etch 150 mA, 400 V, 15° Angle, 100 seconds.

o The transfer of the M1 resist pattern into the GMR film is a timed process,

timed for ca. 150% to 200% of the minimum time needed to perforate the

GMR film.

o Subsequently, the etch trenches are filled with an oxide deposition, e.g.,

transfer of the substrate into the Iontech Ion Beam Sputtering system. Pre-

etch for enhanced adhesion 50 mA, 500 V, 30 seconds. Deposition with

SiO2 target, 100 mA, 1000 V, 90° Angle, 7 minutes (ca. 55 nm of SiO2).

o Resist lift-off. Due to being overcoated with SiO2, the resist is difficult to

remove. 24 h soak in acetone, followed by 10 minutes of sonication in

acetone should remove the resist. If available, the sensor edges should then

be cleaned up (de-burred) with a CO2-snow blaster or sonication in a

suitable abrasive slurry. The wafer should then be cleaned with acetone,

methanol, and IPA.

Page 119: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

107

• Mask M2 (Leads): The leads need to be fabricated with lift-off patterning. The

wafer should be prepared with HMDS, and a bi-layer photoresist should be used.

The recipe for applying and patterning a bi-layer resist is much more challenging

than for a single layer resist, so here are some pointers:

o A high quality HMDS prime is required, otherwise the LOL2000

underlayer resist will dissolve too rapidly during development.

o An LOL2000 underlayer is applied by spin coating in the headway spinner.

The technique should be practiced on dummy wafers. Dried LOL2000

residue on the supply bottle may ruin the coating, so pick up from center of

bottle with a pipette or syringe. DO NOT aspirate LOL2000 with a syringe

with a rubber plunger, because the plunger’s silicone oil lubrication will

interfere with the coating process and result in visibly non-uniform

LOL2000 films. Instead, an Eppendorf Repeater® should be used for

aspirating and dispensing LOL2000. Use at least 5 cc of LOL2000 per 4”

wafer (over-apply) to avoid dry streaks.

o The LOL2000 layer needs to be thoroughly baked to achieve a small and

consistent resist undercut profile. The recommended LOL2000 baking

temperature of 150°C is usually too low (undercut too large). Instead, a

bake of 30 minutes at ca. 180°C should be used to achieve a lateral

undercut of ca. 1 µm.

o The upper resist layer of the bi-layer resist can then be applied with a

standard recipe (1 µm 3612 resist). To achieve a good and consistent

undercut profile, it is important that the upper resist layer develops very

quickly. For this reason, the M2 mask should be generously overexposed –

by as much as possible without degrading the features (e.g., 3 seconds).

The resist development is equally critical, because the width of the lateral

undercut depends heavily on the LOL2000 bake temperature and developer

exposure duration. 2x 15 seconds in regular developer are recommended,

Page 120: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

108

followed by an immediate and thorough clean in DI water and drying with

N2 (no bake). To remove traces of resist in the exposed areas, the wafer

should be briefly exposed to an O2 plasma (e.g., Drytek 1, descum recipe,

30 seconds).

o The lead metallization can now be applied by evaporation or sputtering.

The typical lead structure is 5 nm Ta, 300 nm Au, 5 nm Ta, but Cr could

also be used in place of Ta to promote Au adhesion. However, it is very

important that the Ta layer on the GMR sensor is broken or roughened

(with in-situ ion beam etching or pre-sputtering) right before deposition of

the lead metallization, otherwise the GMR sensor’s native tantalum oxide

surface will result in a very high resistance at the sensor-lead junction. The

best method for lead deposition is Ion Beam Sputter Deposition with in-

situ Ion Beam Etching, e.g., in Iontech:

Tantalum Base Layer: Pre-etch for enhanced adhesion and low

contact resistance 50 mA, 500 V, 30 seconds. Followed by tantalum

deposition 100 mA, 1000 V, 60 seconds (5 nm).

Gold Middle Layer: Pre-etch for enhanced adhesion and low

contact resistance 50 mA, 500 V, 30 seconds. Followed by gold

deposition 100 mA, 1000 V, 7 minutes (300 nm).

Tantalum Top Layer: Pre-etch for enhanced adhesion and low

contact resistance 50 mA, 500 V, 30 seconds. Followed by

Tantalum deposition 100 mA, 1000 V, 60 seconds (5 nm).

o Resist lift-off. Due to being overcoated with the lead metallization, the

resist will take some time to dissolve. A 24 h soak in acetone, followed by

10 minutes of sonication in acetone, should remove the resist. If available,

the sensor edges should then be cleaned up (de-burred) with a CO2-snow

blaster or sonication in a suitable abrasive slurry. The wafer should then be

cleaned with acetone, methanol, and IPA.

Page 121: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

109

• Mask M3 (Sensor Passivation): Since the sensor passivation is thin and only

opened at the bondpads at the chip periphery, where some pattern roughness is

acceptable, a bi-layer photoresist is usually not required. Instead, a simple single

layer resist should be applied (e.g., HMDS treatment, followed by 1 µm 3612

resist, 2 second exposure).

o The sensor passivation is best applied by ion beam sputter deposition in

iontech, with an adequate ion beam cleaning etch prior to each deposition

to enhance adhesion. The typical recipe for a ca. 45 nm oxide-nitride-oxide

passivation in Iontech is as follows:

SiO2 Base Layer: Pre-etch for enhanced adhesion 50 mA, 500 V,

30 seconds. Followed by SiO2 deposition 100 mA, 1000 V, 120

seconds (15 nm).

Si3N4 Middle Layer: Pre-etch for enhanced adhesion 50 mA, 500

V, 30 seconds. Followed by Si3N4 deposition 100 mA, 1000 V,

120 seconds (15 nm).

SiO2 Top Layer: Pre-etch for enhanced adhesion 50 mA, 500 V, 30

seconds. Followed by SiO2 deposition 100 mA, 1000 V, 120

seconds (15 nm).

o Resist lift-off. Due to being overcoated with oxide, the resist may take

some time to dissolve. A 24 h soak in acetone, followed by 10 minutes of

sonication in acetone, should remove the resist. It is not necessary to clean

up the pattern with a CO2-snow blaster. The wafer should be cleaned with

acetone, methanol, and IPA.

Page 122: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

110

• Mask M4 (Lead Passivation): Since the lead passivation is relatively thick, a bi-

layer resist is usually required to pattern it by lift-off. However, since the

passivation is only omitted at the large lead bondpads at the chip periphery, where

some pattern roughness is acceptable, the requirements for this bi-layer resist are

similar, but not as stringent as that for Mask M2 (Leads).

o Follow the bi-layer resist instructions given for Mask M2. Compared to the

M2 mask, the large feature size (bond pads) makes moderate resist defects

in the M4 layer more acceptable (e.g., LOL2000 underlayer defects,

varying or very large undercut, small other resist defects, etc.).

o The lead passivation is best applied by ion beam sputter deposition in

Iontech, with an adequate ion beam cleaning etch prior to each deposition

to enhance adhesion. The typical recipe for a ca. 250 nm oxide-nitride-

oxide passivation in Iontech is as follows:

SiO2 Base Layer: Pre-etch for enhanced adhesion 50 mA, 500 V,

30 seconds. Followed by SiO2 deposition 100 mA, 1000 V, 10

minutes (85 nm).

Si3N4 Middle Layer: Pre-etch for enhanced adhesion 50 mA, 500

V, 30 seconds. Followed by Si3N4 deposition 100 mA, 1000 V, 10

minutes (85 nm).

SiO2 Top Layer: Pre-etch for enhanced adhesion 50 mA, 500 V, 30

seconds. Followed by SiO2 deposition 100 mA, 1000 V, 10

minutes (85 nm).

o Resist lift-off. Due to being overcoated with oxide, the resist will take

some time to dissolve. A 24 h soak in acetone, followed by 10 minutes of

sonication in acetone, should remove the resist. The wafer should then be

cleaned with acetone, methanol, and IPA, and coated with a soft-baked

protective resist for subsequent dicing and storage.

Page 123: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

111

APPENDIX B – TEMPERATURE CORRECTION

Spin Valve Biosensor Signal (MR) DriftS. Osterfeld, Jan-18-2008

-4

-3

-2

-1

0

1

0 100 200 300 400 500

Minutes

Mic

rovo

lts

Spin Valve Biosensor Centertone (Resistance) DriftS. Osterfeld, Jan-18-2008

0

50

100

150

200

0 100 200 300 400 500

Minutes

Mic

rovo

lts

A B

Resistors

Spin Valve Sensors

Figure 57: Example of temperature-induced drift in the magnetoresistive sideband signal (A) and the sense current (B).

Cyclical signal oscillations were observed on an idle spin valve biochip in an

ordinary lab setting (Figure 57a). These oscillations were found to coincide with the

heating/cooling cycles of the air conditioning system. It was also discovered that this

kind of temperature drift affected only real spin valve sensors, but not ordinary

resistors of comparable resistance, as shown in Figure 57a. This was evidence that the

signal drift originated in the sensor, and not in the signal preamplifier electronics.

It was also discovered that the centertone, or sense current applied to the

sensor, exhibited a very similar but much larger amount of drift as shown in Figure

57b. Due to the amplitude modulation scheme described in the principle of operation,

the centertone (500 Hz) is a measure of the average sense current and hence depends

only on the sensor’s average resistance, while the sideband (350 Hz) depends of both

the sense current and the magnetoresistive signal. Upon examination it was found that

the correlation of the sideband (Figure 57a) and centertone (Figure 57b) of spin valve

sensors on the biochip is very consistent. This observation led to the following

development of a temperature correction approach.

Page 124: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

112

Signal (MR) Drift - Corrected with Centertone Signa lS. Osterfeld, Jan-18-2008

-2

-1

0

1

2

0 100 200 300 400 500Minutes

Mic

rovo

lts

Signal (MR) Drift - Corrected with Reference Sensor sS. Osterfeld, Jan-18-2008

-2

-1

0

1

2

0 100 200 300 400 500Minutes

Mic

rovo

lts

A B

Figure 58: The centertone (sense current) drift can indeed be used to correct the signal drift (A). An alternate approach to temperature correction is to subtract a signal from reference sensors (B). A combination of both is also feasible.

The sideband, which contains the magnetoresistance signal, is the primary

signal of interest. While the centertone had been recorded, it had so far largely been

disregarded. However, in light of the observed strong correlation between sideband

and centertone drift, an attempt was made to use the centertone (resistance)

temperature drift to correct the sideband (magnetoresistance) temperature drift. In

particular, if the centertone amplitude is given by CT, and the sideband amplitude by

SB, then temperature correction can be done on an individual sensor basis as follows:

where the exponent α is a correction coefficient which is experimentally

determined, i.e., by cooling the chip with cold spray and optimizing α for optimal

temperature correction. The result of this approach is shown in Figure 58a, which is a

great improvement over the raw data shown in Figure 57a. This approach was

frequently used in real experiments.

However, an alternative approach is to simply subtract a reference signal, i.e.,

from spin valve sensors which are coated with a ~ 300 nm thick oxide. Such sensors

α

=

)(

)(*)()(

0tCT

tCTtSBtSB RawCorrected

Page 125: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

113

will behave exactly like real sensors, except that the thick oxide prevents them from

sensing any nanoparticles. The result of this approach is shown in Figure 58b.

Each approach to temperature correction has advantages and disadvantages.

Using the centertone amplitude to correct the sideband amplitude (Figure 58a) has the

advantage that it works with just a single sensor, and that each sensor can be

individually and separately temperature-corrected. The disadvantage of this approach

is that not only temperature can affect the centertone amplitude. For example, a

slightly unstable contact can also cause the sense current to drift. This would be

misinterpreted as a temperature change, and the corrective response would be

disproportionate. Basically, this type of centertone-based, individual-sensor

temperature correction corrects temperature drift, but it greatly exacerbates other types

of signal drift, e.g., drift which originates from contact resistance changes or shifts in

galvanic potential. If these types of drift occur even in small amounts, then the raw

data frequently looks better than the temperature-corrected data obtained with this

correction method. Another disadvantage of this method is that it only corrects for

temperature drift, but not for drift e.g., in the magnetic tickling field.

Using one or several reference sensors to obtain a correction signal (Figure

58b) has the advantage that it corrects for all sources of signal drift which affect the

entire chip fairly uniformly, such as temperature or magnetic field strength.

Furthermore, unlike the centertone-based method, the reference sensor method does

not further amplify signal errors which stem from unstable contacts or small amounts

of sensor corrosion. This improves the device’s fault tolerance in practice. The

disadvantage of this method is that some sensors need to be dedicated to obtain the

correction signal, which reduces the number of available sensors. Furthermore, if the

reference signal is obtained on a chip which is scanned (time domain multiplexed, one

bank of sensors at a time), then for optimal correction each sensor bank would need to

have its own set of reference sensors to ensure simultaneity of the measurement and

the correction signal.

Page 126: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

114

APPENDIX C – MATHEMATICA CODE

Clear[R,Hx,Hy,Ravg,dRmax,Ht,Hb,Hk,c3,Mp,a,Hpx,Hpy,n,Sensitivity] Hy=Ht+n*Hpy; (* The y-components acting on the free layer *) Hx=Hb+Hk+n*Hpx; (* The x-components acting on the free layer *) R=Ravg+dRmax/2*Hy/Sqrt[Hy^2+Hx^2]; (*Sin[Theta]=Sin[ArcTan[Hy/Hx]=Hy/Sqrt[Hx^2+Hy^2]*) Hpy=c3*Mp[Hb,Ht]*Ht/Sqrt[Ht^2+Hb^2]; (* Particle y-component *) Hpx=c3*Mp[Hb,Ht]*Hb/Sqrt[Ht^2+Hb^2]; (* Particle x-component *) Signal[Hb_,Ht_,Hk_]=n*FullSimplify[Limit[D[R,n],n->0],Ht>0,Hb>0,a>0] Mp[Hb_,Ht_]=Coth[a*Sqrt[Hb^2+Ht^2]]-1/(a*Sqrt[Hb^2+Ht^2]); (* Langevin Function *) a = 0.025; (* Langevin Parameter *) c3 = 0.001; (* Proportionality Constant *) n = 1; (* Particle Coverage Fraction *) dRmax=200; (* Magnetoresistance in Ohms *) Plot[Signal[0,Ht,45],Signal[0,Ht,91],Ht,0,100,AxesLabel->"Tickling Field, Oe (rms)","Signal, microvolts",LabelStyle->Directive[Bold,12]] (* Signal vs. Ht (tickling field) for two different sensors (zero bias field) *)

20 40 60 80 100Tickling Field, Oe HrmsL

50

100

150

200

250

300

S ignal, microvolts

c3 dRmax Hk HHb+ HkL Ht n Mp@Hb, HtD

2 Hb2 +Ht2 IHHb +HkL2 + Ht2M3ê2

Plot3D[Signal[Hb,Ht,15],Hb,0,60,Ht,0,100,AxesLabel->"Hb","Ht","Signal", BoxRatios->1,1,1,MeshFunctions->#3&,#2&,#1&,Mesh->6,3,3, ColorFunction->"Rainbow",LabelStyle->Directive[Bold,12]] (* Signal vs. Hb (bias field) and Ht (tickling field), 3D Plot and Contour Plot *)

Page 127: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

115

BIBLIOGRAPHY

1 Li, Guanxiong, “Spin valve sensor for biomolecular identification: design, fabrication, and characterization”, Ph.D. Dissertation, Stanford University, Department of Materials Science, 2005.

2 Han, Shu-Jen, “CMOS Integrated Biosensor Array Based on Spin Valve Devices”, Ph.D. Dissertation, Stanford University, Department of Materials Science, 2007.

3 Antoniou, A, P D P Pharoah, S Narod, H A Risch, J E Eyfjord, J L Hopper, N Loman, et al. “Average risks of breast and ovarian cancer associated with BRCA1 or BRCA2 mutations detected in case Series unselected for family history: a combined analysis of 22 studies.” American Journal of Human Genetics 72, no. 5 (May 2003): 1117-1130.

4 Fujinami, Robert S, and Thayne L Sweeten. “Letting antibodies get to your head.” Nat Med 9, no. 7 (July 2003): 823-825.

5 Kingsmore, Stephen F. “Multiplexed protein measurement: technologies and applications of protein and antibody arrays.” Nat Rev Drug Discov 5, no. 4 (April 2006): 310-321.

6 Sawyers, Charles L. “The cancer biomarker problem.” Nature 452, no. 7187 (April 3, 2008): 548-552.

7 Schena, M, D Shalon, R W Davis, and P O Brown. “Quantitative monitoring of gene expression patterns with a complementary DNA microarray.” Science 270, no. 5235 (October 20, 1995): 467-470.

8 Caruso, Frank, Elke Rodda, D. Neil Furlong, Kenichi Niikura, and Yoshio Okahata. “Quartz Crystal Microbalance Study of DNA Immobilization and Hybridization for Nucleic Acid Sensor Development.” Analytical Chemistry 69, no. 11 (June 1, 1997): 2043-2049.

9 Johnsson, Bo, Stefan Löfås, and Gabrielle Lindquist. “Immobilization of proteins to a carboxymethyldextran-modified gold surface for biospecific interaction analysis in surface plasmon resonance sensors.” Analytical Biochemistry 198, no. 2 (November 1, 1991): 268-277.

10 Fägerstam, L G, A Frostell-Karlsson, R Karlsson, B Persson, and I Rönnberg. “Biospecific interaction analysis using surface plasmon resonance detection applied to kinetic, binding site and concentration analysis.” Journal of Chromatography 597, no. 1-2 (April 24, 1992): 397-410.

11 Hardenbol, Paul, Fuli Yu, John Belmont, Jennifer MacKenzie, Carsten Bruckner, Tiffany Brundage, Andrew Boudreau, et al. “Highly multiplexed molecular inversion probe genotyping: Over 10,000 targeted SNPs genotyped in a single tube assay.” Genome Research 15, no. 2 (February 2005): 269–275.

Page 128: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

116

12 Zhu, Heng, and Michael Snyder. “Protein chip technology.” Current Opinion in Chemical Biology 7, no. 1 (February 2003): 55-63.

13 Rivas, Luis A., Miriam García-Villadangos, Mercedes Moreno-Paz, Patricia Cruz-Gil, Javier Gómez-Elvira, and Víctor Parro. “A 200-Antibody Microarray Biochip for Environmental Monitoring: Searching for Universal Microbial Biomarkers through Immunoprofiling.” Analytical Chemistry 80, no. 21 (November 1, 2008): 7970-7979.

14 Schweitzer, Barry, Scott Roberts, Brian Grimwade, Weiping Shao, Minjuan Wang, Qin Fu, Quiping Shu, et al. “Multiplexed protein profiling on microarrays by rolling-circle amplification.” Nat Biotech 20, no. 4 (April 2002): 359-365.

15 Prinz, Gary A. “Magnetoelectronics applications.” Journal of Magnetism and Magnetic Materials 200, no. 1-3 (October 1999): 57-68.

16 O'Handley, Robert C. Modern Magnetic Materials: Principles and Applications. Wiley-Interscience, 1999.

17 Wang, Shan X., Seung-Young Bae, Guanxiong Li, Shouheng Sun, Robert L. White, Jennifer T. Kemp, and Chris D. Webb. “Towards a magnetic microarray for sensitive diagnostics.” Journal of Magnetism and Magnetic Materials 293, no. 1 (May 2005): 731-736.

18 Han, S.-J., L. Xu, R.J. Wilson, and S.X. Wang. “A Novel Zero-Drift Detection Method for Highly Sensitive GMR Biochips.” Magnetics, IEEE Transactions on 42, no. 10 (2006): 3560-3562.

19 Guanxiong Li, S.X. Wang, and Shouheng Sun. “Model and experiment of detecting multiple magnetic nanoparticles as biomolecular labels by spin valve sensors.” Magnetics, IEEE Transactions on 40, no. 4 (2004): 3000-3002.

20 Li, G., S. Sun, and S. X. Wang. “Spin valve biosensors: Signal dependence on nanoparticle position.” Journal of Applied Physics 99 (2006): 08P107.

21 Kotitz, R., H. Matz, L. Trahms, H. Koch, W. Weitschies, T. Rheinlander, W. Semmler, and T. Bunte. “SQUID based remanence measurements for immunoassays.” Applied Superconductivity, IEEE Transactions on 7, no. 2 (1997): 3678-3681.

22 Baibich, M. N., J. M. Broto, A. Fert, F. Nguyen Van Dau, F. Petroff, P. Etienne, G. Creuzet, A. Friederich, and J. Chazelas. “Giant Magnetoresistance of (001)Fe/(001)Cr Magnetic Superlattices.” Physical Review Letters 61, no. 21 (November 21, 1988): 2472.

23 Wang, Shan X., and Aleksandr Markovich Taratorin. Magnetic information storage technology. Academic Press, 1999.

24 Baselt, D R, G U Lee, M Natesan, S W Metzger, P E Sheehan, and R J Colton. “A biosensor based on magnetoresistance technology.” Biosensors and Bioelectronics 13, no. 7-8 (October 1, 1998): 731-739.

Page 129: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

117

25 Edelstein, R. L., C. R. Tamanaha, P. E. Sheehan, M. M. Miller, D. R. Baselt, L. J. Whitman, and R. J. Colton. “The BARC biosensor applied to the detection of biological warfare agents.” Biosensors and Bioelectronics 14, no. 10-11 (January 2000): 805-813.

26 Rife, J. C., M. M. Miller, P. E. Sheehan, C. R. Tamanaha, M. Tondra, and L. J. Whitman. “Design and performance of GMR sensors for the detection of magnetic microbeads in biosensors.” Sensors and Actuators A: Physical 107, no. 3 (November 1, 2003): 209-218.

27 Graham, D. L., H. Ferreira, J. Bernardo, P. P. Freitas, and J. M. S. Cabral. “Single magnetic microsphere placement and detection on-chip using current line designs with integrated spin valve sensors: Biotechnological applications.” Journal of Applied Physics 91 (2002): 7786-7788.

28 Schotter, J., P. B. Kamp, A. Becker, A. Pühler, G. Reiss, and H. Brückl. “Comparison of a prototype magnetoresistive biosensor to standard fluorescent DNA detection.” Biosensors and Bioelectronics 19, no. 10 (May 15, 2004): 1149-1156.

29 Joshi, V., Guanxiong Li, S.X. Wang, and Shouheng Sun. “Biochemical stability of components for use in a DNA detection system.” Magnetics, IEEE Transactions on 40, no. 4 (2004): 3012-3014.

30 Li, Guanxiong, Shouheng Sun, Robert J. Wilson, Robert L. White, Nader Pourmand, and Shan X. Wang. “Spin valve sensors for ultrasensitive detection of superparamagnetic nanoparticles for biological applications.” Sensors and actuators. A, Physical 126, no. 1 (2006): 98–106.

31 Shen, Weifeng, Xiaoyong Liu, Dipanjan Mazumdar, and Gang Xiao. “In situ detection of single micron-sized magnetic beads using magnetic tunnel junction sensors.” Applied Physics Letters 86, no. 25 (June 20, 2005): 253901-3.

32 Schmitt, G., J. -W. Schultze, F. Faßbender, G. Buß, H. Lüth, and M. J. Schöning. “Passivation and corrosion of microelectrode arrays.” Electrochimica Acta 44, no. 21-22 (June 1, 1999): 3865-3883.

33 Faßbender, F., G. Schmitt, M. J. Schöning, H. Lüth, G. Buß, and J. -W. Schultze. “Optimization of passivation layers for corrosion protection of silicon-based microelectrode arrays.” Sensors and Actuators B: Chemical 68, no. 1-3 (August 25, 2000): 128-133.

34 Li, Guanxiong, Vikram Joshi, Robert L. White, Shan X. Wang, Jennifer T. Kemp, Chris Webb, Ronald W. Davis, and Shouheng Sun. “Detection of single micron-sized magnetic bead and magnetic nanoparticles using spin valve sensors for biological applications.” Journal of Applied Physics 93, no. 10 (May 15, 2003): 7557-7559.

35 Sandhu, A., and H. Handa. “Practical Hall sensors for biomedical instrumentation.” Magnetics, IEEE Transactions on 41, no. 10 (2005): 4123-4127.

Page 130: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

118

36 Koh, Ai Leen, Wei Hu, Robert John Wilson, Shan Xiang Wang, and Robert Sinclair. “TEM analyses of synthetic anti-ferromagnetic (SAF) nanoparticles fabricated using different release layers.” Ultramicroscopy 108, no. 11 (October 2008): 1490-1494.

37 Fu, Aihua, Wei Hu, Liang Xu, Robert J Wilson, Heng Yu, Sebastian J Osterfeld, Sanjiv S Gambhir, and Shan X Wang. “Protein-functionalized synthetic antiferromagnetic nanoparticles for biomolecule detection and magnetic manipulation.” Angewandte Chemie (International Ed. in English) 48, no. 9 (2009): 1620-1624.

38 De Palma, Randy, Chengxun Liu, Francesca Barbagini, Gunter Reekmans, Kristien Bonroy, Wim Laureyn, Gustaaf Borghs, and Guido Maes. “Magnetic Particles as Labels in Bioassays: Interactions between a Biotinylated Gold Substrate and Streptavidin Magnetic Particles.” The Journal of Physical Chemistry C 111, no. 33 (2007): 12227-12235.

39 Kodera, Yoh, Ayako Matsushima, Misao Hiroto, Hiroyuki Nishimura, Asako Ishii, Tomoo Ueno, and Yuji Inada. “Pegylation of proteins and bioactive substances for medical and technical applications.” Progress in Polymer Science 23, no. 7 (November 1998): 1233-1271.

40 Thorsen, Todd, Sebastian J. Maerkl, and Stephen R. Quake. “Microfluidic Large-Scale Integration.” Science 298, no. 5593 (October 18, 2002): 580-584.

41 Schmidt, M.A. “Wafer-to-wafer bonding for microstructure formation.” Proceedings of the IEEE 86, no. 8 (1998): 1575-1585.

42 Wei, J., H. Xie, M. L. Nai, C. K. Wong, and L. C. Lee. “Low temperature wafer anodic bonding.” Journal of Micromechanics and Microengineering 13, no. 2 (2003): 217-222.

43 McDonald, J. Cooper, David C. Duffy, Janelle R. Anderson, Daniel T. Chiu, Hongkai Wu, Olivier J. A. Schueller, and George M. Whitesides. “Fabrication of microfluidic systems in poly(dimethylsiloxane).” Electrophoresis 21, no. 1 (2000): 27-40.

44 Garra, J., T. Long, J. Currie, T. Schneider, R. White, and M. Paranjape. “Dry etching of polydimethylsiloxane for microfluidic systems.” Journal of Vacuum Science and Technology-Section A-Vacuum Surfaces and Films 20, no. 3 (2002): 975-982.

45 Leclerc, E., Y. Sakai, and T. Fujii. “PDMS dry etching for higher integration of cells culture microchambers: application to Hep G2 cells.” In Microprocesses and Nanotechnology Conference, 2002. Digest of Papers. Microprocesses and Nanotechnology 2002. 2002 International, 276-277, 2002.

46 Schuck, Peter, and Allen P. Minton. “Analysis of Mass Transport-Limited Binding Kinetics in Evanescent Wave Biosensors.” Analytical Biochemistry 240, no. 2 (September 5, 1996): 262-272.

47 Kusnezow, Wlad, Yana V. Syagailo, Sven Rüffer, Konstantin Klenin, Walter Sebald, Jörg D. Hoheisel, Christoph Gauer, and Igor Goychuk. “Kinetics of antigen binding to antibody

Page 131: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

119

microspots: Strong limitation by mass transport to the surface.” Proteomics 6, no. 3 (2006): 794-803.

48 Kusnezow, Wlad, Yana V Syagailo, Igor Goychuk, Jörg D Hoheisel, and David G Wild. “Antibody microarrays: the crucial impact of mass transport on assay kinetics and sensitivity.” Expert Review of Molecular Diagnostics 6, no. 1 (January 2006): 111-124.

49 Schmid, H., and B. Michel. “Siloxane Polymers for High-Resolution, High-Accuracy Soft Lithography.” Macromolecules 33, no. 8 (April 1, 2000): 3042-3049.

50 Thalladi, V., J. Love, G. Whitesides, and T. Odom. “Generation of 30-50 nm structures using easily fabricated, composite PDMS masks.” Journal of the American Chemical Society 124, no. 41 (2002): 12112-12113.

51 Connell, T G, N Curtis, S C Ranganathan, and J P Buttery. “Performance of a whole blood interferon gamma assay for detecting latent infection with Mycobacterium tuberculosis in children.” Thorax 61, no. 7 (July 1, 2006): 616-620.

52 Li, Yiwen, Nidhi Nath, and W. Monty Reichert. “Parallel Comparison of Sandwich and Direct Label Assay Protocols on Cytokine Detection Protein Arrays.” Analytical Chemistry 75, no. 19 (October 1, 2003): 5274-5281.

53 Birken, Steven, Peter Berger, Jean-Michel Bidart, Matthias Weber, Adrian Bristow, Rob Norman, Catharine Sturgeon, and Ulf-Hakan Stenman. “Preparation and Characterization of New WHO Reference Reagents for Human Chorionic Gonadotropin and Metabolites.” Clin Chem 49, no. 1 (January 1, 2003): 144-154.

54 Sturgeon, C.M., and A.R. Ellis. “Standardization of FSH, LH and hCG--Current position and future prospects.” Molecular and Cellular Endocrinology 260-262 (January 2, 2007): 301-309.

55 Korhonen, Juha, Henrik Alfthan, Pekka Ylostalo, Johannes Veldhuis, and Ulf-Hakan Stenman. “Disappearance of human chorionic gonadotropin and its alpha- and beta-subunits after term pregnancy.” Clin Chem 43, no. 11 (November 1, 1997): 2155-2163.

56 Osterfeld, Sebastian J., Heng Yu, Richard S. Gaster, Stefano Caramuta, Liang Xu, Shu-Jen Han, Drew A. Hall, et al. “Multiplex protein assays based on real-time magnetic nanotag sensing.” Proceedings of the National Academy of Sciences 105, no. 52 (December 30, 2008): 20637-20640.

57 Gaster, Richard S, Drew A Hall, Carsten H Nielsen, Sebastian J Osterfeld, Heng Yu, Kathleen E Mach, Robert J Wilson, et al. “Matrix-insensitive protein assays push the limits of biosensors in medicine.” Nat Med 15, no. 11 (November 2009): 1327-1332.

58 Xu, Liang, Heng Yu, Michael S Akhras, Shu-Jen Han, Sebastian Osterfeld, Robert L White, Nader Pourmand, and Shan X Wang. “Giant magnetoresistive biochip for DNA detection and HPV genotyping.” Biosensors and Bioelectronics 24, no. 1 (September 15, 2008): 99-103.

Page 132: A MICROFLUIDIC BIOCHIP BASED ON MAGNETORESISTIVE … › file › druid:px491tp4561...BASED ON MAGNETORESISTIVE DETECTION OF NANOPARTICLES A DISSERTATION ... This file is an open-access

120

59 Mak, Andy C., Sebastian J. Osterfeld, Heng Yu, Shan X. Wang, Ronald W. Davis, Olufisayo A. Jejelowo, and Nader Pourmand. “Sensitive Giant Magnetoresistive-based Immunoassay for Multiplex Mycotoxin Detection.” Biosensors and Bioelectronics, accepted for publication, DOI 10.1016/j.bios.2009.11.028.

60 Li, Guanxiong , and S. X. Wang. “Analytical and micromagnetic modeling for detection of a single magnetic microbead or nanobead by spin valve sensors.” IEEE Transactions on Magnetics 39 (September 1, 2003): 3313-3315.

61 Liu, Y., W. Jin, Y. Yang, and Z. Wang. “Micromagnetic simulation for detection of a single magnetic microbead or nanobead by spin-valve sensors.” Journal of Applied Physics 99 (2006): 08G102.

62 Kakay, A., M. W. Gutowski, L. Takacs, V. Franco, and L. K. Varga. “Langevin granulometry of the particle size distribution.” Journal of Physics A Mathematical and General 37, no. 23 (2004): 6027-6041.