You are on page 1of 101

Near-Surface Slip Flow and Hindered Colloidal

Diusion at the Nano-Scale


by
Pong-Yu (Peter) Huang
MFL TR 2006-01
Microuidics Laboratory
Division of Engineering
Brown University
Providence, RI 02912
October 2006
This report is the PhD thesis of Peter Huang, supervised by Prof. Kenneth Breuer in the Division
of Engineering, and submitted to the Graduate School at Brown University in September 2006.
For more information contact: kbreuer@brown.edu
i
Abstract
Aqueous boundary solution slip at a solid surface and hindered diusion of suspended near-
surface particles are experimentally investigated with a three-dimensional total internal reection
velocimetry (3D-TIRV) technique. The sub-micrometer penetration depth of evanescent wave oers
low-background-noise images where a 10-nm displacement can be accurately measured. At shear
rates less than 1800 sec
1
, a shear-induced slip length of less than 100 nanometers is observed
at a glass surface of sub-nanometer roughness. Surface hydrophobicity is also found to aid slip.
However, the notion that electrokinetic and electrostatic forces acting on charged tracer particles
lead to slip length measurement inaccuracy is disputed by experiments conducted with aqueous
solutions consisting of electrolytes, which report slip length of similar magnitudes. Anisotropic
hindered diusion of suspended near-surface particles is demonstrated with simultaneous three-
dimensional measurements. Hindered diusion coecients of 1.5-m radius particles within 300
nm from the solid wall are shown to be in close agreement with previously published theories.
Lastly, eects of hindered diusion and wall presence on the accuracies of TIRV and other near-
surface measurement methods are studied via Brownian dynamics simulations. The results reveal
that hindered diusion causes a minor bias toward values obtained at the wall if a small interval is
allowed between successive image acquisitions, while particle drop-outs lead to over-estimation of
uidic velocity values if the time interval between image acquisitions is large. It is also found that
the apparent velocities of near-wall particles are not time-invariant. Still, true uidic velocity values
can be obtained from the particle apparent velocities with proper scaling. Such scaling correction is
applicable to particle tracking velocimetry measurements, but not to correlation based velocimetry
measurements for their lack of accounting of particle drop-outs. Shear eect on near-wall particle
translation in a nite imaging depth results in an asymmetric apparent velocity distribution, which
is another potential source of error for correlation based velocimetries. These ndings lead to a
conclusion that particle tracking based velocimetry methods, such as TIRV, are more superior than
correlation based velocimetry methods are in making near-surface uidic measurements.
Copyright, c _, Pong-Yu (Peter) Huang, 2006
i
Pong-Yu (Peter) Huang was born in Taipei, Taiwan and attended schools there until the
age of fteen. After graduating high school from International School Manila in 1996,
he attended Cornell University where he received a Bachelor of Arts degree in physics in
2000. Upon graduation, he enrolled at the Division of Engineering, Brown University,
to pursue a graduate degree under the guidance of Professor Kenneth Breuer. In 2002,
he completed a Master of Science degree in engineering, and continued to pursue a
Doctor of Philosophy degree since 2003. He currently resides in the New England region
with his wife and plans to pursue an academic career. Mr. Huang enjoys travelling and
anything related to baseball (except sabermetrics), and holds great interests in economic
perspectives of events in the world. He is a proud fan of the New York Mets since 1999.
ii
This dissertation consists of experimental studies of boundary slip and hindered diusion in
the nano-scale. Included are an overall review and summary of the whole dissertation in the
rst chapter, and a collection of journal papers, each as a subsequent chapter, with details of
experimental methods and results. The contents of the chapters are:
Chapter 1. Introduction. Including basic concepts of slip ow, hindered diusion and total
internal reection velocimetry, reviews of previously published works, and discussions of the main
results obtained.
Chapter 2. Paper 1. Near-surface velocimetry using evanescent wave illumination by Song-
wan Jin, Peter Huang, Jinil Park, J Y. Yoo and Kenneth S. Breuer. Experiments in Fluids, Vol.
37, pp. 825-833, 2004.
Chapter 3. Paper 2. Direct measurement of slip velocities using three-dimensional total
internal reection velocimetry by Peter Huang, Jerey S. Guasto and Kenneth S. Breuer. Journal
of Fluid Mechanics, Vol. 566, pp. 447-464, 2006.
Chapter 4. Paper 3. Direct measurement of slip length in electrolyte solutions by Peter
Huang and Kenneth S. Breuer. To be submitted to Physics of Fluids, 2006.
Chapter 5. Paper 4. Direct measurement of anisotropic near-wall hindered diusion using
total internal reection velocimetry by Peter Huang and Kenneth S. Breuer. To be submitted to
Physical Review E, 2006.
Chapter 6. Paper 5. Simulations of hindered diusion in shear ow and its implications for
near-wall velocimetry by Peter Huang and Kenneth S. Breuer. To be submitted to Physics of
Fluids, 2006.
Chapter 7. Concluding remarks and suggested studies.
iii
I would like to rst thank my mentor, Professor Kenneth Breuer, for his guidance and support
through my graduate study. Without his encouragement, constructive criticisms and inexhaustible
knowledge, the projects described in this dissertation would not have been fruitful. I also would
like to thank my dissertation readers, Professor Eric Lauga and Professor Anubhav Tripathi, for
their insightful comments during dissertation revisions. Numerous discussions with Professor Tri-
pathi and Professor Thomas Powers during the years on chemistry and statistical mechanics are
acknowledged and very much appreciated.
A special thank goes to my fellow graduate student and my friend Jerey Guasto. Without
his creative ideas and brilliant suggestions many subtleties of experimental methods and results
would not have been found. I also would like to thank Dr. Songwan Jin, for his introduction of the
total internal reection uorescent microscopy setup. In addition, I am very grateful to Professor
MinJun Kim, Professor Sylvain Cloutier, Jahn Torres, Teng-Fang Kuo, Brian Burke, Jinkee Lee
and Matt Kerby, for their generous help in micro-fabrication, optical system design and setup,
nano-scale measurements and sharing of lab equipments.
As a lowly ranked graduate student, the administrative and research supports of Mr. Jerey
Brown, Mr. Michael Jibitsky, Mr. Brian Corkum, Mr. Charles Vicker and Ms. Virginia Novak
are especially appreciated. Without their kind assistance many research projects would not have
moved forward as smoothly as they did.
Finally I would like to thank my parents, my sister and my extended family for their love,
support and best wishes over the years. My greatest gratitude goes to my wife, for her love,
patience and encouragement throughout the years of graduate study, and for putting up with me
during the ups and downs of experiments.
iv
Contents
1 Introduction 1
1.1 BOUNDARY SLIP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 HINDERED DIFFUSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 TOTAL INTERNAL REFLECTION VELOCIMETRY . . . . . . . . . . . . . . . . 4
1.4 SUBTLETIES OF NEAR-WALL SHEAR FLOW . . . . . . . . . . . . . . . . . . . . 6
2 Paper 1: Near-surface velocimetry using evanescent wave illumination 8
2.1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 EXPERIMENTAL PROCEDURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 RESULTS AND DISCUSSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3.1 Cross-stream Velocity Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3.2 Streamwise Velocity Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.3 Comparisons with Numerical Simulations . . . . . . . . . . . . . . . . . . . . 18
2.4 CONCLUDING REMARKS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3 Paper 2: Direct measurement of slip velocities using three-dimensional total
internal reection velocimetry 22
3.1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.2 THEORETICAL CONSIDERATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2.1 Total Internal Reection Microscopy . . . . . . . . . . . . . . . . . . . . . . . 23
3.2.2 Emission Intensity of Fluorescent Particles . . . . . . . . . . . . . . . . . . . 24
3.2.3 Near-Surface Shear Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3 EXPERIMENTAL PROCEDURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3.1 Materials and Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3.2 Velocimetry Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.4 RESULTS AND DISCUSSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.4.1 Validation of Intensity Calibration Curve . . . . . . . . . . . . . . . . . . . . 31
3.4.2 Hindered Diusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.4.3 Velocity Distributions of Particles . . . . . . . . . . . . . . . . . . . . . . . . 32
3.4.4 Measurements of Apparent Slip Velocities . . . . . . . . . . . . . . . . . . . . 34
3.5 CONCLUDING REMARKS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4 Paper 3: Direct measurement of slip length in electrolyte solutions 39
5 Paper 4: Direct measurement of anisotropic near-wall hindered diusion using
total internal reection velocimetry 43
5.1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
v
5.2 THEORY OF HINDERED DIFFUSION . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.3 EXPERIMENTAL PROCEDURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.4 RESULTS AND DISCUSSION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.5 SUMMARY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
6 Paper 5: Simulations of hindered diusion in shear ow and its implications for
near-wall velocimetry 50
6.1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6.2 THEORIES AND COMPUTATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.2.1 The Langevin Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.2.2 Eects of Shear on Particle Velocities . . . . . . . . . . . . . . . . . . . . . . 53
6.2.3 Hindered Diusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.2.4 Implementation of Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.3 RESULTS AND DISCUSSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.3.1 Sedimentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
6.3.2 Particle Displacement Due to Hindered Diusion . . . . . . . . . . . . . . . . 55
6.3.3 Particle Drop-ins and Drop-outs . . . . . . . . . . . . . . . . . . . . . . . . . 58
6.3.4 Horizontal Apparent Velocity Distributions . . . . . . . . . . . . . . . . . . . 60
6.3.5 Time Evolution of Apparent Velocity Distributions . . . . . . . . . . . . . . . 62
6.4 CONCLUDING REMARKS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
7 Concluding Remarks and Suggested Studies 68
A Calibration of beam incident angle 70
B Shear-induced lift force (or more precisely, lack of ) on near-wall submicron-
particles 73
C Evanescent wave image of micron-sized uorescent particles 75
vi
List of Figures
1.1 Schematic of total internal reection uorescence microscopy . . . . . . . . . . . . . 5
2.1 Schematic of objective-based TIRFM . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2 Schematic of the experimental setup . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Particle images of dierent illumination methods: wideeld (direct, ood) illumination 12
2.4 Particle images of dierent illumination methods: near-wall TIRFM image . . . . . . 13
2.5 Schematic of near-wall particles moving near the surface illustrating the observation
range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.6 Distribution of particle velocity vectors of 200-nm particles . . . . . . . . . . . . . . 14
2.7 Cross-stream Brownian motion of particles in an innite medium and near wall . . . 15
2.8 Experimentally-measured distribution of streamwise velocities of 200-nm diameter
particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.9 Apparent particle velocity vs. shear rate . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.10 Streamwise velocity distribution of 200-nm particles determined from Monte Carlo
simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.11 Streamwise apparent velocity predicted by simulation . . . . . . . . . . . . . . . . . 20
3.1 Objective-based total internal reection uorescence microscopy . . . . . . . . . . . . 24
3.2 The emission intensity distribution of particles in a uniform concentration eld . . . 25
3.3 The predicted position distribution of particles with 0.5 < I
e
/I
e
0
< 1 in a uniform
concentration eld . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.4 The ratio of statistical apparent velocity of particles and mean uid velocity under
no slip . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.5 Image acquisition system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.6 Fluorescent particle intensity as a function of its distance to the glass surface . . . . 30
3.7 Observed uorescent particle intensity distribution and its probability density func-
tion (PDF) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.8 Ratio of hindered diusion coecients (D
exp
) for Brownian motion parallel to a surface 32
3.9 Distribution of observed particle streamwise velocities under various shear rates . . . 33
3.10 Distribution of observed particle streamwise velocities after scaling . . . . . . . . . . 34
3.11 Experimental apparent velocity of particles in a shear ow over a hydrophilic surface 35
3.12 Experimental apparent velocity of particles in a shear ow over a hydrophobic surface 36
3.13 The additional slip length due to surface hydrophobicity . . . . . . . . . . . . . . . . 36
4.1 Objective-based total internal reection velocimetry (TIRV) system . . . . . . . . . 40
4.2 Measured slip lengths of aqueous solutions . . . . . . . . . . . . . . . . . . . . . . . . 42
5.1 Schematic of total internal reection velocimetry . . . . . . . . . . . . . . . . . . . . 45
vii
5.2 Intensity calibration of 1.5-m radius uorescent particles in evanescent eld . . . . 46
5.3 Hindered diusion correction vs. particle/glass gap size . . . . . . . . . . . . . . . . 47
6.1 A schematic of the simulation geometry . . . . . . . . . . . . . . . . . . . . . . . . . 52
6.2 Sedimentation of large particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
6.3 Sedimentation of small particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.4 Particle displacement distribution due to hindered diusion . . . . . . . . . . . . . . 58
6.5 Spatial distribution of small particles due to hindered diusion . . . . . . . . . . . . 59
6.6 Spatial distribution of particles from a nite layer at various times . . . . . . . . . . 59
6.7 Percentage of particles remaining in imaging range . . . . . . . . . . . . . . . . . . . 60
6.8 Apparent velocity distribution of particles remaining in a nite imaging range . . . . 61
6.9 Normalized apparent velocity distribution of particles in a nite imaging range . . . 61
6.10 Apparent velocity distribution of particles from various imaging ranges . . . . . . . . 62
6.11 Collapsed apparent velocity distribution of particles at various imaging range . . . . 63
6.12 Time evolution of apparent velocity distribution . . . . . . . . . . . . . . . . . . . . 63
6.13 Time evolution of mean particle apparent velocity . . . . . . . . . . . . . . . . . . . 64
6.14 Rescaled time evolution of mean particle apparent velocity . . . . . . . . . . . . . . . 65
6.15 Schematic of potential paths for particle translation . . . . . . . . . . . . . . . . . . 66
A.1 Schematic of laser beam refraction at the glass/air interface . . . . . . . . . . . . . . 71
A.2 Photo and schematic of beam incident angle measurement . . . . . . . . . . . . . . . 71
A.3 Plot of laser beam angle vs. converging lens position . . . . . . . . . . . . . . . . . . 72
C.1 Schematic of COMSOL simulation geometry . . . . . . . . . . . . . . . . . . . . . . . 76
C.2 COMSOL simulation of total internal reection . . . . . . . . . . . . . . . . . . . . . 76
C.3 Intensity decay of evanescent eld . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
C.4 COMSOL simulation of a particle with d/ = 0.39 in evanescent eld . . . . . . . . 77
C.5 COMSOL simulation of a particle with d/ = 6 in evanescent eld . . . . . . . . . . 78
C.6 COMSOL simulation of a particle with d/ = 12 in evanescent eld . . . . . . . . . 78
C.7 Overall evanescent energy in the suspended particles . . . . . . . . . . . . . . . . . . 79
C.8 Cross-sectional intensity prole of a 3-m particle (d/ = 6) . . . . . . . . . . . . . . 80
C.9 Digital image of a 3-m particle (d/ = 6) in contact with glass surface . . . . . . . 81
C.10 Cross-sectional intensity prole of a 6-m particle (d/ = 12) . . . . . . . . . . . . . 82
C.11 Digital image of a 6-m particle (d/ = 12) in contact with glass surface . . . . . . . 82
viii
List of Tables
6.1 Sample values of the Peclet number and the sedimentation coecient . . . . . . . . 53
6.2 Representative values of the non-dimensional time between consecutive image acqui-
sition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
ix
Chapter 1
Introduction
Since the introduction of the rst microuidic device, these miniaturized uidic manipulation sys-
tems have been regarded as one of the most promising technologies of the late twentieth century.
In particular, investigations into its application in biotechnology has been the most intense. Exam-
ples of such applications include immunosensors [1], reagent mixing [2], content sorter [3] and drug
delivery [4]. Microuidic devices are very attractive in biotechnology over conventional technology
because they require small sample volume and produce rapid results. Indeed, for chemical reac-
tions whose reaction sites are at solid surfaces, the high surface-area-to-volume-ratio characteristic
of microuidics oers a much higher eciency [5]. On the other hand, the high surface-area-to-
volume-ratio also means that near-surface phenomena will have a much larger inuence on the bulk
of the uid content, for example the viscous drag on the channel walls. In both cases, understand-
ing of interactions between the uid content and the solid boundary is critical in designing and
analyzing microuidic devices.
Under most circumstances, the solid boundary is viewed as rigid and inert such that physical
and structural changes due to uidic forces are nonexistent. Thus the majority of interesting
surface-induced physical phenomena occur in the near-surface region of the uid phase and can be
categorized into two groups: (1) changes of the uid mechanical characteristics due to the presence
of the solid surface; (2) interactions between the dissolved molecules, suspended particulates and
the solid surface. Examples of physical phenomena in the former group include electrokinetic ow
[6], slip ow [7] and surface chemistry directed ow [8, 9], while particle or cell adhesion [10] and
detachment [11], increased hydrodynamic drag [12], electrostatic interactions [13] and eects of
depletion layers [14] are eects of the latter group.
1.1 BOUNDARY SLIP
One of the uid mechanics questions that have remained unanswered over centuries is the slip ow
and its origin. Unlike polymer solutions whose slip behaviors have been thoroughly investigated
[15, 16, 17], the existence of boundary slip in a Newtonian liquid has been a controversial topic. Even
though the idea of a slip ow has been considered by scientists as early as Newton [18], empirical
data up to the late twentieth century had deemed it negligible. The strong experimental support
of such assumption could be due to the fact that the magnitude of slip velocity was much smaller
than the measuring instrument accuracy and thus remained undetectable. Another possibility is
the relatively insignicant impact of slip on the uid bulk ow in the macroscopic scale. Westin et
al. [19] calculated that the ratio of slip velocity to bulk velocity becomes very small if the channel
height is more than a few hundred micrometers. Therefore when macroscopic ow measurements
1
2
were conducted slip eect could be hardly noticeable.
In the past decade slip ow began to receive more attention again, because some earlier ex-
periments [20, 21] and molecular dynamics simulations [22, 23, 24] have reported slip boundary
conditions in Newtonian liquid. Since then many experimental attempts were made to conrm the
existence and the magnitude of slip velocities, with varying degrees of accuracy and uncertainty. In
some of these experiments, slip velocity was indirectly inferred from measurement of other physical
quantities. For instance, the relationship between applied pressure drops and measured ow rates
was used by Choi et al. [25] to infer slip velocity, while the forces required to move two surfaces
separated by a thin lm of the test uid were employed Zhu & Granick [26], Neto et al. [27] and
Cottin-Bizonne et al. [28] to calculate slip velocities. Others took a direct approach of studying
slip by measuring near-surface ow velocities. Examples of the direct measurements include uo-
rescence recovery after photobleaching (FRAP) by Pit et al. [29] and micro-scale particle image
velocimetry (PIV) by Tretheway & Meinhart [30] and Joseph & Tabeling [31].
These experimental studies produced various magnitudes and dependencies of boundary slip.
The commonly used formulation to characterize boundary slip is the slip length, , which is based
on the Navier hypothesis that the velocity of a uid at a solid surface is proportional to the shear
stress. That is,
=
U
slip

s
, (1.1)
where U
slip
is the slip velocity and
s
is the uid shear rate at the solid surface. Experimentally
measured slip lengths were reported to range from micrometers [20, 30, 32], hundreds of nanometers
[29, 33] to less than 100 nanometers [25, 26, 27, 28, 31]. Disagreement also exist on whether the
slip behavior is shear-dependent [21, 25, 26, 27] or shear-independent [20, 28, 29, 31, 32]. Of
particular interest to microuidics is aqueous solution slip over dierent surface conditions as water
is the most common solvent in biological and chemical applications. Although it is generally agreed
among research ndings that surface hydrophobicity aids slip eect [25, 30], again the magnitude
of such an eect remains controversial.
The ability of total internal reection velocimetry (TIRV, to be introduced later) in performing
near-surface measurements makes it a perfect technique for direct observation of slip velocities in a
shear ow. In Paper 1: Near-surface velocimetry using evanescent wave illumination (chapter 2), we
report on our rst attempt to measure slip velocities of deionized water using the TIRV technique.
In this report, it is shown that velocities of uid elements within a few hundred nanometers from
a solid surface can be accurately measured, demonstrating the potential of the TIRV technique in
making direct measurements of slip velocities. Although a quantitative conclusion of slip length
magnitude could not be drawn due to the uncertainty in determination of the exact imaging range,
a qualitative conclusion is reached that surface slip length of an aqueous solution would be small
and surface chemistry-dependent, and a slip length above a few hundred nanometers is unlikely.
The accuracy of slip measurements is further improved with the implementation of three-
dimensional tracking in the TIRV technique. In Paper 2: Direct measurement of slip velocities
using three-dimensional total internal reection velocimetry (chapter 3), we report on a slip length
of deionized water being less than 100 nm, and an additional slip length of 16 nm is found at-
tributable to surface hydrophobicity. In conducting this experiment, several physical factors that
are suspected of inducing slip are also examined. With measurements conducted at glass sur-
faces of sub-nanometer surface roughness, the obtained slip length values agree with the surface-
roughness dependency previously reported by Granick et al. [34] The issue of apparent slip caused
by nano-bubbles at the glass surface is additionally disputed with experiments conducted with
sub-nanometer surface roughness and degassed test uids.
3
Still, questions remained on whether other physical forces, particularly the electrostatic and
electrokinetic eects proposed by Lauga et al. [7, 35], would lead to false slip results in TIRV mea-
surements. We address such a concern in Paper 3: Direct measurement of slip length in electrolyte
solutions (chapter 4), by repeating the slip measurements with electrolyte solutions whose ionic
concentrations have been predicted to reduce the apparent slip eect to sub-molecular level. It is
observed that the presence of electrolytes has no eect on the measured slip lengths, suggesting
that the observed slip velocities are most likely not due to electrostatic and electrokinetic eects,
but are consequences of true boundary slip.
1.2 HINDERED DIFFUSION
Another near-surface phenomenon with great uid mechanical and mass transport implications is
hindered diusion of colloidal particles. The concept of near-wall hindered diusion stems from an
increased frictional coecient on a near-wall object. Einstein proposed that the diusion coecient
of a colloidal object is the ratio of the uid thermal energy to the objects frictional coecient in
Stokes ow. Thus if a particle experiences an increase in uidic friction as it approaches a solid
wall, its diusive translation or Brownian motion will be hindered. Brenner [36] and Goldman et al.
[37, 38] were the rst to analytically solve the increased frictional coecients of a near-wall sphere
with a lubrication formulation. Their results were later re-conrmed by Chaoui & Feuillebois [12]
with an spherical harmonics expansion method.
Besides being a uid mechanics theory, hindered diusion has additional implications in chemical
engineering and drug delivery systems since diusion is the dominant mass transport mechanism
in microuidics. Because of its signicance, advancement of near-surface measurement techniques
had lead to several experimental studies of hindered diusion. Employed experimental techniques
include evanescent light scattering o freely suspended particles by Bevan & Prieve [39], evanescent
light-scattering spectroscopy by Hosoda et al. [40], catch-and-release video microscopy with optical
tweezers by Lin et al. [41], evanescent light scattering coupled with optical radiation pressure
by Oetama & Walz [42] and three-dimensional ratiometric total internal reection uorescence
microscopy with nanoparticles by Banerjee & Kihm [43]. Most of these studies have reported
hindered diusion coecients agreeing to values derived from the theories of Brenner and Goldman
et al.
In the theories of Brenner and Goldman et al., the increased friction factor is a function of
particle radius and gap size between the particle and the solid wall. Unlike in the uid bulk,
the uidic friction also becomes anisotropic in the near-wall region. In other words, the hindered
diusion coecient is dierent for directions normal and parallel to the solid surface. Few studies
have attempted to measure such anisotropicity simultaneously due to experimental limitations. Lin
et al. [41] demonstrated such anisotropicity for particles whose gap-size-to-radius ratio is greater
than one, and found hindered diusion coecient values agreeing with theories. Another study
that reported anisotropic hindered diusion coecients for gap-size-to-radius ratio greater than
one was conducted by Banerjee & Kihm [44]. However, their study reported that the measured
diusion coecient in the direction normal to the solid wall agreed with the theoretical values only
when large particles were tested.
By applying the 3D-TIRV technique to freely suspended micron-sized uorescent particles, we
aim to simultaneously observe the three-dimensional anisotropic hindered diusion for gap-size-to-
radius ratio much less than one. The theory under this condition is vastly dierent from that of the
gap-size-to-radius ratio greater than one, and based on our literature survey no such experimen-
tal conrmation of the theory has been reported. In Paper 4: Direct measurement of anisotropic
4
near-wall hindered diusion using total internal reection velocimetry (chapter 5), we demonstrate
that the three-dimensional tracking technique of TIRV can be adopted into 3D displacement mea-
surements of freely suspended 1.5-m radius particles. The displacement measurements reveal that
the hindered diusion coecients are in close agreement with the theoretical values predicted by
the asymptotic solutions of Brenner and Goldman et al. for gap-size-to-radius ratio much less than
one, and hindered diusion anisotropicity is simultaneously demonstrated in all data sets.
1.3 TOTAL INTERNAL REFLECTION VELOCIMETRY
With so much interest in near-surface phenomena, researchers have developed various techniques
to study them. Optical microscopy has been widely used to observe interactions in the micrometer
scale. However, as fabrication technology advances, the denition of near-surface has also evolved
from microscopic and to the nano-scale. Traditional optical techniques are no longer sucient now
because the visible wavelength is the physical limit of optical resolution ( 0.5 m). A frequently
used technique to overcome this obstacle is evanescent wave imaging [45]. In this technique, a beam
of collimated laser light is brought to a solid/liquid interface through the solid phase at a large
incident angle, as shown in gure 1.1. Since solids typically have higher indices of refraction than
uids do, total internal reection occurs at the interface if the incident angle is greater than the
critical angle predicted by Snells Law. In fact, at the nano-scale this reection is not total, but
a small fraction of energy, called evanescent waves, extends into the uid phase. In the uid phase
the evanescent wave energy or intensity, I, decays exponentially with distance, z, away from the
two-medium interface by
I(z) = I
0
e
z/p
, (1.2)
where I
0
is the intensity at the interface and p is known as the evanescent wave penetration depth.
The penetration depth, which characterizes the length scale of the evanescent eld, can be calculated
from
p =

0
4
_
n
2
s
sin
2
n
2
l
_

1
2
, (1.3)
where
0
is the wavelength of the incident light, is the beam incident angle and n
s
and n
l
are
the refractive indices of the solid and liquid, respectively. For visible light, the evanescent wave
energy decays to zero within a few hundred nanometers. Thus by using the evanescent wave as
an illumination source, one can observe dynamics of uids and suspended particulates within a
few hundred nanometers from the solid surface, while contents outside of this layer would remain
invisible. Consequently near-surface phenomena can be observed with clarity and without the
interfering noise contributed from objects in the uid bulk.
Since the 1970s, biologists have combined uorescence and evanescent wave imaging in studying
near-surface biological events [46, 47]. Termed total internal reection uorescence microscopy
(TIRFM), this technique allows observation of uorescently labelled samples placed in the evanes-
cent eld and under a high magnication microscope. Fluorescence is a quantum optical eect
of a unique group of molecules called uorophores, whose quantum states can be excited by a
specic range of visible light and release photons of a longer wavelength after initial excitation.
Their molecular sizes are excellent for probing near-surface regions at the nano-scale. Examples of
TIRFM include the previously mentioned uorescent recovery after photobleaching (FRAP) and
uorescence resonance energy transfer (FRET). In FRAP, the uorophores of a chosen spot in-
side the microscope eld of view is purposely photobleached with a high power laser beam. The
light-emitting uorophores in the surrounding area will diuse into this dark spot immediately
following the photobleaching. By measuring the diusivity of these light-emitting uorophores, one
5

z
liquid
solid
~ penetration depth
Figure 1.1: Schematic of total internal reection uorescence microscopy (TIRFM). A uorescent
particle suspended in water was placed in an evanescent eld. If the illumination beam incident
angle, , is greater than the critical angle predicted by Snells law, total internal reection occurs
at the solid/liquid interface. The evanescent energy then illuminates the encapsulated uorophores
inside the particle.
can study the targeted molecule mobility. For instance, FRAP has been reported in measurement
of protein mobility and activity in living cells [48]. FRET, on the other hand, takes advantages
of a quantum energy transfer that occurs between matched uorescent molecules of two dierent
wavelengths [49]. Because such energy transfer can only occur if the two molecules are less than
10 nm apart, they oer an opportunity to examine interactions in the molecular level.
Surprisingly, a long time had passed before physical scientists nally caught up with the merits
of evanescent wave imaging. At rst, evanescent energy scattered o non-uorescent suspended
particles was used in quantitative measurements of near-surface dynamics. Such experimental
studies include measurement of surface charge densities on suspended particles by von Grunberg
et al. [50], quantifying double-layer repulsion between a colloidal particle and a glass substrate by
Flicker et al. [51], eects of particle size on near-wall particle depletion by Kok et al. [14], charged
particle adsorption by Kun & Fendler [52], and hindered diusion by Bevan & Prieve [39]. Only in
the past decade were uorescent tracer particles used in microscopic dynamics measurements.
Particle-based velocimetry has long been used in uid visualization and measurement [53]. It
is based on an intuitive and for most part correct assumption that the seeding tracer particles are
carried by the uid surrounding them, and therefore their translational velocities must be that of the
local uid elements. Apparent velocities of the tracer particles are calculated based on displacements
of the tracer particles and the time between successive particle imaging. When particle-based
velocimetry methods were adopted to study microuidics, sub-micron uorescent tracer particles
were used to minimize light scattering and imaging noise while attaining spatial resolutions of
tens of nanometers [54]. Santiago et al. [55] were the rst to demonstrate a correlation-based
microscopic particle image velocimetry (PIV) to map out a velocity eld around a Hele-Shaw cell.
In their experiment, cross-correlation analysis was performed on successive images of uorescent
tracer particles to identify the most probable ensemble displacements. Since their report, PIV
has been the work horse among diagnostic techniques of microuidics.
When applying to near-surface measurements, PIV does suer a major drawback. Its resolu-
tion in the direction perpendicular to the imaging focal plane is limited by the focal depth of the
6
microscope objective lens and is at least 0.5 m [31]. Consequently, PIV could not distinguish
physical features within 0.5 m from a solid surface. Zettner & Yoda [56] overcame PIVs de-
ciency by combining it with evanescent wave illumination. The short eective range of evanescent
wave allowed them to measure uid velocities that were within 280 nm from a solid surface. Li
et al. [57] further extended this nano-PIV technique into a multilayer correlation analysis, where
quantitative measurement can be performed on events that occur within 100 nm from the surface.
Zettner & Yodas nano-PIV is undoubtedly a novel technique. However it still has two signicant
limitations. Firstly, rapid Brownian motions of nanoparticles could lead to a large amount of particle
drop-in and drop-out, or events where particles diuse in and out of the evanescent eld between
consecutive frames. Since the statistical analysis of nano-PIV does not make an attempt to treat the
drop-ins and drop-outs, signicant measurement errors could persist. Secondly, because diusion
is treated as measurement noise in the cross-correlation analysis of the nano-PIV, it is unsuitable
if diusive dynamics is of interest.
In Paper 1: Near-surface velocimetry using evanescent wave illumination (chapter 2), we present
a technique, called total internal reection velocimetry (TIRV), as an alternative to nano-PIV
in near-surface measurements. By combining TIRFM with tracking of individual tracer particle,
Brownian motions of the particle ensemble can be monitored and statistically analyzed, leading
to a more accurate accounting of all contributing particles. In Paper 2: Direct measurement of
slip velocities using three-dimensional total internal reection velocimetry (chapter 3), we extend
the TIRV technique to three-dimensional, further improving its accuracy and applicability. It is
observed that the peak intensity of a uorescent particle decays exponentially in the same length
scale as the evanescent wave energy, and thus a particles peak intensity can be used to infer its
distance from the solid surface.
As mentioned previously, we use the TIRV technique to measure aqueous solution boundary
slip and near-surface hindered diusion of spherical particles. In Paper 2: Direct measurement
of slip velocities using three-dimensional total internal reection velocimetry (chapter 3) and in
Paper 3: Direct measurement of slip length in electrolyte solutions (chapter 4), 100- and 200-nm
radius uorescent particles are employed in 3D-TIRV experiments to measure slip lengths over
hydrophilic and hydrophobic surfaces. The analysis of the image data reveals that a statistical
approach is needed in making near-surface velocimetry measurement even for particles whose size
variation is as small as a few percent, owing to the sensitivity of evanescent wave imaging. In Paper
4: Direct measurement of anisotropic near-wall hindered diusion using total internal reection
velocimetry (chapter 5), the TIRV technique is extended to observing anisotropic near-wall hindered
diusion of 1.5-m radius uorescent particles. Under evanescent wave imaging, the peak intensity
of these large particles follow an exponential decay relation that is also identical to the exponential
decay of evanescent energy, and thus the same three-dimensional particle tracking principle applies.
Both slip and hindered diusion studies demonstrate the applicability and versatility of the TIRV
technique in making near-wall particle velocity and displacement measurements.
1.4 SUBTLETIES OF NEAR-WALL SHEAR FLOW
As previous discussion have emphasized, Brownian motion of particles in the near-surface region is a
non-negligible mechanism and has signicant implications in the accuracy of near-wall velocimetry.
Many of the issues that need to be considered, however, could not be easily assessed through
experimental data. Thus numerical approaches oer insightful alternatives to explore the subtleties
of near-wall ow.
The main concern in near-wall velocimetry is the ensemble behavior of tracer particles in the
7
time interval between consecutive image acquisitions, and more importantly, how such ensemble
behavior aects the accuracy of velocity measurements. Two equally viable numerical approaches
have been proposed to study Brownian motion-related problems [58]. In the Fokker-Planck ap-
proach, the time evolution of the particle ensemble phase space conguration function is computed
through numerical solution of the momentum-position partial dierential equations. The ensemble
behaviors are then obtained from probability analysis using the phase space conguration function.
Recently, Sadr et al. [59] had taken the Fokker-Planck approach to study the diusion-induced bias
in near-wall velocimetry.
The other approach takes advantage of the Langevin equation in which the displacement of
a particle consists of a deterministic component (such as displacement of the suspending uid)
and a stochastic component (such as Brownian motion). With a proper time step size, initial and
boundary conditions, ensemble behavior of particles can be obtained through statistical analysis of
a large number of repeated simulations. Also called Brownian Dynamics (BD) simulations for its
direct emulation of Brownian motions, this approach has been used in studying the binding rates
between attached molecules [60], colloidal particle deposition in a microchannel ow [61], accuracy
of potential energy proles determined by evanescent wave scattering [62] and the accuracy of
nano-PIV [63]. Its intuitive resemblance to a physical velocimetry experiment has gained some
popularity in colloidal studies.
We are most interested in the ensemble behaviors of near-surface particles in a shear ow, and
the eects of the wall presence and hindered diusion on the measured velocities because these
are issues that we have noticed during our TIRV measurements of boundary slip and anisotropic
hindered diusion. In Paper 5: Simulations of hindered diusion in shear ow and its implications
for near-wall velocimetry (chapter 6), the relative strengths of near-wall shear and diusive forces
are explored through Brownian dynamics simulations. It is observed that the shear force dominates
over diusion at Pe > 3. In addition, asymmetric shape and width of apparent velocity distributions
are attributed to the shear eect, and we believe it can lead to signicant inaccuracy in nano-PIV
if not treated appropriately.
The same Brownian dynamics simulation also provides a tool to assess the eects of hindered
diusion and wall presence on various near-wall velocimetry techniques. With diusive time and
particle radius as time and length scales, respectively, it is found that if the time interval between
successive particle image acquisitions is short, hindered diusion causes an under-estimation of
uid velocity due to biased sampling of tracer particles very close to the wall. However, if the
time interval between successive particle image acquisitions is long, a signicant portion of the
tracer particles would sample planes of higher velocities, resulting in an over-estimation of overall
uid velocity. Particle drop-ins and drop-outs, are found to be a potential source of measurement
error for nano-PIV because of the rapid Brownian motion of nanoparticles. The combined eect of
the observations described above is that near-wall particle-based velocimetry measurement is not
time-invariant. Still, the results found in the Brownian dynamics simulation oer a way where the
true uidic velocities can be inferred from particle apparent velocities through proper scaling. This
scaling method, however, is available only for TIRV but not for nano-PIV, making TIRV the more
superior method in near-surface measurements.
Chapter 2
Paper 1: Near-surface velocimetry
using evanescent wave illumination
Total Internal Reection Velocimetry (TIRV) is used to measure particle motion in the near-wall
region of a microuidic system. TIRV images are illuminated with the evanescent eld of an
incident laser pulse and contain only particles that are very close to channel surface. Sub-micron
sized uorescent particles suspended in water are used as seed particles and their images are analyzed
with a PTV algorithm to extract information about apparent slip velocity. At relatively low shear
rates (less than 2500 sec
1
), a velocity, proportional to the shear rate was observed. The statistical
dierence between velocities measured over hydrophilic and hydrophobic surfaces was found to be
minimal. The results suggest that the slip length, if present, is less than 25 nm, but uncertainty
regarding the exact character of the illumination eld prevents a more accurate measurement at this
time. Numerical simulations are presented to help understand the results and to provide insight into
the mechanisms that result in the experimentally observed distributions. Issues associated with the
accuracy of the experimental technique and the interpretations of the experimental results are also
discussed.
2.1 INTRODUCTION
Over the past few years, studies regarding near-wall uid motion have intensied because of the
emergence of micro-electro-mechanical-systems (MEMS), where the uid-surface interaction is im-
portant to the design, fabrication and performance of these devices. In particular, the no-slip
boundary condition, which has long been experimentally veried in macroscopic ow, now receives
a great amount of attention in microuidics. Moreover, many biological applications of microu-
idic devices require surface coatings that change the wetting ability of aqueous or biological uids.
Thus, a better understanding of uid-surface interaction becomes necessary in designing applicable
devices.
Many experimental techniques have been applied to study the eect of slip on both hydrophilic
and hydrophobic surfaces. In general, slip measurement techniques can be classied into indirect
and direct methods. In the indirect methods, slip velocity is inferred from other measurements, for
example, the ow rate and pressure within the ow duct [20, 21, 25], or from forces required to move
the surfaces [64]. Examples of the direct measurement methods include hot lm anemometry [32],
total internal reection-uorescence recovery after photobleaching (FRAP) [29] and micro-PIV [30].
There are also other groups of researchers who attempted to understand the slip phenomenon from
8
9
a theoretical perspective through molecular dynamics simulations [22, 23, 24]. The results from
these studies are quite diverse, and apparent slip lengths ranging from nanometers to microns
have been reported. Similarly, the dependence of the slip on the applied shear rate is also a matter
of debate with some results suggesting a constant slip length, while others reporting a slip length
dependent on shear rate and/or surface hydrophobicity.
Recently, a novel technique termed total internal reection uorescence microscopy (TIRFM)
has been applied to near-wall uid measurements. The TIRFM technique has long been used in
biological studies, such as cell-substrate contacts, vesicle fusion and single-molecule observation
[65, 66, 67]. It takes advantage of the fact that total internal reection occurs when light travels
from a dense medium (such as a glass cover slip) with refractive index n
1
into a less dense medium
(such as the sample uid) with refractive index n
2
. Total internal reection occurs if the incident
angle is larger than the critical angle
c
= sin
1
(n
2
/n
1
), which can be derived from Snells Law.
When total internal reection occurs, some fraction of the light energy penetrates into the less dense
medium and propagates parallel to the surface. This parallel light wave is known as the evanescent
wave. One characteristic of the evanescent wave is that its intensity, I, decays exponentially with
respect to the distance from the two-media interface. Thus, images taken under evanescent wave
illumination have low background noise, showing clear features of sample structure that is near the
interface. There have also been several recent experimental studies of uid mechanics by applying
evanescent wave illumination [29, 56, 68]. In particular, Zettner and Yoda [56] used the PIV
technique combined with evanescent wave illumination to measure near-surface velocity eld in
rotating Couette ow.
In this paper, we describe the Total Internal Reection Velocimetry (TIRV) technique, in which
the TIRFM technique is applied to observe particle motion in the close vicinity of microchannel
walls with dierent surface wetting properties. A particle tracking velocimetry (PTV) technique is
applied to track particle displacements and ow velocities near the wall.
2.2 EXPERIMENTAL PROCEDURES
TIRFM requires the creation of total internal reection of a laser beam at the glass-sample interface,
and the presence of uorescent molecules or particles suspended in the sample uid for imaging
and measurement. Two dierent optical congurations are commonly used to create an evanescent
eld close to the glass-sample interface. The rst method utilizes a prism to guide the laser beam
into the glass cover slip, which then serves as a waveguide, allowing total internal reection to
occur. The other method, which is prismless, has the illuminating laser beam directed through a
high numerical aperture (NA) microscope objective at an angle that creates total internal reection
at the glass-sample interface, as shown in Fig. 2.1. The maximum incident angle which can be
achieved by using an objective lens is given by

max
= sin
1
(NA/n
glass
) . (2.1)
As mentioned before, the intensity of an evanescent wave decays exponentially with distance z
away from the two-media interface
I(z) = I
0
e
z/d
, (2.2)
where I
0
is the light intensity at the interface and d is known as the evanescent wave penetration
depth. The penetration depth, which characterizes the evanescent eld, can be calculated as:
d =

0
4
_
n
2
1
sin
2
n
2
2
_

1
2
, (2.3)
10
~ Penetration depth
Fluorescent particle solution
Objective lens
Immersion oil
Glass (n = 1.51)

Water (n = 1.33)
z
x
U
Figure 2.1: Schematic of objective-based TIRFM.
where
0
is the wavelength of the incident light and n
1
and n
2
are the indices of refraction for glass
and water, respectively. Based on these calculations, one can choose appropriate laser wavelength
and microscope objective or prism to achieve desired penetration depth.
In the present experiment, illustrated in Fig. 2.2, we have adopted the prismless TIRFM
method. A dual head Quantel Brilliant Q-Switched Nd:YAG laser was used to generate a pair
of 532 nm, 5 nsec laser pulses to illuminate the ow. The laser heads were timed to re with a
pre-determined pulse separation (1.1 to 3.3 ms, depending on shear rate), illuminating two images
for use in a Particle Tracking Velocimetry (PTV) analysis. The beams were focused o-axis at the
back focal plane of a Nikon PL Apo NA 1.4 100X oil immersion objective lens mounted on a Nikon
Eclipse TE200 inverted epi-uorescent microscope. Because the laser beam is focused o-axis, the
beam incident angle at the glass-uid interface can be controlled by changing the radial location
of the laser beam. With water as the sample uid, the experimental setup had a critical angle

c
= 61.9

. At incident angles equal to or larger than


c
, the laser beam could be totally reected
o the water-glass interface, creating an evanescent wave eld. With a water-glass interface,
0
=
532 nm and NA = 1.4, it can be calculated that the penetration depth is approximately 150 nm
when the incident angle is = 63.4

.
The test cell consisted of a precision glass cover slip on which a polydimethylsiloxane (PDMS)
microchannel was attached. The PDMS channels were fabricated using standard soft lithography
techniques [69] and measured 311 m deep, 2501 m wide and 25 mm long. They were bonded
to the test surface after exposure to an oxygen plasma which renders both the glass and PDMS
surfaces highly hydrophilic for a short time during which they bond on contact. Following bonding,
the channels were immersed in deionized-water to maintain their hydrophilic surface characteristics.
To create hydrophobic microchannels, it is necessary to coat the glass surface with a transparent
and hydrophobic substance because glass is naturally hydrophilic. We coated glass by immersing
glass into a 5-mM octadecyltrichorosilane (OTS) solution (dissolved in hexadecane) at room tem-
perature for several hours. A self-assembled monolayer (SAM) of OTS formed on the glass surface,
and was made permanent after the glass was rinsed in hexane and ethanol, dried in nitrogen, and
baked at 100

C on a hotplate for a few hours. The coating process was performed in a nitrogen gas
chamber to avoid bulk polymerization. The coated glass was bonded to a cleaned, patterned PDMS
channel to create a hydrophobic microchannel. It has been reported that surface roughness can
11
ICCD
Nd-YAG laser
532 nm
Syringe pump
100X, NA = 1.4
Objective lens
Dichroic mirror and
Barrier filter
PDMS channel
Convex lens
Pulse generator Computer
1360 x 1036 x 12 bit
532 nm, 5 nsec pulse
1.1 - 3.3 ms separation
Figure 2.2: Schematic of the experimental setup.
strongly inhibit boundary slip by transferring momentum if liquid molecules are trapped inside the
holes of the monolayer [29]. Thus, an atomic force microscope (AFM) measurement of glass surface
roughness was also performed. The polished Pyrex glass wafer was measured to have a surface
RMS roughness of 2 nm, while the same wafer had a surface roughness of 4 nm after OTS-coating.
Monodisperse Fluorescent Polymer Microspheres (Duke Scientic) with diameters of 200 nm
(5%) and 300 nm (3%), suspended in pure water at 0.02% and 0.04% volume fractions respec-
tively, were used as seed particles in this experiment. These microspheres have a peak absorption
wavelength of 542 nm and emit a red visible light at 612 nm. Particle solutions were owed through
the test channels using a Harvard Apparatus 22 syringe pump, at ow rates ranging from 0.6 to
6.0 L/min (0.35%). An air gap between the piston and the uid in the syringe and long tubes
were used to dampen pulsations from the syringe motion. Near-wall particle images were projected
onto a Q-Imaging Intensied Retiga CCD camera, which has a capacity of recording 1360-pixels-
by-1036-pixels 12-bit images. Under 100X magnication, each image pixel corresponds to 64.3 nm
in the ow channel and the diraction-limited spot size is 30.9 m [70]. For 200- and 300-nm
particles, the image diameters are 36.8 and 43.1 m, respectively [55]. Typically 200 image pairs
were recorded for each ow condition. The recorded image pair was nally analyzed with a PTV
algorithm written in our group using the MATLAB programming environment. The choice of Par-
ticle Tracking Velocimetry (PTV), rather than Particle Image Velocimetry (PIV) is an important
one and this will be re-addressed later on.
Measurements were taken at several shear rates. For the results presented here, the shear
rate was estimated from the applied volume ow rate, Q, assuming steady laminar ow of an
incompressible uid between two parallel plates without wall slip

wall
=
_
du
dy
_
wall
=
6Q
h
2
w
, (2.4)
12
Figure 2.3: Particle images of dierent illumination methods: wideeld (direct, ood) illumination.
where h is the channel depth and w is the channel width. If there is signicant slip at the wall, the
shear rate could be slightly lower than is estimated here [25]. However, for the current experiments,
any possible slip velocity is so small that this uncertainty can be neglected (this presumption will
be conrmed by the results in the following section).
2.3 RESULTS AND DISCUSSION
Figures 2.3 and 2.4 compare particle images of wideeld (direct, ood) illumination and evanescent
eld illumination. A large amount of background noise is observed in the case of direct illumina-
tion. Such background noise is attributed to the light emitting from the out-of-focus particles in
the bulk of the uid and such images are common in micro-PIV measurements but lead to poor
cross-correlations during PIV analysis. Evanescent eld illumination, on the other hand, eliminates
much of the background noise because the evanescent light penetration depth is restricted to ap-
proximately 150 nanometers, and particles in the bulk uid are not illuminated. This characteristic
allows easy detection of only particles that are close to the channel surface.
To measure particle velocities near the channel surface, several hundred image pairs at various
low shear rates were taken for subsequent PTV image analysis. Bright particles with intensities
above a pre-determined threshold value were identied in each image. In this study, the intensity
threshold was set at 50% of the brightest particle intensity (usually about 1500-2000 counts from
a 12 bit CCD camera). This threshold value allowed us to detect almost every particle which can
be identied with the naked eye and restricts the analysis to particles that are well within the
evanescent eld and in focus. The number of particles identied in one image is not very sensitive
to the exact threshold value. Typically about 10-20 particles were identied in each image, although
approximately 50% of these were not tracked between images due to out-of-plane diusion which
causes particles to enter and leave the observation region.
It is important to point out that the determination of an intensity threshold sets an observation
range, as illustrated in Fig. 2.5. The lower bound of the observation range is the particle radius,
13
Figure 2.4: Particle images of dierent illumination methods: near-wall TIRFM image.
Glass Cover Slip
Fluid Velocity
Observation
Range
Figure 2.5: Schematic of near-wall particles moving near the surface illustrating the observation
range.
representing a particle in contact with the channel surface. As the particle moves farther from the
wall, and hence into a region of lower illumination intensity, the emitted intensity also falls. Thus,
the intensity threshold chosen sets an upper bound on the observation range which is dierent from
the penetration depth of the evanescent eld. The exact size of the observation range was not
known in these experiments and this uncertainty turned out to be the major source of uncertainty
in these experiments and is discussed later on.
Although tracking the particle displacements in the x-y plane (parallel to the surface) was
straightforward, tracking the particles motion in the z -axis (normal to the surface) is not easy.
Ovryn [71] developed a method to determine the three-dimensional position of a spherical particle,
based on resolving the spatial variations in the scattering pattern. However in the evanescent eld,
illumination intensity varies exponentially with distance from the wall (as shown in Eq. 2.2), and
consequently it is not easy to directly quantitatively associate the uorescent scattering of a particle
in a non-uniform illumination eld with the particles distance from the wall. For this reason we
14
100 50 0 50 100
40
20
0
20
40
60
80
100
120
140
Cross-stream velocity (m/sec)
S
t
r
e
a
m
w
i
s
e

v
e
o
c
i
t
y

(

m
/
s
e
c
)

Figure 2.6: Distribution of particle velocity vectors of 200-nm particles with = 469 sec
1
.
did not attempt to track the particles motion normal to the surface. There is no reason why this
should not be possible in the future.
The intensities of pixels surrounding the peak intensity were tted to a two-dimensional Gaus-
sian curve [72, 73, 74]. Because the particles are known to be spherical with diameter standard
deviation less than 5%, a two-dimensional Gaussian curve t closely approximates the Lommel func-
tion, and allowed us to locate the particle center coordinates with sub-pixel resolution. Particles
were matched between images using a custom particle tracking algorithm and displacement vectors
were calculated from the displacements of particle centers. A threshold displacement was imposed
to detect error vectors, which were dened as displacements larger than deemed reasonable. These
large displacement vectors usually result from matching two dierent particles mistakenly. At each
testing condition, the total number of validated displacement vectors ranged from 700 to 1500. Us-
ing the geometric scale and the applied pulse separation, a velocity vector can be calculated from
each displacement vector. Figure 2.6 shows an example of a collection of velocity vectors obtained
for a single shear rate. The Brownian motion is particularly strong due to the particles small size.
2.3.1 Cross-stream Velocity Analysis
Analysis of the displacements of particles in the cross-stream direction was used as to check the
consistency of the experiment results. In this direction, in the absence of any mean ow, one would
expect that the particle motion would be purely driven by thermal uctuations. Indeed, we found
that the distributions of spanwise velocities had a Gaussian distribution with mean and skewness
close to zero. The width of the particle displacement distribution, , was used to compute the local
diusion coecient, D:
D =

2
t
2
, (2.5)
15
0 500 1000 1500 2000 2500
0
0.5
1
1.5
2
2.5
x 10
13
Shear rate (1/sec)
D
i
f
f
.

C
o
e
f
f
.

x

R
a
d
i
u
s

(
c
m
3
/
s
e
c
)
Figure 2.7: Cross-stream Brownian motion of particles in an innite medium and near wall (ex-
periment). Solid symbol represents 200-nm particles and open symbol represents 300-nm particles
(, hydrophilic surface; , hydrophobic surface). The dashed line denotes particle diusion in an
innite medium calculated on the basis of the Stokes-Einstein relation.
and Fig. 2.7 shows the cross-stream diusion coecient obtained for 200-nm (solid symbol) and
300-nm (open symbol) particles at a variety of dierent streamwise shear rates over both hydrophilic
and hydrophobic surfaces. The diusion coecient is scaled by the particle radius, r, and the dotted
lines represent scaled particle diusion in an innite medium calculated based on the Stokes-Einstein
relation:
D
0
r =
k
B
T
6
, (2.6)
where D
0
is the Stokes-Einstein diusion coecient, k
B
is the Boltzmann constant, T is the medium
temperature, and is the uid viscosity. As expected, the measured spanwise diusion coecients
are independent of the streamwise shear rate and the diusion coecient, when appropriately
scaled, is independent of particle radius. However, as indicated by the dotted line in Fig. 2.7,
the measured values for D are approximately 30% that of the bulk diusion expected. This is
conrmation of the presence of hindered diusion of particles near a solid surface [37, 38]. If we use
the observed value of D with the theory for hindered diusion [37, 38] to calculate the apparent
particle distance from the wall, dened as the distance between the particle center and the wall, we
nd it to be 110 nm for the 200-nm particles and 165 nm for the 300-nm particles. This is consistent
with the high intensity threshold chosen which denes a small observation range. Unfortunately,
since we did not carefully monitor the observation range, a more precise estimate is not available
for this data.
2.3.2 Streamwise Velocity Analysis
Figure 2.8 shows an example of the distribution of the streamwise velocities, found by projecting the
velocity distributions onto the mean ow direction. Several features are worth noting. Firstly, the
location of the peak of the distribution increases with increasing shear rate, as one would expect.
16
200 0 200 400 600 800
0
0.01
0.02
Velocity ( m/sec)
P
D
F
234 (1/sec)
1172 (1/sec)
2344 (1/sec)
Figure 2.8: Experimentally-measured distribution of streamwise velocities of 200-nm diameter par-
ticles over a range of shear rates.
Less expected, and in contrast to the distributions for the spanwise velocities, is the observation
that the distribution widens with increasing shear rate, and that the skewness clearly shifts from
0.15 at the low shear rate (barely noticeable) to a clearly observable positive value of 0.85 (a
stronger positive tail) at the highest shear rate presented. In this case the most-probable velocity
is quite dierent from the average velocity, raising a fundamental question about the denition of
the actual uid velocity in terms of the observed particle velocity distribution. Figure 2.9 shows
the variation of the mean velocities as a function of shear rate, the error bars depict the standard
deviation about the mean. The velocities fall on a straight line through the origin, and there are no
discernible dierences between the hydrophilic and hydrophobic surfaces and, more interestingly,
between the 200- and 300-nm diameter particles.
The reason for the observed increase in the distribution width and skewness can be explained
in terms of the observation range from which particle motions are recorded. At each shear plane,
the true velocity distribution for tracer particles has a Gaussian shape with a mean equal to the
ow velocity at that plane, and standard deviation determined by the temperature and particle
size. However, any micro-PTV or PIV system, including the current TIRV integrates through the
observation range and thus samples particle motion from a range of shear planes over which the
mean velocity changes. Thus, for a xed observation range, the range of mean velocities sampled
increases as the shear rate increases, leading to the observed broadening of the velocity distribution
for the streamwise velocity. A similar argument can be used to explain the increasing skewness
at higher shear rates. Here, we note that, due to the presence of the wall, a particle is more
likely to move upwards (and thus experience a higher local velocity) than to move downwards (and
experience a lower local velocity). Thus, higher velocity uctuations are more probable than lower
ones in the vicinity near the wall (a similar argument is used to explain positive skewness in the
streamwise velocity distribution in a turbulent wall-bounded ow).
The lack of any dierences between the mean velocities obtained for 200-nm and 300-nm par-
ticles is more troubling. One would expect that the 300-nm particles would sample shear planes
at higher local velocities than those sampled by 200-nm particles, and thus the average velocities
17
0 500 1000 1500 2000 2500
0
100
200
300
400
Shear rate (1/sec)
A
p
p
a
r
e
n
t

v
e
l
o
c
i
t
y
(

m
/
s
e
c
)
Figure 2.9: Apparent particle velocity vs. shear rate (, 300-nm particle, hydrophilic surface; ,
300-nm particle, hydrophilic surface; , 200-nm particle, hydrophilic surface; , 200-nm particle,
hydrophilic surface).
should be larger for the larger tracer particles. The fact that this is not observed remains a some-
what puzzling result and one that needs to be resolved in future measurements. However, there are
several factors that are not yet well understood that could contribute to this counter-intuitive ob-
servation, including uncertainty in the measurement of the illumination observation range, particle
motions in addition to pure translation (such as rolling motion), and other forces on the particles
(e.g. colloidal).
The drift velocities measured can result from two eects: (i) the nite thickness of the observa-
tion range and (ii) the possible existence of a slip velocity at the surface. We can assess the relative
contributions of these eects using a very simple theoretical model. If we assume a constant shear
very near the surface, we can write the local velocity prole as:
u(y) = u
s
+ y, (2.7)
where u
s
is the slip velocity. If we use a Navier representation for the slip velocity, u
s
= l

, (where
l

is the slip length) and compute the average velocity in the mean depth of the observation range,
, we nd that the average, or apparent slip velocity is
u = (l

+) , (2.8)
The slope of the data in Fig. 2.9 is approximately 100-nm, representing this combination of
both the penetration depth and the slip length. Given that the observation range is estimated to be
approximately 100 - 150 nm, this suggests that the slip length, if present, is likely no more than 10
nm, consistent with the results of Choi et al. [25] for these low shear rates. At the present, our lack
of quantitative knowledge about the evanescent wave penetration depth (which probably varied
from experiment to experiment) and the observation range makes a more precise estimate for the
slip length impossible at this time. However, with the appropriate characterization and consistent
control of the observation range, this measurement will be possible and such experiments are
currently underway. Finally, we note that a slip length of order 1000 nm (as proposed by Tretheway
18
and Meinhart [30], and Zhu and Granick [64]) is completely unsupported by these data, as it would
predict apparent slip velocities at least an order of magnitude larger than observed, and would be
independent of any reasonable estimate for the observation range.
2.3.3 Comparisons with Numerical Simulations
The experimental results were compared with a numerical simulation of near-wall Brownian motion
in shear ow. The motivation for this was to generate an intuitive model with which to help
understand the experimental results. This simple two-dimensional numerical model was developed
to simulate the Brownian motion of particles with radius, r, in a no-slip steady-state Couette ow
between parallel innite plates separated by a distance h, the experimental channel depth. At
the beginning of the simulation, 10,000 particles were uniformly distributed between r and , the
assumed upper bound of the observation range. At each time step, each particle was moved by
a linear combination of the mean shear ow (evaluated at the particles geometric center) and
Brownian motion (sampled from a normal distribution with a zero mean and a standard deviation
of

2Dt). Other eects such as particle rotation, etc were not considered at this time.
The Brownian motion was modied to account for the hindered diusion both parallel and
perpendicular to the wall [36, 37, 38]. For the Brownian motion normal to the wall [36], the
modied diusion constant, D

, is such that
D

D
0
=
_
4
3
sinh

n=1
n(n + 1)
(2n 1) (2n + 3)
_
2 sinh (2n + 1) + (2n + 1) sinh 2
4 sinh
2
_
n +
1
2
_
(2n + 1)
2
sinh
2

1
__
1
,
(2.9)
where D
0
is the Stokes-Einstein diusion coecient, = cosh
1
(z/r) and z is the distance between
the geometrical center of a particle and the wall. For the Brownian motion parallel to the wall, the
modied diusion constant, D

, is such that [37]


D

D
0
=1
9
16
_
h
r
_
1
+
1
8
_
h
r
_
3

45
256
_
h
r
_
4

1
16
_
h
r
_
5
+O
_
h
r
_
6
,
(2.10)
by the method of reection [37, 38] for (z r)/r > 1, and
D

D
0
=
2
_
ln
_
hr
r
_
0.9543

_
ln
_
hr
r
_
2
4.325 ln
_
hr
r
_
+ 1.591
, (2.11)
which is the asymptotic solution [37, 38] for (z r)/r < 1.
After 1000 time steps, particles that remained within the observation range were counted and
their displacements from their initial positions were calculated to obtain a particle displacement
distribution, similar to that obtained in the physical experiments. To emulate the experiment,
plate separation distance, particle radii, ow shear rate and the total time (1000 time steps) were
chosen to be equal to measurement conditions, and the assumed observation range was chosen to
be 100 - 175 nm from the wall.
19
200 0 200 400 600 800
0
0.005
0.01
0.015
0.02
Velocity(m/sec)
P
D
F
234 (1/sec)
1172 (1/sec)
2344 (1/sec)
Figure 2.10: Streamwise velocity distribution of 200-nm particles determined from Monte Carlo
simulation at observation depth = 175 nm.
Problems can arise when the simulated particle is predicted to pass through the solid surface
during the nite time step. In the present simulation two dierent wall boundary conditions were
tested to deal with this situation. The rst condition applied a specular reection, in which particle
was simply reected o the wall and back into the ow. The second condition emulated near-wall
lubrication condition and simply allowed the particle to remain at the wall until Brownian motion
perpendicular to the wall transported it back into the bulk ow, one or more time steps later. Only
a few particles needed to be treated with these heuristic boundary conditions, and it was found
that the two models gave almost indistinguishable results, with less than 1% statistical dierence
in the mean and standard deviation of the resultant particle distributions. Figure 2.10 shows
the streamwise velocity distributions of 200-nm particles at various shear rates, while Fig. 2.11
shows the variation of the mean velocities and their standard deviations, denoted as error bars, as
a function of shear rate. The rather simplistic nature of the simulation is such that one should
not make direct, quantitative, comparisons with the experimental results (Fig. 2.8 and 2.9). In
addition, the choice of observation range for the computation has not yet been optimized to match
the simulation results with those of the experiments. Nevertheless, the similarities between the two
gures are striking, showing similar increases in the mean, standard deviation and (less obvious,
but present) skewness of the distributions as the shear rate increases.
Unlike the experimental results in Fig. 2.9 which did not show any dierence in measuring
velocity using dierent particle size, there is a clear dierence in the estimated velocity using
dierent particle size in Fig. 2.11. As discussed earlier this dierence due to particle size is to
be expected for a xed observation range, and this result adds weight to our suspicion that the
observation range in the experiments were not consistent, and that future experiments should pay
more attention to this important issue.
20
0 500 1000 1500 2000 2500
0
100
200
300
400
500
Shear rate (sec
1
)
A
p
p
a
r
e
n
t

v
e
l
o
c
i
t
y

(

m
/
s
e
c
)
Figure 2.11: Streamwise apparent velocity predicted by simulation (, 200-nm particles; , 300-nm
particles) at the observation depth of 175 nm.
2.4 CONCLUDING REMARKS
We have demonstrated that Total Internal Reection Velocimetry (TIRV) can be a very powerful
technique for the study of near-wall ows in microuidic systems. TIRV images have very low
background noise that allows accurate PIV and PTV measurements of ow velocities within a few
hundred nanometers of the surface. By using sub-micron seed particles, we have demonstrated
the ability to measure both velocity distributions and mean velocities in the plane parallel to the
surface.
The determination of velocities in the streamwise and spanwise directions is extremely accurate,
and comparable to standard microPIV techniques, and is limited by the same constraints - the
accuracy of particle centroid location, diraction-limited imaging of sub-micron particles, etc [70].
However, the technique also promises the ability to use intensity information to obtain wall-normal
displacements as well as precise estimates for the slip velocity at the surface. The present imple-
mentation did not (unfortunately) control with sucient accuracy or consistency the characteristics
of the evanescent wave penetration depth which is very dependent on the incident beam angle. For
this reason, we were not able to measure vertical particle motion and our current resolution of the
slip length measurement is limited such that we can only state that a slip length, if present, is of
order 10 nm or less. However, the estimate for the slip length is limited only by the accuracy with
which we can measure the observation range and so we expect to be able to make measurements
with uncertainties less than 5 nm in the near future.
The accuracy of the TIRV system can still be improved by several means that are subject to
future investigation. First, by using a shorter illumination wavelength, the penetration depth of
evanescent eld can be decreased. Secondly, smaller seed particles can move closer to the channel
surface than larger particles, although they also suer from higher Brownian diusion and so the
particle drop-out will increase. Both changes will allow direct measurements of ow velocities that
are closer to the channel surface, where slip might contribute more signicantly to the local ow
velocity.
21
Finally, and most importantly, the technique does raise fundamental questions regarding the
use of passive tracer particles to estimate the velocity eld in microuidic systems. Our results
reveal several features of particle motion in the near vicinity of the surface, including non-Gaussian
Brownian motion (due to the nite thickness of the observation range) and hindered diusion due
to near-wall lubrication eects. There are also eects other than the translational hydrodynamic
drag that can contribute to the particle motion such that the velocity measurements might be
compromised. One is the possibility of particle rotation; a second is the possibility of particle-
wall colloidal interactions and the relative roles of gravity, thermal noise and electrostatic eects
(both the bead and the surface have surface charges). This is an area that has been studied by
a number of researchers, including Flicker et al. [51], Crocker and Grier [75] and Behrens and
Grier [76] who have identied colloidal forces that are highly dependant on bead diameter, surface
charge, ionic strength (and for pure water such as used in these experiments, the rate at which
ions leak from the glass into solution). A full treatment of these issues is beyond the scope of the
current paper, and indeed, the experimental techniques described here to can be used in further
studies of such phenomena. Perhaps the most important consequence of this is that techniques that
treat ensembles of particles (such as PIV which looks at the collective motion of particles in an
interrogation region, and often averages cross-correlations before peak detection) might be suspect
close to walls due to particle interference eects which cannot be neglected and where the particle
motion is not represented by Gaussian statistics (and hence the average particle velocity is dierent
from the most probably particle velocity). For this reason, tracking each particle independently
and then computing statistics from the individual particle motions might yield a more accurate
view of the near-wall velocity eld.
Chapter 3
Paper 2: Direct measurement of slip
velocities using three-dimensional
total internal reection velocimetry
The existence and magnitude of slip velocities between deionized water and a smooth glass surface
is studied experimentally. Sub-micron uorescent particles are suspended in water and imaged
using Total Internal Reection Velocimetry (TIRV). For water owing over a hydrophilic surface,
the measurements are in agreement with previous experiments and indicate that slip, if present, is
minimal at low shear rates, but increases slightly as the shear rate increases. Surface hydrophobicity
is observed to induce a small slip velocity, with the slip length reaching a maximum of 96 nm at a
shear rate of 1800 s
1
. Issues associated with the experimental technique and the interpretation of
results are also discussed.
3.1 INTRODUCTION
The century-old assumption of the no-slip boundary condition between a liquid and a solid has been
challenged by recent experimental results and molecular dynamic simulations and has been the
subject of many recent investigations exploring ows over both wetting and non-wetting surfaces.
Although there is considerable disagreement regarding the existence of slip over a hydrophilic
surface, it is generally believed that surface hydrophobicity aids slip, while the exact mechanism
is not yet understood (for a review of this topic, see [7]). Experimental studies have reported
a wide range of slip lengths, ranging from micrometers [20, 30, 32, 64], hundreds of nanometers
[29, 33], to tens of nanometers or smaller (including no-slip) [25, 26, 27, 28, 31]. Molecular dynamics
simulations, on the other hand, suggest small slip lengths, mostly less than 100 nm [22, 23, 24, 77,
78, 79].
The experiments have been conducted under various direct and indirect measurement techniques
with varying accuracy and uncertainty. In indirect measurements, slip velocity is inferred from
other ow quantities, such as the relationship between ow rate and pressure drop [25] or the
measurement of the forces required to move a pair of crossed cylindrical surfaces separated by a
thin lm of the test uid [26, 27, 28, 64]. In contrast, alternative techniques such as micro-particle
image velocimetry (PIV) have resulted in direct measurements of near-surface uid velocity using
tracer particles [30, 31].
Although the direct measurements have a strong appeal, there are two major drawbacks inher-
22
23
ent to conventional micro-PIV. The rst is that the velocity is determined from the motion of a
collection of particles in an interrogation area (IA) and, in order to obtain a reasonable estimate of
the velocity, the IA must be measuring several hundred nanometers in size [30]. Hence the determi-
nation of phenomena which might measure a few nanometers is very dicult to resolve. A second
problem is that the size of measurement volume along the optical axis for ood-illuminated micro-
PIV techniques is dened by the focal depth of the optical system, which is typically of the order
of 0.5 to 1 micron [30, 31]. Although Joseph & Tabeling [31] reported a slip length measurement
with 100-nm resolution, micro-PIV would still be insucient to provide an armative conclusion
if the true slip length is below 100 nm.
One attractive approach that solves both of these problems is the use of Total Internal Reection
Velocimetry (TIRV) [80] which uses total internal reection of an incident laser pulse to generate a
highly localized illumination of the surface ow, and relies on tracking the motion of single tracer
particles rather than the collective particle motion that is used in PIV. Total internal reection
uorescence microscopy (TIRFM) has long been used by biologists to study near-surface dynamics
[65, 66, 67]. Zettner & Yoda [56] combined micro-PIV and TIRFM to study near-surface Couette
ow while Pit et al. [29] measured slip velocity of hexadecane over a sapphire substrate using
TIRFM and uorescence recovery after photobleaching. We rst reported on the TIRV technique
in the context of slip velocities for near-wall microows [80], but were unable to make a conclusive
determination of the size or character of any slip lengths due to diculties in the determination
the location of tracer particles within the narrow evanescent illumination region.
The present manuscript reports on experiments in which the TIRV technique has been sub-
stantially improved to the point that quantitative statements can be made regarding the nature of
slip velocities over hydrophilic and hydrophobic surfaces. By measuring the intensity of tracer par-
ticles uorescence, and in conjunction with a statistical model for the optical and hydrodynamic
behaviour of small particles near a surface, we are able to track their motion within a narrow region
above the solid surface and to compare that motion to the motion predicted for dierent levels of
boundary slip.
3.2 THEORETICAL CONSIDERATIONS
3.2.1 Total Internal Reection Microscopy
The evanescent eld and its application to uorescence microscopy has been thoroughly documented
in literature [e.g. 40, 44, 47, 56, 63, 80]. In summary, an evanescent eld can be created near a
solid-liquid interface where total internal reection occurs, as illustrated in gure 4.1. The eld
intensity, I, decays exponentially with distance, z, away from the two-medium interface by
I(z) = I
0
e
z/p
, (3.1)
where I
0
is the intensity at the interface and p is known as the evanescent wave penetration depth.
The penetration depth, which characterizes the length scale of the evanescent eld, can be calculated
from
p =

0
4
_
n
2
1
sin
2
n
2
2
_

1
2
, (3.2)
where
0
is the wavelength of the incident light, is the beam incident angle and n
1
and n
2
are
the indices of refraction of the solid and liquid, respectively.
24
Figure 3.1: Objective-based total internal reection uorescence microscopy.
3.2.2 Emission Intensity of Fluorescent Particles
For spherical tracer particles with a uniform volumetric uorophore distribution in an evanescent
eld, Kihm et al. [44] calculated the particles emission intensities as an exponential function
of their distances to a substrate surface. Through experience we have found, however, that the
intensity of uorescence emission from a large number of nominally identical particles can vary due
to a variety of factors, including statistical variations in illumination intensity, quantum eciency of
the imaging system and, most importantly, a distribution in the physical size of the particles. If we
parameterize all of these variations by an eective emission radius, r, we can model the particles
probability density function (PDF) of its eective emission radius with a Gaussian distribution
P(r) = P
0
exp
_

_
r
a
1
_
2

2
r
_
, (3.3)
where a is the mean eective emission radius of the batch,
r
is the characteristic variation (standard
deviation) and P
0
is a normalization constant. Based on the calculation of Kihm et al. [44] and
the assumption that emission intensity of a uorescent particle is proportional to the number of
uorophores within its volume, the emission intensity, I
e
, of a particle in an evanescent eld is
I
e
(r, h) =
_
r
a
_
3
_
I
e
0
e
(ha)/d
_
, (3.4)
where h is the distance from the particles centre to the substrate surface at which total internal
reection occurs, I
e
0
is the emitted intensity of a single particle with r = a located at h = a (ie.
touching the substrate surface), and d is an intensity decay length that need not be the same
as the evanescent penetration depth. The intensity decay length can be obtained experimentally
by statistically measuring intensities of particles as a function of particles distances away from
the substrate. Combining (3.3) and (3.4), one gets the joint PDF of the emitted intensity and a
particles distance to the substrate as
P(I
e
, h) = exp
_

_
3
_
_
I
e
I
e
0
_
e
(ha)/d
1
_
2

2
r
_

_
. (3.5)
25
0 0.5 1 1.5 2 2.5
0
0.5
1
1.5
Normalized Intensity
P
D
F

r
= 0.05

r
= 0.1

r
= 0.15

r
= 0.2
Figure 3.2: The emission intensity distribution of particles in a uniform concentration eld. The
normalized intensity in the horizontal axis is dened as I
e
/I
e
0
, and the calculation is performed
with a normalized observation range of 0 < h/a < 3 and a normalized intensity decay length of
d/a = 1.8.
When an ensemble of tracer particles are imaged, they would certainly be located at a range of
distances from the substrate surface. The lower limit of this imaging range is h = a, representing a
particle touching the substrate surface. The upper limit of the imaging range is determined by the
sensitivity of the recording medium, such as an intensied CCD camera used in this study. For a
given imaging range of h
1
< h < h
2
, the PDF of a tracer particles intensity is given by
P(I
e
, h
1
< h < h
2
) =
_
h
2
h
1
P(I
e
, h)c(h) dh, (3.6)
where c(h) is the concentration prole of particles in the uid. A sample plot of this distribution
(3.6) is shown in gure 3.2 for a range of values of
r
. For a low variation in eective emission
radius (
r
small), we see that the particles in the dened imaging range are sharply dened in
intensity, with the brightest particles located at the surface (I
e
/I
e
0
= 1) and the faintest particles
at the outer edge of the evanescent eld (I
e
/I
e
0
= 0.2). However, as the variation in particle radius
increases, the transitions become more blurred, and some larger particles close the wall contribute
to the long tail at high observed intensities, while small particles close to the edge of the imaging
range are represented by the smooth transition close to I
e
/I
e
0
= 0.
When performing image analysis, however, one need not analyse all imaged particles. Instead,
using particle intensity as a guide, one can choose to identify and analyse only the particles within
a certain desirable intensity range, ignoring particles that appear too bright or too dim. Equation
(3.5) could also describe the particle position distribution within a normalized intensity range of
< I
e
/I
e
0
< by
P (h, < I
e
/I
e
0
< ) = c(h)
_

P(I
e
, h) d
_
I
e
I
e
0
_
. (3.7)
26
100 150 200 250 300 350
0
2
4
6
8
x 10
3
Particle Center Distance to Wall (nm)
P
D
F
Figure 3.3: The predicted position distribution of particles with 0.5 < I
e
/I
e
0
< 1 in a uniform
concentration eld. In this prediction, a = 100 nm, d = 181 nm and
r
= 0.18.
An example of this position distribution (3.7) is shown in gure 3.3 for a normalized intensity
range of 0.5 < I
e
/I
e
0
< 1. The most obvious feature of gure 3.3 is that the distribution is non-
uniform with the majority of particles contributing to this range of intensities coming from close
to the substrate surface (< 250 nm). In addition, one can observe that the distribution has a long
positive tail that gradually decays to zero at far away into the uid bulk. Because of the large
particle eective radius variation in this example, a large particle far away (for example, > 300
nm) can exhibit in the same the same uorescence intensity as that of smaller particles close to
the substrate surface. These observations reinforce the fact that a particles intensity cannot be
monotonically related to its distance to the solid surface, and attention should be given to any
intensity-based quantitative analysis by recognizing the non-uniform position distribution and the
contribution of particles far away from wall.
3.2.3 Near-Surface Shear Flow
Although we are interested in determining the velocity of a pure uid near a solid surface, we
actually measure the velocity of small spheres near the surface, and it is critical to recognize the
dierences between these two quantities. It is well known that shear and near-surface hydrodynamic
eects can cause a tracer particle to rotate and translate at a velocity lower than the local velocity
of the uid in the same shear plane [12, 38]. Goldman et al. [38] proposed that a particles
translational velocity, U, in a shear ow with local shear rate, S, is given by
U
hS
1
5
16
_
h
r
_
3
, (3.8)
valid for large h/r, and
U
hS

0.7431
0.6376 0.2 ln
_
h
r
1
_, (3.9)
27
1 1.5 2 2.5 3 3.5 4
0.4
0.5
0.6
0.7
0.8
0.9
1
h
2
/r

V
h
1
/r = 1
h
1
/r = 1.5
Figure 3.4: The ratio of statistical apparent velocity of particles and mean uid velocity under no
slip in an imaging range of h
1
< h < h
2
, which is non-dimensionalized by particle radius.

V is
computed with linear uid velocity prole, which is a good approximation near a surface. In the
computation of

U, c(h) is assumed to be uniform.
valid for small h/r. Although there exists no analytical approximation for intermediate values of
h/r, Pierres et al. [60] proposed a cubic approximation to numerical values presented by Goldman
et al. [38]:
U
hS

_
r
h
_
exp
_
0.68902 + 0.54756
_
ln
_
h
r
1
__
+ 0.072332
_
ln
_
h
r
1
__
2
+0.0037644
_
ln
_
h
r
1
__
3
_
.
(3.10)
The apparent velocity,

U, of a large ensemble particles chosen from a normalized intensity range of
< I
e
/I
e
0
< and located in an imaging range of h
1
< h < h
2
is given by the average of the local
velocity integrated over the imaging range:

U =
1
h
2
h
1
_
h
2
h
1
U (h, a, S (h)) P(h, < I
e
/I
e
0
< ) dh. (3.11)
This average apparent velocity,

U, can be compared to the mean velocity, of the pure uid,

V ,
calculated for the same imaging range. The ratio of the two is shown in gure 3.4, from which
we see that the average apparent velocity of particles is, not surprisingly, signicantly lower than
that of pure uid velocity when the imaging range is very close to the substrate surface. This
dierence must be incorporated into the data analysis, and failure to recognize this eect can lead
to misinterpretation of the measured data and underestimation of slip lengths.
It should be noted that (3.11) is based on a no-slip condition at the uid/solid interface. If
there exist a slip velocity, U
slip
, at the solid boundary, the apparent velocity of the same ensemble
28
of particles would be

U
app
= U
slip
+

U = S
w
+

U, (3.12)
where the slip velocity is characterized as the product of the slip length, , and the wall shear rate,
S
w
.
3.3 EXPERIMENTAL PROCEDURES
3.3.1 Materials and Setup
A schematic of the image acquisition system is shown in gure 3.5. Fluorescent particles were
illuminated with an objective-based TIRFM method, as detailed in [80]. A pair of 532-nm, 5-
nsec laser pulses (Quantel Twins Q-Switched Double Pulse Nd:YAG Laser) were directed through
a Nikon PL Apo NA 1.45 100X TIRF oil immersion objective at an angle that created total
internal reection at a glass-water interface, thus illuminating the near-surface region of water with
an evanescent eld (gure 4.1). Fluorescent images of near-surface particles were projected onto
a Q-Imaging Intensied Retiga CCD camera (ICCD), capable of recording 1360-pixels-by-1036-
pixels 12-bit images. Under 100X magnication, each image pixel corresponds to 64.3 nm in the
ow channel and the diraction-limited spot size is 19.7 m (calculated based on [70]). A TTL
pulse generator (Berkeley Nucleonics Model 555) was used to synchronize laser pulses and image
acquisitions. The energy of each illuminating laser pulse was recorded simultaneously with each
image acquisition and this was used to adjust the resultant images and thus account for pulse-to-
pulse variations in illumination intensity.
Test channels were fabricated using a polydimethylsiloxane (PDMS) (Dow Corning Sylgard 184)
molding technique [69] and bonded onto polished glass wafers. The dimensions of the test channels
were 50 1 m deep, 250 1 m wide and 15 mm long. To create hydrophilic microchannels,
the pre-bonding PDMS surface was exposed to an oxygen plasma and immediately immersed in
deionized-water after bonding to a glass wafer. Hydrophobic channels, on the other hand, were
fabricated by bonding untreated PDMS channels onto octadecyltrichorosilane(OTS)-coated glass
wafers. Detailed coating procedures can be found in [80]. It has been reported that water has a
contact angle of 120 degrees with a similarly prepared OTS-coating [30, 81]. The roughness of the
test surface was measured using an atomic force microscope prior to bonding. The hydrophilic and
hydrophobic surfaces had an RMS surface roughness of 0.47 nm and 0.35 nm respectively.
Monodisperse Fluorescent Polystyrene Microspheres (Duke Scientic) with diameters of 200 nm
( 5%) suspended in puried water (Fluka) at 0.04% volume fraction were used as tracer particles
in this experiment. These microspheres have a peak absorption wavelength of 542 nm and emit a
red light at 612 nm. There are three reasons that 200-nm diameter particles were selected as probes
in this experiment. First, smaller size uorescent particles have much larger size variation, making
relation of particle center distance from the solid wall and their uorescent intensities less reliable.
Secondly, smaller particles do not contain enough uorophores to provide signicant signals with
a 5-nsec laser illumination. Lastly, 200-nm particles do not diuse too far during the inter-frame
time, with diusion out of imaging range being our major concern.
A Harvard Apparatus 22 syringe pump was used to maintain particle solution ow rates, ranging
from 1 to 10 L/min (0.35%). To maintain steady pumping motions of the Harvard Apparatus
stepper motor, we purposely introduced a large volume of air between the piston and the uid in
the syringe. This volume of air acted as a pressure shock absorber to damp out the oscillatory
forcing of the syringe pump screw. In addition, a long tubing (> 1 m long) was used to connect
29
ICCD
Nd-YAG laser
532 nm
Syringe pump
100X, NA = 1.45
Objective lens
Dichroic mirror and
Barrier filter
PDMS channel
Convex lens
Pulse generator
Computer
1360 x 1036
x 12 bit
532 nm, 5 nsec pulse
2 - 5 ms separation
Power
Attenuator
Half-wave
plate NI_DAQ
Mirror
Convex Lens
10 micron pinhole
Convex Lens
Polarizing
Beamsplitter
Pulse Energy
Meter
Figure 3.5: Image acquisition system.
the syringe and the channel inlet. Particle ow motion was observed to be very steady under this
setup.
1200 images were captured at each ow rate, as well as in quiescent uid. Once all necessary
images of particle motion were captured, the test channel was lled with rhodamine-B solution.
Images of the rhodamine-B solution were subsequently captured under the same evanescent illumi-
nation, and averaged to obtain an image characterizing the spatial distribution of the illuminating
laser beam.
It has been suggested that nanobubbles and a gas bubble layer on hydrophobic surfaces could
be alternative sources of apparent slip [7, 34, 82]. Prior to each experiment, the uid was degassed
by placing the solution in vacuum for at least 30 minutes. This procedure has been reported to
signicantly eliminate nanobubble formation [83].
It is intuitive to assume that one can infer the location of a particle in the evanescent illumination
eld based on the uorescence intensity. Kihm et al. [44] used a ratiometric uorescence intensity
to track particle motions three-dimensionally. Although this is theoretically feasible, in practice,
successful use of this technique requires precise knowledge of the illumination beam incident angle
and a solution of Maxwells equation for an evanescent eld in a three media system (glass, water
and polystyrene), which can be dicult to express explicitly. Thus, an experimental method was
devised to obtain a ratiometric relation between particle emission intensity and its distance to the
glass surface. We attached individual uorescent particles to ne tips of glass micropipettes, which
were translated perpendicularly to the glass substrate in the evanescent eld with a 0.4-nm precision
stage (MadCity Nano-OP25). At each distance to the surface, 30 images of the attached particles
were captured. The pixel values of the imaged particles were then averaged, and subsequently tted
to a two-dimensional Gaussian function to nd their centre intensities. The process was repeated
30
0 50 100 150 200
0
0.2
0.4
0.6
0.8
1
hr, (nm)
N
o
r
m
a
l
i
z
e
d

C
e
n
t
e
r

I
n
t
e
n
s
i
t
y
Figure 3.6: Fluorescent particle intensity as a function of its distance to the glass surface. The
particles used here are 200 nm in diameter. The solid line is a least-square exponential t to the
data whose decay length, , is 181 nm.
several times with dierent particles. The particle intensities are shown in gure 3.6, along with a
exponential t to the measured data:
I
peak
= Ae
(hr)
+B, (3.13)
where A, B and are constants determined by a least-squares t. The tted curve predicts an
intensity decay depth of 181 nm, very close to the evanescent penetration depth that we would
predict based on the imaging lens, wavelength and refractive indices of the system.
3.3.2 Velocimetry Analysis
The analysis of particle image pairs was performed using a custom particle-tracking program written
in a MATLAB environment. Prior to velocimetry analysis, images of particles stuck on the glass
substrate had to be removed from each image pair to prevent undesirable detection by the Particle
Tracking Velocimetry (PTV) algorithm. Although the presence of these stuck particles could distort
the velocity eld, they generally had a very low density and were usually placed at the edge of an
image and far away from where the velocimetry measurements were performed. For each image, the
signals due to stuck particles and background noise were removed by subtracting a reference image
formed from the rolling average of 10 previous and 10 subsequent image pairs. To account for laser
pulse energy uctuation, every image was scaled by the measured laser illumination energy.
Velocimetry analysis started with nding the positions and peak intensities of all particles
via threshold-identication and Gaussian tting of the intensities of the pixels surrounding the
peak. Subsequently each particles displacement was obtained with a limited-range nearest-neighbor
matching, and the displacement was decomposed into streamwise and cross-stream components
(for details of the tracking algorithm, see [80]). The particle seeding density was low enough such
31
0 0.5 1 1.5 2
0
0.5
1
1.5
2
P
D
F
Intensity Ratio (I
e
/I
e
0
)
Figure 3.7: Observed uorescent particle intensity distribution and its probability density function
(PDF). The solid line represents the theoretical prediction (3.6) for an imaging range of 100 to 400
nm and a 15% intensity statistical variation with a uniform concentration distribution of particles.
that this tracking was unambiguous. Intensities of the trackable particles were converted to a
normalized intensity ratio obtained by dividing the raw intensity counts by the average of all the
stuck particles initial intensity counts. Finally, particles within a desired normalized intensity range
were selected for each ow rate and their mean streamwise velocity was dened as the apparent
velocity of the ensemble. For the hindered diusion analysis, the same analysis steps were followed
except that diusion coecients of particles in quiescent uid were obtained from the variance of
displacements instead of velocities.
3.4 RESULTS AND DISCUSSION
3.4.1 Validation of Intensity Calibration Curve
Figure 3.7 shows the distribution of particle intensities recorded during a typical data acquisition
(in this case, in quiescent uid). The accompanying solid line is the theoretical prediction (3.6),
assuming an imaging range of 100 - 400 nm, and an eective emission radius variation of
r
= 0.15.
The good agreement between the observed and predicted intensity variation conrms our calibration
results and the statistical theory of the observed particle intensities discussed earlier.
3.4.2 Hindered Diusion
Near-surface tracer particles are known to exhibit anisotropic hindered diusion due to hydrody-
namic eects, and the hindered diusion coecient in the direction parallel to the substrate surface
32
0 0.5 1 1.5 2
0
0.2
0.4
0.6
0.8
1
(ha)/a
D
e
x
p
/
D
0
Data 1
Data 2
Method of Reflection
Figure 3.8: Ratio of hindered diusion coecients (D
exp
) for Brownian motion parallel to a surface.
D

is [37]
D

D
0
= 1
9
16
_
h
r
_
1
+
1
8
_
h
r
_
3

45
256
_
h
r
_
4

1
16
_
h
r
_
5
+O
_
h
r
_
6
, (3.14)
where D
0
is the Stokes-Einstein diusivity of a spherical particle in the uid bulk. This Method
of Reection approximation is generally valid for (h r)/r > 1. For (h r)/r < 1, Goldman et
al. [38] proposed an asymptotic solution,
D

D
0
=
2
_
ln
_
hr
r
_
0.9543

_
ln
_
hr
r
_
2
4.325 ln
_
hr
r
_
+ 1.591
. (3.15)
The combination of our particle-tracking algorithm and the intensity calibration curve was
applied to analyse near-surface particle Brownian motion parallel to the glass substrate. The
results are compared to the Methods of Reection calculation (3.14) and shown in gure 3.8. The
experimental data shows reasonably good agreement with the theory, including the fall in the
eective diusion coecient close to the surface. The results further improve our condence in the
overall experimental method and subsequent analysis.
3.4.3 Velocity Distributions of Particles
Figure 3.9 shows distributions of particle velocities at three dierent shear rates. Note that the
streamwise velocity distributions widen as shear rate increases, and their asymmetric shapes show
an increasing skewness as the shear rate increases. The increasing distribution width and skewness
at high shear rates are both results of sampling particle displacements from an imaging range with
nite thickness [80]. Physically, at higher shear rates, the nite-thickness imaging range contains
particles that are translated by a wider range of local velocities, hence the increasing width of the
33
200 0 200 400 600 800 1000
0
0.005
0.01
0.015
0.02
Particle Velocity (m/sec)
P
D
F
S = 183 sec
1
S = 732 sec
1
S = 1465 sec
1
Figure 3.9: Distribution of observed particle streamwise velocities under various shear rates. Data
include measurements of ow over a hydrophobic surface.
apparent velocity distributions. In addition, because the particles are conned by the solid wall,
they are more likely to move away from the wall and into a faster moving shear plane. Thus the
distribution exhibits a more pronounced positive skewness at the higher shear rates.
The eect of shear with a nite-thickness imaging range can be characterized if the streamwise
particle velocity in gure 3.9, V , is scaled by

V =
V V )
V )
, (3.16)
where

V is the scaled streamwise particle velocity and V ) is the mean value of each velocity
distribution. The scaled velocity distributions are shown in gure 3.10. At each shear rate, the
relatively contributions of velocities from shear ow and Brownian motion can be characterized
by the Peclet number, Pe. Here the Peclet number is dened as the ratio of the shear-induced
displacement, L, to the diusive displacement, s, during the interframe time period, t,
Pe =
L
s
=
V ) t

2D
0
t
. (3.17)
Because V ) is shear-dependent only, the shear-induced eects in all velocity distributions in
gure 3.10 are scaled to equal magnitudes. It can be observed that small Peclet numbers are
associated with greater distribution widths while the distributions collapse onto a single skewed
distribution at high Peclet numbers. The explanation is the following: at low Peclet numbers the
distributions widths scale inversely to shear, suggesting a diusion dominance and thus the large
widths of the velocity distributions are due to Brownian motion. As the Peclet number increases, the
eect of shear ow on particle motion increases and when Pe > 3, particle motions are completely
shear-dominant, causing the distributions to overlap. It should be noted that the overlapping
34
1.5 1 0.5 0 0.5 1 1.5
0
0.5
1
1.5
Scaled Particle Velocity
P
D
F
Pe = 1.20
Pe = 1.93
Pe = 3.28
Pe = 4.16
Pe = 4.43
Pe = 5.52
Figure 3.10: Distribution of observed particle streamwise velocities after scaling the velocities
distributions in gure 3.9. The velocities are scaled according to (3.17). Pe of 1.20, 3.28 and 4.43
correspond to the shear rates of 183, 732 and 1465 sec
1
in gure 3.9, respectively.
distributions have the same distribution width, asymmetric shape and the same positive tail, which
are absent in small Peclet number cases. These observations conrm the explanation that in gure
3.9, the increasing distribution width and skewness at high shear rates are both eects of shear.
3.4.4 Measurements of Apparent Slip Velocities
For each ow case, the intensity distribution of all trackable particles were compared to the
theoretical prediction (3.6) to identify the unknown parameters such as h
1
, h
2
and
r
through
optimization, as demonstrated in section 3.4.1 and gure 3.7. With these parameters known,
we calculated the theoretical apparent velocities of a particle ensemble, chosen from a normalized
intensity range, by using equation (3.12) and varying the slip length conditions. Subsequently, these
velocities were compared to the experimental apparent velocities (mean values of the streamwise
velocity distribution) of particles from the same normalized intensity range to extract experimental
slip velocities and slip lengths.
Figures 3.11 and 3.12 show the experimental and theoretical apparent velocities as a function
of wall shear rate over hydrophilic and hydrophobic surfaces, respectively. In both gures, the
solid lines represent the theoretical apparent velocities assuming zero slip (equation (3.11)). The
dashed and dotted lines are the expected apparent velocities if one were to assume a 50-nm and
a 100-nm slip length, respectively (calculated using (3.12)). In the case of the hydrophilic surface
(gure 3.11), the measured slip length ranges from 26 to 57 nm with a slightly increasing trend as
wall shear rate increases. The data obtained for ow over the hydrophobic surface (gure 3.12),
on the other hand, suggests a larger slip length ranging from 37 to 96 nm, with a more observable
increasing trend of wall shear rate dependence at higher shear rates. A quantitative comparison of
the two cases shows a slip length attributed to surface hydrophobicity ranging from -7 nm at low
35
0 500 1000 1500 2000
0
100
200
300
400
500
600
700
Wall Shear Rate (1/sec)
A
p
p
a
r
e
n
t

V
e
l
o
c
i
t
y

(

m
/
s
e
c
)
No Slip (Theoretical)
Slip length = 50 nm
Slip length = 100 nm
Exp. data
Figure 3.11: Experimental apparent velocity of particles in a shear ow over a hydrophilic surface
(0.21 < I
e
/I
e
0
< 0.78), and its comparison to slopes of expected apparent velocities if there exists
a 0, 50 or 100-nm slip length. The error bar of each experimental data point represents a 95%
condence interval. In all cases the 95% condence intervals are within 3% of the reported mean
velocity values.
shear rates to 54 nm at the highest tested shear rate, with an average value of 16 nm in between
(gure 3.13). The additional slip caused by surface hydrophobicity is in agreement with many
experimental results [25, 26, 28, 31], but is in sharp disagreement with others [30, 33, 64].
There are three possible sources of measurement uncertainty in gure 3.11 and 3.12: determina-
tion of (a) the wall shear rate, (b) the apparent velocities of particle ensembles and (c) the imaging
range. Because wall shear rates are calculated based on microchannel dimensions and ow rates
which are controlled by the syringe pump, we estimate that the uncertainty in the wall shear rates
is less than 3%. If there is signicant slip at the wall, the shear rates could be slightly lowered than
the reported values. However, the largest observed slip length is less than 1% of the channel height,
and thus the slip eect on wall shear rate calculation can be safely neglected. The determination of
apparent velocities of particle ensembles also has extremely low uncertainly because central limit
theorem constrains the statistical mean of a large sample (a few thousand trackings) to a very low
uncertainty (< 3% of the reported velocity values for 95% condence intervals). In addition, with
subpixel resolution of the particle centre detection, our PTV algorithm is able to track particle
displacements with high accuracy (< 0.1 pixel, [84]). The largest source of uncertainty is the deter-
mination of the imaging range. The imaging range is determined statistically based on the particle
intensity decay length obtained from gure 3.6 where intensity measurements were performed in
increments of 20 nm, which could serve as the upper limit of imaging range uncertainty. However,
it should be noted that, with an uncertainty of 20 nm taken into consideration, the largest mea-
sured slip length still will not exceed 150 nm much smaller than other experimental studies which
reported slip lengths greater than a few hundred nanometers.
One should note that the corrections due to shear-induced rotation (gures 3.11 and 3.12) do
36
0 500 1000 1500 2000
0
100
200
300
400
500
600
Wall Shear Rate (1/sec)
A
p
p
a
r
e
n
t

V
e
l
o
c
i
t
y

(

m
/
s
e
c
)
No Slip (Theoretical)
Slip length = 50 nm
Slip length = 100 nm
Exp. data
Figure 3.12: Experimental apparent velocity of particles in a shear ow over a hydrophobic surface
(0.5 < I
e
/I
e
0
< 1.0), and its comparison to slopes of expected apparent velocities if there exists
a 0, 50 or 100-nm slip length. The error bar of each experimental data point represents a 95%
condence interval. In all cases the 95% condence intervals are within 3% of the reported mean
velocity values.
0 500 1000 1500 2000
20
0
20
40
60
Wall Shear Rate (1/sec)
A
d
d
i
t
i
o
n
a
l

S
l
i
p

L
e
n
g
t
h

(
n
m
)
Figure 3.13: The additional slip length due to surface hydrophobicity.
37
not account for the possibility of slip between the uid and the tracer particle, and we are unaware
of any work that extends the theory of Goldman et al. [38] to include the eect of surface slip.
Physically, surface slip is unlikely to be an issue, because the tracer particles have relatively rough
surfaces. Nevertheless, we can qualitatively estimate its eect. Slip between the uid and the
tracer particle will modify the particle velocity in two ways. Firstly, the reduced shear force will
result in the particle velocity lagging the local uid velocity. Secondly, however, slip will reduce
the shear-induced rotation of the particle. Since these two mechanisms act in opposition to each
other, we believe that, as a rst approximation, we can ignore this eect.
Granick et al. [34] and Lauga et al. [7] both pointed out that large surface roughness can
signicantly decrease slip. The slip length we measured with sub-nanometer surface roughness,
however, agrees well with the surface roughness dependency reported by Zhu & Granick [26]. It
should also be noted that Joseph & Tabeling [31], with ow experiments conducted under similar
surface roughness, hydrophobic coating and shear rates, reported slip length very close to the
present results. It has also been suggested that electrokinetic eects could signicantly aect
velocimetry measurements of suspended particles close to a surface. Lauga [35] reported that a
streaming potential can induce electrophoretic motion of particles if particle surfaces are charged,
and applying this theory to the current conditions results in an apparent slip length of at most
18 nm.
Electrostatic interaction between tracer particles and surface could also aect accuracy of ve-
locimetry measurements [7]. If the tracer particles and the substrate surface are similarly charged,
electrostatic repulsion can create a depletion layer near the substrate surface that eectively shifts
the imaging range. Because the slip length calculations in gure 3.11 and 3.12 are based on the
assumption that particle concentration is uniform near surface, a depletion layer could be a source
of apparent slip if it is not recognized. Lumma et al. [33] suggested that the depletion layer
thickness can be as large as 0.9 m for a pure water and glass combination. Such a thickness seems
unreasonably large as it is much bigger than the evanescent eld penetration depth and conse-
quently few uorescent particles would be observed. In contrast, Flicker et al. [51] experimentally
measured the depletion layer thickness to be in the order of 10 to 30 nm. For our experimental
setup, calculations suggest that, based on the overall particle seeding density, if the particles are
uniformly distributed in the uid one should be able to see around 19 particles in a 26 m-by-26
m image eld. This is in close agreement with our experimentally-derived images, and we can
conclude that the depletion layer thickness must be much smaller than our imaging range and
thus eects due to a depletion layer thickness of this magnitude would not be detectable with our
current imaging technique. However, recognizing the possibility of such depletion layers existence,
we re-analysed the theoretical curves in gure 3.11 and 3.12 by assuming an imaging range farther
away from the substrate surface. We found such a shift would result in higher theoretical apparent
velocities of particle ensembles and therefore lower slip lengths for all shear rates, suggesting that
our reported slip lengths are most likely the upper bound of the true slip length value.
The electrostatic and electrokinetic interactions are an interesting pair as their signicance de-
creases with increasing ionic strength of the test uid. Perhaps a more accurate direct measurement
of the true surface slip can be obtained if aqueous solutions of moderately high ionic strengths were
used.
3.5 CONCLUDING REMARKS
The results presented here represent direct measurements of local velocities in the near wall region
of a uid-solid boundary. Despite the relative simplicity of the experiment, several subtle issues
38
need to be carefully addressed, including the spatial non-uniformity of the illumination region, the
eects of the local shear on the dynamics of seeding particles, and the statistical variations in the
eective visibility of the seeding particles. The current experiment conrms previous measurements
in our own group [25] as well as some other recent measurements [26, 31] that there is minimal slip
over hydrophilic surfaces, and that hydrophobic surfaces do appear to introduce a discernable, but
small boundary slip.
Although the present technique does allow for measurement of particle motions very close to the
solid surface, it does suer from some limitations and there is room for improvement. Most notably,
the tracking in the wall-normal direction is currently limited to an accuracy of approximately 10%
of the evanescent penetration depth (about 20 nm in the current case). If particles with a higher
degree of monodispersity can be reliably found, and if the imaging system noise can be reduced
(both of which are technical improvements on the immediate horizon), we believe that the technique
can be used to track particles with an order of magnitude improvement in accuracy. However, even
with the current levels of uncertainty, there is no possibility that the ow exhibits a slip length
more than 150 nm (gure 3.12) over the range of shear rates tested.
While the debate of slip versus no-slip will likely continue in the foreseeable future, we believe
that a consensus is forming from more recent experiments that the slip eect is not as extreme
as some studies might have suggested. As many authors have pointed out, measurement and
interpretation of slip data in the nanoscale can be very tricky because other physical phenomena,
such as electrokinetics and nanobubbles, might lead to observations of an apparent slip caused by
other complicating factors. All of these suggest that close attention must be to paid to documenting
the experimental conditions, and that this might well reveal the reasons behind the range of reported
apparent slip phenomena.
Chapter 4
Paper 3: Direct measurement of slip
length in electrolyte solutions
The emergence of microuidics over the past decade has brought out increased research interests
in liquid boundary slip, which for centuries has been regarded as negligible. Advancements in
ow measurement techniques have resulted in many recent experiments reporting aqueous slip
length with a wide range of values and various degrees of dependency on applied shear rate, surface
hydrophobicity and surface smoothness [20, 21, 25, 26, 28, 30, 31, 32, 33]. A total internal reection
velocimetry (TIRV) technique was recently reported to be an eective method in measuring liquid
velocities in the region of less than 300 nm from solid surfaces [80]. By conducting a three-
dimensional TIRV (3D-TIRV) experiment, Huang et al. reported that slip length can be obtained
within a 20-nm uncertainty [85]. They reported a small but non-negligible slip length for de-ionized
aqueous solutions in shear ows over both hydrophilic and hydrophobic surfaces. However, some
questions remained as to whether electrostatic and/or electrokinetic forces acting on the tracer
particles could lead to false slip observation [7, 35]. In this letter, slip length measurement of ionic
aqueous solutions using TIRV is presented and compared to the previous results of Huang et al. as
an attempt to answer these open questions.
The TIRV technique uses total internal reection of an incident laser beam to generate a highly
localized illumination of the near-boundary liquid phase, and relies on tracking motions of individual
tracer particles to determine uid velocity vectors in the planes parallel to a solid surface (as
demonstrated in gure 4.1). The exponentially decaying evanescent eld leads to determination
of tracer particles positions in the direction normal to the solid surface based on their uorescent
intensities. Subsequently slip velocities and slip length can be inferred from the measured apparent
velocity vectors by applying a statistical model for optical and hydrodynamic behaviors of small
particles near a solid/liquid interface. A detailed description of the 3D-TIRV theoretical work,
including a proof of concept and a discussion of its subtleties, was presented by Huang et al. [85].
A brief summary is presented in this letter to facilitate the readers understanding and resultant
discussion.
When a uorescent particle of radius a is illuminated by an evanescent eld of incident wave-
length
0
(a <
0
), its emission intensity, I
e
, exhibits the same exponentially decaying characteristic
as the evanescent eld energy does, namely,
I
e
(z) = I
e
0
e
h/p
, (4.1)
where h is the distance between the particle center and the solid phase, I
e
0
is the emission intensity
of the particle when it is in contact with the solid phase (h = a), and p is the evanescent wave
39
40
Figure 4.1: Objective-based total internal reection velocimetry (TIRV) system.
penetration depth. The penetration depth, which characterizes the length scale of the evanescent
eld, can be calculated from
p =

0
4
_
n
2
1
sin
2
n
2
2
_

1
2
, (4.2)
where n
1
and n
2
are the indices of refraction of the solid boundary and liquid, respectively. However,
a variety of factors, including statistical variations in illumination intensity, quantum eciency of
the imaging system and, most importantly, non-uniformity of the particle sizes, can lead to varia-
tions in the observed uorescent intensities of an ensemble of nominally identical particles. Thus,
a statistical analysis is carried out by characterizing these variations through a single parameter
r, or eective emission radius, with which a joint probability density function (PDF), P (h, I
e
), of
the particle distance to the solid surface and emitted uorescent intensity can be obtained.
If near-wall velocities of a pure uid are to be measured by tracking motions of tracer particles,
it is important to recognize that shear and near-surface hydrodynamic eects can cause a tracer
particle to rotate and translate at a velocity, U, lower than the velocity of the local uid element in
the same shear plane [12, 38, 60]. Factors that determine the particle translation velocity include
particle radius, a, particle distance to the solid surface, h, and the local uid shear rate, S (h).
Therefore corrections need to be made when one convert uid velocities into particle velocities. It
should be noted that in the presented analysis, the corrected particle translation velocity does not
account for the possibility of slip between the uid and the tracer particle, and thus any calculation
based on U represents a no-slip case.
Because tracer particles are used to probe near-surface shear ow, assurance is needed that the
particle ensemble uniformly samples all shear layers. It has been suggested that shear-induced lift
can be a source of particle migration away from the solid wall [86]. However, based on the theory
presented by Cherukat & McLaughlin [87], the shear-induced lift under the presented experimental
conditions is insignicant, and thus alleviating this concern (see Appendix B).
The ensemble-averaged velocity,

U, of a large number of uniformly distributed particles emitting
in an intensity range of < I
e
< and located in an imaging range of h
1
< h < h
2
, can be
calculated by

U =
1
h
2
h
1
_
h
2
h
1
U (h, a, S (h)) P(h, < I
e
< ) dh. (4.3)
Because equation (4.3) is based on a no-slip boundary condition, if there exists a slip velocity, U
slip
,
41
at the liquid/solid substrate interface, the same particle ensemble will exhibit an apparent velocity
of

U
app
= U
slip
+

U =
w
+

U, (4.4)
where U
slip
is characterized as the product of the slip length, , and the wall shear rate,
w
. It
should be noted that

U
app
can be experimentally measured by nding an ensemble-averaged velocity
of trackable tracer particles while
w
is computed from ow rate and channel geometry. With

U
calculated by using equation (4.3), one slip length can be obtained for each wall shear rate.
Monodispersed Fluorescent Polystyrene Microspheres (Duke Scientic) with diameters of 100
(10%) and 200 nm (5%) suspended in puried water (Fluka) at 0.005-0.04% volume fraction
were used as tracer particles in this experiment. These microspheres have a peak absorption
wavelength of 542 nm and emit at 612 nm. Prior to each experiment, the uid was degassed
by placing the solution in vacuum for at least 30 minutes. This procedure has been reported to
signicantly eliminate nanobubble formation [83], which could be an alternative source of apparent
slip [7, 34, 82].
Test channels were fabricated using a polydimethylsiloxane (PDMS) (Dow Corning Sylgard 184)
molding technique [69] and bonded onto polished glass wafers. The dimensions of the test channels
were 50 1 m deep, 250 1 m wide and 15 mm long. Hydrophilic and hydrophobic microchan-
nels were created by oxygen plasma treatment of the PDMS mold and by bonding untreated PDMS
channels onto octadecyltrichorosilane(OTS)-coated glass wafers, respectively. The hydrophilic and
hydrophobic glass surfaces had an RMS surface roughness of 0.47 nm and 0.35 nm respectively.
A continuous-wave 514-nm Argon-ion laser beam (Coherent) was directed through a Nikon PL
Apo NA 1.45 100X TIRF oil immersion objective at an angle that created total internal reection at
a glass-water interface, thus illuminating the near-surface region of water with an evanescent eld
(gure 4.1). Fluorescent images of near-surface particles were captured by a Q-Imaging Intensied
Retiga CCD camera (ICCD). Approximately 1000 images were captured at each ow rate. The
ows were driven by a Harvard Apparatus 22 syringe pump, with special attention paid to removing
pulsations and maintaining steady pumping motions. Detailed description and schematics of the
experimental setup have been presented in [85].
The analysis of particle image pairs was performed using a custom particle-tracking algorithm.
In summary, velocimetry analysis started with nding the positions and peak intensities of all
particles via threshold-identication and Gaussian tting of the intensities of the pixels surrounding
the peak. Subsequently each particles displacement was obtained with a limited-range nearest-
neighbor matching. Finally, particles within a desired intensity range were selected for each ow
rate and their mean streamwise velocity was dened as the apparent velocity of the ensemble and
compared to theoretical no-slip values. Slip lengths were then obtained based on equation (4.4).
Figure 4.2 shows the measured slip lengths of aqueous solutions under various experimental
parameters. The rst observation to be made is that all measurements report slip lengths of
approximately 100 nm or less. These results further conrm previously reported measurements
[25, 26, 31, 85] that there is minimal slip over smooth solid surfaces. Slip on a hydrophobic surface
appears to be slightly higher than on a hydrophilic surface. A quantitative comparison of all
measurements shows that the slip length attributed to surface hydrophobicity averages to 22 nm at
the tested shear rates. The additional slip caused by surface hydrophobicity is in agreement with
many experimental results [25, 26, 28, 31], but is in sharp disagreement with some others [30, 33].
Electrokinetic eects [35] and electrostatic repulsion between tracer particles and glass surface
[7] have both been proposed as sources that might cause an apparent slip and could aect the
accuracy of slip velocity measurements. Their signicance, however, is expected to decrease with
increasing ionic strength of test uids. Lauga [35] suggested that for an aqueous solution with 1
42
0 500 1000 1500 2000
0
50
100
150
Wall Shear Rate (1/sec)
A
p
p
a
r
e
n
t

S
l
i
p

L
e
n
g
t
h

(
n
m
)
200nm, DIwater, Hydrophilic
200nm, DIwater, Hydrophobic
200nm, 0.1mM NaCl, Hydrophobic
200nm, 1mM NaCl, Hydrophobic
100nm, DIwater, Hydrophobic
Figure 4.2: Measured slip lengths of aqueous solutions. The error bar on each data point repre-
sents a 95% condence interval of each ensemble calculation. The uncertainty of each slip length
measurement (not drawn for plot clarity) is less than 20 nm, estimated based on penetration depth
calibration and particle radius variation [85]. Data of 200-nm, DI-water, Hydrophilic and 200-
nm, DI-water, Hydrophobic have been previously reported by Huang et al.
mM NaCl concentration, the apparent slip length would become sub-molecular. As shown in gure
4.2, variations in aqueous ionic concentration do not change measured slip lengths signicantly.
This observation suggests that the measured slip lengths are most likely not due to ionic eects,
but are consequences of true boundary slip.
Another important observation is that the slip lengths do not vary signicantly over the tested
range of shear rates. This observation agrees with experimental studies conducted under similar
ranges of moderate shear rates [28, 29, 31]. Still, this result does not suggest that slip length is
independent of shear rates, as the conducted measurements span only one order of magnitude of
shear rates. Zhu & Granick [26], Neto et al. [27] and Choi et al. [25] all tested over larger ranges
of shear rates (> 2 orders of magnitude) and reported shear-dependent slip lengths. It is physically
plausible that the shear dependence of slip length is nonlinear and the shear eect is signicant
only at higher shear rates. To assess this, a TIRV system equipped with a very sensitive high-speed
camera would be needed for measurements at high shear rates.
Chapter 5
Paper 4: Direct measurement of
anisotropic near-wall hindered
diusion using total internal reection
velocimetry
By applying the three-dimensional total internal reection velocimetry (3D-TIRV) technique to
freely suspended micron-sized uorescent particles, we aim to simultaneously observe the three-
dimensional anisotropic hindered diusion for gap-size-to-radius ratio much less than one. We
demonstrate that the three-dimensional tracking technique of TIRV used for nanoparticles can be
adopted into 3D displacement measurements of freely suspended 1.5-m radius particles. The dis-
placement measurements reveal that the hindered diusion coecients are in close agreement with
the theoretical values predicted by the asymptotic solutions of Brenner [36] and Goldman et al. [37]
for gap-size-to-radius ratio much less than one, and hindered diusion anisotropicity is simultane-
ously observed in all data sets.
5.1 INTRODUCTION
Advancement in colloidal sciences have lead to many applications of micro- and nano-particles in
engineering research (eg. particle imaging velocimetry) and biotechnology (eg. drug delivery). At
these length scales Brownian motions are signicant and understanding these thermal agitations
are essential in making practical use of these tiny particles. Inside a uid bulk the diusivity of
an isolated particle follows the Stokes-Einstein relation, which is the ratio of uid thermal energy
to the particles hydrodynamic mobility. In the presence of a nearby solid boundary, however, a
particle would experience an increased, anisotropic drag which hinders its mobility. Brenner [36]
and Goldman et al. [37] were the rst to developed drag force correction factors for near-wall
spheres under no inertial eects and a no-slip boundary condition. These correction factors have
since evolved into theories of hindered diusion [88].
Following these works, many experimental studies have been conducted to observe hindered dif-
fusion and to verify the correction factors. Demonstrated experimental techniques include evanes-
cent light-scattering [39, 40, 42], total internal reection uorescence microscopy (TIRFM) [43, 44]
and combined optical tweezers and digital video microscopy [41]. In most of the above mentioned
studies, simultaneous three-dimensional measurements of the anisotropic diusivity could not be
43
44
conducted due to experimental limitations. One exception is a ratiometric-TIRFM measurement
conducted by Banerjee & Kihm [43], who reported anisotropic diusivities that only partially agree
with the theories, probably due an inaccurate accounting of particle positions. Another exception
is a study reported by Lin et al. [41], who demonstrated that the anisotropic hindered diusion
could extend far away from the wall.
In the theories of Brenner and Goldman et al. the hindered mobility of a near-wall particle is
a function of the particles radius, a, and the gap size between the particle and the wall, h. They
proposed analytical solutions of the hindered mobility that are vastly dierent for h/a 1 and
for h/a 1. Based on our literature survey, no study has demonstrated the anisotropic hindered
diusion for h/a 1 with a simultaneous three-dimensional measurement as studies of Banerjee
& Kihm and Lin et al. both attempted to measure hindered diusion at h/a 1. In this paper we
present our experimental study aimed to validate the anisotropic mobility correction coecients for
h/a 1 through a direct measurement of hindered diusion using three-dimensional total internal
reection velocimetry (3D-TIRV).
5.2 THEORY OF HINDERED DIFFUSION
When an isolated particle is at the vicinity of a solid boundary, its Brownian motion is hindered
due to an increase in hydrodynamic drag. The presence of the solid wall also breaks the symmetry
of particle dynamics, resulting in anisotropic Brownian motion. Brenner [36] successfully solved
the lubrication equation of particle motion normal to the solid wall and proposed an innite series
solution of the drag correction factor,

. He reported that the correction factor is a function of


the particle radius, a, and the gap size between the particle and the solid wall, h. Later Bevan &
Prieve [39], using a regression method, reported that

D
0

6h
2
+ 2ah
6h
2
+ 9ah + 2a
2
(5.1)
is a close approximation to Brenners innite series solution. For the diusivity correction factor in
the direction parallel to the solid wall,

, an exact analytical solution was never found. Instead,


Goldman et al. [37] oered an asymptotic solution,

D
0
=
2
_
ln
_
h
a
_
0.9543

_
ln
_
h
a
_
2
4.325 ln
_
h
a
_
+ 1.591
, (5.2)
for particles very close to the wall, dened by h/a 1.
5.3 EXPERIMENTAL PROCEDURES
To experimentally verify equations (5.1) and (5.2), diusion of micron-sized uorescent particles in
an evanescent eld was tracked with digital imaging microscopy. The technique, termed 3D-TIRV,
have been demonstrated by Huang et al. in measuring motions of near-wall nanoparticles [89].
Unlike evanescent light-scattering experiment of non-uorescent particles, TIRV allows particle
centers to be identied on the captured uorescent images, and consequently particle motions
can be tracked via the displacements of particle centers. In addition, uorescent intensity of the
particles can be used to infer the instantaneous separation distance between the particle and the
two-media interface, as commonly done in traditional TIRFM experiments [44, 90].
45

z
x
water (n = 1.33)
glass (n = 1.51)
~ penetration depth
Figure 5.1: Schematic of TIRV. A red uorescent microparticle suspended in water is placed in
a green evanescent eld created within the focal depth of a high numerical aperture microscopy
objective. A collimated laser beam through the objective is used as the illumination light source.
At the illumination beam incident angle, , greater than 61.9

, total internal reection occurs at the


glass/water interface. Because the particle radius is much larger than the evanescent penetration
depth, the evanescent energy illuminates only the lower portion of the encapsulated uorophores.
Our 3D-TIRV setup was created on an inverted epi-uorescent microscope (Nikon TE-2000). A
continuous-wave, collimated 514-nm Argon-ion laser beam was directed through a Nikon PL Apo
NA 1.45 100X TIRF oil immersion objective at an angle that created total internal reection at
a glass/water interface. The water phase was an aqueous solution consisted of 10mM NaCl and
individual 1.5m-radius (5%) uorescent polystyrene particles (refractive index 1.59, Spherotec
Inc.) suspended at 0.02% volume fraction. The low volume fraction, and thus low particle number
density, ensured the isolated-particle assumption of the theories was met. Before each measurement
the solution was sonicated in intervals of 30 seconds for 10 minutes to disperse coagulated particles.
For image acquisition, 200 L of the solution was injected into a closed reservoir formed by a piece
of carved polydimethylsiloxane (PDMS) sandwiched between glass coverslips. Subsequently TIRV
imaging of hindered particle Brownian motions was performed at one of the glass/solution interface.
The uorescent images of the particles were projected onto a Q-Imaging Intensied Retiga CCD
camera (ICCD), and recorded at a 10-Hz frequency. Overall, 2000 to 2500 sample images were
captured for each experiment for reliable statistics.
A particles 3D position was determined through threshold-identication and Gaussian tting of
the intensities of the pixels surrounding the peak. This method has been previously demonstrated
in TIRV with nanoparticles. Still, a calibration experiment was necessary to verify its validity for
1.5-m particles, especially in determining the relationship between the particle/wall gap size and
the particles peak uorescent intensity. To do so, individual 1.5-m radius uorescent polystyrene
particles were attached to polished conical tips of thin graphite rods. By mounting the assembly to
a 0.4-nm precision translation stage, the uorescent particles were traversed through the evanescent
eld and their uorescent images were digitally captured at various distances to the glass surface.
The result of this calibration is shown in gure 5.2. The close agreement between particle intensitys
exponential decay length and the evanescent eld penetration depth suggests that particle-glass gap
size can be reliably inferred from the particle peak intensity. We also found that the precision of
center identication was approximately 0.5 pixels, or 32 nm.
46
0 50 100 150 200 250 300
0
0.2
0.4
0.6
0.8
1
Gap size, h (nm)
I
/
I
0
Particle 1
Particle 2
Particle 3
Exp. fit (d = 142 nm)
Figure 5.2: Gaussian-tted peak intensities, I, of 1.5-m radius uorescent particles in evanescent
eld, taken at 0 to 275 nm away from the glass surface. I
0
is the tted intensity at h = 0. Each data
point represents the particle peak intensity averaged over 10 images, while the length of the error
bar spans two standard deviations. The laser incident angle is 64

and the resulting evanescent


eld penetration depth is 148 nm. All data series exhibit an exponential decay tendency versus
gap size. With an exponential decay length of 142 nm, the best t of the intensity data shows close
agreement to the penetration depth.
Once positions of all imaged particles were determined, particle displacements were calculated
between consecutive frames. The hindered diusion coecient in the direction parallel to glass
surface, D

, was calculated from the equation


R)
2
= D

t (5.3)
where R is the averaged radial displacement of particles, t is the time between consecutive
image acquisitions (= 100 millisecond), and ) represents ensemble average. Similarly, the hindered
diusion coecient in the direction normal to glass surface, D

, was calculated from


(z)
2
) z)
2
= 2D

t (5.4)
where z is the particle displacement normal to the glass surface. Finally, to obtain the hindered
diusion correction factors,

and

, diusion coecient of the same 1.5-m particles were


measured in the uid bulk, using regular uorescent digital microscopy.
5.4 RESULTS AND DISCUSSION
Shown in gure 5.3 are hindered diusion correction factors measured by 3D-TIRV. In addition to
experimental results, hindered diusion coecients from a Brownian dynamics simulation [89] are
also plotted for comparison. These results are compared to the asymptotic solution of Goldman et
47
0 0.05 0.1 0.15 0.2
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
h/a
D
/
D
0
GCB
MOR
Bevan
D
X
(Sim.)
D
Z
(Sim.)
D
X
(Exp.#1)
D
Z
(Exp.#1)
D
X
(Exp.#2)
D
Z
(Exp.#2)
Figure 5.3: Hindered diusion correction factor (D
x
/D
0
, D
z
/D
0
) vs. non-dimensional gap size
between particles and glass surface (h/a). D
0
is the diusion coecient measured in the uid
bulk while D
x
= D

and D
z
= D

. Each data point is obtained from an ensemble of particles


found within 0.005 or 0.01 of the targeted gap size values. GCB, MOR and Bevan repre-
sent asymptotic solution of Goldman et al., the Bevan approximation and Method of Reection
solution, respectively. Exp. represents experimental data while Sim. means data obtained
from Brownian dynamics simulation. Each error bar represents the 95% condence interval of
measurement.
al. (equation (5.2)), Bevans approximation (equation (5.1)) and a Method of Reection (MOR)
solution [37]. Theoretically equation (5.2) is more appropriate for the measured gap sizes of h/a 1
while the method of reection solution is more accurate at h/a > 1, but nevertheless they are both
plotted for a quantitative comparison. The correction factor in the direction parallel to glass surface,

D
x
/D
0
, increases from approximately 0.2 at h/a 0 to 0.4 at h/a 1.3, while the correction
factor in the direction normal to glass surface,

= D
z
/D
0
, increases from 0 to 0.1 in the same
h/a range, demonstrating the anisotropicity of near-wall hindered diusion. It is observed that the
measured

values at h/a < 0.05 are very close to that predicted by equation (5.2) but signicantly
deviate from the MOR solution, conrming the validity of the asymptotic solution at h/a 1.
The transition from the asymptotic solution to the method of reection solution is demonstrated
at h/a > 0.1 as the measured

values fall between the two solutions. The measured

agrees
well with Bevans approximation, except at 0.05 < h/a < 0.1 where measured

is slight larger
than the predicted values possibly due to insucient or slightly biased particle sampling. A close
agreement between the measured and the predicted

values at extremely small h/a has also been


reported by Oetama & Walz [42], who conducted one-dimensional hindered diusion measurement
in the range h/a < 0.025. Finally, the experimental data also agree well with the simulation results,
conrming the validity of our measurement technique.
As in most near-wall colloidal measurements, we realize the importance of recognizing other
acting physical forces that could lead to measurement inaccuracy. One of these forces is gravitational
48
pull or sedimentation of the particles. At a rst glance, it is perceivable that sedimentation could
lead to bias in diusion measurement normal to the glass boundary. In the uid bulk of density
f
and dynamic viscosity , the terminal settling velocity, w
s
, of a particle with radius a and density

p
is
w
s
=
2 (
p

f
) ga
2
9
, (5.5)
where g is the gravitational acceleration. For a 1.5-m radius polystyrene particle of density 1050
kg/m
3
, this terminal settling velocity would be 245 nm/sec, equivalent to a settling displacement
of 24.5 nm during the imaging inter-frame time of 0.1 seconds. However, when the particle ap-
proaches the solid boundary, its terminal velocity is expected to decrease with a correction factor
identical to

for the same hydrodynamic reason of decreased mobility. Based on equation (5.1),
within the evanescent penetration depth the particles would settle for a distance of less 2.45 nm
in the inter-frame time of 0.1 seconds, or equivalently a non-dimensional gap size of 0.0016 for
our measurement gap size of h/a < 0.14. Therefore in the time scale of our Brownian motion
measurements, sedimentation can be considered as a negligible factor.
Besides hydrodynamic interaction, electrostatic repulsion can also exist between the polystyrene
particles and the glass surface as they both carry negative surface charges when immersed in
aqueous solution [51]. Indeed when the same uorescent particles were suspended in de-ionized
water and observed under evanescent wave imaging, we found few particles came close to glass
surface. However the length scale of electrostatic repulsion can be easily decreased by increasing
the electrolyte concentration of the suspending aqueous solution. By adding 10 mM of NaCl to the
solution, we decreased the electric double layer thickness surrounding the particles and the glass
surface to a Debye length 3 nm, or h/a 0.002, and thus signicantly reduced the eect of
electrostatic repulsion.
Another colloidal force that could potentially aect measurement accuracy is van der Waals
force between the particles and the glass. However, for a particle/glass gap size of 10 nm and
greater, van der Waals force scales with h
3
[91], making its eective range even shorter than that
of electrostatic repulsion. Thus it is safe to assume that van der Waals force is also insignicant
for the range of h/a under consideration.
The hydrodynamic mobility correction factors proposed by Brenner and Goldman et al. assume
a no-slip boundary condition at particle surfaces. If uid slip exists at the particle surfaces, it
is expected that the theories of Brenner and Goldman et al. would under-estimate near-surface
particle mobility and the hindered diusion coecients. Several recent theoretical and experimental
studies, however, oer insights into whether slip eect should be taken into consideration. First,
by conducting slip measurements of water owing over hydrophobic surfaces, Zhu & Granick [26]
concluded that apparent slip lengths are only a few nanometers if the rms surface roughness is
3.5 nm or larger. Because the polystyrene microspheres used in the current hindered diusion
experiment are hydrophilic and have surface roughness of at least 10 nm [92], the eective slip
length at particle surfaces, , is expected to be of only a few nanometers in magnitude, if not sub-
nanometer. Thus under the current experimental conditions, h/ 10. Consequently, based on
the theory of particle mobility presented by Vinogradova [93], slip of such magnitude would have an
insignicant eect on the particle mobility for the particle diameter and particle-wall gap size under
consideration. Therefore the no-slip assumption is considered valid in the current experiment, and
is additionally veried by the agreement between experimental data and theoretical values shown
in gure 5.3.
49
5.5 SUMMARY
A direct measurement of hindered diusion was conducted with total internal reection velocimetry.
We believe that this is the rst simultaneous three-dimensional measurement of hindered diusion
coecients for h/a 1. The anisotropic characteristic of hindered diusion is experimentally
conrmed and the correction factors are found to agree with the theories of Brenner and Goldman
et al. The results conrm the increase of hydrodynamic drag when a particle approaches a solid
boundary, and such correction shall be applied to not only diusion but also other translational
motion of particles where the drag force is of concern.
Chapter 6
Paper 5: Simulations of hindered
diusion in shear ow and its
implications for near-wall velocimetry
The ensemble behaviors of near-surface particles in a shear ow, and the eects of the wall presence
and hindered diusion on the measured apparent velocities are explored through Brownian dynamics
simulations with diusive time and particle radius as time and length scales, respectively. It is
observed that the shear force dominates over diusion at Pe > 3, and asymmetric shape and width
of apparent velocity distributions are attributed to the shear eect. It is also found that if the
time interval between successive particle image acquisitions is short, an under-estimation of uid
velocity can result from biased sampling of tracer particles very close to the wall. However, if the
time interval between successive particle image acquisitions is long, a signicant portion of the
tracer particles would sample planes of higher velocities, resulting in an over-estimation of overall
uid velocity. Particle drop-ins and drop-outs, are found to be a potential source of measurement
error for nano-PIV because of the rapid Brownian motion of nanoparticles. The combined eect
of the observations described above is that near-wall particle-based velocimetry measurement is not
time-invariant. Still, the results found in the Brownian dynamics simulation oer a way where the
true uidic velocities can be inferred from particle apparent velocities through proper scaling. This
scaling method, however, is available only for TIRV but not for nano-PIV, making TIRV the more
superior method in near-surface measurements.
6.1 INTRODUCTION
Colloidal particle-based image velocimetry has long been used as an experimental method in mea-
suring uid velocity prole and other physical quantities. Its accuracy heavily relies on the assump-
tion that tracer particles will accurately trace local uid velocities. In the past decade, particle
image velocimetry (PIV) has been extended to the micro- [55] and nano-scale [56] where size of
tracer particles has also decreased from microns to nanometers. At these physical length scales,
Brownian motion of tracer particles can be quite signicant, thus violating the assumption of par-
ticles travelling at velocities equal to that of the surrounding uids. Consequently, conducting an
accurate velocimetry experiment requires a good understanding and a careful treatment of Brow-
nian motion via either a thorough error analysis [55], an ensemble correlation averaging technique
[54], a statistical analysis of the particle ensemble [80, 85] or a method of statistical tracking [94].
50
51
In the vicinity of a solid wall, a colloidal particle will undergo a hindered Brownian motion
due to an lubricational increase of its hydrodynamic drag. This is known as hindered diusion and
has been reported both theoretically [12, 36, 37, 38] and experimentally [39, 41, 42, 43, 95]. In most
cases, such a spatially dependent Brownian motion would lead to greater diculties in accessing
measurement uncertainties. Only a few researchers managed to turn the disadvantages into merits
by proposing usage of the extent of hindering to reconstruct near-wall shear ow velocity prole
[96] and to probe the no-slip boundary condition [97].
To study the ensemble behavior of hindered and non-hindered Brownian motions, there are
two equally appropriate approaches [58]. One is the Fokker-Planck approach, which solves the
momentum-space partial dierential equations to obtain a time evolution of spatial conguration
function of the particle ensemble. Recently, Sadr et al. [59] had taken the Fokker-Planck approach
to study the hindered diusion of particles in the direction normal to the wall and its eect on
the bias of velocimetry. The other approach is a Brownian dynamics (BD) simulation, which is
to simulate motion of each particle in a large ensemble through a stochastic dierential equation.
In this approach, both deterministic (such as uid ow and sedimentation) and stochastic (such
as Brownian motion) processes contribute to particle displacements. The grand simulation then
yields a spatial distribution of the particle ensemble after each time step. Details provided by this
method are not only intuitive in understanding ensemble behavior but also bear a close resemblance
to physical velocimetry experiments.
The usefulness of BD simulations has been demonstrated in studying the impact of hindered
Brownian diusion on the accuracy of nano-PIV [63], on the colloidal particle deposition in a
microchannel ow [61], and on the accuracy of potential energy proles determined via total internal
reection microscopy [62]. What has yet to be examined is the eect of hindered diusion and the
presence of the wall on the accuracy of velocimetry. In addition, it would be of interest to examine
how the relative strength of hindered diusion and shear ow leads to skewed near-wall apparent
velocity distributions rst reported by Jin et al. in their total internal reection velocimetry (TIRV)
experiment [80], and the relationship between particle diusion, time between image acquisitions
and the accuracy of velocity measurements based on ensemble averages. In this paper the theories
of near-wall particle dynamics will rst be presented and followed by a detailed description of
the BD simulation algorithm. We then discuss the simulation results and their implications, and
conclude with a summary and our thoughts on the accuracies of various velocimetry methods.
6.2 THEORIES AND COMPUTATIONS
6.2.1 The Langevin Equation
In Brownian dynamics simulation, particle displacements are computed based on a stochastic
Langevin equation [58]. For a particle in a shear ow as shown in gure 6.1, its displacement
between simulation time step i and step i + 1 with a step size t is,
x

i+1
x

i
= Ut +
dD
x
dx

t +
D
x
k
B

F
x
t +N
_
0,
_
2D
x
t
_
, (6.1a)
z

i+1
z

i
=
dD
z
dz

t +
D
z
k
B

F
z
t +N
_
0,
_
2D
z
t
_
, (6.1b)
where (x

, z

) is the particles center position, k


B
is the Boltzmann constant and is the uid
temperature. D
x
and D
z
are the diusion coecients in the directions parallel and normal to the
solid surface, respectively. In the uid bulk, both D
x
and D
z
would be equal to the Stoke-Einstein
52
Solid
z
x
a
z
x
S
U
Fluid
Figure 6.1: A schematic of the simulation geometry. A colloidal particle of radius a is freely
suspended in a uid but near a solid boundary. The uid is undergoing a shear ow of shear rate
S, and thus the particle translates due to both uid ow and its Brownian motion. A no-slip
boundary condition is assumed at the uid/solid interface. Even though existence of slip has been
reported [85], its presence would not alter the simulation results and is thus neglected.
diusivity, D
0
. As it has been mentioned earlier, near-wall diusion is anisotropic and hindered,
and thus D
x
< D
0
and D
z
< D
0
. It should be noted that because D
x
is only a function of z

,
dD
x
/dx

= 0 in equation (6.1a). The z

-dependency of D
x
and D
z
is discussed in section 6.2.3.
F
x
and F
z
represent external forces acting on the particle in the x- and z-directions, respectively,
and N
_
0,

2Dt
_
denotes the stochastic displacement of Brownian motion randomly sampled
from a normal distribution with a zero mean and a standard deviation

2Dt. In the current
BD simulation scheme there exists no force in the x-direction and F
x
= 0. Density mismatch
between tracer particles and the suspending uid, on the other hand, leads to either sedimentation
or oatation of the particles. The gravitation sinking force acting on the particles, F
g
, is thus the
only external force in the z-direction and
F
g
=
4
3
a
3
g, (6.2)
where is the density dierence between the tracer particles and the suspending uid, and g is
the gravitational acceleration.
The equations of motion (6.1a) and (6.1b) are non-dimensionalized by choosing the particle
radius, a, as the length scale and time required for the particle to diuse a distance of one radius,
a
2
/D
0
, as the time scale. Therefore equations (6.1a) and (6.1b) become
X

i+1
= X

i
+F
_
Z

i
_
Pe Z

i
T +N
_
0,
_
2
x
(Z

i
) T
_
(6.3a)
and
Z

i+1
= Z

i
+
d
z
dZ

i
T +G
z
_
Z

i
_
T +N
_
0,
_
2
z
(Z

i
) T
_
(6.3b)
53
S = 10
0
sec
1
10
1
sec
1
10
2
sec
1
10
3
sec
1
G =
a = 100 nm Pe = 0.0046 0.0463 0.4630 4.630 5.042 10
5
a = 1 m 4.630 46.30 463.0 4630 -0.5042
Table 6.1: Sample values of the Peclet number, Pe, and sedimentation coecient, G.
t = 0.1 msec t = 1 msec t = 10 msec t = 100 msec
a = 100 nm 0.022 0.22 2.2 22
a = 1 m 0.000022 0.00022 0.0022 0.022
Table 6.2: Representative values of the non-dimensional time between consecutive image acquisi-
tion, T.
where X

/a, Z

/a, and T D
0
t/a
2
are the non-dimensional x

, z

and t, respectively and


F
_
Z

U
z

S
,
x
_
Z

D
x
(z

)
D
0
,
z
_
Z

D
z
(z

)
D
0
as described in sections 6.2.2 and 6.2.3 while
Pe
Sa
2
D
0
, G
4a
4
g
3k
B

.
The Peclet number, Pe, represents the relative strength of shear force and diusion, and the
sedimentation coecient, G, characterizes the gravitational pull on the particles. Representative
values of Pe and G are shown in table 6.1.
In image-based velocimetry measurements, another important control parameter is the time
between consecutive image acquisitions, t. Some sample values of non-dimensional time between
consecutive image acquisition, T, for 1-m and 100-nm particles are shown in table 6.2.
6.2.2 Eects of Shear on Particle Velocities
It is well known that shear and near-surface hydrodynamic eects can cause a tracer particle to
rotate and translate at a velocity lower than the local uid velocity in the same shear plane [12, 38].
Goldman et al. [38] proposed that the translational velocity, U, of a particle with radius a in a
shear ow of a local shear rate S is given by
U
z

S
F
_
Z

_
1
5
16
_
Z

_
3
, (6.4)
which is valid for large Z

, and
U
z

S
F
_
Z

0.7431
0.6376 0.2 ln (Z

1)
(6.5)
for small Z

. Although there exists no analytical approximation for intermediate values of Z

,
Pierres et al. [60] proposed a cubic approximation to numerical values presented by Goldman et
al. in [38]:
U
z

S
F
_
Z

_
1
Z

_
exp
_
0.68902 + 0.54756
_
ln
_
Z

1
_
+0.072332
_
ln
_
Z

1
_
2
+ 0.0037644
_
ln
_
Z

1
_
3
_
.
(6.6)
54
Because a particles translational velocity depends on the shear rate and the distance to the wall,
it needs to be updated in equation (6.1a) at every time step of the simulation.
It has also been suggested that shear-induced lift can be a source of particle migration away
from the wall [86]. However, based on the theory presented by Cherukat & McLaughlin [87], the
shear-induced lift is insignicant for micro- and nano-particles in a low shear ow regime, and thus
will be neglected (for details, see Appendix B).
6.2.3 Hindered Diusion
Near-wall tracer particles are known to exhibit anisotropic hindered Brownian motion due to hy-
drodynamic eects, and their hindered diusion coecient in the direction parallel to the solid
surface, D
x
, is [37]
D
x
D
0

x
_
Z

_
= 1
9
16
_
Z

_
1
+
1
8
_
Z

_
3

45
256
_
Z

_
4

1
16
_
Z

_
5
+O
_
Z

_
6
, (6.7)
where D
0
is the Stokes-Einstein diusivity of an isolated spherical particle in the uid bulk. This
Method of Reection approximation is more accurate for Z

> 2. For Z

< 2, Goldman et al.


[38] proposed an asymptotic solution,
D
x
D
0

x
_
Z

_
=
2 [ln (Z

1) 0.9543]
[ln (Z

1)]
2
4.325 ln (Z

1) + 1.591
. (6.8)
In the direction normal to the wall, the modied diusion constant, D
z
, is such that [36]
D
z
D
0

z
_
Z

_
=
_
4
3
sinh

n=1
n(n + 1)
(2n 1) (2n + 3)
_
2 sinh (2n + 1) + (2n + 1) sinh 2
4 sinh
2
_
n +
1
2
_
(2n + 1)
2
sinh
2

1
__
1
,
(6.9)
where = cosh
1
(Z

). Bevan & Prieve [39] reported that the equation

z
_
Z

_
=
6 (Z

1)
2
+ 2 (Z

1)
6 (Z

1)
2
+ 9 (Z

1) + 2
(6.10)
well approximates equation (6.9) and is much less computationally intensive than the innite series.
Like the particle translational velocity, the dependence of these hindered diusion coecients on
Z

requires them to be updated at every time step of the simulation.


6.2.4 Implementation of Simulation
Since the experiments are conducted using dilute suspensions, the BD simulations were conducted
under an assumption of no particle-particle hydrodynamic interactions. At the beginning of each
simulation, a particle was situated at X

= 0 and randomly placed in a pre-determined range of


1 < Z

< 10. The simulation then progressed for a total of T/T steps as prescribed by equations
(6.3a) and (6.3b), with F and s updated and the position (X

, Z

) of the particle recorded after


each time step. Because the solid wall was located at Z = 0, the smallest Z

value a particle could


have was Z

= 1, where the particle would be in contact with the wall. A boundary condition
was needed in the event that a particle attempted to enter the solid wall during a simulation step.
55
Peters & Barenbrug [98, 99] have studied the eciencies of dierent boundary conditions for BD
simulations. Here we chose a simple and yet eective specular reection to prevent a particle from
entering the wall. With the time step taken small enough, the use of the boundary condition was
seldom triggered and less than 0.001% of the simulated displacements required applications of the
boundary condition.
The single particle simulation was then repeated 10
5
times to obtain a large ensemble. It was
also repeated for various values of parameters Pe and G to study their eects. The initial positions
of particles and the seeding of our random number generator were made identical for all runs of
simulations to ensure that the dierences in results found between runs were the consequences of
simulation parameters only.
Finally, it is important to mention the selection of the size of the computational time step,
T, which has two physical constraints [58]. First, the time step must be much greater than the
particle relaxation time, mD
0
/k
B
, where m is mass of one particle. In nondimensional terms it
is equivalent to
T
mD
0
2
k
B
a
2
O
_
10
6
_
(6.11)
for a > 100 nm. Secondly, numerical accuracy requires that the time step must be short enough
such that the diusion coecients and their gradients are essentially constant during the time step
(that is, T 1). Therefore T was chosen to be 10
4
, which satises both constraints and was
numerically ecient.
6.3 RESULTS AND DISCUSSIONS
6.3.1 Sedimentation
Accurate particle velocimetry measurements require unbiased sampling of uid velocities by particle
ensembles. A common bias of this sort is sedimentation of particles between image acquisitions.
Shown in gures 6.2 and 6.3 are BD simulations of particle distribution for 1.5-m and 100-nm
radius particles, respectively, sampled after various times T. In the case of 1.5-m particles, it can
be observed that the distribution started out uniformly, but evolves with time and eventually settle
into a steady, non-uniform spatial distribution after a transient period. The 100-nm particles,
on the other hand, remain dispersed through the volume at all times. In both case the nal
spatial distribution agrees with the Boltzmann distribution prediction [91]. These results are not
surprising, but do serve to validate our simulation. With sedimentation being negligible for small
particles, we concentrate on presenting results of sub-micron particle simulations, as these particles
are commonly used in micro-PIV, nano-PIV and TIRV as tracers.
6.3.2 Particle Displacement Due to Hindered Diusion
The rst observation made is the displacement probability density function (PDF) of hindered
Brownian motion in the direction perpendicular to the wall, or the z-direction. It is well known
that the one-dimensional displacement PDF of an isolated particle in the uid bulk is a Normal
distribution N
_
0,

2D
0
t
_
. However, with the presence of the wall hindered diusion is expected
to break the symmetry of the displacement PDF, and is demonstrated in gure 6.4. Particles that
start o farther away from the wall can diuse a longer distance within a given amount of time,
as demonstrated by the wider displacement distribution widths. In addition, these displacement
PDFs are more symmetric and collapse onto each other. The displacement PDFs of near-wall
56
1 2 3 4 5 6 7 8 9 10
0
0.2
0.4
0.6
0.8
1
Z
P
D
F
(a)
1 2 3 4 5 6 7 8 9 10
0
0.5
1
1.5
2
2.5
Z
P
D
F
(b)
T = 0
T = 0.5
T = 1
T = 2
T = 3
T = 5
T = 5
T = 6
T = 7
T = 8
T = 9
T = 10
Boltzmann Distribution
Figure 6.2: Sedimentation of particles at G = 2.5, which is equivalent to a 1.5m-radius
polystyrene particle (density 1050 kg/m
3
) suspended in pure water. The particles are uniformly
distributed in the volume (1 < Z

< 10) at T = 0 and allowed to diuse only within this volume.


PDF stands for probability density function.
57
1 2 3 4 5 6 7 8 9 10
0.04
0.06
0.08
0.1
0.12
Z
P
D
F
(a)
1 2 3 4 5 6 7 8 9 10
0.04
0.06
0.08
0.1
0.12
Z
P
D
F
(b)
T = 0
T = 0.5
T = 1
T = 2
T = 3
T = 5
T = 5
T = 6
T = 7
T = 8
T = 9
T = 10
Boltzmann Distribution
Figure 6.3: Sedimentation of particles at G = 5 10
5
, which is equivalent to a 100nm-radius
polystyrene particle (density 1050 kg/m
3
) suspended in pure water. The particles are uniformly
distributed in the volume (1 < Z

< 10) and allowed to diuse only within this volume.


58
2 1 0 1 2
0
0.5
1
1.5
2
Z Z
0
P
D
F
Z
0
= 1.5
Z
0
= 2
Z
0
= 6
Z
0
= 11
Z
0
= 51
Z
0
= 101
Figure 6.4: Displacement distribution of small particles from various initial positions, or Z
0
. All
particles are allowed to diuse for an identical time T = 0.1, during which time no particle is
able to reach the wall.
particles, on the other hand, are signicantly skewed with a more pronounced tails in the positive
values and are narrower. This observation suggests that hindered diusion could lead to a bias of
measurements toward values obtained in the near-wall region.
Further proof of the hindered diusion eect is shown in gure 6.5. Particles start to diuse
away from the initial position of Z

= 2 as time evolves, with the spatial distribution skewed toward


the wall (Z = 0), which is a direct consequence of hindered diusion. After long periods of time
when the particles are allowed to reach the wall through diusion (T > 0.3), the skewness of the
spatial PDF is even more pronounced, leading to a high concentration of particles at the wall. Such
time-dependence of the particle spatial distribution could lead to time-dependent apparent velocity
measurements and is further discussed in section 6.3.5.
6.3.3 Particle Drop-ins and Drop-outs
The accuracy of velocimetry measurements relies on the ability to detect the same particles in
consecutive image acquisitions. This is especially dicult in near-wall microscopy measurements
because near-wall sub-micron particles can diuse away from the near-wall region quickly. One
such situation is demonstrated in gure 6.6. Suppose the imaging range of near-wall microscopy is
1 < Z

< 5, which is a good estimate for evanescent wave imaging using 100-nm radius particles.
The particles, initially uniformly distributed within the imaging range, will start to leave the
imaging range as time evolves. The portion of the particles that are not present in the imaging
range at the second image acquisition are said to have dropped out. Clearly, more particles
would drop out with longer inter-acquisition time. For the overall concentration of the particles to
remain uniform, many outside particles would have to drop in to take over the space. In PIV,
images of the dropping in particles at the second image acquisition would lead to additional errors
59
1 2 3 4 5
0
0.5
1
1.5
Z
P
D
F
T = 0
T = 0.1
T = 0.3
T = 0.5
T = 1
T = 5
Figure 6.5: Spatial distribution of small particles at various times due to hindered diusion. All
particles are located at Z

= 2 at T = 0.
0 5 10 15
0
0.05
0.1
0.15
0.2
0.25
Z
P
D
F
T = 0
T = 0.0001
T = 0.001
T = 0.01
T = 0.1
T = 1
T = 10
Figure 6.6: Spatial distribution of particles located within the range of 1 < Z

< 5 at T = 0. The
distributions are taken after various inter-acquisition times, T.
60
0 2 4 6 8 10
30
40
50
60
70
80
90
100
T
%

R
e
m
a
i
n
i
n
g

i
n

I
m
a
g
i
n
g

R
a
n
g
e 1.0 < Z
0
< 3.0
1.0 < Z
0
< 5.0
3.0 < Z
0
< 5.0
Figure 6.7: Percentage of particles in the imaging range for dierent inter-acquisition time. The
imaging range is the same as the initial particle distribution range, Z
0
.
in measurements.
As mentioned previously, the major concern for PIV is the percentage of particle drop-ins and
drop-outs because all imageable particles at the time of image acquisition are analyzed without
any eort to distinguish the entering and the leaving. The percentage of particles that do not
drop out for various inter-acquisition time is shown in 6.7, which is in agreement with a separate
numerical study conducted by Sadr et al. [59] Not surprisingly, a longer times between consecutive
acquisition results in a lower percentage of good particles, which are particles that do not drop
out. In dimensional terms, more than 90% of 100-nm particles remain in the imaging range of
100 < z < 300 nm (10% drop-out) if image acquisitions are taken at 1 msec apart. However, this
percentage drops to 65% if the inter-acquisition time is 10 msec. Based on this result, near-wall PIV
measurement with sub-micron particles should be conducted with a inter-acquisition time as short
as possible to ensure accuracy. If an acceptable percentage of drop-outs is 95%, the acquisitions
should be taken with T < 0.2, or t = 0.5 msec for 100-nm radius tracer particles. The drop-
outs, however, present less of a problem for Particle Tracking Velocimetry (PTV) as the technique
makes an eort to distinguish the particle drop-ins and drop-outs.
6.3.4 Horizontal Apparent Velocity Distributions
Subsequent analysis concentrates on horizontal motion of particles that do not drop out, emulating
physical TIRV measurements. In particular, the focus is on the eect of shear on apparent velocity
measurements. Shown in gure 6.8 is the distribution of non-dimensional apparent velocity, V =
X

/T, for particles under various Peclet number, Pe, or equivalently the shear rates. It is
clearly observable that the apparent velocity PDFs widen and skew as shear rate increases. This
agrees well with experimentally measured PDFs reported by Jin et al. [80] and Huang et al. [85].
Jin et al. suggested that the increasing skewness and widening of the apparent velocity distribution
61
0 10 20 30 40 50
0
0.2
0.4
0.6
0.8
V
P
D
F
Pe = 0.5
Pe = 1
Pe = 3
Pe = 5
Pe = 10
Figure 6.8: Apparent velocity distribution of particles remaining in the imaging range 1 < Z < 3
after T = 10. Larger Pe leads to an apparent velocity distribution with larger mean, width, and
skewness.
3 2 1 0 1 2 3
0
0.5
1
1.5
(VV
avg
)/V
avg
P
D
F
Pe = 0.3
Pe = 0.5
Pe = 1
Pe = 3
Pe = 5
Pe = 10
Pe = 30
Figure 6.9: Normalized apparent velocity distribution of particles remaining in the imaging range
1 < Z < 3 after T = 10. V
avg
is the mean of each apparent velocity PDF. The normalized
PDFs collapse onto each other at Pe > 3. In general, larger Pe has a narrower but more skewed
normalized PDF.
62
0 50 100 150
0
0.01
0.02
0.03
0.04
0.05
V
P
D
F
1 < Z
0
< 2
2 < Z
0
< 3
3 < Z
0
< 4
4 < Z
0
< 5
5 < Z
0
< 6
Figure 6.10: Apparent velocity distribution of particles at various imaging ranges. All apparent
velocity distributions are measured at T = 10 and Pe = 10. Particles that start o farther away
from the surface move faster because they are carried by uids at higher velocity planes, and their
distributions are more symmetric due to less inuence of the wall and hindered diusion.
violates the symmetric PDF assumption of PIV analysis and could lead to measurement errors.
Upon further analysis, it is observed that the apparent velocity PDFs can collapse onto a single
distribution for Pe > 3, as shown in gure 6.9. It is noted that the collapsed PDFs for Pe > 3
have a smaller distribution width but a larger skewness, while the distributions of Pe < 3 are wider
but more symmetric. The explanation is that at low shear rates, the horizontal displacement of
the particle ensemble is dominated by diusion, leading to a wider normalized PDF. At high shear
rates, on the other hand, the nite thickness of the imaging range and shear eect give rise to
the skewness and distribution widening. This physical explanation is supported by the fact that
both skewness and distribution width are scalable by the shear rate only at Pe > 3. The same
shear-dominance eect at Pe > 3 has been reported experimentally by Huang et al. [85], and the
agreement of BD simulations further validates the presented conclusion.
Another source of apparent velocity distribution skewness is the hindered diusion and the
presence of the wall. One can observe in gure 6.10 that the apparent velocity PDFs of imaging
ranges farther away from the wall are wider and less skewed than that of imaging ranges closer to
the wall. The dierence is even more obvious if each PDF is re-scaled by its distribution mean,
V
avg
, as plotted in gure 6.11. The combined eect of the hindered diusion and the shear-induced
rotational slow-down in the near-wall region causes the apparent velocity PDFs to skew, while at
far away the wall presence is not felt and the symmetric distributions again show scaling similarity.
6.3.5 Time Evolution of Apparent Velocity Distributions
Finally, the time evolution of apparent velocity distributions was examined and shown in gure
6.12. For a xed Pe, it is observed that diusion is quite dominant at small T, leading to apparent
63
100 50 0 50 100
0
0.01
0.02
0.03
0.04
0.05
V V
avg
P
D
F
1 < Z
0
< 2
2 < Z
0
< 3
3 < Z
0
< 4
4 < Z
0
< 5
5 < Z
0
< 6
6 < Z
0
< 7
7 < Z
0
< 8
Figure 6.11: Collapsed apparent velocity distribution shown in gure 6.10. The PDFs are scaled
by subtracting their corresponding distribution mean, V
avg
.
300 200 100 0 100 200 300 400
0
0.005
0.01
0.015
0.02
0.025
V
P
D
F
T = 0.0001
T = 0.01
T = 0.01
T = 0.1
T = 1
T = 10
Figure 6.12: Time evolution of apparent velocity distribution. All apparent velocity PDFs are
obtained at Pe = 30 and with particles within the imaging range of 1 < Z

< 3 at T = 0. Note
that the apparent velocity distribution narrows and skews with increasing inter-acquisition time,
T.
64
0 2 4 6 8 10
0.9
0.95
1
1.05
1.1
1.15
T
V
a
v
g
/
V
s
h
e
a
r
Pe = 0.3
Pe = 0.5
Pe = 1
Pe = 3
Pe = 5
Pe = 10
Pe = 30
Figure 6.13: Time evolution of mean particle apparent velocity. Each data point represents the
mean value of one apparent velocity distribution shown in gure 6.12. All mean apparent velocities
are obtained from particles that remained at imaging range of 1 < Z < 3. V
shear
is the calcu-
lated velocity of the particles in a near-wall shear ow if there is no Brownian motion, and is the
target quantity that velocimetry is trying to measure. Thus V
avg
/V
shear
can be interpreted as the
accuracy of velocimetry measurement. Note that data of all Pe show a minimum velocity ratio at
approximately T = 1.
65
10
3
10
2
10
1
10
0
10
1
0.9
1
1.1
1.2
1.3
1.4
1.5
T/W
2
V
a
v
g
/
V
s
h
e
a
r
W = 2
W = 4
W = 6
W = 8
W = 10
Figure 6.14: Rescaled time evolution of mean particle apparent velocity. For each data series, the
imaging range is set at 1 < Z

< W and the non-dimensional time is rescaled with 1/W


2
. Note that
the velocity ratios, V
avg
/V
shear
, collapse onto each other under this rescaled time scheme, except
near the minimum. Minimal velocity ratios for all Ws occur near an alternative non-dimensional
time scale of T/W
2
10
1
. All data series have Pe = 30.
velocity distributions that are signicantly wider than that of at large T. Due to increased shear
dominance, the apparent velocity distribution narrows as time increases. Such change can be
explained by the fact diusive velocity distribution wideth scales with T
1/2
,
X)

T V ) =
X)
T

1

T
, (6.12)
while the shear-induced ow velocity distribution width scales linearly with T. Thus diusion
accounts for a large fraction of the measured apparent velocity at short inter-acquisition time,
while the majority of the measure velocity is due to the shear ow at large T. This explains
the narrowing of the apparent velocity distributions in gure 6.12 as T increases. Clearly, if the
apparent velocity distributions evolve with time, it would be of interest to see if the mean value of
the velocity distributions evolve as a function of time as well.
The goal of vecimetry measurements is to accurately estimate uid velocities using the tracer
particle apparent velocity as an estimator. If V
shear
is dened as the mean velocity of all uidic
planes inside the imaging range and V
avg
is the mean value of an apparent velocity distribution such
as the ones shown in gure 6.12, then V
avg
/V
shear
can be interpreted as velocimetry measurement
accuracy with V
avg
/V
shear
= 1 being the ideal measurement. The time dependence of the velocity
ratio V
avg
/V
shear
is shown in gure 6.13. It can be observed that all V
avg
/V
shear
ratios exhibit the
form of a concave-up function of T, with a minimum occurring at approximately T = 1. The
velocity ratio would keep increasing as T increases. Figures 6.12 and 6.13 demonstrate that both
the shape and the mean of an apparent velocity distribution are not time-invariant.
With further analysis, it can be observed that the values of T at which the minima of
V
avg
/V
shear
ratio occur depend on the thickness of the imaging range. In their paper Sadr et al. [59]
suggested an alternative time scale, T/W
2
, if the imaging range is dened at 1 < Z

< W. This
alternative time scale, therefore, can be interpreted as the time it takes for a particle in contact
66
path 2
path 1
Imaging
range
Figure 6.15: Schematic of 2 potential paths for particle translation. Between image acquisitions,
particles that have taken dierent paths could have travelled dierent distances while sampling
dierent velocities along the way. Two potential paths of a particle are shown in this gure. If
the particle takes path 1, it samples the velocities of the lower planes and thus translates a short
horizontal distance. On the other hand if the particle takes path 2, it samples the higher velocities
of planes that are out of the imaging range, and thus translates a larger horizontal distance.
with the wall to diuse out of imaging range due to Brownian motion. Using this non-dimensional
time scale, gure 6.14 demonstrates that the velocity ratio V
avg
/V
shear
also has a scaling similarity.
In addition, it should be noted that in both gures 6.9 and 6.14, V
avg
/V
shear
0 at T 0 as
the instantaneous apparent velocities of particles would be the closest to the local uidic velocities.
So why is the velocity ratio minimal at T/W
2
10
1
, and increases monotonically afterward?
The answer to this question lies in gure 6.5 and is illustrated in gure 6.15. For small T, the
skewed distribution of particles toward the wall suggests that the imaged particles are more likely
to move closer to the wall. That is, path 1 in gure 6.15 is more likely to be taken than path
2. Since drop-outs are less signicant for small T, the consequence is that the tracer particles
sample toward the lower velocity planes, and thus the mean apparent velocity decreases. On the
other hand for large T, some of the particles that have left the imaging range are given enough
time to return at the second acquisition, such path 2 illustrated in gure 6.15. The larger the T,
the higher probability that particles can sample farther away from the surface and still manage to
return to the imaging range at the time of second image acquisition. Because they have sampled
the uidic velocities of the higher shear planes, these contributions of uidic velocities outside of
the imaging range are reected in the higher values of mean particle apparent velocities.
Clearly, the physical reasoning described in the previous paragraph suggests that the thickness of
the imaging range would be an important parameter, which justies the scaling of non-dimensional
time T with 1/W
2
. Indeed Sadr et al. used the Fokker-Planck approach to study the PDF
of uidic planes that an ensemble of particles would have sampled between image acquisitions in
nano-PIV. They obtained a ratio of averaged Z

value sampled by the particles to the mid-distance


of the imaging range, (W + 1) /2. The time dependence of this ratio is in striking resemblance to
that shown in gure 6.14. In both cases, the turning point in the relative contribution of the two
groups of particles (the ones that move closer to the surface and the ones that sample the higher
shear plane and return) occurs at T/W
2
10
1
. This suggests that the percentage of particle
drop-out and the positions whose uidic velocities are sampled by the particles play a crucial role
in the accuracy of velocimetry measurements.
Figures 6.13 and 6.14 can also serve as guides for choosing the appropriate T for near-wall
particle tracking velocimetry. First, one can observe in gure 6.13 that the velocity ratio is Pe-
67
invariant except near the minimum where uctuations exist. Thus the selection of T such that
the velocity ratio would be minimum should be avoided. Such avoidance is further reinforced by the
observation that the velocity ratio in gure 6.14 is not W-invariant in regions near the minimum.
It can be clearly observed that the dip in velocity ratio depends on the values of W, where large
W values lead to a lower minimal velocity ratios, while the velocity ratio shows no dependence on
W at both the higher and lower ends of the time scale T/W
2
. These observations suggest that
choosing an inter-acquisition time of T/W
2
O
_
10
1
_
would most likely yield measurement
inaccuracy.
6.4 CONCLUDING REMARKS
Brownian dynamics simulation oers an attractive approach to the study of near-wall particle dy-
namics. We demonstrate that wall-induced hindered diusion and translational slow-down of a
colloidal particle in a simple shear ow can be accurately captured by Brownian dynamics sim-
ulations, with results conrmed by experimental data. In near-wall colloidal dynamics, physical
length and time scales are particle radius and diusive time, and the problem can be mathematically
closed by specifying a Peclet number and a sedimentation constant. In the direction perpendic-
ular to the wall, hindered diusion skews the displacement distributions toward to wall. In the
direction parallel to the wall, colloidal particle motion is dominated by shear when Peclet number
is greater than 3, with hindered diusion also contributing to changes in the velocity distribution
width and its skewness. Furthermore, velocity distributions evolve as a function of time, reporting
mean velocities that are time-dependent with a minimum at T/W
2
10
1
.
This study also provides an assessment tool to the relative accuracy of near-wall PIV and PTV.
Because of a large amount of drop-ins and drop-outs when nanoparticles are used, the inaccuracy
of PIV in near-wall measurement could be signicant. PIV will remain accurate as long as the
time between acquisition is much smaller than the diusion time of tracer particles. As tracer
particles get smaller and their diusion become greater, the incapability of imaging equipments to
make consecutive image acquisitions with an extremely short inter-acquisition time would limit the
applicability of PIV in near-wall measurements.
On the other hand, PTV and SPTV (Statistical Particle Tracking Velocimetry) [94] are much
more reliable measurements because PTV attempts to exclude particle drop-ins and drop-outs
while SPTV treats the drop-ins and drop-outs as statistical noise. The potential error of both
methods would stem from biased sampling of the velocities at various shear planes. However, gure
6.14 oers a way to correct for this bias. By using it as a scaling guide, the correct shear-induced
velocity of the particles can be retrieved from the measured mean velocity values. An alternative
method is to calculate which inter-acquisition time would result in the measured mean velocity
being equal to the shear-induced velocity before experiments are conducted. Neither correction
method can be applied to PIV measurements.
Chapter 7
Concluding Remarks and Suggested
Studies
In this dissertation we have presented our experimental studies of aqueous solution boundary slip
and anisotropic hindered diusion, and described the total internal reection velocimetry (TIRV)
method used in these experiments. A Brownian dynamics simulation study is also presented to
compare the eectiveness and accuracy of TIRV and nano-PIV. Furthermore, these studies open
up opportunities for future experimental works in the areas of uid mechanics and colloidal science.
First of all in pursuing a better understanding of boundary slip, water is by no means the
only test uid of interest. Water and similar solvents such as alcohols have a polar molecular
conguration, and their chemical properties are vastly dierent from the non-polar group of solvents,
such as hexane and toluene. It would certainly be of scientic interest to see if the solvent molecular
polarity would have an eect on boundary slip, and if the solid surface charges would interact with
the weak charges carried by the solve molecules.
The TIRV technique can certainly be extended to near-surface measurements in a gas phase.
Although the existence of boundary slip at a gas/solid interface has been well documented, the
TIRV technique oers a method to make direct measurement of the gas slip velocities. In addition,
TIRV can also be applied to experimentally investigate ow characteristics within the gas boundary
layer. Of course, uorescent aerosols would be required as tracer particles, and we believe such
technology is possible in the near future.
Returning to the liquid phase, particle or cell adhesion and desorption are both areas of great
interests in colloidal science and biology, and the TIRV technique is ideal for studying these phe-
nomena. The area of interest include both specic and non-specic binding of suspensions and
the solid boundary. To investigate specic chemical binding, the particles and the solid boundary
surface can be coated with molecules that bind correspondingly, such as ligands and receptors to
study blood cell adhesion. Besides binding, kinetics of surface-based chemical reactions can also be
studied with the same mechanism.
Understandably, advancement of technology would soon deem particles of 100 nanometer in-
sucient in probing the near-wall region as scientists would like to look even closer to the solid
surface. Thus a new breed of tracer particles or molecules in the order of 1 nanometer would be
needed. In fact, quantum dots have been viewed as the next generation of tracer probes, and their
usage under evanescent wave imaging settings have been demonstrated by Pouya et al. [100] and
Guasto et al. [94]. These quantum dots are small (a few nanometer in diameter), more uniform in
size than polystyrene uorescent particles, and available in water or organic solvent soluble forms,
thus making them an ideal candidate to probe regions that are even closer to the wall with a variety
68
69
of liquids. Perhaps a more accurate measurement of the slip velocities can be achieved by using
quantum dots in TIRV.
The Brownian motion of these quantum dots is undoubtedly much more signicant and the par-
ticle tracking method in the TIRV will be incapable of dealing with the large proportion of particle
drop-ins and drop-outs between image acquisitions. An additional limitation of the particle-
tracking based TIRV is its requirement of low particle seeding density to avoid tracking ambiguity.
In some experimental situations, however, a high seeding density might be desirable to preserve
the particle-particle interactions and/or to increase measurement eciency. The statistical particle
tracking velocimetry (SPTV) developed by Guasto et al. [94] is a perfect solution to these prob-
lems. They proposed that the true statistics of the desired physical displacement can be extracted
by matching each detected particle to all others, provided that experimental parameters such as
camera noise, imaging depth, particle size and diusivity, and seeding density are known. The
SPTV thus eliminates the needs to keep track of particle drop-ins and drop-outs and to unam-
biguously matching detected particles. It is certainly more advantageous than other particle-based
velocimetries, especially in near-surface measurements.
Appendix A
Calibration of beam incident angle
In our TIRV setup, the total internal reection was created at the glass/water interface by directing
a laser beam through a microscope objective lens. Knowing the beam incident angle is critical in
determining evanescent eld penetration depth. In this appendix, the measurement method of the
illumination beam incident angle is presented.
As demonstrated in equation (2.3), it would be ideal if one can measure the beam incident angle
directly when determining the penetration depth. In practice, because the total internal reection
bends the reected beam back into the objective lens, as shown in gure 2.1, such measurement is
extremely dicult.
The illumination laser beam is directed into the objective lens by passing through the a con-
verging lens. The bending of the beam is achieved by translating the converging lens [47]. Thus
a geometric relation should exist between the the lens translation distance and the beam incident
angle, and this relation should be true for all incident angles despite of total internal reection. To
determine this relation, the laser beam is directed through the objective as used in a non-TIRF
setup. The beam is then allowed to refract at a glass/air interface, as shown in gure A.1. Snells
law determines the refraction angle by
n
1
sin = n
2
sin , (A.1)
where n
1
and n
2
are refractive indices of glass and air, respectively. The refracted beam in air is
then projected to a far-away wall or ceiling where the refractive angle can be measured accurately
(shown in gure A.2).
By examining the data, we found a linear relationship between the lens position, x, and sin .
Using equation (A.1), an empirical linear relation between x and sin was found (shown in gure
A.3). The best t line was then extended to greater than critical angle to determine incident
angle. The empirical equation was found to accurately predict the lens position of critical angle.
70
71
Glass
Air

Figure A.1: Schematic of laser beam refraction at the glass/air interface.

Figure A.2: Photo and schematic of beam incident angle measurement. The refraction angle is
the same as that of gure A.1.
72
8.5 9 9.5 10 10.5 11
0.2
0
0.2
0.4
0.6
0.8
x (mm)
s
i
n
(

)

sin() = 0.33 x + 3.5
Data
Linear fit
Figure A.3: Plot of laser beam angle vs. converging lens position.
Appendix B
Shear-induced lift force (or more
precisely, lack of ) on near-wall
submicron-particles
It has been suggested that a shear-induced lift force can act on suspended particles in a linear
shear ow, making them more likely to move away from a solid boundary. This would result in a
non-uniform distribution of particles near our glass solid surface, where our total internal reection
velocimetry (TIRV) measurements take place. Because our statistical analysis is based on the
assumption of uniform particle distribution, the shear-induced lift force would lead to signicant
inaccuracies. However, we believe that such a lift force is insignicant in our ow regime, and in
this appendix, an established theory is presented to support our argument.
The subject of lift forces acting on a small sphere in a wall-bounded linear shear ow has been
thoroughly studied by Cherukat & McLaughlin [87]. Here we will present only the theory that
applies to the ow conditions under consideration. Suppose that a free-rotating rigid sphere of
radius a is in a Newtonian incompressible uid of kinematic viscosity and is in the vicinity of a
solid wall. In the presence of a linear shear ow, the free-rotating sphere can can travel at a velocity
U
sph
that is dierent from the uid velocity, U
G
, of the shear plane located at its center [38]. We
can dene a characteristic Reynolds number based on the velocity dierence, U
s
= U
sph
U
G
, with
Re
s
=
U
s
a

. (B.1)
Another characteristic Reynolds number based on shear rate can be dened as
Re
G
=
Ga
2

, (B.2)
where G is the wall shear rate. In this geometry, the wall can be considered as located in the inner
region of ow around the particle if Re
s
and Re
G

2
, where a/h and h is the distance
between the particles center and the wall. For near-wall particle velocimetry described in chapters
3 and 4, Re
s
Re
G
10
7
while O(1), and thus the inner region theory of lift force applies.
For a at wall located in the inner region of ow around a free-rotating particle, the lift force,
F
L
, which is perpendicular to the wall, is scaled by [87]
F
L
Re
s
I, (B.3)
73
74
where I is a coecient that can be numerically estimated by
I =
_
1.7631 + 0.3561 1.1837
2
+ 0.845163
3

_
3.21439

+ 2.6760 + 0.8248 0.4616


2
_ _
Re
G
Re
s
_
+
_
1.8081 + 0.879585 1.9009
2
+ 0.98149
3

_
Re
G
Re
s
_
2
.
(B.4)
Again for the velocimetry conditions described in chapters 3 and 4, I O
_
10
2
_
. Therefore
F
L
Re
s
I
_
10
7
_ _
10
2
_
1, (B.5)
and the lift force acting on near-wall particles is insignicant and can be neglected for all practical
purposes.
Appendix C
Evanescent wave image of
micron-sized uorescent particles
Imaging of uorescent nanoparticles under TIRFM have been well reported by many groups, partic-
ularly in the exponential decay relation between particle intensity and its distance to the substrate
surface [44, 85, 90]. But for micron-sized uorescent particles in evanescent eld, there has been no
report on their imaging characteristics. In this appendix, numerical simulations were performed to
investigate some of the imaging characteristics associated with uorescent particles under evanes-
cent wave imaging.
The numerical simulation was conducted in the commercially packaged COMSOL Multiphysics
software environment. The simulation geometry is simplied from three-dimensional to two-
dimensional but is still a representative of near-surface imaging, and shown in gure C.1. Polystyrene
particles (refractive index 1.59) of three dierent diameters, d/ = 0.39, 6 and 12 where is the
incident beam wavelength, are immersed in the water phase (refractive index 1.33) and placed in
the vicinity of the glass phase (refractive index 1.51). A Gaussian beam enters from the lower-left
boundary and reects o the water/glass interface at a 64.54

angle, while scattering boundary


conditions are applied to all other external boundaries. Simulations were repeated for various gap
sizes, h, between the particle and the glass, with h/ ranging from 0 to 2. The eectiveness of the
COMSOL Multiphysics package in modelling total internal reection and evanescent eld was rst
demonstrated in the absence of a polystyrene particle, and is shown in gure C.2. The simulation
model perfectly captures total internal reection and an exponential decay of evanescent energy
density in the water phase is found to be in agreement with theories (shown in gure C.3).
Time-averaged energy densities of particles in evanescent eld are shown in gures C.4, C.5 and
C.6, respectively. In gure C.4, it can be observed that the presence of a particle with diameter
d/ = 0.39 does not distort the energy eld. In contrast, larger particles such as d/ = 6 and
12 can distort the energy eld quite signicantly. Their presence can be seen to create a shadow
in the reected beam. These lost energy leaks into the interior of the large particles through
a frustrated-total internal reection mechanism, illuminating regions beyond the penetration
depth of evanescent wave. These energy continue to reect and refract at the polystyrene/water
interface, resulting in a high energy density in the lower-right side of the particle, while some
additional energy escapes into the right half of the water phase.
Using the subdomain integration function of COMSOL, one can nd out the overall energies
inside the polystyrene particles. This quantity is important because in TIRFM experiments, only
the light emitted by the uorophores inside the polystyrene particles is imageable, and the uo-
rophores are excited only by the light energy that enters the particles. The overall energies that
75
76
Incident Gaussian Beam
Glass Phase
Water Phase
Polystyrene Particle
Figure C.1: Schematic of COMSOL simulation geometry. A polystyrene particle is suspended
in the water phase and located in the vicinity of glassinterface. Scattering (or oating) boundary
condition is applied to all external boundaries at the water and glass phases, except at the boundary
inside the glass phase where the incident Gaussian beam enters. Because the incident angle of the
Gaussian beam is 64.54

, a total internal reection occurs at the glass/water interface. The particle


shown in the gure has a diameter of d/ = 6.
Figure C.2: COMSOL simulation of total internal reection. Plotted in the gure is time-averaged
total energy density of total internal reection at a glass/water interface. The geometry of this
simulation is identical to that of gure C.1, except the presence of a suspended particle.
77
0 0.5 1 1.5 2
0
0.2
0.4
0.6
0.8
1
h/
N
o
r
m
a
l
i
z
e
d

I
n
t
e
n
s
i
t
y


Simulation
Theory
Figure C.3: The exponential intensity decay of evanescent eld in gure C.2. Close agreement
between simulation results and theoretical calculations is obtained.
Particle
Figure C.4: COMSOL simulation of a particle with d/ = 0.39 in evanescent eld. Plotted in the
gure is time-averaged total energy density, and the geometry of this simulation is identical to that
of gure C.1, except that the particle diameter is with d/ = 0.39 and the gap size between the
particle and the glass surface is h/ = 0.1.
78
Figure C.5: COMSOL simulation of a particle with d/ = 6 in evanescent eld. Plotted in the
gure is time-averaged total energy density, and the geometry of this simulation is identical to that
of gure C.1, except that the gap size between the particle and the glass surface is h/ = 0.1.
Figure C.6: COMSOL simulation of a particle with d/ = 12 in evanescent eld. Plotted in the
gure is time-averaged total energy density, and the geometry of this simulation is identical to that
of gure C.1, except that the particle diameter is d/ = 12 and the gap size between the particle
and the glass surface is h/ = 0.1.
79
0 0.5 1 1.5 2
0
0.2
0.4
0.6
0.8
1
h/
N
o
r
m
a
l
i
z
e
d

I
n
t
e
n
s
i
t
y


d/ = 0.39
d/ = 6
d/ = 12
Evanescent Field
Figure C.7: Overall evanescent energy in the suspended particles at various gap sizes. The in-
tensities are normalized by the intensity observed when the particle is in contact with the glass
surface. The evanescent eld intensity is obtained from the cross-sectional time-averaged total
energy density prole of gure C.2.
enter the particles at gap sizes h/ ranging from 0 to 2 is shown in gure C.7. As it has been
demonstrated by Huang et al. [85] and Sarkar et al. [90], the overall intensity of a small particle
(d/ = 0.39) follows an exponential decay function that is identical to the evanescent energy decay.
This forms the basis of using the overall intensity of nanoparticles for 3D displacement tracking.
The overall intensity of larger particles, however, does not follow the same exponential decay func-
tion of the evanescent eld, particularly at small gap sizes. This is probably due to the fact that
a larger radius of curvature provides more surface area for light to leak into the particles through
the frustrated-total internal reection mechanism. Still, a near-exponential dependency of large
particle intensity on the gap size is observed in gure C.7.
A more relative question is what will a uorescent particle with diameter greater than the
illumination wavelength look like under a high numerical aperture (NA), high magnitude objective
whose focal depth, , is roughly / = 2. One of the ways to nd out is to convolute the energy
density inside the particles with diraction equation and point spread function. However, such a
convolution is computationally complicated and expensive. A simpler way is to look at the intensity
prole when the uorescent energy is projected onto the focal plane, with which one could still gain
some qualitative understanding of the image. For a point light source that is a distance z o the
focal plane, its intensity is [101]
I (z) =
_
sin (u/4)
u/4
_
2
I
0
(C.1)
where I
0
is the point source emission intensity and
u =
2

_
a
f
_
2
z. (C.2)
Here is the wavelength, a is the radius of the objective lens and f is the objective focal length. In
80
4 2 0 2 4
0
0.2
0.4
0.6
0.8
1
I
/
I
0
x/
Figure C.8: Cross-sectional intensity prole of a 3-m particle (d/ = 6). The blue squares
represent the cross-section intensity prole indicated as a red line in gure C.9. The red dotted
line, however, represents the projected intensity prole based on results of COMSOL simulation of
a particle of the same size and equation (C.3). I
0
is the peak intensity and x/ = 0 corresponds to
the particle geometric center. The direction of the evanescent wave propagation is toward positive
x/.
getting equation (C.1), it is assumed that both a and f are much larger than . a/f is commonly
referred to as the f-number of lens and for a microscope objective, it is equal to the numerical
aperture [73]. Therefore, for a particle in contact of the wall and its center is located at x = 0 and
z = a, its projected intensity at point x = x
0
on the z = z
0
plane is [101]
I
p
(x
o
) =
_
2ab
b
I (x
0
, z) zdz (C.3)
where b = a
_
a
2
x
2
0
and I (x
0
, z) is calculated using equation (C.1). This projection is obviously
a simplied calculation as it assumes negligible diraction. This leads to physical inaccuracy as we
will demonstrate subsequently.
The intensity projection of a particle with d/ = 6 is shown in gure C.8 as a red dotted line,
where the focal plane is set at the glass surface. In the projection, the intensity exhibits a peak value
near the particle center, while a longer tail is observed in the direction of the laser beam propagation.
A TIRFM image of a 3-m particle (equivalently d/ = 6) is shown in gure C.9. A peak intensity
near the center can be clearly observed in the image. For comparison, a cross-sectional intensity
prole of gure C.9 is also shown in gure C.8 as blue squares. This intensity prole is similar in
shape when compared to that of red dotted line from simulation. A peak intensity is found near the
particle center, while a slightly longer tail is found in the direction of evanescent wave propagation.
However, the intensity prole from actual imaging of a particle appears to be wider than that of the
simulated result. This is due to the fact that the computation leading to the projection (red dotted
line) in gure C.8 neglects diraction of light, which is important in microscopy. Nevertheless, the
81
1 m
Figure C.9: Digital image of a 3-m particle (d/ = 6) in contact with glass surface. The intensity
prole along the red cross-section shown is plotted in gure C.8.
resemblance of of the two proles is striking and the proles are in qualitative agreement.
The same imaging projection is also performed for a 6-m particle (d/ = 12) and shown in
gure C.10. With a particle of this size, we can again see that there is a bright peak near its center,
while the asymmetric tail on the right side is even more pronounced. A similar observation can be
made on an actual image of a 6-m particle (gure C.11) as well as on a cross-sectional intensity
prole (gure C.10, blue squares). A large bright center is clearly visible while the right half of
the particle is signicantly brighter. In addition, toward the right tip, there are small, alternating
regions of bright spots, which can be also be observed in the right tail of the red simulation intensity
prole in gure C.10. Again, the modelling results and the actual image show qualitative agreement
despite the simplicity of the integration model in equations (C.1) and (C.3). However, as observed in
gure C.8, the intensity prole obtained from the simulation results and equation (C.3) is narrower
than that of an actual image because of the neglected diraction in numerical integration.
82
6 4 2 0 2 4 6
0
0.2
0.4
0.6
0.8
1
x/
I
/
I
0
Figure C.10: Cross-sectional intensity prole of a 6-m particle (d/ = 12). The blue squares
represent the cross-section intensity prole indicated as a red line in gure C.11. The red line,
however, represents the projected intensity prole based on results of COMSOL simulation of a
particle of the same size and equation (C.3). I
0
is the peak intensity and x/ = 0 corresponds to
the particle geometric center. The direction of the evanescent wave propagation is toward positive
x/.
2 m
Figure C.11: Digital image of a 6-m particle (d/ = 12) in contact with glass surface. The
intensity prole along the red cross-section shown is plotted in gure C.10.
Bibliography
[1] Adam Bange, H. Brian Halsall, and William R. Heineman. Microuidic immunosensor sys-
tems. Biosensors and Bioelectronics, 20:24882503, 2005.
[2] Samuel K. Sia and George M. Whitesides. Microuidic devices fabricated in
poly(dimethylsiloxane) for biological studies. Electrophoresis, 24:35633576, 2003.
[3] Jan Kruger, Kirat Singh, Alan Oneill, Carl Jackson, Alan Morrison, and Peter OBrien.
Development of a microuidic device for uorescence activated cell sorting. Journal of Mi-
cromechanics and Microengineering, 12:486494, 2002.
[4] W. Mark Saltzman and William L. Olbricht. Building drug delivery into tissue engineering.
Nature Reviews, 1:177186, 2002.
[5] David J. Beebe, Glennys A. Mensing, and Glenn M. Walker. Physics and applications of
microuidics in biology. Annual Review of Biomedical Engineering, 4:261286, 2002.
[6] George E. Karniadakis and Ali Beskok. Micro ows: fundamentals and simulation. Springer,
2002.
[7] Eric Lauga, Michael P. Brenner, and Howard A. Stone. Microuidics: The no-slip boundary
condition. In J. Foss anc C. Tropea and A. Yarin, editors, Handbook of Experimental Fluid
Dynamics, chapter 15. Springer, New York, 2006.
[8] Bin Zhao, Jerey S. Moore, and David J. Beebe. Surface-directed liquid ow inside mi-
crochannels. Science, 291:10231026, 2001.
[9] Bin Zhao, Jerey S. Moore, and David J. Beebe. Principles of surface-directed liquid ow in
microuidic channels. Analytical Chemistry, 74:42594268, 2002.
[10] Zbigniew Adamczyk, Katarzyna Jaszczolt, Barbara Siwek, and Pawel Weronski. Irreversible
adsorption of particles at random-site surfaces. Journal of Chemical Physics, 120:11155
11162, 2004.
[11] Kai-Chien Chang and Daniel A. Hammer. Inuence of direction and type of applied force on
the detachment of macromolecularly-bound particles from surfaces. Langmuir, 12:22712282,
1996.
[12] M. Chaoui and F. Feuillebois. Creeping ow around a sphere in a shear ow close to a wall.
Quarterly Journal of Mechanics and Applied Mathematics, 56:381410, 2003.
83
84
[13] Poppo J. Wit, Albert Poortinga, Jaap Noordmans, Henry C. van der Mei, and Henk J.
Busscher. Deposition of polystyrene particles in a parallel plate ow chamber under attractive
and repulsive electrostatic conditions. Langmuir, 15:26202626, 1999.
[14] P. J. A. Hartman Kok, S. G. Kazarian, B. J. Briscoe, and C. J. Lawrence. Eects of particle
size on near-wall depletion in mono-dispersed colloidal suspensions. Journal of Colloid and
Interface Science, 280:511517, 2004.
[15] Remco Tuinier and Takashi Taniguchi. Polymer depletion-induced slip near an interface.
Journal of Physics: Condensed Matter, 17:L9L14, 2005.
[16] Yoram Cohen and A.B. Metzner. Apparent slip ow of polymer solutions. Journal of Rheology,
29:67102, 1985.
[17] Howard A. Barnes. A review of the slip (wall depletion) of polymer solutions, emulsions and
particle suspensions in viscometers: its cause character and cure. Journal of Non-Newtonian
Fluid Mechanics, 56:221251, 1995.
[18] Sydney Goldstein. Fluid mechanics in the rst half of this century. Annual Review of Fluid
Mechanics, 1:129, 1969.
[19] K. Johan A. Westin, Kenneth S. Breuer, Chang-Hwan Choi, Peter Huang, Zhiqiang Cao,
Bruce Caswell, Peter D. Richardson, and Merwin Sibulkin. Liquid transport properties in sub-
micron channel ows. In Proceedings of 2001 ASME International Mechanical Engineering
Congress and Exposition, 2001.
[20] Erhard Schnell. Slippage of water over nonwettable surfaces. Journal of Applied Physics,
27:11491152, 1956.
[21] N. V. Churaev, V. D. Sobolev, and A. N. Somov. Slippage of liquids over lyophobic solid
surfaces. Journal of Colloid and Interface Science, 97:574581, 1984.
[22] Peter A. Thompson and Sandra M. Troian. A general boundary condition for liquid ow at
solid surfaces. Nature, 389:360362, 1997.
[23] Jean-Louis Barrat and Lyderic Bocquet. Large slip eect at a nonwetting uid-solid interface.
Physical Review Letters, 82:46714674, 1999.
[24] Marek Cieplak, Joel Koplik, and Jayanth R. Banavar. Boundary conditions at a uid-solid
surface. Physical Review Letters, 86:803806, 2001.
[25] Chang-Hwan Choi, Johan A. Westin, and Kenneth S. Breuer. Apparent slip ows in hy-
drophilic and hydrophobic microchannels. Physics of Fluids, 15:28972902, 2003.
[26] YingXi Zhu and Steve Granick. Limites of the hydrodynamic no-slip boundary condition.
Physical Review Letters, 88:106102, 2002.
[27] C. Neto, V. S. J. Craig, and D. R. M. Williams. Evidence of shear-dependent boundary slip
in newtonian liquids. The European Physical Journal E, 12:S71S74, 2003.
[28] C. Cottin-Bizonne, B. Cross, A. Steinberger, and E. Charlaiz. Boundary slip on smooth hy-
drophobic surfaces: intrinsic eects and possible artifacts. Physical Review Letters, 94:056102,
2005.
85
[29] R. Pit, H. Hervet, and L. Leger. Direct experimental evidence of slip in hexadecane: solid
interface. Physical Review Letters, 85:980983, 2000.
[30] Derek C. Tretheway and Carl D. Meinhart. Apparent uid slip at hydrophobic microchannel
walls. Physics of Fluids, 14:L9L12, 2002.
[31] Pierre Joseph and Patrick Tabeling. Direct measurement of the apparent slip length. Physical
Review E, 71:035303(R), 2005.
[32] Keizo Watanabe, Yanuar, and Hiroshi Mizunuma. Slip of newtonian uids at solid boundary.
JSME International Journal, 41:525529, 1998.
[33] D. Lumma, A. Best, A. Gansen, F. Feuillebois, J. O. Radler, and O. I. Vinogradova. Flow
prole near a wall measured by double-focus uorescence cross-correlation. Physical Review
E, 67:056313, 2003.
[34] Steve Granick, Yingxi Zhu, and Hyunjung Lee. Slippery questions about complex uids
owing past solids. Nature Materials, 2:221227, 2003.
[35] Eric Lauga. Apparent slip due to the motion of suspended particles in ows of electrolyte
solutions. Langmuir, 20:89248930, 2004.
[36] Howard Brenner. The slow motion of a sphere through a viscous uid towards a plane wall.
Chemical Engineering Science, 16:242251, 1961.
[37] A. J. Goldman, R. G. Cox, and H. Brenner. Slow viscous motion of a sphere parallel to a
plane wall - i: motion through a quiescent uid. Chemical Engineering Science, 22:637651,
1967.
[38] A. J. Goldman, R. G. Cox, and H. Brenner. Slow viscous motion of a sphere parallel to a
plane wall - ii: Couette ow. Chemical Engineering Science, 22:653660, 1967.
[39] Michael A. Bevan and Dennis C. Prieve. Hindered diusion of colloidal particles very near
to a wall: revisited. Journal of Chemical Physics, 113:12281236, 2000.
[40] M. Hosoda, K. Sakai, and K. Takagi. Measurement of anisotropic brownian motion near an
interface by evanescent light-scattering spectroscopy. Physical Review E, 58:62756280, 1998.
[41] Binhua Lin, Jonathan Yu, and Stuart A. Rice. Direct measurements of constrained brownian
motion of an isolated sphere between two walls. Physical Review E, 62:39093919, 2000.
[42] Ratna J. Oetama and John Y. Walz. Simultaneous investigation of sedimentation and
diusion of a single colloidal particle near an interfce. The Journal of Chemical Physics,
124:164713, 2006.
[43] Arindam Banerjee and Kenneth D. Kihm. Experimental verication of near-wall hindered
diusion for the brownian motion of nanoparticles using evanescent wave microscopy. Physical
Review E, 72:042101, 2005.
[44] K. D. Kihm, A. Banerjee, C. K. Choi, and T. Takagi. Near-wall hindered brownian diusion of
nanoparticles examined by three-dimensional ratiometric total internal reection uorescence
microscopy (3-d r-tirfm). Experiments in Fluids, 37:811824, 2004.
86
[45] Daniel Axelrod. Total internal reection uorescence microscopy. In Methods in Cell Biology,
volume 30, chapter 9, pages 245270. Academic Press, Inc., 1989.
[46] Chris Rowe Tiatt, George P. Anderson, and Frances S. Ligler. Evanescent wave uorescence
biosensors. Biosensors and Bioelectronics, 20:24702487, 2005.
[47] Daniel Axelrod. Total internal reection uorescence microscopy in cell biology. Trac,
2:764774, 2001.
[48] Eric A. J. Reits and Jacques J. Neefjes. From xed to frap: measuring protein mobility and
activity in living cells. Nature Cell Biology, 3:E145E147, 2001.
[49] Taekjip Ha. Single-molecule uorescence resonance energy transfer. Methods, 25:7886, 2001.
[50] H. H. von Grunberg, L. Helden, P. Leiderer, and C. Bechinger. Measurement of surface
charge densities on brownian particles using total internal reection microscopy. Journal of
Chemical Physics, 114:1009410104, 2001.
[51] Scott G. Flicker, Jennifer L. Tipa, and Stacy G. Bike. Quantifying double-layer repulsion
between a colloidal sphere and a glass plate using total internal reection microscopy. Journal
of Colloid and Interface Science, 158:317325, 1993.
[52] Robert Kun and Janos H. Fendler. Use of attenuated total internal reection-fourier transform
infrared spectroscopy to investigate the adsorption of and interactions between charged latex
particles. Journal of Physical Chemistry, 108:34623468, 2004.
[53] P. Buchhave. Particle image velocimetry. In Lars Lading, Graham Wigley, and Preben
Buchhave, editors, Optical diagnostics for ow processes, pages 247270. Plenum Press, New
York, 1994.
[54] S. T. Wereley and C. D. Meinhart. Micron-resolution particle image velocimetry. In K. Breuer,
editor, Microscale Diagnostic Techniques, pages 51112. Springer, 2005.
[55] J. G. Santiago, S. T. Wereley, C. D. Meinhart, D. J. Beebe, and R. J. Adrian. A particle
image velocimetry system for microuidics. Experiments in Fluids, 25:316319, 1998.
[56] C. M. Zettner and M. Yoda. Particle velocity eld measurements in a near-wall ow using
evanescent wave illumination. Experiments in Fluids, 34:115121, 2003.
[57] Haifeng Li, Reza Sadr, and Minami Yoda. Multilayer nano-particle image velocimetry. Ex-
periments in Fluids, 41:185194, 2006.
[58] Donald L. Ermak and J.A. McCammon. Brownian dynamics with hydrodynamic interactions.
Journal of Chemical Physics, 69:13521360, 1978.
[59] Reza Sadr, Christel Hohenegger, Haifeng Li, Peter J. Mucha, and Minami Yoda. Diusion-
induced bias in near-wall velocimetry. Submitted to Journal of Fluid Mechanics, 2006.
[60] Anne Pierres, Anne-Marie Benoliel, Cheng Zhu, and Pierre Bongrand. Diusion of micro-
spheres in shear ow near a wall: use to measure binding rates between attached molecules.
Biophysical Journal, 81:2542, 2001.
87
[61] H. N. Unni and C. Yang. Brownian dynamics simulation and experimental study of colloidal
particle deposition in a microchannel ow. Journal of Colloid and Interface Science, 291:28
36, 2005.
[62] David S. Sholl, Michael K. Fenwick, Edward Atman, and Dennis C. Prieve. Brownian dy-
namics simulation of the motion of a rigid sphere in a viscous uid very near a wall. Journal
of Chemical Physics, 113:92689278, 2000.
[63] Reza Sadr, Haifeng Li, and Minami Yoda. Impact of hindered brownian diusion on the
accuracy of particle-image velocimetry using evanescent-wave illumination. Experiments in
Fluids, 38:9098, 2005.
[64] YingXi Zhu and Steve Granick. Rate-dependent slip of newtonian liquid at smooth surfaces.
Physical Review Letters, 87:096105, 2001.
[65] N. L. Thompson and B. C. Langerholm. Total internal reection uorescence: applications
in cellular biophysics. Current Opinion in Biotechnology, 8:5864, 1997.
[66] Jerey S. Burmeister, Lauri A. Olivier, W. M. Reichert, and George A. Truskey. Application
of total internal reection uorescence microscopy to study cell adhesion to biomaterials.
Biomaterials, 19:307325, 1998.
[67] Derek Toomre and Dietmar J. Manstein. Lighting up the cell surface with evanescent wave
microscopy. Trends in Cell Biology, 11:298303, 2001.
[68] J. Yamada. Evanescent wave doppler velocimetry for a walls near eld. Applied Physics
Letters, 75:18051806, 1999.
[69] David C. Duy, J. Cooper McDonald, Olivier J. A. Schueller, and George M. Whitesides.
Rapid prototyping of microuidic systems in poly(dimethylsiloxane). Analytical Chemistry,
70:49744984, 1998.
[70] Carl D. Meinhart and Steve T. Wereley. The theory of diraction-limited resolution in
microparticles image velocimetry. Measurement Science and Technology, 14:10471053, 2003.
[71] B. Ovryn. Three-dimensional forward scattering particle image velocimetry applied to a
microscopic eld-of-view. Experiments in Fluids, 29:S175S184, 2000.
[72] R. J. Adrian and C. S. Yao. Pulsed laser technique application to liquid and gaseous ows
and the scattering power of seed materials. Applied Optics, 24:4452, 1985.
[73] M. G. Olsen and R. J. Adrian. Out-of-focus eects on particle image visibility and correlation
in microscopic particle image velocimetry. Experiments in Fluids, 29:S166S174, 2000.
[74] C. J. Bourdon, M. G. Olsen, and A. D. Gorby. Validation of an analytic solution for depth of
correlation in microscopic particle image velocimetry. Measurement Science and Technology,
15:318327, 2004.
[75] J. C. Crocker and D. G. Grier. When like charges attract: The eects of geometrical conne-
ment on long-range colloidal interactions. Physical Review Letters, 77:18971900, 1996.
[76] S. H. Behrens and D. G. Grier. The pair interaction of charged colloidal spheres near a
charged wall. Physical Review E, 64:050401, 2001.
88
[77] T. M. Galea and Phil Attard. Molecular dynamics study of the eect of atomic roughness on
the slip length at the uid-solid boundary during shear ow. Langmuir, 20:34773482, 2004.
[78] Gyoko Nagayama and Ping Cheng. Eects of interface wettability on microscale ow by
molecular dynamics simulation. International Journal of Heat and Mass Transfer, 47:501
513, 2004.
[79] Cecile Cottin-Bizonne, Jean-Louis Barrat, Lyderic Bocquet, and Elisabeth Charlaiz. Low-
friction ows of liquid at nanopattened interfaces. Nature Materials, 2:237240, 2005.
[80] S. Jin, P. Huang, J. Park, J. Y. Yoo, and K. S. Breuer. Near-surface velocimetry using
evanescent wave illumination. Experiments in Fluids, 37:825833, 2004.
[81] Chang-Hwan Choi, K. Johan A. Westin, and Kenneth S. Breuer. To slip or not to slip - water
ows in hydrophilic and hydrophobic microchannels. In Proceedings of IMECE 2002, pages
IMECE200233707, 2002.
[82] Eric Lauga and Michael P. Brenner. Dynamic mechanisms for apparent slip on hydrophobic
surfaces. Physical Review E, 70:026311, 2004.
[83] Xue H. Zhang, Xiao D. Zhang, Shi T. Lou, Zhi X. Zhang, Jie. L. Sun, and Jun Hu. De-
gassing and temperature eects on the formation of nanobubbles at the mica/water interface.
Langmuir, 20:38133815, 2004.
[84] Michael K. Cheezum, William F. Walker, and William H. Guilford. Quantitative comparison
of algorithms for tracking single uorescent particles. Biophysical Journal, 81:23782388,
2001.
[85] Peter Huang, Jerey S. Guasto, and Kenneth S. Breuer. Direct measurement of slip velocities
using three-dimensional total internal reection velocimetry. Journal of Fluid Mechanics,
566:447464, 2006.
[86] Michael R. King and David T. Leighton Jr. Measurement of the inertial lift on a moving
sphere in contact with a plane wall in a shear ow. Physics of Fluids, 9:12481255, 1997.
[87] Pradeep Cherukat and John B. McLaughlin. The inertial lift on a rigid sphere in a linear
shear ow eld near a at wall. Journal of Fluid Mechanics, 263:118, 1994.
[88] Nasser A. Frej and Dennis C. Prieve. Hindered diusion of a single sphere very near a wall
in a nonuniform force eld. Journal of Chemical Physics, 98:75527564, 1993.
[89] Peter Huang and Kenneth S. Breuer. Simulations of hindered diusion in shear ow and its
implications for near-wall velocimetry. In preparation, 2006.
[90] Atom Sarkar, Ragan B. Robertson, and Julio M. Fernandez. Simultaneous atomic force
microscope and uorescence measurements of protein unfolding using a calibrated evanescent
wave. Proceedings of the National Academy of Sciences, 101:1288212886, 2004.
[91] Richard A. L. Jones. Soft Condensed Matter. Oxford University Press, Oxford, England,
2002.
89
[92] Lakkapragada Suresh and John Y. Walz. Direct measurement of the eect of surface roughness
on the colloidal forces between a particle and at plate. Journal of Colloid and Interface
Science, 196:177190, 1997.
[93] Olga L. Vinogradova. Slippage of water over hydrophobic surfaces. Internal Journal of
Mineral Processing, 56:3160, 1999.
[94] Jerey S. Guasto, Peter Huang, and Kenneth S. Breuer. Statistical particle tracking ve-
locimetry using molecular and quantum dot tracer particles. in press, 2006.
[95] Peter Huang and Kenneth S. Breuer. Direct measurement of anisotropic near-wall hindered
diusion using total internal reection velocimetry. In preparation, 2006.
[96] Christel Hohenegger and Peter J. Mucha. Statistical reconstruction of velocity proles for
nano particle image velocimetry. To be published, 2006.
[97] Eric Lauga and Todd M. Squires. Brownian motion near a partial-slip boundary: A local
probe of the no-slip condition. Physics of Fluids, 17:103102, 2005.
[98] E. A. J. F. Peters and Th. M. A. O. M. Barenbrug. Ecient brownian dynamics simulation
of particles near wall. i. reecting and absorbing walls. Physical Review E, 66:056701, 2002.
[99] E. A. J. F. Peters and Th. M. A. O. M. Barenbrug. Ecient brownian dynamics simulation
of particles near wall. i. sticky walls. Physical Review E, 66:056702, 2002.
[100] Shahram Pouya, Manoochehr Koochesfahani, Preston Snee, Moungi Bawendi, and Daniel
Nocera. Single quantum dot (qd) imaging of uid ow near surfaces. Experiments in Fluids,
39:784786, 2005.
[101] Max Born and Emil Wolf. Principles of Optics: Electromagnetic theory of propagation inter-
ference and diraction of light. Cambridge University Press, 6th edition, 1980.

You might also like