characterizing magnetic particle transport for

108
CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR MICROFLUIDIC APPLICATIONS ASHOK SINHA Dissertation submitted to the Faculty of Virginia Polytechnic Institute and State University in partial fulfillment for the requirements of the degree of Doctor of Philosophy In Engineering Mechanics Doctoral Committee Dr. Ishwar K. Puri, Chair Dr. Mark S. Cramer Dr. Muhammad Hajj Dr. Michael W. Hyer Dr. Mayuresh J Patil September 25, 2008 Blacksburg, Virginia Keywords: Microfluidics, magnetic microparticles, manipulation, magnetic separation

Upload: others

Post on 02-Oct-2021

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

MICROFLUIDIC APPLICATIONS

ASHOK SINHA

Dissertation submitted to the Faculty of Virginia Polytechnic Institute and State University in partial fulfillment for the requirements of the degree of

Doctor of Philosophy In

Engineering Mechanics

Doctoral CommitteeDr. Ishwar K. Puri, Chair

Dr. Mark S. Cramer Dr. Muhammad Hajj Dr. Michael W. Hyer Dr. Mayuresh J Patil

September 25, 2008 Blacksburg, Virginia

Keywords: Microfluidics, magnetic microparticles, manipulation, magnetic separation

Page 2: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

MICROFLUIDIC APPLICATIONS

Abstract

Magnetic particles with active functional groups offer numerous advantages for

use in μ-TAS (Micro Total Analytical Systems). The functional site allows chemical

binding of the particle with the target species in the fluid sample. Selection of the

functional group establishes the target molecule and vice versa under assumptions of

highly specific biding. The particles hence act as mobile reaction substrates with high

surface to volume ratios owing to their small size. The concept of action at a distance

allows their use as agents for separation in microchannels based on relatively simple

design. It is possible to manipulate magnetic particles and bound target species using an

externally applied magnetic field. Hence, the particles can be effectively separated from

the flow of a carrier fluid. Magnetic fields create dipolar interactions causing the particles

to form interesting structures and aggregates. Depending upon the applied field, the

microstructure evolution of the aggregate is interesting in its own right, e.g. related to

improvements in material properties and bottom-up self assembly. The shape of the

aggregates can be determined a priori if the interaction between the particles is well

characterized. The dominant competing forces that influence magnetic particle dynamics

in a flow are magnetic and viscous. There are a number of physical parameters such as

viscosity, magnetic susceptibility, fluid velocity, etc. which are varied to study their

individual effects.

Initially dilute suspensions are studied experimentally and numerically using a

particle based dynamics approach. Once established, a force model for particle

ii

Page 3: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

iii

interaction is proposed for concentrated suspensions. Since the Lagrangian particle

tracking algorithm returns positions of the particles, the proceeding work focuses on

studying the dynamics of these particles.

Thereafter, a mathematical model is proposed for functionalization between

magnetic and non-magnetic particles. Having characterized the collection of magnetic

particles, the effect of relative concentrations is investigated on the collection of the non-

magnetic types.

Page 4: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

iv

Dedicated to my father, and to my (late) mother who always wanted to attend my defense

and grill me with tough questions.

Page 5: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

v

Acknowledgements

My deepest gratitude is due to Professor Ishwar K. Puri, without whose role, this

thesis and the work presented herein would be far inferior to its present form. The

journey, though long and arduous, has been extremely rewarding in all senses through his

excellent mentoring. His positive outlook on problems of research, and everyday in

general, has been inspiring and an unfaltering guiding light.

Sincere thanks are due to Professors Mark S. Cramer, Michael W. Hyer,

Muhammad R. Hajj and Mayuresh J. Patil for serving as critical analysts of my thesis and

members of my doctoral committee.

I have thoroughly enjoyed and valued working with Dr. Ranjan Ganguly. His role

as a co-mentor and much more, has been immensely helpful. The enormity of his

contribution to this work is priceless.

The staff at ESM has been of immense help in ensuring that my concerns remain

purely academic.

The general spirit in the lab was persistently kept elevated by the best colleagues

Anindya, Soumik, Sayan and Ganesh. The profusion of constructive criticism reflects in

every aspect of this work. Sayan’s camaraderie especially has taught me lessons for life

that aren’t taught in academic environments.

I consider myself to be extremely fortunate to have had such an exciting and

enriching experience, and be part of the ESM family.

AS.

Page 6: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

vi

Table of contents

Chapter 1: Magnetic Particles, their properties and fundamentals of manipulation..... 1

1.1. Magnetism and Microfluidics...................................................................................................... 1 1.2. Motivation ................................................................................................................................... 5 1.3. Background ................................................................................................................................. 5 1.4. Forces on a magnetic particle .................................................................................................... 16 1.5. Conclusion................................................................................................................................. 21 1.6. Thesis organization.................................................................................................................... 21

Chapter 2: Magnetic Separation ................................................................................. 23 2.1. Magnetic Separation.................................................................................................................. 23 2.2. Related work ............................................................................................................................. 26 2.3. Challenges ................................................................................................................................. 28 2.4. Thesis motivations and objectives............................................................................................. 29 2.5. Concepts Proposed .................................................................................................................... 29

Chapter 3: Magnetic Separation in dilute suspensions ............................................... 32 3.1. Introduction ............................................................................................................................... 32 3.2. Background ............................................................................................................................... 33 3.3. Methodology ............................................................................................................................. 34 3.4. Conclusion................................................................................................................................. 46

Chapter 4: Magnetic separation from concentrated suspensions incorporating dipolar interactions 47

4.1. Introduction ............................................................................................................................... 47 4.2. Dipolar Interactions in Concentrated Suspensions .................................................................... 49 4.3. Overlap Prevention Model ........................................................................................................ 53 4.4. Experimental Setup ................................................................................................................... 57 4.5. Results and Discussion.............................................................................................................. 61 4.6. Conclusion................................................................................................................................. 72

Chapter 5: Capture of Non-magnetic particles through interaction with functionalized magnetic particles ............................................................................................................. 74

5.1. Introduction ............................................................................................................................... 74 5.2. Model ........................................................................................................................................ 75 5.3. Results and Discussion.............................................................................................................. 80 5.4. Conclusions ............................................................................................................................... 87

Chapter 6: Conclusions............................................................................................... 88 6.1. Current Trends........................................................................................................................... 89 6.2. Societal Impacts ........................................................................................................................ 90 6.3. Outlook and Perspectives .......................................................................................................... 90 6.4. Recommendations for future work ............................................................................................ 91

Bibliography ..................................................................................................................... 92

Page 7: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

vii

List of Figures

Figure 1.1: (a) Structure of a magnetic microparticle. (b) SEM micrograph 3 μm size

particles ............................................................................................................................... 6

Figure 1.2: Hysteresis behavior of a ferromagnetic material (Becker,1982).................... 11

Figure 1.3: Hysteresis-free B-H curve for superparamagnetic particles (Becker,1982)... 12

Figure 1.4: Effect of particle size on saturation magnetization as predicted by the

Langevin equation............................................................................................................. 14

Figure 1.5: Dependence of particle magnetization on the applied magnetic field and the

volume fraction of magnetite, calculated using Langevin function. The parameters

considered were: Msat = 450,000 Am-1, di = 10nm, Kb = 1.38μ10-23JK-1, T= 300K and

μo=10-7 Hm-1 ..................................................................................................................... 14

Figure 1.6: Velocity profile in the cross section of a channel with a large aspect ratio. The

color bar represents a normalized velocity being highest in the central section and rapidly

diminishing only close to the edges. ................................................................................. 18

Figure 2.1: Schematic diagram of flow-through separation ............................................. 25

Figure 3.1: Schematic of the 2D domain .......................................................................... 37

Figure 3.2: Image of capillary and electromagnet during the experiments. ..................... 38

Figure 3.3: Magnetic field (Guass) in the domain of interest. .......................................... 38

Figure 3.4: Velocity profile in the microchannel.............................................................. 38

Figure 3.5: Particle trajectories ......................................................................................... 41

Figure 3.6: Variation of Ф with П1 for constant values of П2 = ymag/h (= 0.75). ............. 42

Figure 3.7: Variation of Ф with respect to П2 for a specified value of П1 (=12250).

Region shows that increasing values of ymag have a strong influence on Φ. In contrast,

in region Diminishing values of h have a strong influence on Φ................................. 44

Figure 4.1: (a) chains of 1 μm diameter magnetic micro-spheres under a homogeneous

magnetic field. (b) Pair-wise particle dipole-dipole interaction model. ........................... 50

Figure 4.2: Variation of t with θ for three different (L/2) cases. All other parameters in ψ

are kept fixed..................................................................................................................... 52

Page 8: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

viii

Figure 4.3: Parametric dependence of aggregation time t as functions of η, L, B, χ and a.

The simulations are performed for θ = 45o. All the plots collapse on a single curve when

the aggregation time is plotted against the group variable Ψ=(ηL5)/(χB2a5). ................... 52

Figure 4.4: Increase in V for r < R as predicted by Eq. 2. As R →1, the collision model

shifts from a soft to a hard sphere model.......................................................................... 54

Figure 4.5: (a) Figure 2(a): Schematic of particle-particle dipole interaction model and of

the (b) simulation domain. ................................................................................................ 55

Figure 4.6: (a) Side view (z=0 plane), (b) End view (x=l/2 plane) and (c) Bottom view

(y=-h/2 plane) of the profile of the magnetic field created by a point dipole. The color bar

represents the magnetic field strength in milliTesla (mT) ................................................ 55

Figure 4.7: Schematic of the separation scheme............................................................... 58

Figure 4.8: Schematic of imaging system. Inset shows the position of the needle magnet

with respect to the channel cross section .......................................................................... 58

Figure 4.9: (above): Experimental setup showing position of magnet relative to the

section of the microchannel, (below) variation By with increasing separation between

magnet and channel........................................................................................................... 60

Figure 4.10: Time varying aggregation of magnetic microspheres extracted from the flow

by the localized magnetic trap for y = 0mm .................................................................... 61

Figure 4.11: Particle aggregation varying with time and distance of needle tip from

channel .............................................................................................................................. 62

Figure 4.12: The position of the electromagnet tip relative to the microcapillary ........... 63

Figure 4.13: (a) Spatial distribution of By for I = 0.5A. Field measurement is made by a

Hall probe Gaussmeter. (b)Variation of the magnetic field component By coil along the

electromagnet core axis (where Bx= Bz= 0) for four different currents up to 1mm above

the electromagnet tip......................................................................................................... 65

Figure 4.14 (a) Experimentally-obtained images of growing aggregates at various time

intervals (seconds) for a = 2 μm, I = 0.5A, U= 1mm/s, s = 125 μm, c = 1012/m3. Bar =

200 μm. (b) Side view, (c) End view and (d) Bottom view of representative snapshots of

the simulated particle aggregation at steady state............................................................. 66

Figure 4.15: Aggregate growth rate with s=125μm and variable I. The inset shows the

relationship between the collection rate and current intensity.......................................... 68

Page 9: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

ix

Figure 4.16: Effect of change in dipole strength m on the number of collected particles.

The inset shows linear relation between slope of Figure 8(a) and dipole strength m....... 69

Figure 4.17: Linear relation of P with Π. The inset shows variation of P with changing m.

Simulation conditions are a = 500 μm, χ=0.1, Uo= 10 μm/s, s = 15 μm, φ = 0.7, c =

2×1015/m3 .......................................................................................................................... 70

Figure 5.1: Schematic of the computation domain ........................................................... 76

Figure 5.2: (a-c) Quality of the magnetic field generated due to the point dipole. (d-f)

Snapshots of the simulation for the base case................................................................... 77

Figure 5.3: Effect of varying R on the collision model................................................... 78

Figure 5.4: Resultant interaction model predicting three regimes, attraction, repulsion and

no interaction .................................................................................................................... 79

Figure 5.5: Aggregate size vs. time. The curves passing through the shaded area labeled

as M refer to nM, while those passing through the corresponding N area represent nN. The

inset shows the behavior of c(t). (b) Time averaged aggregate size n* during washaway

phase and its dependence on c0. The secondary axis shows the corresponding time taken

for the aggregate to reach critical size. Uo = 10 μm/s, m = 9×10-12 A-m2........................ 81

Figure 5.6: (a) The critical aggregate size prior to washaway vs. the relative concentration

of M and N-particles. (b) The rapid increase in the standard deviation of an aggregate

during washaway phase for the two particle types ........................................................... 84

Figure 5.7: (a) T Variation in P*N and P*

N as a function of (a) dipole strength m, Uo = 10

μm/s (b) Peak centerline velocity U0, m = 9×10-12 A-m2. c0 = 0.36 and is held fixed in

both cases (a) and (b) ........................................................................................................ 87

Page 10: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

1

Chapter 1: Magnetic Particles, their properties and fundamentals of manipulation

1.1. Magnetism and Microfluidics

1.1.1. History of magnetism Magnetism dates back to historic times (Whittaker, 1910). Over the ages, the

concepts of magnetism have time and again been utilized in ever emerging applications

with tremendous success. Today, magnetism is well-known to be coupled with electricity

and prevalent in most aspects of human life. It is a science well developed and

characterized by the Maxwell laws.

1.1.2. Microfluidics Microfluidics, on the other hand, is a field of scientific research that has only

recently become popular (Gravesen et al., 1993). This is mainly due to the development

in MicroElectroMechanicalSystems (MEMS), which has been a spin-off from the

semiconductor manufacturing technologies (Ho and Tai, 1998). Thus effective techniques

for fabrication of miniaturized features have only recently been combined with

microfluidic fabrication. The modes of generating magnetic fields have evolved

gradually, from using cheap and simple external permanent magnets, to sophisticated

embedding of microelectromagnets during the fabrication process.

1.1.3. First appearance of microfluidics The first commercial microfluidic based systems were introduced for life science

applications (Regnier et al., 1999). In particular, microfluidic technologies have been

developed to carry out chemical and biochemical analyses. In biotechnology and

biochemical processing, the need to manipulate fluids moving in narrow channels has

stimulated several new research areas, such as the development of new microfabrication

Page 11: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

2

methods for fluidic systems, implementation of novel components for the assembly of

complex microfluidic devices and the study of the fundamental behavior of fluids in

narrow-bore channels (Whitesides, 2001). Since the introduction of photolithographic

techniques for the fabrication of chemical and biochemical microdevices (Manz et al.,

1992), the number of applications has increased exponentially (Vladislav Dolník, 2000).

In pharmaceutical and bioanalytical research, components for microdevices have widely

been developed for proteomics, genomics, clinical diagnostics and drug discovery.

1.1.4. Miniaturization In general there is a movement towards smaller, simpler, smarter devices in most

areas of our lives. Perhaps the most obvious examples are found in the area of electronics

such as cell phones, computers, copiers, printers, and other “information” technologies.

The impact of Personal Digital Assistants (PDAs) and personal laser printers and copiers

is remarkable when considered from the perspective of just 20 years ago. Moreover, these

developments towards miniaturization have been largely driven by developments in the

power and density of integrated electronic circuits (Moore, 1965). Optical elements and

their controls found in bar code scanners, digital cameras, and personal CD players have

followed the same trend of smaller, simpler, and smarter.

1.1.5. Microfluidic chips As a result, a new technology and a new discipline has made its appearance:

microfluidics. As microfluidic chips become closer in size to the cells and molecules on

which they work, so too must the channels, pumps and valves that manipulate the

solutions of reagents and substrates. A chip intended to handle thousands of micro

quantities of samples must be able to store, aliquot and mix these reagents and then

provide some means of measuring the activity of interest.

Page 12: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

3

1.1.6. Multidisciplinary nature of microfluidics Microfluidics typically requires a multi-disciplinary approach for building

effective systems through the interface of physics, chemistry, engineering and

biochemistry. With the added advantage of reduced size and weight, these devices are

capable of performing sample handling with an efficiency paralleled by laboratory scale

analytical devices. Flow and reaction channels are of the order of ten’s of microns. The

reduced size of the system implies reduced sample sizes (of the order of nanoliters or

less) and warrants reduced sampling time (Greenwood and Greenway, 2002). Typical

analytical microdevices are glass-, silicon- or polymer-based planar chips ranging in

overall size from the millimeter to the centimeter scale with individual structures (e.g.

separation channels) at the micrometer scale.

1.1.7. Advantages of magnetic field over electric field Electric fields have previously been utilized in microfluidic applications, such as

capillary electrophoretic separations, electroosmotic pumping and dielectrophoretic

trapping (H. A. Stone, 2001). Magnetic fields, however, were initially employed

relatively rarely, despite the great advantages they could offer. For example, objects

inside a microfluidic channel can be manipulated by an external magnet that is not in

direct contact with the fluid. Target molecules can be isolated from a sample by attaching

them to small magnetic particles which are then recovered using an external magnetic

field. In contrast to electric manipulation, magnetic interactions are generally not affected

by surface charges, pH, ionic concentrations or temperature.

1.1.8. Fusion of magnetism and microfluidics

The fusion of Magnetism and microfluidics thus has been relatively recent. The

potential of conveniently incorporating particulate nature in the relatively calm Stokes

Page 13: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

4

flow of microchannels is no longer a daunting challenge, liable to blocking the channels.

The basic structure of the magnetic field remains as it was long ago; however, the

generating magnets and electromagnets are now more realistically feasible than ever

before in a variety of sizes and shapes.

Microfluidic researchers have been somewhat reluctant to place objects such as

magnetic particles into microchannels. Furthermore, ‘useful’ magnetic microparticles

functionalized with antibodies or other biomolecules have only become off-the-shelf

products in the last few years.

The avenues created by this fusion have been all-pervasive in a variety of fields,

not only limited to physics and engineering. In fact, one of the first applications of

magnetic particles was in clinical biosciences, even before microchannels first made their

appearance. Magnetic particles have been used as chaperones to manipulate other non-

magnetic objects through their surface chemical interactions and subsequent magnetic

collection.

Some new applications stemming from this fusion include pumping and mixing of

fluids, as well as the incorporation of switches and valves into lab-on-a-chip devices

(Pamme, 2006). Magnetic forces are used for transport, positioning, separation and

sorting of magnetic as well as non-magnetic objects. Bio-assays have been performed on

the surface of magnetic particles trapped inside a microchannel. More recently, on-chip

detection techniques based on magnetic forces have been investigated and basic research

of magnetic behavior, not possible on the large scale, has been undertaken in the confined

space of microchannels. The theme of the following chapters will develop upon some

Page 14: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

5

experimental and numerical approaches to systematically characterize the behavior of

these magnetic microparticles when used in microfluidic environments.

1.2. Motivation The use of superparamagnetic iron oxide nanoparticles is interesting in its own

right for a number of reasons. These particles have no residual magnetism upon removal

of the magnetic field. With no magnetic memory, the particle motion is highly

predictable. This is an advantage, especially in complicated processing methods (Gould,

2004).

Another advantage of action at a distance allows their use in microchannels with

relatively simple design. It is possible to manipulate the particles using an external field.

Hence, the particles may be effectively separated from a carrier fluid which otherwise

flows steadily. This allows using reagents from various sources, directed to the site, as

needed by the process while arresting the motion of the particles with a magnetic field.

Further, as mentioned previously, due to low residual magnetism, the particles readily

redisperse into the flow upon removal of the field. Motion of an ensemble of magnetic

microspheres in a microfluidic channel can also be controlled in a manner that the

particle-fluid hydrodynamic interaction augers mixing.

The inherent benefits offered by magnetic handling includes: reduced reagent

costs, elimination of labor intensive steps, easy automation, high purity and decreased

processing time compared to conventional methods.

1.3. Background

1.3.1. Structure of particles and their conventional applications Micron sized magnetically responsive particles generally consist of nanometer

size (fine) particles distributed through a non magnetic matrix. Figure 1.1(a) is a

Page 15: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

6

schematic view of such magnetic particles, where fine magnetic nanoparticles are

encapsulated in a supporting matrix to form the composite microparticle. In biological

applications the matrix is composed from inert organic polymers, silica, or ceramics. The

role of this matrix is not only to assemble the particle core, but also to protect the

magnetic nanoparticles against the external environment influences such as oxidation.

Moreover, in view of biological use, the matrix material must present the desired surface

chemistry that allows, for instance, coating the particles’ surface with specific affinity

biomolecules. Selection of the functional group establishes the target molecule and vice

versa under assumptions of highly specific binding. These functional groups are shown as

surface extensions in the schematic of Figure 1.1(a). The particles act as mobile reaction

substrates with high surface to volume ratios owing to their small size.

(a) (b)

Figure 1.1: (a) Structure of a magnetic microparticle. (b) SEM micrograph 3 μm size particles

The magnetic nanoparticles are frequently composed from magnetic iron oxide

and are in the 1−100 nm size range. Magnetite γ−Fe2O4, or more complex alloys

MnFe2O4 or (MO)n·Fe2O3 (where M represents another component such as Co, Li, Zn, Ni

etc.), represent a well-known and important class of iron oxide materials(Sun et al.,

2004). The particular interest in this class of materials originates from their easy synthesis

Page 16: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

7

and mass production potential, as well as from the possibility to control the magnetic

properties by adjusting the molecular composition. The nanoparticles are generally

produced in aqueous suspensions, called ferrofluids or magnetic fluids. The particles are

covered with a shell of surfactant molecules that will stabilize the suspension and hinders

the particles from flocculation and sedimentation. The present technology allows the

realization of substantially monodisperse (less than 10% dispersion) and highly stable

ferrofluid suspensions (Liu et al., 2000, Kang et al., 1996). Figure 1.1(b) is a SEM

micrograph showing monodisperse magnetic particles of 3 μm diameter (from Kisker

Biotech), composed of magnetite dispersed in a polymer matrix. The nanostructure

composition of magnetic microparticles leads to new properties and functions that are not

achievable in simple bulk materials (Koneracká et al., 1999).

1.3.1.1. Biological applications In the field of biochemistry and biomedical diagnostics, magnetic particles with

bioactive coatings play a dominant role in the preparation, separation and detection of

biological molecules (such as DNA and proteins) because of their advantageous

characteristics, which makes them highly efficient in mild operating conditions (Šafařík

and Šafaříková, 2002). The first characteristic is related to their controllable size, ranging

from a few nanometers up to tens of nanometers, which places them at dimensions that

are smaller than or comparable to those of a cell (10 − 100 μm), a virus (20 − 450 nm), a

protein (5 − 50 nm) or a gene (2 nm wide and 10 − 100 nm long).

1.3.1.2. Drug discovery/nucleic acids exploration Magnetically responsive particles are used as a magnetic solid-phase support for

tagging and carrying target biomolecules. The main application area of magnetic

separation in drug discovery and diagnostics is sample preparation including high

Page 17: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

8

throughput genome isolation for sequencing or polymerase chain reactions to carry out

genotyping (Hawkins et al., 1997). For instance, the separation of nucleic acids from the

complex mixtures in which they are often found is necessary before other processes like

sequencing, amplification, hybridization and detection. The presence of large amounts of

cellular or other contaminating material like proteins in such complex mixtures often

impedes many of the reactions and techniques used in molecular biology. After an

incubation time, the target nucleic acids are bound onto the surface of the magnetic

particles. The particles/nucleic acids complex is then separated magnetically from the

liquid solution.

Magnetic particles are now increasingly used as carriers for binding proteins,

enzymes and drugs (Schuster et al., 2000). They have been used for various immuno-

assays including fluorescence or enzyme immunoassays (Yazdankhah et al., 1999). In

these assays, magnetic particles with primary or secondary antibodies are used for

separation and quantification of antigens. Nowadays, the use of magnetic particles for

cell separation has become a very popular method in cell biology (SafarIk and

SafarIková, 1999). The general approach involves the use of paramagnetic particles

coated with antibodies against the specific target cell surfaces. For instance, magnetic

particles are proving to be an extremely efficient sampling tool in the detection and

removal of circulating tumor cells from peripheral blood (Udo Bilkenroth, 2001,

Luxembourg et al., 1998). In such immunomagnetic cell separations, both positive and

negative selection can be performed. In case of positive selection targets, cellular subsets

are magnetically labeled and subsequently separated, while negative selection involves

target purification by removing all other contaminating cells.

Page 18: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

9

1.3.2. Magnetic characteristics Instead of the ferromagnetic behavior observed in bulk magnetic materials, the

nanometer magnetic composition of a magnetic particle can lead to the apparition of a

superparamagnetic behavior (Ganguly, 2005). The latter is characterized by the fact that

the particles do not retain any magnetism after removal of the magnetic field. This feature

is very important since it permits an easy magnetization using an external magnetic field

and a redispersion that prevents the agglomeration of the particles after the magnetic field

is removed.

Another issue in manipulating magnetic particles at the micro-scale is the

agglomeration of the particles. For instance, manipulating magnetic particles in a liquid

flow microchannel necessitates a continuously applied high magnetic field and magnetic

field gradient. In such conditions, the particles being superparamagnetic, agglomerate

affecting the system behavior.

1.3.2.1. Field Effects at the microscale When utilizing external magnetic fields, the particles and the soft magnetic

structures of the channel are magnetized. The latter thereby provide local magnetic field

gradients that lead to the capture of the magnetic particles. This is similar to the high

gradient magnetic separators (Dijkhuis and Kerkdijk, 1981) (HGMSs) used for large

scale magnetic separation, which comprise a separation column filled with a steel wool

matrix in a large external magnetic field. However, a significant difference to large scale

HGMSs is that in most microsystems the magnetic material is located outside the

separation channel and thus not in direct contact with the fluid sample. The magnetic

forces due to microscopic magnetic structures are inherently short-ranged. One solution

to overcome this limitation has been to integrate systems for magnetic separation with

Page 19: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

10

microfluidic mixer structures that force the particles closer to the magnetic structures

resulting in substantial improvements of the capture efficiency.

1.3.2.2. Magnetism Magnetic fields are produced due to electric current. For a solenoid of length l and

having n turns that carries a current I through the conductor the magnetic field inside the

coil (having n/l turns per unit length and air or vacuum in its core) is represented by

B=μ0nI/l along the axis of the coil. The quantity nI/l is purely a function of the coil

geometry and the current, and is denoted by the variable H(A/m). The introduction of a

magnetizable material inside the solenoid enhances the magnetic field. This happens

because the H produced by the current carrying coil induces a magnetic dipole moment

M per unit volume in the material, and it adds to H to produce the magnetic field

B=μ0(M + H). M=(m/Vm) is the induced magnetic moment density of the material, Vm

being the material volume. For a linear homogeneous and isotropic material the

relationship between M and H is linear and given by:

M = χmH (1.1)

where χm is a function of |H|, material property, and temperature, and is called

the magnetic susceptibility (or simply the susceptibility). Thus, the magnetic field inside

a material can be represented as

B = μ0(1 + χm)H (1.2)

H ≡ μ0μrH (1.3)

where μ0 = 4π×10−7 Henry/m, χm is a material dependent quantity, called the

magnetic susceptibility and where μr is the relative magnetic permeability of the medium.

Static magnetic fields as described by the Maxwell equations are represented as

Page 20: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

11

∇ × M = 0 => ∇ × H = 0 (1.4)

∇ ⋅ B = 0 = > ∇ ⋅ H = 0 (1.5)

In the absence of an external magnetic field, the individual particle dipoles are

randomly oriented due to thermal agitation. However, upon application of H, individual

particles experience instant magnetization inducing the alignment of the thermally

disoriented magnetic moments of these particles along the induced H.

In ferromagnetic materials, however, the relationship between the three vector

fields H, B and M is generally non-linear and history-dependent. Therefore, simple

relationships such as those of Eq. (1.2) and Eq. (1.3) are not justified. In this case, the

material is characterized by a magnetic hysteresis loop, as shown in Figure 1.2.

Figure 1.2: Hysteresis behavior of a ferromagnetic material (Becker, 1982)

A hysteresis loop is characterized by the coercive field Hc, the remnant

(spontaneous) magnetization Br, and the saturation magnetization Bs (the maximum

magnetization of a magnetic particle in a strong field).

Page 21: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

12

1.3.3. Hysteresis An important effect to emphasize in magnetic particle behavior is the size effect

on the hysteresis shape. Figure 1.2 presents the qualitative behavior of the coercivity of

nanoparticles as the particle size changes. In large particles (of the order of a micron or

more), energetic considerations favor the formation of domain walls which results in a

multi-domain structure of the particle (Koneracká et al., 2002). In this case, a small field

is sufficient to move the domain walls, which explains the relatively low coercive field.

As the particle size decreases, the formation of domain walls becomes energetically

unfavorable and the particle shows a single domain structure. In this case, any change of

magnetization can only occur through coherent spin rotation, an operation that

necessitates generally a larger magnetic energy.

Figure 1.3: Hysteresis-free B-H curve for superparamagnetic particles (Becker, 1982)

As the particle size continues to decrease, the thermal energy becomes of the

same order as the magnetic energy, so that the spins are increasingly affected by the

thermal fluctuations (Kittel, 1946). At this scale (of the order of tens of nanometers or

less), where the spin of the particle is free to fluctuate in response to thermal energy, the

Page 22: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

13

coercive field is substantially equal to zero, leading to an anhysteretic, but still sigmoidal,

B−H curve as shown in Figure 1.3. Such behavior, related to the nanosized structure of

the particles, is termed as superparamagnetic behavior (Brown, 1963).

Since the nanoparticles immobilized in the non magnetic matrix show

superparamagnetic nature, magnetization of the particles and the ferrofluid bulk obey the

Langevin’s formula

|M| = Msatφ(cot(α)-1/ α) (1.6)

where ϕ is the volume fraction of magnetic particles in the fluid, Msat the bulk

saturation magnetization of the particle material and α is the Langevin parameter, which

is defined as follows

30 sat

B

M4 a3 k T

μπα =

H (1.7)

Figure 1.4 and Figure 1.5 present the behavior of saturation of a

superparamagnetic particle with varying size and volume fraction respectively.

Page 23: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

14

0.0

0.2

0.4

0.6

0.8

1.0

0.0E+00 1.0E+05 2.0E+05 3.0E+05 4.0E+05 5.0E+05H(A/m)

M/ ϕ

Msa

t

2a = 5nm2a = 10nm2a = 15nm2a = 20nm

Figure 1.4: Effect of particle size on saturation magnetization as predicted by the Langevin equation

0

5000

10000

15000

20000

0 25 50 75 100 125 150H(kA/m)

Mp(

A/m

)

φ=1%

φ=2%

φ=3%

φ=4%

φ=5%

Figure 1.5: Dependence of particle magnetization on the applied magnetic field and the volume fraction of magnetite, calculated using Langevin function. The parameters considered were: Msat = 450,000 Am-1, di =

10nm, Kb = 1.38μ10-23JK-1, T= 300K and μo=10-7 Hm-1

Page 24: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

15

1.3.4. The composite microparticle The previous analysis shows that the underlying physics of nanocomposite

magnetic particles is determined by the domain size effects. For single domain

nanoparticles, the magnetic behavior is mainly related to the energy ratio β = KmV /kBT

between the magnetic barrier energy of spin reversal Em = KmV and the thermal energy

ET = kBT, where Km is the magnetic anisotropy, V the volume of the particle, kB

Boltzmann constant and T the temperature (Verdes et al., 2002). The magnetic properties

dependence on β can be more empirically expressed by a characteristic temperature,

called blocking temperature (TB), below which β > 1. Below this temperature, the spin of

the fine particles will not fluctuate (be blocked) and the microparticle will present the

classical ferromagnetic hysteresis behavior of Figure 1.2. This temperature depends

obviously on the fine particles size but it can also be affected by their concentration

within the matrix and the interparticle interactions.

For weakly interacting particles within the matrix, the magnetic induction of the

microparticles is the sum of the magnetic inductions of the nanoparticles of which it is

composed. Having a large volume fraction (ϕ) is, therefore, crucial to obtain particles

with a good magnetic response (large susceptibility) to an external magnetic field. With

current fabrication technologies, a volume fraction around 50% is achievable.

Another factor that considerably influences the microparticles’ susceptibility is

the demagnetization effect. Roughly, this effect is related to the self-generated magnetic

field in the surrounding space, which has a tendency to demagnetize the microparticle.

Due to this effect, the magnetic particles will have an effective magnetic susceptibility

given by

χeff =χm(1 + Nd χm) (1.8)

Page 25: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

16

where χm = (∂B/∂H|H=0) is the magnetic susceptibility of the composite particle

and Nd the demagnetization factor, equal to 1/3 for a spherical particle . As a result, the

magnetic susceptibility of the microparticles is generally small, and χeff ≤ 1 for

superparamagnetic particles.

1.4. Forces on a magnetic particle

1.4.1. Magnetic forces The magnetic force is obtained using an “effective” dipole moment approach

where the magnetized particle is replaced by an “equivalent” point dipole with a moment

m. For magnetic particles that are diluted in a non magnetic suspension such as pure

water, one can approximate equation 1.6 by considering χeff ≤ 1 and therefore μr ≈ 1.

Fm = (m· ∇ )B (1.9)

It is interesting to emphasize in equation (2.9) the role played by the magnetic

field gradient ( ∇ B) which will attract the magnetic particles to the regions where the

magnetic field is higher. The higher the value of this gradient, the more the force exerted

on the particle is important. This fact is the basis of magnetic particle manipulation for

various applications.

Forces in a force field such as electric or magnetic are short range forces. For

instance a charged particle in an electric field experiences a force that decays as 1/r2. On

the other hand magnetic forces diminish rapidly as 1/r7. It is evident that the magnetic

dipole-dipole interaction is proportional to 1/r7. As the interparticle separation increases,

this effect rapidly decreases (but here r denotes the distance between two magnetic

particles). It is clear from this discussion that magnetic forces are much stronger at the

microscale than at larger scales, making magnetic applications attractive for microscale

or microfluidic applications.

Page 26: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

17

Fm effm ax x y z i y j z x y z k

x y z= −

+ + + + + +

+ +

FHGG

IKJJ4

5 4 52

03

2 2 2 3 2 2 2

2 2 2 5μ χ πc h c h

c h$ $ $

(1.10)

Using the analytical expression for magnetic body force on an isolated single

particle in Eq. 1.10, using parameters typical of microfluidic environments, we get Fm ~

10-12N.

1.4.2. Viscous force on a particle Fluid flow in small channels exhibits different behavior from that at the

macroscopic scale. There are several excellent comprehensive reviews on fluid

mechanics in micro/nanoscale (Lofdahl and Gad-el-Hak, 1999, Stone et al., 2004, Squires

and Quake, 2005). The Reynolds number (Re) is the most often mentioned dimensionless

number in fluid mechanics. The Re number, defined by ρUL/μ, represents the ratio of

inertial forces to viscous ones. In most circumstances involved in micro- and

nanofluidics, the Re number is at least one order of magnitude smaller than unity, ruling

out any turbulence flows in micro/nanochannels. Microfluidic channel dimensions

typically range from 1 to 1000 mm in width and height. Frequently, only one channel

dimension is within that range; this is sufficient to provide laminar flow characteristics.

Inertial force plays an insignificant role in microfluidics, and as systems continue to scale

down, it will become even less important. For such small Re number flows, the

convective term of Navier–Stokes equations can be dropped. Without this non-linear

convection, simple micro/nanofluidic systems have laminar, deterministic flow patterns.

They have parabolic velocity profile in pressure-driven flows, plug-like velocity profile

in electroosmotic flows, or a superposition of both. One of the benefits from the low Re

number flow is that sensitive material can be transported easily without shearing in lab-

on-a-chip devices. It is worth noting, however, that other mechanisms, such as capillary

Page 27: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

18

effects, viscoelasticity, and electrokinetic effects may complicate the flow phenomena

since their non-linearity increases in small scales (Squires and Quake, 2005).

For laminar flow the exact solution for flow is given by the infinite Fourier series

summation resulting from solving the pressure Poisson equation (Ichikawa et al., 2004).

The velocity distribution in a cross section of the rectangular channel with the boundary

condition of u=0 at y=±h/2 or z=±w/2 as

u fn

y z Ghn y

hn b

h

n zh

i,cosh

coshcos $b g

b g

b gb g

=FHGIKJ −

+LNM

OQP

+LNM

OQP

RS||

T||

UV||

W||

+LNM

OQP=

∑2

308

32 1

2 1

2 12

2 1η π

π

ππ

(1.11)

The resulting velocity profile across the cross section of a microchannel is shown

in Fig. 1.6.

Figure 1.6: Velocity profile in the cross section of a channel with a large aspect ratio. The color bar represents a normalized velocity being highest in the central section and rapidly diminishing only close to

the edges.

The concept of magnetic manipulation consists of the use of magnetic force to

separate or to transport magnetic particles suspended in a liquid. Efficiently manipulating

magnetic particles is, however, not easily realized, since the magnetic force must

overcome the hydrodynamic drag force imposed by the surrounding liquid. In the case of

a spherical magnetic particle, the viscous drag force is given by (Happel, 1983)

Fd = 6πηa (uf-up) = 6πηaΔv (1.12)

yz

Page 28: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

19

where η is the viscosity of the fluid (water in our case), a is the radius of the

magnetic particle and vf ,v0 are the velocities of the surrounding fluid and the particles

respectively.

1.4.3. Scaling of some other forces in comparison with Fm and Fd and assumptions The possible short-range interactions between particles (double-layer and entropic

repulsion, London–van der Waals interaction, adsorption or chemical bridging) have been

neglected in this study because they are negligible when compared with the strong

magnetic interactions taking place during the aggregation process (Kleef, 1984).

Other forces commonly encountered in particle flows are gravity, added mass,

and Brownian effects. We shall show through scaling analysis that in the present

scenario, magnetic and viscous forces are the dominant forces. The other forces are

insignificant in comparison.

1.4.3.1. Gravity For a 1 μm diameter microsphere, the particle drag force ≈ 8.4 pN with a fluid

viscosity of 0.00089 Pa-S and a relative particle velocity of 1 mm/s. The gravitational

force ≈ 0.004 pN (assuming ρp=1800 kg/m3 and ρl=1000 kg/m3) is negligible in

comparison. The downward velocity of a reference particle through fluid is about 4×10-5

m/s, or two to three orders of magnitude less than the magnetically driven motion.

Gravity does become important with larger particles.

1.4.3.2. Diffusion Typically the diffusivity of water soluble molecules (e.g., dyes or ions) is ~10-9

m2/s. For larger particles that have a 1-10 μm diameter (e.g., microparticles) D ~ 10-13-10-

14 m2/s, and their corresponding diffusion time ~ 107-108 s. Diffusion of particles of this

size is insignificant. The Nernst–Einstein relation predicts a diffusion constant (Warnke,

Page 29: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

20

2003) of 5.4×10-9cm2/s. This value yields diffusion fluxes approximately six orders of

magnitude lower than the flux resulting from the magnetic force.

1.4.3.3. Brownian effects To investigate the importance of Brownian effects, we consider the dimensionless

dipole strength

λ πμ χ= 0

3 202

9a Hk TB

(1.13)

which compares the magnetic energy interaction between the particles (where a is

the particle radius and χ is the magnetic susceptibility) in an external magnetic field (H0)

to the thermal energy. For instance, for a suspension of particles with a = 0.5 μm, χ = 0.1

subjected to a field strength H0 = 2000 A/m at room temperature, λ = 100>>1. In this

case, the magnetic interactions between particles greatly exceed the thermal energy, so

Brownian motion does not play a significant role. When the magnetic field strength

decreases and approaches zero, the factor λ becomes lower and consequently the thermal

Brownian motion will dominate the magnetic interaction, initiating resuspension.

Additionally, the characteristic Brownian walk is given by

l

d k T

B

B

=

π

πη

0

63

(1.14)

The characteristic walk length turns out to be on the order of 100 nm. This is

negligible in our case where the particle size itself is O[3] larger than the Brownian walk.

Eq. 1.13 compares the effects of Brownian motion with viscous stabilization.

Page 30: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

21

1.4.3.4. Inertia The particle inertia force can be likewise neglected for the small particle mass

(~0.94 pg) unless the particle acceleration exceeds an improbably large value

(~1000m/s2).

When the magnetic field is turned on, the particle will accelerate, since the

viscous drag force Fd is smaller than the magnetic force Fm. When the velocity reaches a

critical value for which the magnetic (2.9) and hydrodynamic forces (2.10) are equal, the

particle (with magnetic volume Vm = (4/3) πa3) will move with a constant velocity: Δv.

Since particle motion within the fluid reaches equilibrium almost immediately, the

magnetic force on the particle is equal to the force exerted by that particle on the fluid.

1.5. Conclusion These physical properties of magnetic particles offer many potential applications

and enable new features in the fields of biology, biotechnology and other bio-medical

disciplines (Grady, 2002, Jiles, 1991). The fundamental underlying physics of magnetic

particles in liquids have been discussed in this chapter. Further details, covering different

aspects of the subject, can be found in one of the many extensive textbooks and review

papers (Morrish, 1965). In addition, contemporary bench-top applications are listed.

In microfluidic environments, these magnetic microspheres offer a viable option

for MEMS based biosensors by miniaturizing the above listed techniques.

1.6. Thesis organization

After a brief introduction to the structures and physics of magnetic microparticles,

Chapter 2 discusses magnetic separation and the various methods associated with it.

Chapter 3 presents results pertaining to investigating dilute suspensions. Treating dilute

suspensions as a suspension of non-interacting isolated particles, a parametric

Page 31: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

22

investigation reveals their dynamics in a Stokes flow. The proposed model for a point

dipole is studied in comparison with experimentally obtained particles trajectories.

Inter-particle interactions are included in the dynamical mode in Chapter 4. The

presence of a magnetic entity in a magnetic field distorts the field locally. This lays the

basis for magnetic interaction between particles. Results obtained from a numerical

analysis are presented in comparison with experimental separation in a microcapillary.

In Chapter 5, a mathematical model is presented for functionalization and

interaction between magnetic and non-magnetic particles. The collection of target species

is characterized numerically.

Chapter 6 concludes the thesis with sections on current trends, societal impacts

and recommendation for future work.

Page 32: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

23

Chapter 2: Magnetic Separation 2.1. Magnetic Separation

Magnetic particles can be manipulated using an external magnetic field,

independently from the motion of the surrounding biological liquid. In fact, when

subjected to an external magnetic field, magnetic particles acquire a strong magnetic

moment that effectively reacts to an external magnetic field gradient. The principle of

operation of magnetic separation devices is the interaction between magnetic forces and

competing gravitational, hydrodynamic, and interparticle forces within the magnetic

separator.

2.1.1. Industrial separation For a long time, magnetism has stood out as an interesting and important driving

force to separate and manipulate magnetic components from non-magnetic components

within a mixture (Oberteuffer, 1974). Enrichment of low grade iron ore, removal of

metallic impurities from large volumes, waste water treatment or removal of weekly

magnetic impurities from kaolin clay are typical examples of industrial applications of

magnetic separation.

2.1.2. Clinical Separation Since the 1980’s, magnetic separation techniques found potential applications in

the bioscience fields (Olsvik et al., 1994). The idea consisted of using magnetically

labeled biochemical components for the purification and the extraction of target

biomolecules using a high external magnetic field gradient. To achieve these goals,

spherical magnetic particles (also termed as magnetic particles) with bio-active surface

coatings were used to capture (or to tag) a specific biological entity on a magnetic

Page 33: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

24

particle surface. Subsequently the complex magnetic particle/target biomolecule would

be manipulated using an external magnetic filed. The technique has been shown to be

efficient for the isolation of certain cells, bacteria and viruses from fluids such as blood,

and this principle has found several clinical applications. Subsequent to the enrichment

process, identification is then accomplished by routine optical methods. Thus, magnetic

separation techniques for biomolecules manipulation have enjoyed a great resurgence of

interest and have become widely used in biology and other bio-medical disciplines.

2.1.3. Modes of Magnetic separation Achieving separation is not easily realized since the magnetic force must

overcome the hydrodynamic drag force imposed by the fluid. Magnetic separation

devices are composed from an arrangement of permanent magnets or electromagnets in

proximity with the fluid handling parts. A variety of magnetic separators have been

proposed that can take the form of very simple particle concentrators composed of a test

tube manually brought in the proximity of a permanent magnet, to very complicated and

expensive automated devices. Available solutions can be divided in two categories,

namely batch and flow-through magnetic separators.

2.1.3.1. Batch-type magnetic separators Batch-type magnetic separators are usually made from strong rare-earth

permanent magnets that are brought in proximity with a test tube containing a liquid

solution to cause attraction and aggregation of particles at the tube wall followed by

removal of the liquid. Some separators have a removable magnet plate to facilitate easy

washing of magnetic particles. Test tube magnetic separators enable separation of

magnetic particles from volumes ranging approximately from 5 μl to 50 ml. However, as

the sample volume increases, the separation and accumulation of the particles becomes

Page 34: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

25

slow and difficult. The first issue in magnetic separation is the generation of a high

magnetic field and magnetic field gradient extending over the active (test tube) zone. This

is a difficult task to achieve in a large volume. The separation velocity of the magnetic

particles decreases with the increase of the distance between the magnetic poles.

Separation becomes increasingly inefficient as the sample volume (fixed by the volume

of the internal space formed by the magnets arrangement) becomes larger.

2.1.3.2. Flow-through magnetic separators A flow through-based magnetic separator system is often preferable to increase

the separator efficiency for large liquid volumes, as it offers the possibility of

continuously pushing the sample through a relatively "small" region with a magnetic field

gradient.

Figure 2.1: Schematic diagram of flow-through separation

The latter must be strong in order to overcome the drag force imposed by the flow

carrying the magnetic target. For this purpose, a series of permanent magnets may be

arranged along the channel as a source of a high radial magnetic force that will attract the

magnetic target toward the channel walls. Moreover, flow-through systems have been

demonstrated to be efficient solutions for the separation of magnetic substances based on

differences of their magnetophoretic mobility (Moore et al., 2001). In this case,

separation is achieved by a differential migration of magnetically susceptible species

Page 35: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

26

across the flow streamlines. This concept is exploited in a process known as ‘fluid flow

fractionation’, in which the fluid is split at the outlet into fractions containing tagged cells

or proteins with differing magnetophoretic mobilities (Hoyos et al., 2000).

The flow in the channel in Figure 2.1 carries magnetic substances (black circles)

that will be deviated and captured by the magnetic force generated by the magnets, while

the non-magnetic substances (white circles) continue to flow freely in the channel. There

are three electromagnets placed along length of the channel.

2.2. Related work Many areas in microfluidic applications involve manipulation of particles in a

controllable manner. In the design of microscale devices for cell sorting, cell analysis or

cell removal from a sample, it is important to predict the motion of the individual

particles in response to the local flow conditions. For this purpose, various principles and

methods have been developed in microsystems, such as the optical tweezers (Furst,

2005), electrokinetic methods (Binyamin et al., 2003), acoustic means and

dielectrophoresis (Ravula et al.).

More recently, focus has shifted toward fabricating self-assembled structures

using paramagnetic particles carried by liquids in microchannels. Simulation of such

particulate microflows could play an important role in the development and active control

of dynamically reconfigurable self-assembled structures.

In order to determine the mechanics of paramagnetic microparticles it is desirable

to characterize the particles and the forces that act on them. In the absence of a magnetic

field the particles form a colloidal suspension, stabilized against permanent aggregation

by addition of surfactants or surface treatment of the particles. Brownian motion tends to

Page 36: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

27

disperse the particles but they aggregate to form aggregates once a magnetic field is

applied.

The behavior of particles interacting through induced magnetic fields is complex

and their dynamics are not well understood. For example, the long-range nature of the

particle interaction can persist even when they form chains or columns. While there has

been some progress toward understanding the dynamics of magnetic chains and columns,

a majority of the studies have focused on empirical curve fitting. Experimental imaging

has shown (Ivey et al., 2000) various microstructure formation depending upon applied

field. There is an abundance of experimental studies (Furst et al., 1998) conducted to

characterize the dynamic behavior of these particles with regards to changes in

experimental operating parameters.

Previous experimental work in systems subject to external rotating magnetic

fields (Helgesen et al., 1990, Meakin and Skjeltorp, 1993) demonstrate that these systems

may show a very rich dynamics depending on parameters, such as the amplitude and

frequency of the rotating field, and the viscosity of the carrier fluid. Video microscopy

shows (Melle et al., 2003) a crossover Mason number (ratio of viscous to magnetic

forces) above which a rotating field prevents the particle aggregation to form chains.

Models of dynamic collision have been proposed based on experiments performed to

study effects of viscous dissipation and the increased pressure in the interparticle gap as

compared with dry collisions (Yang and Hunt, 2006).

Existing models for predicting influence of the dipole-dipole interaction on the

dynamics of monodispersive ensembles of magnetic nanoparticles have been studied by

Monte Carlo simulations (Andersson et al., 1997, Satoh et al., 1996). A continuum based

Page 37: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

28

model (Warnke, 2003, Grief and Richardson, 2005, Christian Mikkelsen, 2005) provides

bulk properties of the fluid flow but lacks information on individual particle dynamics.

Simplistic Stoksian dynamics simulations have been conducted but are limited in

either geometry (Liu et al., 2005) or in the complexity of the applied field. Other studies

have focused on studying isolated (Furlani, 2007) particles hence missing the dynamics

of interparticle interactions.

2.3. Challenges Manipulating magnetic particles within a microsystem is one part of the arduous

effort toward the miniaturization of biochemical analysis. The use of particles in

microfluidic systems further increases the chemical interaction area per unit volume in

surface-based biochemical assays. At the same time, it allows down-scaling of

conventional analysis techniques to the micron scale. The issues of microscale

manipulation of magnetic particles are mainly related to both the integration of magnetic

inductive components within the microfluidic circuits and the generation of a well

controlled magnetic field and magnetic field gradient. The simplest magnetic particle

handling system is achieved by simply placing an external magnet or an electromagnet in

proximity with a microfluidic chip. However, an effective control and manipulation of

magnetic particles on a small scale requires a local control of the magnetic field and

magnetic field gradient. These conditions are hardly achieved by using a simple external

permanent magnet or electromagnet, and the integration of magnetic "micro-inductive

components" within the microfluidic systems seems to be necessary to achieve a more

accurate manipulation of magnetic particles.

Miniaturization, in conjunction with integration of multiple functionalities can

enable the construction of structures that exceed the performance of traditional

Page 38: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

29

macroscopic systems, can provide a plethora of new functionalities and offer the potential

of low-cost mass production (Jensen, 2001).

2.4. Thesis motivations and objectives The motivation for this thesis is to characterize transport and manipulation of

magnetic particles in microfluidic flows. Through the proposed concept, some solutions

are sought to the problems related to the integration of the magnetic inductive

components within microfluidic network. The understanding of the physical aspects of

particle aggregation is a critical point in any future advance in microfluidic magnetic

particles manipulation. To address this question, specific attention will be given to the

physics of magnetic particle agglomerations structures, when a suspension of the particles

is subjected to an external field.

2.5. Concepts Proposed

2.5.1. Magnetic Separation in dilute suspensions Sensor miniaturization is motivated by the need to create smaller comprehensive

µ-TAS devices that are capable of simultaneous sample handling and analytical

measurements. These devices typically require that an analyte be brought into contact

with a reagent in microfluidic channels. One method to collect analytes is by bonding

them with functionalized particles and subsequently introducing these conjugates into a

reagent flow. Functionalized magnetic particles can be remotely steered using an

externally generated magnetic field and collected for processing.

The motion of magnetic microparticles is examined in a microfluidic flow under

the influence of a nonuniform external magnetic field in order to characterize their

collection. The operating parameter space for optimal magnetic particle collection

efficiency is characterized through a coupled experimental and numerical investigation in

Page 39: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

30

which simulated particle trajectories are compared with measurements and validated. The

parameters include the magnetic field strength, particle size, magnetic susceptibility, host

fluid velocity and viscosity, and the characteristic length scale. The efficiency of a

magnetic collector in a microfluidic device depends upon two dimensionless numbers

that compare the magnetic and particle drag forces. The magnetic force depends upon the

imposed magnetic field and its gradient provided dipole-dipole interactions with

neighboring particles are neglected (which is appropriate for dilute particle suspensions).

2.5.2. Magnetic separation from concentrated suspensions incorporating dipolar interactions

The magnetophoretic separation and field-induced aggregation of magnetic

particles under the combined influence of magnetic and fluid shear forces in a

microfluidic channel is investigated. Magnetic forces arise due to an externally applied

field gradient as well as due to induced dipolar interactions in a non-dilute suspension.

This interaction between the particles aids the separation process. By varying the physical

parameters, a strategy is demonstrated to control the collection of these particles. In order

to characterize this phenomenon, a parametric analysis is conducted via numerical

simulation. A dimensional group composed of these same physical variables emerges as

the key parameter to predict the dynamics of the aggregate. Additionally, experimental

images verify some key aspects of the model. Through experimental and numerical

investigation, the aggregate is shown to grow linearly with time. A parametric

investigation reveals the relationship of the aggregate size with a dimensional parameter

composed of the relevant physical variables. The results are important in design of

magnetophoretic separators of MEMS-based cell sorters, biosensors, nucleic acid

Page 40: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

31

isolators, etc. where it is important to know the magnetophoretic separation efficiency

and the integrity of the separated cluster of magnetic microparticles.

2.5.3. Capture of Non-magnetic particles through interaction with functionalized magnetic particles

Immunomagnetic separation (IMS) is an immunology application in which

magnetic particles with surface immobilized antibodies specifically bind with antigens.

The possibility of massively parallel IMS microfluidic networks could potentially enable

more robust remote monitoring. However, there are no comprehensive models available

to characterize the interaction between target species and magnetic particles in

microfluidic channels. We investigate the mechanics of magnetic separation of target N-

particles by magnetic M-particles using three-dimensional numerical simulations.

Interparticle and clumping effects are accounted for explicitly. In the absence of a well-

characterized viscoleastic model of the forces in functionalizing strands, a velocity effect,

producing unconditional binding similar to that in covalent bonds, is employed. M-

particles are attracted towards the point dipole and aggregate close to it during buildup.

Particles on the periphery of the aggregate experience a holding force due to the magnetic

field and a depleting drag force. Depending upon the conditions, washaway occurs

through the shearing of individual particles or larger aggregate chunks. From the

perspective of practical device design, the separation section concentrates the interacting

M- and N-particles into an aggregate. Some bound M-N conjugates are released from this

aggregate through washaway and conveyed to subsequent sections for analysis. While the

M-N ratio in the washed away material is constant, the washaway frequency and the

amount of washed-away material determine the rapidity of sensing and the signal

strength.

Page 41: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

32

Chapter 3: Magnetic Separation in dilute suspensions 3.1. Introduction

Single-particle micromanipulation techniques traditionally have been developed

for biophysical applications, but they are being increasingly employed in materials

science applications such as rheology and polymer dynamics. Continuing developments

and improvements in single-molecule manipulation techniques afford new opportunities

in a broad range of fields. An innovative method of single-particle manipulation is

presented.

Along with experimental and theoretical studies, numerical simulation has been

an indispensable tool in almost every research and application field for many years. It

also aids in the design of microfluidic and nanofluidic devices. Simulations allow

researchers to rapidly determine how design change will affect chip performance, thereby

reducing the number of prototyping iterations. However, there are several factors that

complicate the numerical simulation of micro/nanoscale phenomena and thus distinguish

it from the macroscale counterpart. Undoubtedly, hydrodynamic interactions in the

context of magnetic separation in microfluidic systems employing dilute suspensions is

important. However, a comprehensive Eulerian-Lagrangian formulation for such a

problem is outside the scope and the time line of this study.

The focus of the subsequent work is to manipulate matter (particulate as well as

fluid) at the microscale. By particulate matter, it is implied that the magnetic particles

functionalized with appropriate functional groups can bind with other species present in

Page 42: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

33

the fluid media. Subsequently the external field can guide the assembled combination

through the microchannel through various processing steps.

Because of the small physical dimensions of microfluidic flows, it is difficult to

insert sensors that do not disrupt the flow. Thus experimental techniques often measure

bulk, instead of local properties. To obtain information about individual particle

dynamics, a simplified Stokesian dynamics approach is proposed. The outcome will also

predict the formation of microstructure or aggregates when the particles interact amongst

themselves in the microfluidic environment. Given the proper operating variables, it

should be possible to predict the formation of one of the two dominant formations –

chains (corresponding to a uniform field) and hexagonal aggregates (possible due to

anisotropic field).

3.2. Background The collection efficiency of a magnetic collector depends upon two dimensionless

numbers that compare the magnetic and particle drag forces (Sinha et al., 2007).

Experimental investigations have so far only provided qualitative demonstrations of how

microfabricated electromagnets can generate a strong enough force field to ensure

particle capture (K. Smistrup, 2005, Rida and Gijs, 2004) or separation (Friedman and

Yellen, 2005), or enable MEMS-based cell sorters (Jiang et al., 2006). While useful,

these investigations do not provide a quantitative particle-tracking model that is

applicable regardless of configuration and geometry. Previous simulations of magnetic

particle transport in microfluidic contexts (Smistrup et al., 2005) have also not focused

on individual particle trajectories or detailed parametric analyses. An analysis on the

microfluidic transport and capture of 1μm diameter magnetic microspheres provided by

continuum approach (Christian Mikkelsen, 2005) is questionable, since in water such a

Page 43: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

34

treatment is generally valid only for Fe3O4 particles diameters smaller than 40 nm

(Furlani and Ng, 2006). Limited parametric investigations on particle transport in the

context of microfluidics and magnetic drug targeting have also been conducted by other

groups (Furlani, 2007) but such studies have not identified the key nondimensional

parameters that influence magnetophoretic particle capture in different but geometrically

similar systems.

Studying the motion of single particles is equivalent to studying how dilute

suspensions behave. In order to obtain attain unambiguous information about the motion

of isolated particles in a microchannel flow an external magnetic field is used to direct

and steer the particles across the microchannel. Numerical simulations are performed to

predict particle collection, and are validated through experiments. Next, parametric

simulations illustrate the effects of flow and magnetic parameters on the particle

collection efficiency in the microchannel. The investigation concerns both the transport

of an isolated particle or of very dilute suspensions for which the dipole-dipole

interaction among adjacent particles is negligible.

3.3. Methodology

3.3.1. Theoretical Without resort to a detailed analysis of the separation process, an understanding

of the characteristics of magnetic separators is possible by consideration of a simple

force-balance model. A mathematical model is developed for predicting the transport and

capture of the carrier particles taking into account the dominant magnetic and fluidic

forces. The fluidic force is predicted using Stokes’ law for the drag on a sphere in a

laminar flow field.

Page 44: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

35

As a first step, the translational motion of magnetic particles is examined in a

microfluidic flow under the influence of a nonuniform external magnetic field. Their

collection is characterized in terms of the magnetic field strength, particle size, magnetic

susceptibility, host fluid velocity and viscosity, and the characteristic length scale. The

Newtonian relation for particle motion in terms of its absolute velocity up, is given by the

well known relation

ma = Fm + Fd + Fg (3.1)

The right hand side of Eq. 3.1 contains the magnetic, drag, and gravity-induced

forces, i.e.,

F H Hm a= FHGIKJ +LNM

OQP

⋅∇μ π χχ0

343 1 3

b g (3.2)

F u ud f pa= −6πη d i (3.3)

F gg p la= FHGIKJ −

43

3π ρ ρd i (3.4)

The particle inertia force can be likewise neglected for the small particle mass

(~0.94 pg) unless the particle acceleration exceeds an improbably large value (~1000

m/s2). The particle drag force thus balances the magnetic force, i.e.,

Fm = – Fd (3.5)

u F up m lay z= +

16πη

,b g (3.6)

From Eq. 3.6, the particle position at any instant is

x u xt t dt tpt

t t

+ = ++

zδδ

b g b g (3.7)

Page 45: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

36

which we solve for using a forward differencing time-marching scheme. For

solution stability, the differencing time interval δt is chosen such that the particle does

not travel more than 1/1000th of the channel height in one time step, i.e.,

( u p t a× ≤δ / 10b gμm.

3.3.2. Experimental Setup The experiments follow the motion of magnetic microspheres suspended in a

distilled water flow within a microchannel. A pressure-driven flow is established in

borosilicate capillary tubes of rectangular (100 µm x 2000 µm) cross section and 100 mm

length by maintaining a fixed level difference between the reservoirs at the ends of the

channel. (A 7 mm height difference institutes a peak centerline velocity of 1 mm/s.) The

host water flow is laden with a very dilute concentration of 1 µm diameter magnetic

polystyrene microspheres (PMSi-H1.0-5 Kisker-Biotech). The magnetic particles are

steered by applying a nonuniform magnetic field that is established by an electromagnet

with a tapered tip (that approximates a point dipole) and an adjustable current coil. A

schematic diagram of the experimental configuration is presented in Figure 1.1. The

hatched section represents the experimental viewing window. The bottom offset includes

the channel thickness and displacement of the outer channel edge from the electromagnet

tip.

The magnetic field distribution is mapped using a standard Gaussmeter and is

shown in Figure 3.3. Since the microfluidic channel has a dimension comparable to the

spatial resolution of the Gaussmeter probe (~1mm), the field distribution inside the

channel could not be evaluated directly from its readings. Instead, magnetic field

measurements are made over a larger 5 mm×5 mm domain and the local field strength

Page 46: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

37

and gradient at the channel location are subsequently obtained through interpolation. The

measured field strength is 1300 Gauss at y = 0, and rapidly diminishes to 300 Gauss at

the top of the 500 µm wide region of interest. The strength and distance of an equivalent

point dipole used in the numerical simulations is so chosen that its field and gradient at

the channel mid-plane match the corresponding measured values. Although the

electromagnet used in the experiment has a finite size, the point dipole assumption

accurately predicts the experimentally established field in the region close to the tip when

the dipole offset distance ymag (see Figure 3.1) and strength are suitably adjusted. Direct

images are obtained using a digital zoom microscope coupled with an additional

objective lens at 100X magnification.

Figure 3.1: Schematic of the 2D domain

Y

Z

-1.0 E-3m

0

1.0 E-3 m

umax = 1mm/s

500µm

y = 1000µm

ymag = 750µm

Magnetic dipole

rmag

Page 47: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

38

Figure 3.2: Image of capillary and electromagnet during the experiments.

Figure 3.3: Magnetic field (Guass) in the domain of interest.

Figure 3.4: Velocity profile in the microchannel

3 mm

Y

Z

rmag

Micro-capillary

Page 48: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

39

3.3.3. Particle Tracking The transport of magnetic microspheres in microfluidic channels is influenced by

both hydrodynamic and the magnetic forces. Dipole-dipole interactions between

neighboring particles are neglected, which is appropriate for dilute particle suspensions.

The Brownian force is generally negligible for larger particles. The model considers one-

way coupling between the fluid and the particles in the dilute suspension, i.e., only the

particle trajectory is influenced by drag due to the fluid and not vice versa.

A point dipole of strength m is used to represent the magnetic field where the

position vector r′r is measured from the dipole. Here, the dipole aligned along the y

direction ( ˆm mj=r

) is placed at a location ymag below the channel as shown in Figure 3.2

producing the field shown in Figure 3.3. This location corresponds to the position of the

tapered magnetic tip in the experiments. We assume a fully developed pressure-driven

flow that has a gradient –dp/dx = G through the rectangular channel that has a height h

and a width b. The fluid velocity (Ichikawa et. al, 2004) is given by the relation is

illustrated in Figure 3.4.

u fn

y z Ghn y

hn b

h

n zh

i,cosh

coshcos $b g

b g

b gb g

=FHGIKJ −

+LNM

OQP

+LNM

OQP

RS||

T||

UV||

W||

+LNM

OQP=

∑2

308

32 1

2 1

2 12

2 1η π

π

ππ

(3.8)

For the experimental capillary tube aspect ratio of 1:200, the velocity profile is

nearly parabolic in the x-wise direction and exhibits an approximately top-hat profile in

the cross stream y-direction. or a 1 μm diameter microsphere, the particle drag force ≈8.4

pN with a fluid viscosity of 0.00089 Pa S and a relative particle velocity of 1 mm/s. The

Page 49: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

40

gravitational force ≈0.004 pN (assuming ρp=1800 kg/m3 and ρl=1000 kg/m3) is negligible

in comparison.

Fm effm ax x y z i y j z x y z k

x y z= −

+ + + + + +

+ +

FHGG

IKJJ4

5 4 52

03

2 2 2 3 2 2 2

2 2 2 5μ χ πc h c h

c h$ $ $

(3.9)

where (x′,y′,z′) denotes the coordinates with respect to the origin of the dipole (located at

(xmag, -ymag, zmag)). ( )1eff i i31⎡ ⎤χ = χ + χ⎣ ⎦ denotes the effective susceptibility of the spherical

particles.

Figure 3.5 presents a comparison between predictions and measurements for

particles released from the inlet at five separate positions. The simulations consider

particles of 1 μm size, particle magnetic susceptibility of 0.1, fluid viscosity of 0.00089

Pa s, and average fluid velocity of 1 mm/s through the channel. The corresponding pairs

of measured and simulated trajectories are numbered 1 through 5. The measured

trajectories show that the optimal combination of modeled values of m and the magnet

displacement from the channel centerline rmag are 2.28×10-7 A-m2 and 1.75×10-4 m,

respectively. For all cases, the predictions and measurements are in generally good

agreement, although there are small local deviations, since the model of the magnetic

field is an idealization while, due to its greater complexity, the experimentally established

field cannot be adequately described through a simple analytic relation. Predictions of the

field gradient experienced by particles that lie closer to the magnet slightly exceed their

experimentally established values. Hence, the simulated trajectories 1 and 2 fall

somewhat short of their measured particle capture locations. The simulated point dipole

accurately predicts the actual field gradients experienced along trajectory 3 and thus

Page 50: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

41

capture location. Similarly, the gradient predictions are slightly larger along trajectories 4

and 5 that are relatively further removed from the dipole origin.

Figure 3.5: Particle trajectories

Having validated the simulations, important parameters that influence the particle

trajectories are characterized next. It is intuitive that χeff, m, a, U, η, and a characteristic

length scale L all have an effect, since these are all implicitly contained in the equation of

motion. The capture of magnetic particles can be quantified through capture efficiency

( )κλδγβαeff LηUamχf∝Φ (3.10)

The capture efficiency represents the ratio of the number of particles retained in

the channel compared with the total number introduced into the microchannel in case of a

steady flow of a homogeneous magnetic microsphere suspension. An alternate

-1.0E-03

-9.0E-04

-8.0E-04

-7.0E-04

-6.0E-04

-5.0E-04

0.0E+00 2.0E-04 4.0E-04 6.0E-04 8.0E-04 1.0E-03Z

Y

ExperimentalPredicted

1

2

3

4

5

(m)

(m)

Page 51: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

42

description is obtained by identifying the marginally collected particle trajectory.

Particles below this trajectory are collected but those above it escape the collector, hence

Φ = y0/h. Thus, information about the initial position y0 of the marginally collected

particle at the inlet plane z = 0 is sufficient to predict the (0,1) collection possibility for

any particle above or below this position.

The normalized magnetophoretic velocity term is a function of the dimensionless

ratio of the magnetic to the to the particle drag force, namely,

72

2222

7

2o

d

m1 ]m][s/m][m/Ns[

]1[]Am][A/N[]m[UL

maFF 2

χμ∝=Π (3.11)

Figure 3.6: Variation of Ф with П1 for constant values of П2 = ymag/h (= 0.75).

1

10

100

100 1000 10000 100000

Π1

Φ (

%)

χ m aU η L

Ф = 0.25 П10.87

Ф = 0.55 П10.48

Π1 = 1650

eff

Page 52: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

43

An increase in the magnetic susceptibility and the magnet dipole strength increase

the magnetic body force acting on a particle. For a specified particle magnetic

susceptibility, both the particle magnetization and the local field gradient scale according

to m, which then bears a second order relationship with П1. The magnetic force depends

upon the particle volume ∼a3 while the drag force is proportional to the particle radius a,

i.e., the ratio of the two forces depends on a2. The particle drag force is proportional to η

and the particle relative velocity. The magnetic field strength and its gradient, thus the

magnetic force, decrease with increasing displacement of the particle from the dipole.

The axial viscous drag force on a particle competes with the cross-stream

magnetic force by resisting magnetophoretic particle motion and sweeping the particles

downstream. Therefore, the capture efficiency also intuitively depends on the ratio П1 of

these two forces. This hypothesis is corroborated in Figure 3.6 in which all the data

collapse on a single curve. One would intuitively expect a linear relationship between Φ

and Π1 in a homogeneous velocity field that experiences a constant magnetic field

gradient. However, both the magnetic field and fluid velocity distributions are spatially

nonlinear in our case so that the Φ vs Π1 data in Figure 3.6 require a nonlinear fit with

differing exponents for the different regions discussed previously. While there are several

curvefit options, the one we present provides two piecewise linear (on a logarithmic

scale) curves that follow the correlations.

Φ = 0.025 П10.87 for П1 < 1650 (3.12)

Φ = 0.55 П10.48 for П1 > 1650 (3.13)

Here, for П1 < 1650, α = 0.87, β = γ = 1.74, δ = λ= -0.87, and κ = -6.09 while α = 0.48, β

= γ = 0.96, δ = λ= -0.48, and κ = -3.36 for П1 > 1650. A larger П1 value represents a

Page 53: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

44

stronger magnetic force that results in a high particle capture. Use of Figure 3.6 enables

the selection of П1 for a desired value of Φ in a microfluidic application.

In Figure 3.6 and eq. 3.13, the value of L is selected as the distance of the channel

centerline from the dipole origin, i.e., L = rmag (See Figure 3.1). However, rmag = ymag +

h/2, i.e., it depends on the channel half height (h/2) and the distance of the magnet from

the lower wall ymag. The data in Figure 3.6 represent variation of Ф with П1 for fixed П2

(= 2ymag/h) accomplished by appropriately varying ymag and h. In a family of

geometrically similar channels having different physical dimensions, magnetic and

hydrodynamic similarities require that the ratio 2ymag/h remain constant irrespective of

the values of ymag and h.

Figure 3.7: Variation of Ф with respect to П2 for a specified value of П1 (=12250). Region shows that increasing values of ymag have a strong influence on Φ. In contrast, in region Diminishing values of h

have a strong influence on Φ.

Φ = -2X10-05Π26 - 0.0006Π2

5 + 0.0457Π24 -

0.8633Π23 + 7.3068Π2

2 - 27.242Π2 + 81.517

0102030405060708090

0 5 10 15Π2

Φ (

%)

Page 54: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

45

Thus, for a given value of П2, the local distributions of Fm/Fd inside the channel

are similar for all values of П1. This collapses all the data for various parametric values

on a single curve. However, if П2 is also varied, the data fall on distinct curves, each

corresponding to a specified П2 value.

The nonlinear dependence of Φ with П 2 while maintaining П1 constant is

illustrated in Figure 3.7. Here, ymag and h are varied but all other parameters are held

unchanged. The figure shows that the capture efficiency initially decreases and then

increases as П2 increases. The minima of the curve corresponds to a minimum capture

efficiency Φ = 45% for П2 = 4.0. If the value of L is held constant, an increase in П2

occurs when ymag increases in value while that of h decreases. The data in Figure 3.7 are

divided into two regions corresponding to small and large П2. For small П2, the values of

h are large while those of ymag are relatively small (in the region labeled in the figure).

Here, a small increase in ymag has a pronounced effect on the location of the critically

captured particle (i.e., y0 decreases more rapidly than does h). This explains the observed

decrease in the value of Φ. For П2>4.0, i.e., for relatively larger ymag values and hence

small h values (region in Figure 3.7), changes in ymag have a smaller influence on Ф

than in region . Since the denominator in the capture efficiency expression decreases

more rapidly than does the numerator, the capture efficiency with increasing ymag again

increases in region . Figure 3(b) also provides a basis to design to microfluidic

separation devices that utilize magnetic microsphere separation.

For various combinations of the operating parameters applied to a specified

geometry, the capture efficiency depends solely on П1 that characterizes the ratio of the

magnetic and viscous drag forces on the particles. If the geometry is altered, the values of

Page 55: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

46

Φ in various channels depend on П2, which is the ratio of the distance of the magnet from

the channel to the channel height. These correlations provide a method to rationally

design magnetic particle separators over a wide range of operating conditions.

3.4. Conclusion

We have investigated the motion of magnetic microparticles that are under the

influence of a magnetic field gradient in a microfluidic channel through experiments and

numerical simulations. There is good agreement between both measurements and

simulations. Two dimensionless global parameters П1 and П2 are able to characterize the

capture efficiency (Φ), which in a steady flow of a homogeneous suspension of magnetic

microspheres represents the ratio of the number of particles captured in the device to the

number of particles entering the channel. For various combinations of the operating

parameters applied to a specified geometry, the capture efficiency depends solely on П1

that characterizes the ratio of the magnetic and viscous drag forces on the particles. If the

geometry is altered, the values of Φ in various channels depend on П2, which is the ratio

of the distance of the magnet from the channel to the channel height. These correlations

provide a method to rationally design magnetic particle separators over a wide range of

operating conditions.

Page 56: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

47

Chapter 4: Magnetic separation from concentrated suspensions incorporating dipolar interactions

4.1. Introduction

Magnetic microspheres are composed of superparamagnetic nanoparticles

embedded in micron size silica or polystyrene microparticles. These microspheres can be

contained in colloidal suspensions and suitably functionalized as mobile substrates to

‘tag’ specific molecules or cells. Thereafter, the magnetized conjugates can be readily

separated from the background suspension, e.g., in microfluidic systems for bioassays or

for in vivo applications (SafarIk and SafarIková, 1999, Koneracká et. al, 2002,

Verpoorte). Such magnetophoretic separation has been conducted in microfluidic

geometries (K. Smistrup, 2005, Smistrup et. al, 2005, McCloskey et al., 2003, Inglis et

al., 2004, Lund-Olesen et al., 2006) to enable particle collection (Rida and Gijs, 2004,

Furlani et al., 2007), separation (Yellen and Friedman, 2004), and cell sorting (Jiang et.

al, 2006). Magnetophoretic transport and aggregation of the magnetic particles are the

key steps in all these processes and their study is important for designing the relevant

microfluidic devices.

Practical applications, such as bio-filters and cell sorters operate with high particle

concentrations (Zaytseva et al., 2005). Magnetically immobilized aggregates or plugs of

functionalized beads have been used in microfluidic protein digesters (Smistrup et. al,

2005, Anne Le Nel, 2005, A. Le Nel, 2006) and simultaneous bioassays (Bronzeau and

Pamme, 2008). Even for a dilute suspension, the particle concentration in the vicinity of

an aggregate can be sufficiently high, given the significance of dipole-dipole interaction

(Helgesen et al., 1988).

Page 57: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

48

While the literature discusses microfluidic systems that utilize magnetic particles,

there are few fundamental analyses of the underlying mechanics, and virtually all of these

pertain to dilute solutions. Analytical results exist for diluted suspensions and work well

in predicting the physics for isolated single particles (Furlani et. al, 2007, Nandy et al.,

2008) which have been experimentally verified (Wirix-Speetjens et al., 2006). However,

such analyses cannot be extended to concentrated suspensions and near the aggregates,

where interparticle interaction plays an important role. Since the previous investigations

of such flows have been essentially empirical (Rida and Gijs, 2004, Hayes et al., 2001a),

we provide a more fundamental parametric investigation of the mechanics of particle

collection and aggregation on a microfluidic platform. Particle trajectories and

aggregations are predicted using a Lagrangian particle tracking model (Sinha et. al, 2007)

that incorporates the particle-particle interaction. The simulations are also validated

through experimental visualization. A parametric variation quantifies the aggregate size

as a function of the magnetic field, flow velocity, particle size, magnetic field strength,

etc. The study provides design bases for microfluidic separators and packed-bed filters

involving magnetic microspheres.

For the low Reynolds number flow, the particle aggregation at the collection site

depends upon the forces experienced by the particles due to interparticle magnetic

interactions, the external magnetic field, and the viscous drag (Sinha et. al, 2007). The

interparticle interactions among magnetic microparticles suspended in fluids can be

investigated using the Monte Carlo (Chubykalo et al., 2003), cellular automaton (Chen

and Matthaeus, 1987), distinct element (Fazekas et al., 2005), or continuum mechanics

(Christian Mikkelsen) models. We use a particle-based approach that accounts for the

Page 58: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

49

geometrical factors and forces involved in aggregate formation, which makes it possible

to predict the aggregate microstructure.

4.2. Dipolar Interactions in Concentrated Suspensions

Under a homogeneous magnetic field (B0 = μ0H0, where, μ0 denotes the

permeability in vacuum), an isolated particle experiences no net force. However, in a

system of particles (as in most practical situations), a dipole-dipole interaction occurs due

to an induced magnetic dipole moment m=(4/3)πa3χeffH0 in each microsphere. The

interaction amongst the particles results in the formation of local structures built from the

basic building blocks – the magnetic particles. Depending upon the conditions, roughly

aligned chains parallel to the applied magnetic field are observed. In other situations, an

ordered packing structure is observed. Such ordered structures have strong relevance in

bioMEMS applications, e.g., to produce compact bio-filters and separators. For a two

particle system, the field at the location of the particle is modified from that for a single

particle and is a vector sum

0

1

n

i ijj=

= + ∑H H H (4.1)

Hm m r r

ijj j ij ij

r r= − +

ij ij3 5

3d i ; i → [1,n]; j → [0,n]; i ≠ j, (4.2)

where Hi denotes the resultant value of the field at i. H0 is the constant uniform

field while the second term gives the contribution to the field caused at the location of

particle i due to the existence of particle j (and vice-versa). Particle j acts as an induced

point dipole and is the source of the resultant gradient in the total field at i. Inter-particle

interactions are modelled by considering pair wise interactions. While considering pair

wise interaction, Fij gives the force on particle i due to j (i ≠ j). During the calculation,

Page 59: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

50

particle i is excluded from the domain. Thus Fij represents the force on i due to particle j

(i=1,2; j=1,2). L (characteristic length scale which in Figure 4.2(b) is 2|rij|) is the distance

between the two particles at t = 0.

Attractive forces between magnetic particles are utilized to enable such tests.

Repulsive forces can also exist between magnetic particles. A system of two particles can

experience either attractive or repulsive forces depending upon the geometry of the

system and the orientation of the field.

In order to achieve effective “action-at-a-distance” on the magnetic microspheres,

it is very important that the relevant transport issues are analyzed and characterized

carefully.

(a) (b)

Figure 4.1: (a) chains of 1 μm diameter magnetic micro-spheres under a homogeneous magnetic field. (b) Pair-wise particle dipole-dipole interaction model.

Figure 4.1(a) is a microscope photograph showing chain-columnar structure

formation of magnetic particles under a static and transverse uniform magnetic field. The

exact shape of these structures (also known as magnetorheological structures (MRS)), is

dependent upon the diameter of the particles, their volume fraction in the solution, their

magnetic susceptibility, the external magnetic field, the channel geometry and the

temperature of the system.

ij

rij

Y

X

rj0 ri0

O

H0

θ

mj

mi

B

Page 60: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

51

As θ increases, assembly time (t) increases for a given L. When θ approaches π/2,

the time required for assembly increases asymptotically. At θ=85o, this time increases to

approximately 100 seconds (two orders of magnitude higher) and at θ = π/2, the model

predicts the interparticle force to be purely repulsive (shown in Figure 4.2). Aggregation

time is found to increase with L, and the t-θ trend is found to be similar for any L. For a

given L, the time required for chain formation increases as θ decreases. The block arrow

in Figure 4.2 depicts the tendency for chain formation as opposed to ordered hexagonal

packing.

The trajectory of each particle is tracked using a particle tracking algorithm

detailed in previous work (Sinha et. al, 2007). The variables L and θ in a suspension

cannot be controlled externally in a deterministic manner. Instead, the variables that are

most amenable to external control are fluid viscosity (η), initial separation between

particles (L), strength of the applied field (B), particle size (a) and the magnetic

susceptibility of the particle (χ). The results of a parametric variation are presented in

Figure 4.3. Interestingly, it is observed that all the data of parametric investigation merge

on one linear curve when aggregation time is plotted against the group variable

Ψ=(ηL5)/(χB2a5). This provides for a basis of identifying a single tuneable parameter that

can be monitored to achieve controlled self- aggregation of magnetic microspheres in

applications such as bioMEMS amongst others.

Dipole-dipole interaction in a suspension of magnetic microspheres is

investigated to characterize the particle aggregation using a 2-particle interaction model.

The analysis showed that the tendency of particle chain formation is enhanced when the

particle magnetizations are in-line.

Page 61: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

52

1.E-01

1.E+00

1.E+01

1.E+02

1.E+03

1.E+04

1.E+05

1.E+06

0 10 20 30 40 50 60 70 80 90θ(degrees)

t(s)

L/2= 10E-06L/2= 25E-06L/2= 50E-06

Tendency to form chains

Repulsive at θ = π/2

Figure 4.2: Variation of t with θ for three different (L/2) cases. All other parameters in ψ are kept fixed.

1.E+011.E+021.E+031.E+041.E+051.E+061.E+071.E+081.E+091.E+10

0.0001 0.01 1 100 10000Ψ

t (s)

BaχηL/2

Figure 4.3: Parametric dependence of aggregation time t as functions of η, L, B, χ and a. The simulations

are performed for θ = 45o. All the plots collapse on a single curve when the aggregation time is plotted against the group variable Ψ=(ηL5)/(χB2a5).

The aggregation time depends both on the orientation of the particle

magnetization (θ) and the initial separation (L) of the particles. Aggregation time is also

found to depend on externally tuneable parameters like magnetic field strength, its

gradient, particle size, magnetic susceptibility, host fluid velocity and viscosity, and the

Page 62: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

53

characteristic length scale through a group-variable Ψ=(ηL5)/(χB2a5). For a wide range of

θ values, it has been observed that the aggregation time t ∝ Ψ.

In the case of superparamagnetic particles and when the external field is removed,

the particles immediately lose their magnetic dipole, and the superparamagnetic particles

involved in the chain structures diffuse away in the bulk; the uniform initial suspension is

then recovered. For this reason, mostly superparamagnetic particles are used in biological

applications where the aggregations are considered as an “undesirable” effect that leads,

for instance, to decrease the available active surface of the particles. However, recently

the super-structure formed by equally spaced magnetic chains has been used as an

efficient system for DNA separation and analysis [43]. Controlling magnetic particles

super-structures opens new perspectives for biological applications, a fact that will be

extensively demonstrated by the results of the following chapters.

4.3. Overlap Prevention Model An unphysical overlapping of particles is prevented through a velocity-based

collision barrier (Dance et al., 2004), where Vref (the maximum overlap velocity) and Rref

denote critical parameters with the dimensions of velocity and distance respectively. The

repulsive velocity when rij<Rref,

v rrij

ref ref ij

refij

Va

R rR a

=− −

FHG

IKJ2 4

2 2

2 2

2

. (4.3)

Using the notationV =|-vrij|/Vref, R =Rref/2a, r =rij/2a, Eq. (1) assumes the form

V R r

Rr=

−−

FHG

IKJ

2 2

2

2

1; r :(1,∞); R :(Rref/2a,∞). (4.4)

The behavior of this collision barrier is described through Figure 4.4.

Page 63: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

54

Figure 4.4: Increase in V for r < R as predicted by Eq. 2. As R →1, the collision model shifts from a soft to a hard sphere model.

Approaching particles continue undisturbed till they are separated by a distance

Rref. Once, r < R (or rij<Rref), V increases with decreasing r . The selection of R

determines whether the colliding particles behave as hard or soft spheres. In general, Rref

can be selected arbitrarily under the constraint that any further reduction does not

significantly alter the dynamics predicted by the simulations and so herein we use

R =1.02. At r =1, V =1 for all R , which corresponds to a ‘kissing’ configuration. Some

particles pairs never attain this ‘kissing’ state because all particles are assigned a Vref

based on the fastest approaching particle pair in the domain.

In our simulations, the flow is assumed to be fully developed and pressure-driven

along the x axis in a square microchannel cross section, i.e., h = w. Figure 4.5(b) shows

an point dipole centered at (l/2,-(h/2+s),0) in the x-z plane placed at a distance s below the

channel. For a single loop coil, the dipole strength m = Id2, where d denotes the coil

diameter and I the current flowing through it. Since we are interested in the effects of the

field rather than in its generation, we assume that d is sufficiently small so that the size

0

0.25

0.5

0.75

1

1.25

1.5

0.9 1 1.1 1.2 1.3

r

V

1.11.011.051.21.02

R

Soft collision

Hard collision

Page 64: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

55

effects of this coil are small compared with the field that it produces above it. Thus, such

a finite size coil can be approximated as a point dipole of strength m = mj$ that is aligned

along the y direction and situated at the center of the coil, which is consistent with the

field produced by our electromagnet. The corresponding expression for the generated

field is identical to the treatment of the microparticles as point dipoles in Eq. (1.9). This

allows us to decouple the problem of particle separation from that of magnetostatics, the

latter of which is not our intended focus.

Figure 4.5: (a) Schematic of particle-particle dipole interaction model and of the (b) simulation domain.

Figure 4.6: (a) Side view (z=0 plane), (b) End view (x=l/2 plane) and (c) Bottom view (y=-h/2

plane) of the profile of the magnetic field created by a point dipole. The color bar represents the magnetic field strength in milliTesla (mT)

j

i k rij

yY

xY

rj0 ri0

O

mi

mj

(a)

x

h

y

m

z 0

w

l

s

(b)

y

x

y

z

z

x

(b) (a) (c)

Page 65: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

56

For the microparticles suspended in the fluid inertia effects are negligible (Modak

et al.), since Rea«1, implying that while the fluid exerts drag on each particle, the

particles exert no force on the fluid. Also, because of the extremely small volume (and

mass) gravity force is negligible as compared to the magnetic or viscous force on the

particle. Particle transport is thus accomplished through a balance between the

hydrodynamic and magnetic forces. Hence, Eq. (2) assumes the simple form F Fm d= − so

that the magnetophoretic velocity u F up m fa= +1 6πηb g , i.e., the updated particle

positions are x u xt dtp

t

b g = +z0

0 . The forward differencing time intervals dt in our

simulations are specified so as to limit the particle travel to a distance smaller than 1/20th

of the particle size per time step, i.e., |up|×dt < (a/20).

We use the infinite Fourier series solution of the Stokes equation for the velocity

distribution (Brody et al., 1996). Particles are introduced through the left boundary (the

inlet) and those that reach the right boundary (the outlet) are removed from the

simulations. Particle overlap is prevented by using a velocity barrier approach (Dance et.

al, 2004) and considering a very short time step for simulation. The top, bottom and side

walls are modeled as impermeable with the same velocity barrier.

The extent of particle collection and the aggregate size depend upon the particle

size a, concentration c, and magnetic susceptibility χ, the magnetic volume fraction φ,

dipole strength m, fluid velocity Uo and the distance of the dipole from the channel wall s.

Figure 4.6(a)-(c) present different views of the simulated magnetic field generated with a

point dipole of strength m = 1×10-11 A-m2 that corresponds to H = 3000 A/m or B ≈ 4

mT. Figure 4.6 (a) presents the field in the xy plane along z = 0. The field strength is the

Page 66: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

57

highest (~100 mT) at the point closest to the dipole position and falls rapidly with

increasing distance to ~0.1 mT at the boundary of the domain. Similarly, Figure 4.6 (b)

and (c), that represent the planes x = l/2 and y = -h/2, respectively, show a similar trend,

i.e., that the field has a spherical shape centered on the point dipole.

4.4. Experimental Setup

Borosilicate (pyrex) microslides (100 µm × 1000 µm crosssection, 100mm length,

VitroCom Inc.) were used for the microchannel prototype. These are rectangular

crosssection capillaries the ends of which were bonded to cylindrical soft plastic

reservoirs (30mm height × 10 mm dia.) with epoxy. Small differences in liquid levels

between the two reservoirs were sufficient to generate pressure driven flows (of the order

of 1 mm/s ) in the microchannel. A pressure head of 100 mm was found to be sufficient

to produce a flow with a Reynolds number ~1 in the channel. For the ensuing discussion

we consider cases for which the height difference was maintained at 7mm during the

experiment.

The 1 μm diameter functional magnetic silica microspheres were obtained from

Corpuscular Inc. (PMSi-H1.0-5). Images of the aggregation of the microspheres were

obtained using a CCD Camera (Hitachi KP-D20A/KP-D20B) for the front view (2000μm

feature) of the channel. A stereoscopic zoom microscope with an attached digital camera

that was coupled with an additional objective lens was used to examine the top view

(100μm feature) of the channel as shown in the schematic of Figure 4.8. The entire length

of the microchannel in Figure 4.8 is representative of the separator section shown in

Figure 4.7. The host fluid used was distilled water.

Page 67: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

58

Figure 4.7: Schematic of the separation scheme

Figure 4.8: Schematic of imaging system. Inset shows the position of the needle magnet with respect to the channel cross section

Beads Host fluid

Magnet

Reservoir

100 μm

2000

μm

100 mm

Flow Direction

N S

CCD Camera (Side view)

N S

N

S Mic

rosc

ope

Obj

ectiv

e (T

op V

iew

)

Page 68: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

59

To mimic the effect of a localized dipole, a tempered steel needle was attached

with a rare-earth neodymium-iron-boron cube magnet obtained from National Imports.

The dimensions of the magnet were 25mm X 25mm X 25mm. A two dimensional

schematic of the arrangement is presented in Figure 4.9. The aggregation permanent

needle-magnet produced a sharp gradient in a small region adjacent to and within the

channel. The magnet was mounted on a horizontal uniaxial traverse at the same vertical

height as the channel. Images shown later are representative of one such trap.

A viable microfluidic network can be based on our concept. We found that the

microspheres were highly responsive to the magnetic field and were readily transported

across the laminar streamlines of the microchannel when such a field was applied.

Spontaneous self-assembled chain formation of the microspheres was observed along the

lines of force of the imposed magnetic field. The images confirm the feasibility of

separating of magnetic beads from the host flow. Our future work will involve

characterization of the separation efficiency in terms of flow and magnetic parameters.

The presence of the needle magnet results in a crowding of the lines of force (B)

at the needle tip, significantly altering the local uniform magnetic field due to the magnet

itself, as shown in Figure 4.9. In a uniform magnetic field, the magnetic microsphere

would experience no force. The block arrow indicates the flow direction. ‘y' represents

the distance between the outside of the channel wall and the tip of the needle. The

variation in y is from 0 mm to 1 mm, the former being the value when the needle was in

contact with the outer wall of the microchannel. The magnetic field (By) was measured

using a standard DC Magnetometer (Serial#1145, AlphaLab Inc), results for which are

also presented in Figure 4.9.

Page 69: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

60

Figure 4.9: (above): Experimental setup showing position of magnet relative to the section of the microchannel, (below) variation By with increasing separation between magnet and channel.

38 mm

25 mm

y mm

Channel section

Magnet

Micro-spheres

Experimental domain

B

By

Bx

x

0

5

10

15

20

0 0.01 0.02 0.03 0.04

By (T)

y (m

m

Page 70: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

61

4.5. Results and Discussion Figure 4.10 shows a time varying aggregation of the particles for the case when y

= 0. The block arrow represents the direction of flow in the channel. The images show

large scale collection as well as small scale chain formations due to magnetization of

individual microspheres. As the microspheres join the aggregation, incoming particles

latch onto already trapped particles increasing the density of entrapment. The images of

Figure 4.10 are extracted from the background using Photoshop.

Figure 4.10: Time varying aggregation of magnetic microspheres extracted from the flow by the localized

magnetic trap for y = 0mm

After extraction of the background, a standard image processing software Scion

Image was used to calculate the integrated pixel density. The algorithm integrates the

individual values of each pixel to arrive at a single value for the entire image. If Pij is the

pixel intensity of any arbitrary pixel located at the ith row and jth column

14.5 s. 8s. 8.5s. 9s. 9.5s. 10.5s. 11s. 11.5s. 13s. 13.5s. 14s. 15s. t = 0 s.

50 µm

Page 71: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

62

(4.5)

Figure 4.11 illustrates time and spatial evolution behavior of the aggregation of

the particles. As the channel is moved away from the needle tip, the magnetic field

gradient decreases and the collection effectiveness of the localized dipole is reduced.

Integrated pixel density is normalized from zero to one.

Figure 4.11: Particle aggregation varying with time and distance of needle tip from channel

0

0.05

0.1

0.15

0.2

0.25

0.3

0 10 20 30 40 50 60

Time (sec)

Inte

grat

ed p

ixel

den

i

j

(0,0)

(m,n) (0,n)

(m,0)

Pij

■ y = 0 mm ∆ y = 0.5 mm

{ }*

0 0

* ;0 255(256)

m n

iji j

ijij ij

I P

PP P

= =

=

= ≤ ≤

∑ ∑

Page 72: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

63

The curves of Figure 4.11 represent a situation where both the strength of the field

as well as the gradient change. In order to isolate these two effects and study their

dependence individually, the source of the magnetic field was replaced to include an

electromagnet. The setup was identical to that shown in Figure 4.8.

Figure 4.12: The position of the electromagnet tip relative to the microcapillary

A pressure-driven flow produces a peak centerline velocity of 1 mm/s resulting in

a Reynolds number Rep ~ O(10-3), which shows that viscous forces in the flow dominate

over the fluid inertia. Figure 4.12 shows a representative view of the region in which the

tip of an electromagnet leads to particle capture. The electromagnet tip touches the other

wall of the microcapillary so that the distance between the tip of the electromagnet and

the fluid is limited by the width (125μm) of the microcapillary wall.

The electromagnet consists of 26 AWG enamel coated wire wound around (ca.

125 turns per cm.) a 1 mm diameter iron core and has a length of 4 cm. The core is

tapered at the end so as to generate a large magnetic field gradient at the tip. The resulting

nonuniform magnetic field leads to particle aggregation (and thus collection) on this wall

adjacent to the tip. The coil is oriented with its axis aligned with the y direction as shown

1

2

s 200μm

Page 73: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

64

in the inset such that Bx and Bz, are zero. The coil is activated with currents of 0.25, 0.5,

0.75 and 1 A. Figure 4.13(a) presents the spatial variations of By near the region of

collection.

For the largest current, the induced field, which is measured by a Gaussmeter

probe, has a maximum value of 30 mT in the microcapillary region closest to the tip,

which drops to 10 mT at a 1 mm distance away from it. At lower currents, the trend is

similar but with reduced field strengths. The spatial variation in the y component of the

field in the z=0 plane is illustrated in Figure 4.13(b). By is maximum closest to the

electromagnet being 30 mT and gradually fading to 0.1mT towards the boundary of the

measured space (i.e., at y=50 mm).

Direct images of particle collection are obtained using a digital stereo microscope

(National Model 420T-430PHF-10 Trinocular Stereo Zoom) that is coupled with an

objective lens which provides an overall magnification of 200X. The particle aggregate

size at any instant is determined by evaluating the ratio of the total area occupied by the

beads and the channel section under consideration. This section is a rectangular window

of 1000 μm×1500 μm (i.e., the channel width × 1.5 times that width).

During the experiment, the magnetic microparticles are immobilized on the

channel wall adjacent to the electromagnet tip where the field is highly divergent. Bright

field imaging shows the particles as dark spots in Figure 4.12. Particle collection is

characterized through the ratio ς = npxparticle/npx

image, where npxparticle denotes the number of

pixels in the image representing the collected aggregate and npximage (=720×480) is the

total number of pixels in the image. Our image analysis includes a cutoff pixel evaluation

so that all pixels with luminosity values below this cutoff are assigned a zero value, while

Page 74: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

65

pixels above this cutoff, which represent the particles in the aggregate, are assigned a

value of unity.

(a)

(b)

Figure 4.13: (a) Spatial distribution of By for I = 0.5A. Field measurement is made by a Hall probe Gaussmeter. (b)Variation of the magnetic field component By coil along the electromagnet core axis (where Bx= Bz= 0) for four different currents up to 1mm above the electromagnet tip.

0

5

10

15

20

25

30

35

0 200 400 600 800 1000

Distance from magnet tip - s (μm)

Mag

netic

Fie

ld S

tren

gth

- BY (m

T)

0.25 A0.50 A0.75 A1.00 A

Increasing I

x y

z

4 mm

0 mm 75 mm -75 mm

50 mm

y

x

Page 75: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

66

Figure 4.14 (a) shows instantaneous snapshots showing magnetophoretic capture

and aggregation of magnetic microspheres in the microchannel over a duration of 550 s

for a flow rate of 1mm/s (corresponding to a particle Re of 10-3) and the electromagnet

current of 0.5 A. Very near the lower wall, the velocity is low due to no slip, whereas the

magnetic field gradient is the maximum.

(a)

Figure 4.14 (a) Experimentally-obtained images of growing aggregates at various time intervals (seconds) for a = 2 μm, I = 0.5A, U= 1mm/s, s = 125 μm, c = 1012/m3. Bar = 200 μm. (b) Side view, (c) End view and (d) Bottom view of representative snapshots of the simulated particle aggregation at steady state

Hence the magnetic force experienced by a particle dominates the hydrodynamic

force near the wall adjacent to the dipole, resulting in relatively larger collected

aggregates. As the aggregate size increases, its outer boundary moves further away from

the wall (and also from the dipole). The values of both the magnetic field strength and its

y

x

y

z

z

x

(c) (b) (d)

t=0 s 50 s 100 s 150 s

200 s 250 s 300 s 350 s

400 s 450 s 500 s 550 s

u0

Page 76: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

67

gradient diminish with increasing displacement of this boundary from the wall. At the

same time, the fluid velocity increases with increasing distance from the wall. Thus, with

growing size of the aggregate, the magnetic force (the holding force) on the particle at the

periphery of the aggregate no longer dominates over the shear force (the depleting force)

applied by the flow. This marks boundary of a steady aggregate in the flow. When the

particle separation and aggregation are simulated, similar structures are observed. Figure

4.14 (b), (c) and (d) shows the three views of a stable aggregate. Away from the

aggregate, both the experiment and the simulation show existence of small agglomerates,

which are formed due to the inter-particle interactions of two closely spaced particles.

Such interaction between individual particles to form clumps is very characteristic of

concentrated suspensions under an imposed magnetic field.

Figure 4.15 illustrates the temporal evolution of ς for various values of current I

over a period of approximately ten minutes. The value of ς increases linearly with time,

and with the aggregate growth rate and size increases with I (with the largest value being

ς = 0.05). The collection of the aggregate is a dynamic phenomenon. Particles assemble

into small aggregates in the flow before settling into the larger collected structure. This is

indicated through the rapid instantaneous increase in the slope for the curves in Figure

4.15. Point A in the same figure (for I = 0.25 A) highlights a phase of a high rate of

aggregate growth. Since the particle concentration is kept constant, the inference points to

the addition of a large number of particles to the aggregate. At other times, an aggregate

loses particles due to fluid instabilities, which is indicated through the abrupt

instantaneous decreases in these slopes. Point B in Figure 4.15 corresponding to I = 1.0 A

Page 77: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

68

indicates an abrupt decrease in aggregate size implying instantaneous loss of many

particles.

Figure 4.15: Aggregate growth rate with s=125μm and variable I. The inset shows the relationship between the collection rate and current intensity

The magnetic force F on a particle (Sinha et. al, 2007) is proportional to m2, such

that F∂I2, which is seen in the inset of Figure 4.15, since (dς/dt) ∂ I2. Thus, the particle

collection rate is proportional to the magnetic force experienced by individual particles.

The maximum particle aggregation occurs at the location of the highest field strength

which, for a single point dipole, also corresponds to the location of the highest field

gradient.

Our simulations also predict similar trend of particle aggregate growth rate, as

shown in Figure 4.16. For a representative simulation, the dipole is activated at t = 0 s

after the particle number in the microchannel reaches ∼30, i.e., a concentration of 3×1015

0

0.01

0.02

0.03

0.04

0.05

0.06

0 100 200 300 400 500 600

Time (sec.)

ς

1.00 A0.50 A0.75 A0.25 A

0.E+00

5.E-05

1.E-04

0 0.5 1I (A)

dς/d

tIncreasing I

B

A

Page 78: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

69

particles/m3. As observed experimentally and in the simulations, the aggregate size in

Figure 4.16 is found to grow linearly with time, before reaching a steady value. The slope

of a curve in Figure 4.16 is representative of the aggregation rate. An increasing dipole

strength shows this slope to increase, corroborating the experimental measurements in

Figure 4.15.

Figure 4.16: Effect of change in dipole strength m on the number of collected particles. The inset shows

linear relation between slope of Figure 8(a) and dipole strength m.

The aggregate ceases to grow after a certain time. This is the critical size

characterized by the equality between viscous and magnetic forces, thus segregating

aggregation dynamics into two distinct phases. Aggregate build up and the critical state

are marked as Phase I and Phase II respectively in Figure 4.16. The latter is assumed to

occur when the change in n(t) (the number of particles in the microchannel) is smaller

Dipole strength - m (A-m2)

0

100

200

300

400

500

600

700

0 100 200 300 400 500 600

Time (sec.)

n(t)

1.50E-11

1.00E-11

7.50E-12

8.50E-12

6.50E-12

5.00E-12

Increasing m

0

0.25

0.5

0.75

1

0.0E+00 5.0E-13 1.0E-12m (A-m2)

(dQ

/dt)

II I

Page 79: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

70

than 1%. The time taken to reach this state t* for a 1×10-11 A-m2 dipole strength is

approximately 430 s.

In order to characterize the aggregate size during Phase II, we define nII =

n tt t

( )=

∗∑ . The resulting volume fraction of collection.

P = nII.VMP/VD, (4.6)

where VMP = 4πa3/3 denotes the volume of each microparticle, VD = l.w.h the

microchannel volume. The relation between P and m is linear, as shown in the inset of

Figure 4.17.

Figure 4.17: Linear relation of P with Π. The inset shows variation of P with changing m. Simulation conditions are a = 500 μm, χ=0.1, Uo= 10 μm/s, s = 15 μm, φ = 0.7, c = 2×1015/m3

We analyze the simulations to also determine the dependence of P on χ, Uo, s, φ

and c that yield the relationships P ∝ a2, χ1/2, Uo-2/3, s-7/2, φ0.5 and c0.1. Here, while the

P = (270.18)Π

0.000

0.005

0.010

0.015

0.0E+00 2.0E-05 4.0E-05 6.0E-05

Π

P

ηmaχULφ c

P = (4.7724)m0.9689

0.000

0.002

0.004

0.006

0.008

0.010

0.00 0.05 0.10 0.15 0.20m ×1010 (A-m)

P

Page 80: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

71

nanoparticle volume fraction is held constant, increasing their size also increases the

overall magnetic content in the microparticle. These are collated into a single

dimensional group

70

32

0

212.0

=Πsη

χφμU

mac (4.7)

Meanwhile, the dimensionless ratio of magnetic to viscous force is

70

22

1 ==ΠsηUχμmaφ

FF o

d

m (4.8)

A relation thus emerges relating the dimensional group established here with the

standard dimensionless group Π1.

16/10

1.0

Π=ΠUac

(4.9)

The resulting relationship between P and Π, which is shown in Figure 4.17, is

linear. The relation for П is empirically-derived for a concentrated suspension while that

for П1 follows a scaling law for a single isolated particle. The force on a magnetic

particle, which scales as 1/r7, is a product of the magnitude of the magnetic field (∝1/r3)

and its local gradient (∝1/r4). While the presence of a magnetizable particle in a field

does not appreciably alter the local field strength, it changes the local gradient

significantly. Neighboring particles interact if they are sufficiently close. In a relatively

dilute suspension, the particle separation is large and this close range field distortion has

a weak effect on neighboring particles so that P ∝ Π0.5. However, as the concentration

increases, a larger fraction of particles lie within the distorted region so that the scaling of

P with Π1 must be modified, which results in the expression for Π. Since an increase in

Page 81: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

72

the particle concentration increases the interparticle interaction, it also leads to a larger

collected aggregate size.

4.6. Conclusion The magnetophoresis of magnetic microparticles is characterized in a

microchannel in terms of collection rate. Magnetic microspheres can be suitably

functionalized to ‘tag’ specific molecules or cells in a suspension and the magnetized

conjugates be readily separated. Hence, magnetic microspheres are indispensable for use

as mobile substrates for bioassays, especially with the emergence of microfluidics.

Even in dilute suspensions, the concentration at the location of particle collection

can be sufficiently high due to particle aggregation. In such circumstances, dipole-dipole

interaction amongst the particles must be accounted for. A fundamental parametric

investigation is conducted on the mechanics of particle collection and aggregation on a

microfluidic platform.

An in-house manufactured electromagnet is characterized through magnetic field

measurements and used for separating microparticles suspended in a creeping flow in a

microcapillary. The experiments capture the dynamics of magnetic particles. They

interact amongst each other to form a variety of structures. Additionally, magnetic

separation has been demonstrated in a microcapillary using an externally generated field

and employing the concept of ‘action at a distance’. The time varying dynamic behavior

of the aggregate is captured up to approximately ten minutes showing a linear growth

rate. The rate increases with increasing dipole strength and bears a quadratic relationship.

The rate of collection turns out to directly proportional to the magnetic force on a

particle.

Page 82: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

73

A numerical parametric study analyses the effects of major physical parameters

on magnetic separation. The values of the operating parameters selected are as typical of

microscale systems. The maximum particle aggregation occurs at the location of the

highest field strength which, for a single point dipole, also corresponds to the location of

the highest gradient. The results are consistent with the experimental observations.

The collection characterized by the volume fraction collected P ∝ m, η-1/2, χ1/2, Uo-

2/3, s-7/2, φ0.5 and c0.1 and in terms of a single dimensional group,

70

32

0

212.0

=Πsη

χφμ

U

mac.This relation is empirically-derived for a concentrated suspension

with dominant particle interaction. The resulting relationship between P and Π is shown

to be linear.

Clearly, it is crucial to consider interparticle effects when studying the dynamics

of aggregation. Regardless of whether used in-vivo or in-vitro, the critical transport

questions have been addressed and a dimensional group has been suggested that includes

the combined effects of the significant physical variables in the collection process.

Page 83: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

74

Chapter 5: Capture of Non-magnetic particles through interaction with functionalized magnetic particles

5.1. Introduction Advances in immunology have resulted in a variety of diagnostic techniques and

cellular probes, including for immunomagnetic separation (IMS) (SafarIk and

SafarIková, 1999), where magnetic particles with surface immobilized antibodies

specifically bind with antigens to form conjugates (Olsvik et. al, 1994, Yu et al., 2000).

After binding, a high gradient magnetic field holds the magnetic material in the container

as magnetically unresponsive nontarget cells or molecules are washed away. Sequential

activation and deactivation of the external field enables cyclic resuspension and

collection of the particles during successive wash steps. What is left behind is a

concentrated and purified suspension of the target species that is labeled with magnetic

particles. The primary purpose of magnetic separation is to concentrate the sample to

attain a higher signal to noise ratio than is possible for its untreated counterpart, thus

eliminating the need for sample pretreatment and amplification of the target

concentration. In contrast, physical concentration methods such as filtration and

centrifugation can produce loss of sample integrity.

Batch mode separators are limited to relatively small sample volumes (Rotariu et

al., 2004), an open operating environment, and relatively long reaction times dictated by

the particle and analyte concentrations. These limitations prevent their use for

physiological and environmental monitoring, especially when sample enrichment is

undesirable or impossible. During flow-through separation, a container is continuously

flushed with the sample and functionalized particles. Strategically placed magnets on the

Page 84: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

75

opposing walls of a channel result in the cross-stream migration of the particles, thereby

enhancing their interactions with the target species. This interaction binds the magnetic

particles with the target species. The particle-target aggregate is thereafter magnetically

immobilized as washing steps flush the container with various buffers. The flow-through

method has a potential for high purity and recovery, multiplexing and continuous

operation, thus enabling continuous sensing.

While magnetic particles have only recently been used in

microchannels(Verpoorte, 2003), the possibility of massively parallel IMS microfluidic

networks could potentially enable more robust remote monitoring (Van Emon and L.

Gerlach, 1998). The initial feasibility studies (Jiang et. al, 2006, Bronzeau and Pamme,

2008, Rashkovetsky et al., 1997) in this regard have been mostly empirical (Hayes et al.,

2001b), and there are no comprehensive models available for the interaction between

target species and magnetic particles. We therefore conduct a fundamental parametric

investigation of the mechanics of magnetic separation on a microfluidic platform using

three-dimensional numerical simulations (Sinha et. al, 2007, Nandy et. al, 2008). The

target cell or molecule is assumed to have identical dimensions as its magnetic

counterpart. The covalent bonds that represent biochemical functionalization, which

cannot be broken by fluid shear forces or by particle collisions, are assumed to occur if a

particle and a target come closer than a critical distance (Rotariu et. al, 2004).

5.2. Model A schematic of the simulation domain is presented in Fig. 1. The channel has a

length l and a square cross section defined by w = h. The fluids carrying similar sized

magnetic (M) and nonmagnetic (N) particles enter through the top left (3h/4) and bottom

left (h/4) sections of the channel, respectively. A point dipole is situated at a distance s

Page 85: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

76

below the geometric center of the lower channel boundary and l/2 downstream of its

inlet. The suspension is assumed to contain monodisperse, neutrally buoyant and

frictionless non-Brownian micron sized spherical magnetic particles. The Newtonian

relation for particle motion is 4 3 3b gπ ρa p i im

id

ig&u F F F= + + , where F m Bi

mi i= ⋅∇b g and

F u uid

ifa= −6πη c h denote the viscous drag force and magnetic force respectively. The

particles attain their terminal velocity in the Stokes flow, thereby diminishing the effect

of inertia, which is negligible. A standard velocity profile is used (Ichikawa et. al, 2004),

where we parametrically vary the peak centerline velocity U0.

Figure 5.1: Schematic of the computation domain

The instantaneously induced magnetic moment within each magnetic

microparticle m Hi iV= χ ϕ is limited by the saturation magnetization of the 5 nm

particles embedded within it. Here, V denotes the particle volume, φ the volume fraction

of the magnetic content within it, and B Hi o i= +μ χ1b g the induced field. The presence

of a particle distorts the field in its vicinity so that the force on a neighboring particle

changes, which provides the basis for the interaction between adjacent point dipoles. This

dipole-dipole interaction must be considered while simulating the magnetic force Fm. The

s

l

l/2

N

h

w

m

M

h/4

j

Page 86: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

77

local field due to the neighboring particles

H H m r r mi j ij ij j ij ijj i j

N

r r= + ⋅ −= ≠∑0

2 5

1

1 4 3πb g d ie j /,

, where H0 is the applied field due to the

external dipole m $j .

Figure 5.2: (a-c) Quality of the magnetic field generated due to the point dipole. (d-f) Snapshots of the

simulation for the base case.

An unphysical overlapping of particles is prevented through a velocity-based

collision barrier(Dance et. al, 2004), where Vref (the maximum overlap velocity) and Rref

denote critical parameters with the dimensions of velocity and distance respectively. The

repulsive velocity when rij<Rref,

v rrij

ref ref ij

refij

Va

R rR a

=− −

FHG

IKJ2 4

2 2

2 2

2

. (5.1)

Using the notationV =|-vrij|/Vref, R =Rref/2a, r =rij/2a, Eq. (5.1) assumes the form

V R rR

r=−−

FHG

IKJ

2 2

2

2

1; r :(1,∞); R :(Rref/2a,∞). (5.2)

y

x

y

-z

-z

x

(e) (d) (f)

y

x

y

-z

-z

x

(b) (a) (c)

Page 87: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

78

The behavior of this collision barrier is described through Figure 5.3.

Approaching particles continue undisturbed till they are separated by a distance Rref.

Once, r < R (or rij<Rref), V increases with decreasing r . The selection of R determines

whether the colliding particles behave as hard or soft spheres. In general, Rref can be

selected arbitrarily under the constraint that any further reduction does not significantly

alter the dynamics predicted by the simulations and so herein we use R =1.02. At r =1,

V =1 for all R , which corresponds to a ‘kissing’ configuration. Some particles pairs

never attain this ‘kissing’ state because all particles are assigned a Vref based on the

fastest approaching particle pair in the domain.

Figure 5.3: Effect of varying R on the collision model

Since the nature of viscoleastic forces in functionalizing strands is unclear (Strick

et al., 1996), we employ a velocity effect without resorting to the details of the binding

forces. This effect produces an unconditional binding similar to that observed in covalent

bonds. The velocity barrier collision model now includes an attraction term, i.e.,

0

0.25

0.5

0.75

1

1.25

1.5

0.9 1 1.1 1.2 1.3

r

V

1.11.011.051.21.02

R

Soft sphere model

Page 88: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

79

V R rR

r R rR

rc

c

a

a

=−−

FHG

IKJ −

−−

FHG

IKJ2

1 1

2 2

2

2 2 2

2

2

, Rc=1.01, Ra=1.02. (5.3)

Henceforth, the R conditions related to collision and attraction will be referred to

as Rc and Ra, respectively. Since reducing R below 1.02 does not change the dynamics of

the system, we set Ra= 1.01 and Rc = R .

Figure 5.4: Resultant interaction model predicting three regimes, attraction, repulsion and no interaction

The resulting velocity barrier is described in Fig. 2(b). There are three regimes,

namely, pure repulsion (when rij<R′c), attraction (when R′c>rij>Ra), and the third when no

interaction occurs between particles. A conservative value for Vref is determined by

examining the maximum hydrodynamic approach velocity between interacting particle

pairs (rij>Ra) in the domain. This reduces the effective value of Rc for the model from the

-2.5

-2

-1.5

-1

-0.5

0

0.5

1

1.5

2

2.5

0.99 1.00 1.01 1.02 1.03

r

V

Predicted by Eq. 3

First term of Eq.3; Rc=1.01, V=2

Second term of Eq.3; Ra=1.02, V=1

R’c Ra Rc

Repulsion No interaction Attraction

M M M N N

Page 89: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

80

stipulated 1.01 to 1.005, as shown in Fig. 2(b), which is referred to as R′c. The time step

for the simulation is constrained by the width of the attraction/repulsion zone, namely, δt

= (Ra – R′c)/10. The simulation runs until 24,000δts, where δts(=a/u0) is the Stokes settling

time.

5.3. Results and Discussion Particles are deflected from the flow streamlines towards the activated point

dipole and aggregate as close to it as possible as shown in Figure 5.2 (d-f). Particles lying

on the periphery of the aggregate describe a boundary along which the aggregate-holding

force due to the magnetic field competes with the aggregate-depleting drag force. As the

aggregate size increases, its outer boundary moves progressively further away from the

wall. Consequently, the values of the magnetic field strength and its gradient, both of

which attract particles from the flow into the aggregate, diminish with time. Eventually,

the aggregate-holding magnetic force at the boundary becomes comparable to the flow-

induced shear force so that despite particle addition, aggregate growth ceases. The net

average of particles added to the aggregate and removed from it is equal at this stage.

Nevertheless, due to the granular nature of the aggregate, changes in its size due to

shearing are discrete but never significantly large.

Figure 5.5 presents the number of particles retained in the aggregate (nN, nM) for

various values of c0=c(t=0)=nN/(nN+nM)|t=0. For the cases considered herein, (nN+nM)|t=0

=30, which corresponds to a concentration of ~1015/m3. The value of nN|t=0 is gradually

increased to characterize the effect of changing N- and M-particle concentrations. After

the dipole is activated at t=0, magnetic particles are progressively collected at the channel

wall. When c0 = 0.04 (i.e., the flow contains mostly M-particles), this collection phase

lasts ≈500 seconds until the aggregate reaches a critical size, marking the distinguishing

Page 90: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

81

stage between the aggregate buildup and washaway phases. For c0 = 0.04, the buildup

phase occurs to the left of the grey shaded area labeled M in Figure 5.5. The duration of

this phase depends on various system variables (e.g., the fluid velocity, particle

magnetization and size, etc.). During washaway, the loss of particles at any instant

generally equals particle collection, resulting in a roughly constant particle number for an

aggregate. Using the superscript * to denote time-averaged values during the washaway

phase, n*M = 320 and n*

N ≈ 0 for c0 = 0.04.

Figure 5.5: Aggregate size vs. time. The curves passing through the shaded area labeled as M refer to nM, while those passing through the corresponding N area represent nN. The inset shows the behavior of c(t). (b) Time averaged aggregate size n* during washaway phase and its dependence on c0. The secondary axis shows the corresponding time taken for the aggregate to reach critical size. Uo = 10 μm/s, m = 9×10-12 A-m2

The aggregate buildup rate is linear for M-particles, but quadratic for N-particles.

The enlarging aggregate forms a growing obstacle within the flow as more M-particle are

collected, allowing a larger surface area exposed the flow for the immobilization of N-

0

50

100

150

200

250

300

350

400

450

500

0 200 400 600 800 1000 1200

t (s)

n

0.08 0.370.60 0.04

M

N

2 1 0.001

0.010

0.100

1.000

10.000

0 400 800 1200t (s)

c(t)

c 0=0.600.37

0.08

0.04

Page 91: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

82

particles on it. Due to the relatively small number of N-particles for small c0 values, the

N-particle presence on the aggregate is mostly limited to the outer surface. Increasing the

value of c0 also increases that of nN, thus increasing the probability of interactions with

the nM M-particles on the aggregate surface. The M-M interaction is effective for adjacent

‘kissing’ particles, but depletes substantially as they are separated. When a layer of N-

particles coats the M-particles on the surface of a large enough aggregate through

unconditional M-N interactions, particles that would have been collected due to M-M

interactions are now no longer collected.

The inset of Figure 5.5 shows that the instantaneous M-particle concentration c(t)

in the aggregate decreases as N-particles are also collected, eventually reaching a nearly

constant value, except for periodic particle washaway events. As c0 increases, the rate of

M-particle addition to the aggregate during the buildup phase decreases, while the

corresponding values of d(nN)/dt increases, reaching a maximum for c0 = 0.6 when c ≈ 1.

Figure 5.5 provides insight into the size evolution of an aggregate as c0 is varied. Here,

nmax (=nN+nM) is indicative of the critical size of the aggregate before the onset of

washaway.

The data for c0 = 0.04 are divided into two regimes, marked as 1 and 2. For

Regime 1, the aggregate washaway occurs through the shearing of individual particles

While, in contrast, for Regime 2 larger chunks of the aggregate (that do not necessarily

lie on the aggregate surface) are washed away. In the latter regime, interparticle forces

within the aggregate that lead to bulk effects are as important as the fluid shear. From the

perspective of practical device design, the separation section concentrates the interacting

M- and N-particles into an aggregate. Some bound M-N conjugates are released from this

Page 92: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

83

aggregate through washaway and conveyed to subsequent sections for analysis. While the

M-N ratio in the washed away material is constant, the washaway frequency and the

amount of washed away material determine the rapidity of sensing and the signal

strength. A lower recovery of c(t) in the inset of Figure 5.5 corresponds to a higher

washaway frequency.

0.0E+00

1.0E-03

2.0E-03

3.0E-03

4.0E-03

5.0E-03

0.0 0.2 0.4 0.6 0.8 1.0c 0

P* M +N

MN

M >N M≈N

c 0 = 0.75

P 0

Figure 5.6: The critical aggregate size prior to washaway vs. the relative concentration of M and N-

particles.

The aggregate volume is related to the volume fractions P*M = nM.VP/V and P*

N =

nN.VP/V, where VP = 4πa3/3 denotes the volume of each particle and V = lwh the domain

volume. Figure 5.7(a) presents a relationship between P*M and P*

N (that are obtained

from time averaged values during the washaway phase) while Figure 5.7 presents their

standard deviations σM,N as function of c0. Large c0 values correspond to relatively high

Page 93: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

84

numbers of N-particles for which M-N interactions are highly probable, but M-M

interactions less so.

0.0E+00

1.0E-04

2.0E-04

3.0E-04

4.0E-04

5.0E-04

6.0E-04

7.0E-04

8.0E-04

9.0E-04

1.0E-03

0 0.2 0.4 0.6 0.8 1

c 0

σ

MN

Figure 5.7: The rapid increase in the standard deviation of an aggregate during washaway phase

for the two particle types

The aggregate volume is related to the volume fractions P*M = nM.VP/V and P*

N =

nN.VP/V, where VP = 4πa3/3 denotes the volume of each particle and V = lwh the domain

volume. Figure 5.7 presents a relationship between P*M and P*

N (that are obtained from

time averaged values during the washaway phase) while Figure 5.7 presents their standard

deviations σM,N as function of c0. Large c0 values correspond to relatively high numbers

of N-particles for which M-N interactions are highly probable, but M-M interactions less

so. Washaway occurs more readily in this case as there are not sufficient M-particles to

form a tightly bound aggregate so that the P*M plot approaches 0 as c0→1. On the other

Page 94: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

85

hand, at values close to c0 = 0, negligibly small amounts of N-particles are collected.

Initially, as the value of c0 increases, the value of P*M decreases while that of σM

increases until c0 ≈ 0.75, while P*N increases linearly with increasing c0. At c0 ≈ 0.75,

both P*M and P*

N have equal values, i.e., an equal number of M- and N-particles are

immobilized in the aggregate and its washed away portion. If c0 is increased any further,

both P*M and P*

N decrease. For this c0, range, the relative concentrations of both M- and

N-particles immobilized in the aggregate is roughly equal although the number of M-N

conjugates always decreases with increasing c0. When c0 = 1, the system contains no M-

particles to initiate aggregation so that there is zero collection. Figure 5.7 presents the

change in σM,N with respect to varying c0. Close to c0 = 0.6, the aggregate experiences

complete cyclic washaway, i.e., it is entirely lost periodically, with the consequence that

c(t)→c0. The standard deviation increases with increasing c0 for both M and N-particles

and reaches a maximum value at c0 = 0.6. while all simulations for c0>0.6 produced

washaway, the values of σM,N decrease and approach zero for c0 = 1. The time taken for

the aggregate to include nmax particles increases for larger σM,N.

Figure 5.8 illustrates the influence of the external dipole strength m and the peak

centerline fluid velocity U0 with c0 = 0.36 for all cases. When m is increased, so are H

and its gradient, which favors aggregation and a corresponding increase in the collection

of N-particles. Figure 5.8 shows the variation of P* with m. The average number of

particles n* during the washaway phase also increases with increasing m. The peak

centerline fluid velocity is varied to determine the influence of fluid shear on washaway,

results for which are presented in Figure 5.9. These show that as P*∝U-1, since increasing

U0 also increases shear at the aggregate periphery, producing smaller aggregates.

Page 95: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

86

Increasing the value of U0 greater than 30 μm/s results in trivial collection of both M- and

hence N-particles.

0.000

0.002

0.004

0.006

5.00E-12 7.00E-12 9.00E-12 1.10E-11

m (A-m2)

P*

MN

Figure 5.8: Variation in P*

N and P*N as a function of (a) dipole strength m, Uo = 10 μm/s. c0 = 0.36 and is

held fixed

P *M = 4E-08U 0

-0.9734

P *N = 4E-09U 0

-1.034

0.0E+00

1.0E-03

2.0E-03

3.0E-03

4.0E-03

0.0E+00 1.0E-05 2.0E-05 3.0E-05 4.0E-05

U 0 (m/s)

P*

MN

Page 96: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

87

Figure 5.9: Variation in P*N and P*

N as a function of peak centerline velocity U0, m = 9×10-12 A-m2. c0 = 0.36 and is held fixed.

5.4. Conclusions We provide a binding model for target species with magnetic particles.

Thereafter, we investigate the influence of changing relative concentrations of such N-

and M-particles on the aggregation dynamics in a microchannel flow. Two behaviors are

observed; one when N-particle collection increases with decreasing M-particles and the

other that this collection has a limiting value. The larger the amount of N-particles

collected, the greater the fluctuation in aggregate size up to c0 = 0.75. This corresponds to

the optimal working range of the magnetic separator. Aggregate depletion occurs

periodically through either surface shear or bulk washaway. The number of N- and M-

particles collected just prior to washaway remains approximately constant. The optimal

condition for the release of conjugates from the microchannel occurs at c0 ≈ 0.75 (and not

intuitively at 0.5). From the perspective of practical device design, the separation section

concentrates the interacting M- and N-particles into an aggregate. Some bound M-N

conjugates are released from this aggregate through washaway and conveyed to

subsequent sections for analysis. While the M-N ratio in the washed away material is

constant, the washaway frequency and the amount of washed away material determines

the rapidity of sensing and the signal strength. Varying the magnetic strength and the

fluid velocity, leads to P*∝(m, U-1). Varying U0 and m show self-similar trends in the

aggregation dynamics that show variations only in terms of the relative concentrations of

M- and N-particles.

Page 97: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

88

Chapter 6: Conclusions

In this thesis, the transport and manipulation of magnetic particles has been

characterized in terms of physical operating variables. Microfluidic capture of

paramagnetic particles in dilute, concentrated and multi-specie suspensions is studied.

Experimental as well as theoretical models have been proposed to study the dynamics of

suspensions of magnetic microparticles in a fluid medium. At the outset, a simplistic

particle based Lagrangian model is proposed. Such a simplified model performed well

when compared with experimental tracking of particles in a dilute suspension. Two non-

dimensional numbers are established depicting the competition between magnetic and

viscous effects for particle motion.

For concentrated suspensions, a force based interaction model is proposed.

Experiments have been conducted for a separation system. The time scale for growth of

the aggregate is characterized. The general trend of growth of the aggregate is in good

agreement both numerically and experimentally. A dimensional group is developed that

characterizes the rate of aggregate growth in terms of critical parameters such as dipole

strength, fluid velocity, etc.

For a multi-species model, a mathematical model is presented to represent

functionalized bonding between magnetic and non-magnetic particles. Keeping the

previously varied parameters fixed, the effect of relative changes in concentration is

studied. An optimal relative concentration emerges to maximize non-magnetic particles

capture.

Page 98: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

89

6.1. Current Trends There has been a continuing effort to develop and improve manufacturing

techniques so as to produce microfluidic devices in increasing quantities at lower costs.

As the cost per microfluidic device is lowered, it becomes economically feasible to use

such devices as disposable items. Accordingly, there is a need to develop manufacturing

processes for microfluidic technologies that correspond to "wafer level" semiconductor

processing techniques. Such "wafer level" technologies would allow for simultaneous

rather than sequential processing associated with the production of microfluidic devices

and/or components thereof, thereby reducing their per unit cost.

The driving force for miniaturization is an increasing demand for low-cost

instruments capable of rapidly analyzing compounds in very small sample volumes with

a high level of automation. The concept termed ‘Micro-Total Analysis Systems (μ-TAS)’,

and also ‘Lab-on-a-chip’, aims to develop integrated micro-analytical systems. They

perform complete analysis cycles (e.g. sample pre-treatment, chemical reactions,

analytical separation, detection, and data handling) on the same microdevice.

Microfluidic chips have integrated thousands of individually-addressable valves

and storage chambers (Jonas O. Tegenfeldt, 2004), and recent trends indicate that

microfluidics is following a path similar to Moore's law (Moore, 1965).

As advancements in microfabrication techniques allow for further shrinking lab-

on-a-chip devices from the micron regime to the nanometer regime, much of the non-

traditional physics is still to be discovered and many exciting applications of lab-on- a-

chip devices are yet to be exploited.

The future of biomolecular manipulation depends on three factors: the integration

and further development of single-molecule techniques; progress in the field of

Page 99: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

90

nanotechnology; and the use of high-throughput systems such as microfluidics.

Ultimately, the goal of single-molecule manipulation is to access the machinery of a

living cell.

6.2. Societal Impacts Access to a safe environment is important as a health and development issue at a

national, regional and local level. Investments in environment monitoring and awareness

are expected to yield economic benefits, since the reductions in adverse health effects and

health care costs outweigh the costs of undertaking the interventions. Once developed,

these microfluidic based sensors are expected to provide a preventive care at a personal

level. Only when their usage becomes wide spread will their ultimate promise be

realized.

6.3. Outlook and Perspectives Microfluidic devices are beginning to make significant contributions to

biomedicine and drug discovery. This is reflected in the development of many new

journals and conferences and increased funding for research in this area.

So far, the research in μ-TAS has focused on the development of individual

reaction, separation, and detection schemes on a chip and, thus, the next steps should be

an integration of sample preparation techniques as well as electronics on microdevices.

Rapid and automatic dispensing and handling of small sample volumes are also the issues

that need to be explored more closely. Full potential of micro-analytical systems will be

realized when the whole package is constructed.

With growing popularity and the dramatic progress in their investigation, it is fast

becoming clear that miniaturization with disregard to changes in the underlying physics is

insufficient for building functional units. Miniaturization raises many new challenges.

Page 100: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

91

Furthermore, some of the promises of microfluidics (integration of all laboratory

functions on a chip, the commercialization of truly hand-held, easy-to-use microfluidic

instruments) have yet to be fulfilled.

6.4. Recommendations for future work In this analysis, we have assumed that the cell/particle structure is rigid, and have

ignored its rotation. A more complete analysis would include the effects of rotation and

structural deformation, as well as the influence of motion on the local fluid flow (i.e. the

coupled structure/fluid interaction). However, these effects are beyond the scope of the

present work.

Due to difficulties associated with the microscale experiments, simulations will

play a more and more important role in the development and optimization of such

phenomena. Therefore, considerable research effort should be put to construct the

foundations for the multiscale methodology and to develop computational capability to

realize such multiscale simulations.

Page 101: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

92

Bibliography 1. Whittaker, E. T. 1910. A history of the theories of aether and electricity from the

age of Descartes to the close of the nineteenth century. Longmans, Green, London, New York.

2. Gravesen, P., J. Branebjerg, and O. S. Jensen. Microfluidics-a review. 1993. Journal of Micromechanics and Microengineering 3:168-182.

3. Ho, C. M., and Y. C. Tai. Micro-electro-mechanical-systems (MEMS) and fluid flows. 1998. Annual Review of Fluid Mechanics 30:579-612.

4. Regnier, F. E., B. He, S. Lin, and J. Busse. Chromatography and electrophoresis on chips: critical elements of future integrated, microfluidic analytical systems for life science. 1999. Trends in Biotechnology 17:101-106.

5. Whitesides, G. M., Stroock, A.D. Flexible methods for microfluidics. 2001. Phys. Today 54:42-48.

6. Manz, A., D. J. Harrison, E. M. J. Verpoorte, J. C. Fettinger, A. Paulus, H. Ludi, and H. M. Widmer. Planar chips technology for miniaturization and integration of separation techniques into monitoring systems : Capillary electrophoresis on a chip. 1992. Journal of Chromatography A 593:253-258.

7. Vladislav Dolník, S. L. S. J. Capillary electrophoresis on microchip. 2000. Electrophoresis 21:41-54.

8. Moore, G. E. Cramming more components onto integrated circuits. 1965. Electronics 38.

9. Greenwood, P. A., and G. M. Greenway. Sample manipulation in micro total analytical systems. 2002. TrAC Trends in Analytical Chemistry 21:726-740.

10. H. A. Stone, S. K. Microfluidics: Basic issues, applications, and challenges. 2001. AIChE Journal 47:1250-1254.

11. Pamme, N. Magnetism and microfluidics. 2006. Lab on a Chip 6:24-38.

12. Gould, P. Nanoparticles probe biosystems. 2004. Materials Today 7:36-43.

13. Sun, S., H. Zeng, D. B. Robinson, S. Raoux, P. M. Rice, S. X. Wang, and G. Li. Monodisperse MFe2O4 (M = Fe, Co, Mn) Nanoparticles. 2004. J. Am. Chem. Soc. 126:273-279.

Page 102: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

93

14. Liu, C., B. Zou, A. J. Rondinone, and Z. J. Zhang. Reverse Micelle Synthesis and Characterization of Superparamagnetic MnFe2O4 Spinel Ferrite Nanocrystallites. 2000. J. Phys. Chem. B 104:1141-1145.

15. Kang, Y. S., S. Risbud, J. F. Rabolt, and P. Stroeve. Synthesis and Characterization of Nanometer-Size Fe3O4 and &#x03B3;-Fe2O3 Particles. 1996. Chem. Mater. 8:2209-2211.

16. Koneracká, M., P. Kopcanský, M. Antalík, M. Timko, C. N. Ramchand, D. Lobo, R. V. Mehta, and R. V. Upadhyay. Immobilization of proteins and enzymes to fine magnetic particles. 1999. Journal of Magnetism and Magnetic Materials 201:427-430.

17. Šafařík, I., and M. Šafaříková. Magnetic Nanoparticles and Biosciences. 2002. Monatshefte für Chemie / Chemical Monthly 133:737-759.

18. Hawkins, T. L., K. J. McKernan, L. B. Jacotot, J. B. MacKenzie, P. M. Richardson, and E. S. Lander. DNA SEQUENCING: A Magnetic Attraction to High-Throughput Genomics. 1997. Science 276:1887-1889.

19. Schuster, M., E. Wasserbauer, C. Ortner, K. Graumann, A. Jungbauer, F. Hammerschmid, and G. Werner. Short cut of protein purification by integration of cell-disrupture and affinity extraction. 2000. Bioseparation 9:59-67.

20. Yazdankhah, S. P., L. Slverd, S. Simonsen, and E. Olsen. Development and evaluation of an immunomagnetic separation-ELISA for the detection of Staphylococcus aureus thermostable nuclease in composite milk. 1999. Veterinary Microbiology 67:113-125.

21. SafarIk, I., and M. SafarIková. Use of magnetic techniques for the isolation of cells. 1999. Journal of Chromatography B: Biomedical Sciences and Applications 722:33-53.

22. Udo Bilkenroth, H. T., Dagmar Riemann, Udo Rebmann, Hans Heynemann, Axel Meye,. Detection and enrichment of disseminated renal carcinoma cells from peripheral blood by immunomagnetic cell separation. 2001. International Journal of Cancer 92:577-582.

23. Luxembourg, A. T., P. Borrow, L. Teyton, A. B. Brunmark, P. A. Peterson, and M. R. Jackson. Biomagnetic isolation of antigen-specific CD8+ T cells usable in immunotherapy. 1998. Nat Biotech 16:281-285.

24. Ganguly, R. 2005. Ferrofluid transport analysis for thermal, biomedical and mems applications. In Mechanical Engineering. University of Illinois at Chicago, Chicago.

25. Dijkhuis, J., and C. Kerkdijk. Upgrading of coal using cryogenic HGMS. 1981. Magnetics, IEEE Transactions on 17:1503-1505.

Page 103: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

94

26. Becker, R. 1982. Electromagnetic fields and interactions. Dover Publications, New York :.

27. Koneracká, M., P. Kopcanský, M. Timko, and C. N. Ramchand. Direct binding procedure of proteins and enzymes to fine magnetic particles. 2002. Journal of Magnetism and Magnetic Materials 252:409-411.

28. Kittel, C. Theory of Long Period Magnetic Relaxation. 1946. Physical Review 69:640.

29. Brown, W. F. Thermal Fluctuations of a Single-Domain Particle. 1963. Physical Review 130:1677.

30. Verdes, C., B. Ruiz-Diaz, S. M. Thompson, R. W. Chantrell, and A. Stancu. Computational model of the magnetic and transport properties of interacting fine particles. 2002. Physical Review B 65:174417.

31. Lofdahl, L., and M. Gad-el-Hak. MEMS applications in turbulence and flow control. 1999. Progress in Aerospace Sciences 35:101-203.

32. Stone, H. A., A. D. Stroock, and A. Ajdari. Engineering flows in small devices: Microfluidics Toward a Lab-on-a-Chip. 2004. Annual Review of Fluid Mechanics 36:381-411.

33. Squires, T. M., and S. R. Quake. Microfluidics: Fluid physics at the nanoliter scale. 2005. Reviews of Modern Physics 77:977-950.

34. Ichikawa, N., K. Hosokawa, and R. Maeda. Interface motion of capillary-driven flow in rectangular microchannel. 2004. Journal of Colloid and Interface Science 280:155-164.

35. Happel, J. 1983. Low Reynolds number hydrodynamics : with special applications to particulate media. M. Nijhoff ; Distributed by Kluwer Boston, The Hague ; Boston : Hingham, MA, USA :.

36. Kleef, R. W. 1984. Particle behavior in high field magnetic flocculation and separation.

37. Warnke, K. C. Finite-element modeling of the separation of magnetic microparticles in fluid. 2003. Magnetics, IEEE Transactions on 39:1771-1777.

38. Grady, K. O. Biomedical applications of magnetic nanoparticles. 2002. Journal of Physics D: Applied Physics.

39. Jiles, D. 1991. Introduction to magnetism and magnetic materials. Chapman and Hall, London ; New York :.

40. Morrish, A. H. 1965. The physical principles of magnetism. Wiley, New York.

Page 104: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

95

41. Oberteuffer, J. Magnetic separation: A review of principles, devices, and applications. 1974. Magnetics, IEEE Transactions on 10:223-238.

42. Olsvik, O., T. Popovic, E. Skjerve, K. S. Cudjoe, E. Hornes, J. Ugelstad, and M. Uhlen. Magnetic separation techniques in diagnostic microbiology. 1994. Clin. Microbiol. Rev. 7:43-54.

43. Moore, L. R., A. R. Rodriguez, P. S. Williams, K. McCloskey, B. J. Bolwell, M. Nakamura, J. J. Chalmers, and M. Zborowski. Progenitor cell isolation with a high-capacity quadrupole magnetic flow sorter. 2001. Journal of Magnetism and Magnetic Materials 225:277-284.

44. Hoyos, M., L. R. Moore, K. E. McCloskey, S. Margel, M. Zuberi, J. J. Chalmers, and M. Zborowski. Study of magnetic particles pulse-injected into an annular SPLITT-like channel inside a quadrupole magnetic field. 2000. Journal of Chromatography A 903:99-116.

45. Furst, E. M. Applications of laser tweezers in complex fluid rheology. 2005. Current Opinion in Colloid & Interface Science 10:79-86.

46. Binyamin, G., T. D. Boone, H. S. Lackritz, A. J. Ricco, A. P. Sassi, S. J. Williams, R. E. Oosterbroek, and B. Albert van den. 2003. Plastic microfluidic devices: Electrokinetic manipulations, life science applications, and production technologies. In Lab-on-a-Chip. Elsevier, Amsterdam. 83-112.

47. Ravula, S. K., D. W. Branch, C. D. James, R. J. Townsend, M. Hill, G. Kaduchak, M. Ward, and I. Brener. A microfluidic system combining acoustic and dielectrophoretic particle preconcentration and focusing. Sensors and Actuators B: Chemical In Press, Corrected Proof.

48. Ivey, M., J. Liu, Y. Zhu, and S. Cutillas. Magnetic-field-induced structural transitions in a ferrofluid emulsion. 2000. Physical Review E 63:011403.

49. Furst, E. M., C. Suzuki, M. Fermigier, and A. P. Gast. Permanently Linked Monodisperse Paramagnetic Chains. 1998. Langmuir 14:7334-7336.

50. Helgesen, G., P. Pieranski, and A. T. Skjeltorp. Dynamic behavior of simple magnetic hole systems. 1990. Physical Review A 42:7271.

51. Meakin, P., and A. T. Skjeltorp. Application of experimental and numerical models to the physics of multiparticle systems. 1993. Advances in Physics 42:1 - 127.

52. Melle, S., O. G. Calderón, M. A. Rubio, and G. G. Fuller. Microstructure evolution in magnetorheological suspensions governed by Mason number. 2003. Physical Review E 68:041503.

Page 105: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

96

53. Yang, F. L., and M. L. Hunt. Dynamics of particle-particle collisions in a viscous liquid. 2006. Physics of Fluids 18:121506.

54. Andersson, J. O., C. Djurberg, T. Jonsson, P. Svedlindh, and P. Nordblad. Monte Carlo studies of the dynamics of an interacting monodispersive magnetic-particle system. 1997. Physical Review B 56:13983.

55. Satoh, A., R. W. Chantrell, S.-I. Kamiyama, and G. N. Coverdale. Three Dimensional Monte Carlo Simulations of Thick Chainlike Clusters Composed of Ferromagnetic Fine Particles. 1996. Journal of Colloid and Interface Science 181:422-428.

56. Grief, A. D., and G. Richardson. Mathematical modelling of magnetically targeted drug delivery. 2005. Journal of Magnetism and Magnetic Materials 293:455-463.

57. Christian Mikkelsen, H. B. Microfluidic capturing-dynamics of paramagnetic bead suspensions. 2005. Lab no a Chip 5:1293-1297.

58. Liu, D., M. R. Maxey, and G. E. Karniadakis. Simulations of dynamic self-assembly of paramagnetic microspheres in confined microgeometries. 2005. Journal of Micromechanics and Microengineering 15:2298-2308.

59. Furlani, E. P. Magnetophoretic separation of blood cells at the microscale. 2007. Journal of Physics D: Applied Physics:1313.

60. Jensen, K. F. Microreaction engineering -- is small better? 2001. Chemical Engineering Science 56:293-303.

61. Sinha, A., R. Ganguly, A. K. De, and I. K. Puri. Single magnetic particle dynamics in a microchannel. 2007. Physics of Fluids 19:117102-117105.

62. K. Smistrup, B. G. K., J. L. Reimers, M. Dufva, J. Petersen and M. F. Hansen. On-chip magnetic bead microarray using hydrodynamic focusing in a passive magnetic separator. 2005. Lab on a Chip 5:1315-1319.

63. Rida, A., and M. A. M. Gijs. Dynamics of magnetically retained supraparticle structures in a liquid flow. 2004. Applied Physics Letters 85:4986-4988.

64. Friedman, G., and B. Yellen. Magnetic separation, manipulation and assembly of solid phase in fluids. 2005. Current Opinion in Colloid & Interface Science 10:158-166.

65. Jiang, Z., J. Llandro, T. Mitrelias, and J. A. C. Bland. 2006. An integrated microfluidic cell for detection, manipulation, and sorting of single micron-sized magnetic beads. In 50th Annual Conference on Magnetism and Magnetic Materials. AIP, San Jose, California (USA). 08S105-103.

Page 106: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

97

66. Smistrup, K., O. Hansen, H. Bruus, and M. F. Hansen. Magnetic separation in microfluidic systems using microfabricated electromagnets--experiments and simulations. 2005. Journal of Magnetism and Magnetic Materials 293:597-604.

67. Furlani, E. P., and K. C. Ng. Analytical model of magnetic nanoparticle transport and capture in the microvasculature. 2006. Physical Review E (Statistical, Nonlinear, and Soft Matter Physics) 73:061919-061910.

68. Verpoorte, E. FocusBeads and chips: new recipes for analysis. 2003. Lab on a Chip 3:60N-68N.

69. McCloskey, K. E., J. J. Chalmers, and M. Zborowski. Magnetic Cell Separation: Characterization of Magnetophoretic Mobility. 2003. Anal. Chem. 75:6868-6874.

70. Inglis, D. W., R. Riehn, R. H. Austin, and J. C. Sturm. Continuous microfluidic immunomagnetic cell separation. 2004. Applied Physics Letters 85:5093-5095.

71. Lund-Olesen, T., H. Bruus, and F. M. Hansen. Quantitative characterization of magnetic separators: Comparison of systems with and without integrated microfluidic mixers. 2006. Biomedical Microdevices 9:195-205.

72. Furlani, E., Y. Sahoo, K. Ng, J. Wortman, and T. Monk. A model for predicting magnetic particle capture in a microfluidic bioseparator. 2007. Biomedical Microdevices 9:451-463.

73. Yellen, B. B., and G. Friedman. Programmable Assembly of Colloidal Particles Using Magnetic Microwell Templates. 2004. Langmuir 20:2553-2559.

74. Zaytseva, N. V., V. N. Goral, R. A. Montagna, and A. J. Baeumner. Development of a microfluidic biosensor module for pathogen detection. 2005. Lab on a Chip 5:805-811.

75. Anne Le Nel, N. M., Marcela Slovakova, Claire Smadja, Jean-Michel Peyrin, Myriam Taverna, Jean-Louis Viovy. 2005. Use of self assembled magnetic particles for on-chip protein digestion. In International conference on miniaturize systems for chemistry and life sciences, Micro Total Analysis Systems 2005, Boston.

76. A. Le Nel, L. K., M. Slovakova, N. Minc, C. Smadja, J.M. Peyrin, J.L. Viovy, M. Taverna, in: T. Kitamori, H. Fujita, S. Hasebe 2006. A microfluidic device based on self-organized magnetic beads: application to the on-chip analysis of the prion protein. In Micro Total Analysis Systems 2006. H. F. T. Kitamori, S. Hasebe, editor, Tokyo, Japan. 350.

77. Bronzeau, S., and N. Pamme. Simultaneous bioassays in a microfluidic channel on plugs of different magnetic particles. 2008. Analytica Chimica Acta 609:105-112.

Page 107: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

98

78. Helgesen, G., A. T. Skjeltorp, P. M. Mors, R. Botet, and R. Jullien. Aggregation of Magnetic Microspheres: Experiments and Simulations. 1988. Physical Review Letters 61:1736.

79. Nandy, K., S. Chaudhuri, R. Ganguly, and I. K. Puri. Analytical model for the magnetophoretic capture of magnetic microspheres in microfluidic devices. 2008. Journal of Magnetism and Magnetic Materials 320:1398-1405.

80. Wirix-Speetjens, R., W. Fyen, J. De Boeck, and G. Borghs. Single magnetic particle detection: Experimental verification of simulated behavior. 2006. Journal of Applied Physics 99:103903-103904.

81. Hayes, M. A., N. A. Polson, and A. A. Garcia. Active Control of Dynamic Supraparticle Structures in Microchannels. 2001a. Langmuir 17:2866-2871.

82. Chubykalo, O., R. Smirnov-Rueda, J. M. Gonzalez, M. A. Wongsam, R. W. Chantrell, and U. Nowak. Brownian dynamics approach to interacting magnetic moments. 2003. Journal of Magnetism and Magnetic Materials 266:28-35.

83. Chen, H., and W. H. Matthaeus. New cellular automaton model for magnetohydrodynamics. 1987. Physical Review Letters 58:1845.

84. Fazekas, S., J. Kertesz, and D. E. Wolf. Piling and avalanches of magnetized particles. 2005. Physical Review E (Statistical, Nonlinear, and Soft Matter Physics) 71:061303-061309.

85. Dance, S. L., E. Climent, and M. R. Maxey. Collision barrier effects on the bulk flow in a random suspension. 2004. Physics of Fluids 16:828-831.

86. Modak, N., A. Datta, and R. Ganguly. Cell separation in a microfluidic channel using magnetic microspheres. Microfluidics and Nanofluidics.

87. Brody, J. P., P. Yager, R. E. Goldstein, and R. H. Austin. Biotechnology at low Reynolds numbers. 1996. Biophysical Journal 71:3430-3441.

88. Yu, H., J. W. Raymonda, T. M. McMahon, and A. A. Campagnari. Detection of biological threat agents by immunomagnetic microsphere-based solid phase fluorogenic- and electro-chemiluminescence. 2000. Biosensors and Bioelectronics 14:829-840.

89. Rotariu, O., N. J. C. Strachan, and V. Badescu. A stochastic model simulating the capture of pathogenic micro-organisms by superparamagnetic particles in an isodynamic magnetic field. 2004. Physics in Medicine and Biology 49:3971-3978.

90. Van Emon, J. M., and C. L. Gerlach. Environmental monitoring and human exposure assessment using immunochemical techniques. 1998. Journal of Microbiological Methods 32:121-131.

Page 108: CHARACTERIZING MAGNETIC PARTICLE TRANSPORT FOR

99

91. Rashkovetsky, L. G., Y. V. Lyubarskaya, F. Foret, D. E. Hughes, and B. L. Karger. Automated microanalysis using magnetic beads with commercial capillary electrophoretic instrumentation. 1997. Journal of Chromatography A 781:197-204.

92. Hayes, M. A., N. A. Polson, A. N. Phayre, and A. A. Garcia. Flow-Based Microimmunoassay. 2001b. Anal. Chem. 73:5896-5902.

93. Strick, T. R., J.-F. Allemand, D. Bensimon, A. Bensimon, and V. Croquette. The Elasticity of a Single Supercoiled DNA Molecule. 1996. Science 271:1835-1837.

94. Jonas O. Tegenfeldt, C. P., Han Cao, RichardL. Huang, RobertH. Austin, StephenY. Chou, EdwardC. Cox and JamesC. Sturm. Micro- and nanofluidics for DNA analysis. 2004. Analytical and Bioanalytical Chemistry:1678–1692.