You are on page 1of 30

MASTER THESIS

In Order to Obtain the

RESEARCH MASTER
in
Condensed Matter Physics

:Presented and defended by


Ali Hassan Mansour
On Thursday September 13, 2018

Title
Adhesion of spherical Nano-particles to metallic surfaces
Brownian dynamic simulation

Supervisor
Dr. Ali Atwi
Reviewers
Prof. Abbas Hijazi
Dr. Nader Yaacoub

Lebanese University-Faculty of sciences


Acknowledgment

Writing this report has been fascinating and extremely rewarding. I would like to thank a number
of people who have contributed to the final result in many different ways:
To commence with, I pay my obeisance to GOD, the almighty who has bestowed upon me good
health, courage, inspiration, zeal and the light. After GOD, I would like to express my gratitude
to my supervisor doctor Ali Atwe for the useful comments, remarks and engagement through the
learning process of this master thesis. I especially thank the professors of my master 2 at the
Lebanese University, for their support throughout the academic year especially Prof. Abbas
Hijazi, the head of master 2. I would like newly to acknowledge the reviewers Prof. Abbas Hijazi
and Dr. Nader Yaakoub who taught me a few courses this year at Lebanese University, as the
second readers of this report.

Finally, I must express my very profound gratitude to my parents, my friends and to my


inspiration Dr. Narjes Shoaib for providing me with unfailing support and continuous
encouragement throughout my years of study. This accomplishment would not have been
possible without them. Thank you.

Ali Mansour

“The beginning of knowledge is the discovery of


Something we do not understand.”
Frank Herbert

Keywords: colloidal suspension, Jeffery PDF, numerical simulations, mesopore nanotechnology,


diffusive collision.

1
Contents
TABLE OF CONTENTS........................................................................................................................................... 1
CHAPTER I GENERAL INTRODUCTION.................................................................................................................. 2
I.1 MACROMOLECULAR PARTICLES AND THEIR FORMS.........................................................................................................3
I.2 SPHERICAL NANOPARTICLES AND ITS APPLICATIONS.........................................................................................................4
I.3 JEFFERY’S METHOD.................................................................................................................................................. 4
CHAPTER II COLLOIDAL PARTICLE MODEL DYNAMICS.......................................................................................... 6
II.1 BOEDER EQUATION.................................................................................................................................................6
II.2 JEFFERY’S EQUATION...............................................................................................................................................6
II.3 COLLOIDAL PARTICLE MODEL DYNAMICS IN 3D-SPATIAL FRAMES....................................................................................8
II.4 COLLOIDAL PARTICLE ALGORITHMS IN 2D-SPATIAL FRAMES FOR OPEN PORE CHANNELS......................................................11
II.4.1 For three-dimensional spatial frame (3D)...................................................................................................12
II.4.2 For two-dimensional spatial frame (2D).....................................................................................................13
CHAPTER III: SIMULATION RESULTS FOR SPHERICAL FORMS OF COLLOIDAL PARTICLES.....................................14
III.1 SIMULATION RESULTS FOR RESTITUTION COEFFICIENTS E = 1........................................................................................14
III.2 SIMULATION RESULTS FOR RANDOM RESTITUTION COEFFICIENTS E.................................................................................15
III.3 SIMULATION RESULTS FOR RESTITUTION COEFFICIENTS E = 0.........................................................................................15
III.4 SIMULATION RESULTS FOR RESTITUTION COEFFICIENTS E' = 0........................................................................................18

CHAPTER IV MODELLING ADHESION AT PORE BOUNDARIES UNDER DIFFUSIVE COLLISIONS...............................20


CONCLUSION.................................................................................................................................................... 23
REFERENCES...................................................................................................................................................... 24
TABLE OF FIGURES ……………………………………………………………………………………………………………………………………………26

1
CHAPTER I

Chapter I General Introduction

The dynamics and mobility of macromolecular particles in colloidal suspensions in fluids


streaming inside mesopores, and the diffusive collisions and possible adhesion of proposals
molecule at the solid boundaries of the pores, is a subject of fundamental physical and biological
interest. The study and analysis of these impacts is necessary to comprehend an assortment of
complex phenomena, and towards technical applications, such concerning coatings and catalysis
in industrial process; and in protein diffusion in membranes and in the circulation of red blood
cells and platelets inside the human body in biological processes.

There is a great amount of research work which has explored the dynamics of colloidal particles
suspended in a flowing solution at various scales [1-4]. This is propelled by the biological and
engineering applications which include the stream of a suspension of particles in the fluid. [5-7]

Adhesion has been considered seriously finished the previous 40 years. This science depicts
how material particles of diverse sizes may adhere on solid boundaries. Adhesion phenomena are
essential in zones, for example, chemistry, engineering, and biology. Adhesion forces amongst
particles and surfaces is enthusiasm for a few industrial applications such as particle filtration
[8], and petroleum production. For biological systems such as cells or viruses [9], adhesion to
surfaces is critical for processes like bio-film development or infection.

The adhesion phenomenon has been widely utilized, even in the ancient age, when bitumen and
natural resins were used as adhesives for manufacturing weapons and tools, and to repair broken
pottery [10], and to make the oldest known paintings which are still adhering on their rock
substrate at the Chauvet-Pont-d'Arc cave, in southern France dated at about 32,000 years prior
[11].

The issue of determining the bond between solid bodies in contact is vital for understanding
friction, wear, the agglomeration and dispersion of colloidal particles, and many other
phenomena [12]. As a complex, multifaceted phenomenon, adhesion involves factors such as
surface chemistry, thermodynamics, polymer chemistry and material science, hydrodynamics,
and contact mechanics [13].

Though daily experience does not straightforwardly show us about adhesion, (the basic outline
for the most established houses of prayer depended on the presumption that stones don't stick
together [14] however are basically held set up by gravity), yet it is obvious that the atoms and
molecules, stick together to a great degree well. At any rate in the microscopic world almost
everything sticks well. Nevertheless, something obstructs the consciousness of universal
adhesion in our macroscopic world. The adhesion paradox is tributary to two important aspects
of molecular contact formation, namely contamination and roughness.

2
CHAPTER I

I.1 Macromolecular particles and their forms

The geometry of the macromolecules is perplexing and takes a few structures and shapes. To
facilitate the study of the macromolecules, research scientists have modeled the macromolecules
particles by basic and simple geometric shapes: discs, ellipsoidal, spheres, rod- like particles, and
others compound shapes like in Fig. 1, Fig. 2 and Fig. 3.

Fig. 1. Ellipsoidal particles Fig. 2. Spherical particles Fig. 3. Rod-like particles

Ellipsoidal particles are important because their colloidal behavior such as Brownian motion,
[15], maximum packing, [16] as well as crystal structure [17] can be tuned systematically by
changing the aspect ratio of their axes. Moreover, ellipsoidal particles can be an amazing model
system in condensed matter physics and drug delivery vehicle design. In addition, [18] showed
that ellipsoids can be composed into both the translational and orientational order through
electric field assembly, leading to new structures that are otherwise unattainable using spherical
particles, such arranged structures may have new mechanical properties, Fig. 4.

Fig. 4. A scanning electron microscopy (SEM) image of the titanium dioxide particles; with an aspect
ratio of 4.19.

3
CHAPTER I

3
CHAPTER I

I.2 Spherical nanoparticles and its applications

Recent advances in nanotechnology are as a result of the development of engineered


nanoparticles. Efficiently, metallic nanoparticles have been widely exploited for biomedical
application and among them, gold nanoparticles (AuNPs) are highly remarkable. Consequent
upon their significant nature, spherical nanoparticle attract extreme attention. Their intrinsic
features such as optical, electronic, physicochemical and surface plasmon resonance (SPR);
which can be altered by changing the characterizations of particles such as shape, size, aspect
ratio, or environment; ease of synthesis and functionalization properties have resulted to various
applications in different fields of biomedicine such as sensing, targeted drug delivery, imaging,
photothermal and photodynamic therapy. Moreover, Spherical titanium dioxide (TiO2) is a
multifunctional material used in pigments, cosmetics, and food additives since the early
twentieth century [19] TiO2 has recently shown promise in energy and environmental
applications, such as solar cells [20], photocatalysts, and batteries. TiO2 exhibits unique optical,
electrical, and chemical properties, and is non-toxic and widely abundant.
In addition, most of the magnetic particles studied so far are spherical, which somewhat limits
the possibilities to make these nanoparticles multifunctional.
Furthermore, Spherical Gold Nanoparticles it has interest applications such as:

 Probes for dark-field microscopy, electron microscopy and near-infrared fluorescent


microscopy.
 Diagnostics for lateral flow immunoassay.
 Sensors for colorimetric detection and surface enhanced Raman spectroscopy.
 Drug targeting, delivery and photothermal therapy.

I.3 Jeffery’s method

The inception for current models for the study of the dynamics of colloidal non-spherical
particles in a shear flow at low Reynolds numbers is historically the approach proposed some
time ago by Jeffery. Jeffery acquired a set of differential equations for the rotation of an
ellipsoid, with analytic solutions depending on the initial conditions. This theoretical approach
was confirmed experimentally and has been studied and developed since by several groups and
specialists, both theoretically and experimentally. The overall rotation of the colloidal particles in
the bulk liquid is alluded to as a Jeffery orbit. Bretherton demonstrated that Jeffery’s model
could be applied to any axis-symmetric particle, provided that an equivalent aspect ratio was
introduced.

4
The study of the dynamics of spheroids has been carried out by a few creators. The numerical
simulation of the dynamics for spheroids moving close to a solid surface boundary was done [21]
using Jeffery’s equations. The authors concluded that hydrodynamic forces and torque depend
emphatically on the spheroid orientation and its position relative to a wall boundary; they
revealed tumbling of spheroids that resulted in motion toward the wall, since the spheroids were
turned around the point nearest toward the wall.

The dynamics of the colloidal particles depend to a variety of effects, namely the hydrodynamic
forces coming from the shear flow, the thermal originating from stochastic Brownian movement
and the diffusive collisions at the pore limits.

It is the purpose behind the present work to treat these forces toward a comprehensive analysis of
the dynamics of the colloidal particles, in the two-dimensional spatial frame (2D), in the bulk of
the streaming liquid as well as close to the surrounding solid surface boundaries.

To that end we will develop in this project appropriate algorithm incorporating these random
forces, in addition to the impact of the hydrodynamic forces at the heart of Jeffery’s equations.
Based on these algorithms we at that point complete numerical simulations to analyze the
dynamics of colloidal particles of general ellipsoidal forms, then spherical forms by controlling
the aspect ratio, in dilute suspensions, in the bulk of the flowing liquid solution, and by perfect
atomically flat solid surface boundaries.

Our study is centered on the determination of the probability distribution functions (PDF) for the
positions of the centers of mass of these particles. The conditions at the surface limit for
diffusive collisions are clarified with the utilization of the idea of mechanical restitution, which
shows a closed measure, to create unambiguous algorithms for the molecular dynamics of the
diffusive collisions, and thus unambiguous numerical simulations.

It is notable that it is obstinate to study by the only analytical means the dynamics of the
molecular particles in dilute colloidal suspensions near solid surfaces, as a result of the irregular
nature of the Brownian motion and the similarly random nature of diffusive collisions. The main
reasonable alternative is to do this by numerical simulations with appropriate algorithms.

In chapter 2, we present a technical introduction to the problem and to liquid bulk


macromolecular dynamics under Brownian and hydrodynamic motion, and we show the
developed algorithm for diffusive collisions at the solid surface boundaries of open pore
channels. Chapter 3 presents simulation results for the spatial PDF distributions for colloidal
suspensions, for a wide range of hydrodynamic conditions and for spherical particles in bulk
solution and near the solid surfaces. In Chapter 4, we used the results obtained by Malak Ashkar
master 2 student and our results to determine under given physical conditions the outcome for
adhesion or not.

5
CHAPTER II

Chapter II Colloidal particle model dynamics


An ellipsoidal particle in simple shear flow invests the vast majority of its time adjusted
relatively parallel to the streamlines, however when the concentration of particles in the
suspension increase, the interaction of the particles with different particles and with the boundary
solid surface prompts other induced ordinations. Particle-particle interaction has been the subject
of study for many researchers.

Our present work deals with dilute suspensions of spherical colloidal particles and subsequently
does not consider the interaction effects between the particles; we focus instead on the particle -
solid surface boundary interactions.

II.1 Boeder equation

Boeder, [22] is the first to have studied the orientations of dilute concentrations of
macromolecular particles in the bulk of a streaming liquid from the theoretical perspective, for
rod-like macromolecular particles. He presented a differential equation (BDE) in a 2D-spatial
frame, considering the dynamic effects due to the Brownian and hydrodynamic forces acting on
the particles. The BDE governs the variations of the probability distribution functions (PDF),
P(θ), of the particles as a function of their orientations θ in the bulk with respect to the direction
of the shear flow.

A key normalized amount for the study of the problem is α, which is the ratio γ̇ /Drot, where γ̇ is
the constant shear rate of the linear hydrodynamic flow and Drot is the diffusion coefficient for
the Brownian rotational motion of the macromolecular particle about its center of mass. Diverse
macromolecules in solvents have naturally different Drot. The rotational peclet number α is the
dimensionless ratio that characterizes the relative strengths of hydrodynamic and Brownian
effects.

Some scientists introduced [1, 23-26] the notion of geometric restitution for both Brownian and
hydrodynamic collisions. Since the nature of these impacts is different, random when arising
from diffusion and causal under the hydrodynamic shear, two separate restitution coefficients, e
and e', are considered to describe these collisions, respectively.

II.2 Jeffery’s equation

In particular, the mechanical restitution model at the solid boundaries is generalized to three
dimensions and combined with a fuller implementation of the Jeffery’s which govern the
movement of ellipsoidal forms of the colloidal particle in an unbounded straight flow field.

6
CHAPTER II

It is hence this entire version which is utilized in the present model to compute the PDF
distributions for macromolecular ellipsoidal particles in the neighborhood of solid surface
boundaries.
We consider a simple shear flow acting on an ellipsoidal particle, such that

vx = γ̇ y and vy = vz = 0

γ̇ is the hydrodynamic shear rate. The motion of a solid ellipsoid particle suspended in a simple
shear flow, was computed analytically [27] neglecting the inertia of the fluid and the particle.
The ellipsoidal orientation is characterized by three angles (θ, φ, ψ), which determine the
Cartesian coordinates x, y, z. In contrast, a local coordinate system x', y', z', translates and rotates
with the ellipsoidal particle, see Fig. 5. In Jeffery’s theory it is assumed that the ellipsoidal center
translates with the indistinguishable linear speed as that of the particle centroid.

Fig. 5. Coordinate system for ellipsoidal particle centered at the origin

The angles φ and θ are used to define the unit direction of the primary axis and ψ represents the
rotational angle about the ellipsoidal particle. The differential equation governing the time
evolution of θ and φ are given by

1
θ̇ = - γ̇ ( 2 ) (r 2e sin2 θ +cos 2 θ ) (1)
r +1
e

2
γ̇ r e −1
φ̇ = ( 2 ) sin 2 φ sin 2θ (2)
4 r e +1

In these equations, γ̇ is the shear rate; re is the aspect ratio of the ellipsoidal particles, θ is defined
as the angle of the particle with respect to flow direction, φ as the vorticity axis; φ = π /2 when
the ellipsoidal particle lies in the plane of shear.

6
CHAPTER II

It very well may be seen that the angular velocity θ' is maximum when the ellipsoid is
perpendicular to the flow direction (φ = π /2) and minimum when it aligned to the flow direction
(θ = 0).

From Eqs. (1) and (2), it is observed that the angular velocities depend linearly on the shear rate,
if Eqs. (1) and (2) are integrated with respect to time they may be re-written as

γt˙
¿
θ (t) = tan ¿ ¿ tan r + 1
−1
(3)
e
re

−1 C re
φ (t) = tan ( ) (4)
√r 2e cos2 θ+sin 2 θ
While the rotational equation is
t
γ̇
ψ (t) = ∫ ( −θ̇) (5)
0 2

The constant C characterizes the eccentricity of the orbit executed by the particle and takes the
values between zero and infinity. C = 0 suggests that the particle is aligned along the vorticity
axis (φ= 0), while C→∞ implies that the particle lies in the shear plane φ = π /2, see Fig. 6 for
details. The orientation of the ellipsoidal particle can be described completely by the spherical
coordinate system. The θ and φ defined in a settled Cartesian coordinate system, as shown in
Fig.2, (-π/2 ≤θ≤ + π/2 and 0≤ φ ≤ π), are adequate because the problem is invariant under the
transformation φ→φ + π.

Fig. 6. Calculated example of Jeffery’s different orbit constants C for re→ ∞

II.3 Colloidal Particle Model Dynamics in 3D-Spatial Frames

In the simulations using Jeffery’s equations, the hydrodynamic force has a tendency to align the
ellipsoidal macromolecular particles in the direction of the shear flow, but in the case of a

6
CHAPTER II

spherical particle, their suspension shows a much smaller range of behavior, since the orientation
of these particles make its configuration less likely to be affected by flow fields. To simulate the
effects of hydrodynamic forces in a time interval ∆t between two successive simulation events
labelled sand s+1, we compute the hydrodynamic rotations about the center of mass of the
particles, ∆ θ hyd (s+1 ; s ) and ∆ φhyd ( s+1 ; s), by using the following algorithm:

1 2 2 2
∆ θ h yd ( s+ 1; s)= θ (t s+1 ) - θ (t s) = - γ̇ ( 2 ) (r e sin θ ( t s ) + cos θ(t s ))∆ t (6a)
r +1
e

r 2e +1 sin 2θ ( t s ) sin 2 φ(t s )


∆ φh yd (s +1; s)= φ (t s+1 ) - φ (t s) = γ̇ ( ) ∆t (6b)
r 2e −1 4

In contrast the Brownian forces in the bulk liquid solution make a diffusive rotational motion of
the particles, for which the rotation variables ∆ θ rot (s+1 ; s )and ∆ φrot ( s+ 1; s)can then be
assumed in the algorithm as already for ∆ θ rot in the form

∆ θ rot (s+1 ; s ) =± ∆ θrot (7a)

∆ φrot ( s+ 1; s) = ± ∆ φ rot (7b)

The Brownian rotations, ± ∆ θrot and ± ∆ φ rot, are clock and anticlockwise. In this strategy the
simulation time interval Δt is connected, as previously stated, to an effective variable for the
Brownian rotation diffusion by the following equations

∆ θ 2rot = 2 D θrot ∆ t (8a)


∆ φ2rot = 2 Dφrot ∆ t (8b)

For the small Δt simulation time intervals to be unique, at least for hydrodynamic events, they
must satisfy
1 1
∆t = Δθ2rot = φ2
2 Dθrot 2 D φrot Δ rot (9)

This leads to the simulation relationship

2 Dθrot 2
∆ θ 2rot = ∆ φrot ≡ r θ , φ ∆ φ 2rot (10)
2 Dφrot

There are two ways to treat r θ , φin the numerical simulations, as follows:
Dθrot
r θ , φ= =1 (11a)
Dφrot

r θ , φ = ( sinφ)−2 (11b)

6
CHAPTER II

Eq. (11a) is based on the assumption of independent and random Brownian simulation events,
where the ellipsoidal symmetry cuts the same diffusion coefficient in whatever angular
displacement it makes. This underestimates somewhat the physical relation of ∆ θ rot and ∆ φrot for
hydrodynamic events, which can be derived directly as in Eq. (11b). However, Eq. (11b) would
overestimate this relationship in the global simulation since random Brownian events may be
viewed as disconnected. Eqs. (11a) and (11b) yield respectively

∆ θ rot (s+1 ; s ) =± ∆ θrot (12a)

∆ φrot (s+1; s) = ± sin φ∆ θ rot (12b)

We have used the two Eqs. (12a) and (12b) to run our overall simulations, for a variety of
hydrodynamic conditions and for a specialized aspect ratio. The results do not differ in any
significant manner. It is noticeable however that we gain on the calculation time when
simplifying the simulation conditions, by considering r θ , φ= 1.

The time interval Δt can be now eliminated from the hydrodynamic algorithm using Eq.(11a), so
that equations Eq. (6a) and (6b) become

1 2 2 2 Δθ2rot
∆ θ hyd (s+1 ; s ) = - γ̇ ( 2 ) (r e sin θ ( t s ) + cos θ(t s )) (13a)
re+ 1 2 Dθ rot

r 2e +1 sin 2θ ( t s ) sin 2 φ(t s ) Δθ2rot


∆ φhyd (s+1 ; s)= γ̇ ( 2 ) (13b)
r e −1 4 2 D θ rot

On the other hand, we have the peclet number α = γ̇ / Drot , this yield

2
1 2 2 2 Δθ rot
∆ θ hyd=−α ( 2 )(r e sin θ+cos θ) (14a)
r e +1 2
(r 2e +1)
∆ φhyd =α (14b)
8(r ¿ ¿ e 2−1)(sin 2θsin 2 φ) Δ θ 2rot ¿

In a numerical simulation the decision of a value for ∆ θ rot is dictated by technical criteria
towards establishing efficient simulation runs and negligible scatter. This choice determines
effectively a simulation time interval Δt which is inversely proportional to the diffusion
coefficient from Eq. (9) in this report. For the purpose of the present simulations we typically
take ∆ θ rot 0.03 radians, for a wide range of the Peclet number α. This value for ∆ θ rot may
however be varied for different experimental conditions and different peclet numbers.

For dilute solutions of the particles, it is possible to treat these processes separately since the
macromolecules are free from particle ̲ particle interactions that would some way or another

6
CHAPTER II

couple their rotation and translation diffusion. In this approximation a new Brownian random
variable, Δzc,tr is introduced that displaces the particle in either sense normal to the surface. We
dispense the simulation time step variable again in order to derive an expression for Δzc,tr as a
function of the basic simulation variable ∆ θ rot, namely:

2R
Δzc,tr = ± |∆ θ rot| (15)
3

The factor 2R/3 in Eq. (15) comes from the ratio of the Brownian translation and rotation
diffusion coefficients.

II.4 Colloidal particle algorithms in 2D-spatial frames for open pore


channels

In this section, we present the model for the numerical simulations for the dynamics and PDF
distributions of nearly spherical colloidal particles in open channels, beginning with the
algorithm of ellipsoidal particles. This simulation is done by making the aspect ratio
equal to 1.02. The mesopore system is hence modeled by two infinite solid plates which confine
the fluid flow along a Cartesian direction in the x-y plane parallel to the plates, however which
are sufficiently apart to model an open channel as in Fig. 4. The separation D between the plates
along the y axis is considered macroscopic in its dimensions and greater than the length L of the
ellipsoidal particle (L = 2R), R is the radius of a spherical particle, D >>L. The particles have the
freedom to move dynamically in the flow, in translational and rotational movements, in all
directions in the 3D space, under the combined impact of the Brownian and the hydrodynamic
forces, and of the diffusive collisions at the solid surface boundary.

Fig. 7. Schematic representation of the spherical macromolecular particles near to the


solid surface boundary of one of the two plates which confine the hydrodynamic flow of the fluid.

6
CHAPTER II

Our algorithm contains components that simulate both the hydrodynamic and the Brownian
movements in the bulk when the extremities of the spherical macromolecular particles do not
come into contact with the solid boundaries. The algorithm picks randomly between
hydrodynamic and Brownian events. It is essential, in any case, in this novel situation to develop
the algorithm to account for the collisions when they occur at the boundary surfaces, to test the
sequence of collisions that follow and their outcomes.

Different types of possible collisions may be distinguished, labeled here as A, B, C, and D, to


represent for all possible events in the modified algorithm. Furthermore, the boundary surface is
considered as atomically flat for the purpose of our simulations in this report. The ambiguities
arising from the simulation time steps, where time is a continuous variable, are eliminated here
since we work with ∆ θ rot, a basic simulation random variable.

II.4.1 For three-dimensional spatial frame (3D)

Case A: Corresponds to the situation when a random translation diffusion event brings the
particle extremities into contact with the surface because of the collision. We may then write

L
zcollision = =R (16a)
2

It is then conceivable to write the algorithm for a Brownian diffusion translation between the N
and N+ 1 events, as

zc (N + 1) = zcollision + e [zc(N) − zcollision] (16b)

For a given macromolecular species and given topography of a prepared surface, [28] the
collision due to the random diffusive processes might be thus characterized by a given e
restitution coefficient as in Eq.(16b). The event following the arrival of the particle into N+1
moves the particles into other configurations. A comparable algorithm may be given for a
collision arising due to the rotation diffusion around a solid angle determined from the angles θ
and φ since these processes are considered inherently similar.

Case B: The Brownian rotation diffusion about an angle θ leaves constant the position of the
center of mass between the N and N+1 events, so that zc(N) = zc(N+1). It follows for this case that
the angle of collision can be written as:

θcollition = sin−1 ¿ ¿ (17a)

The Brownian rotation algorithm between N and N+1 event, due to collisions, becomes then

θ(N+1) = θcollision +e [θ(N) - θcollision ] (17b)

6
CHAPTER II

Case C: Corresponds to the Brownian rotation about the angle φ that brings the particle
extremity into collision with the solid surface boundary. In this situation we have a similar
algorithm as for the random Brownian θ rotation, namely

φ collition = sin−1 ¿ ¿ (18a)

φ(N+1) = φ collision+e [φ(N) - φ collision] (18b)

The similarity of the sets of Eqs. (17) and Eqs. (18), for the Brownian events, is coherent with
the choice of Eq. (11a) and Eq. (12a) for the Brownian algorithms.

Case D: corresponds to the situation when the collision is because of hydrodynamic rotations
near the solid surface boundary. The particle in this situation, between the successive N and N+ 1
event turns about the contact point determined by the collision in-between, into a solid angle
described by the coordinate angles θ and φ

π
 φ(N+1) = φ(N) + é[ - φ(N)] (19a)
2
π
 θ(N+1) = θ(N) + é[ - θ(N)] (19b)
2
z c(N+1) = R | sinθ(N+1) sinφ(N+1) | (19c)

The restitution coefficient é characteristic of hydrodynamic events, is dealt with independently


from that for Brownian events, e, although both will be treated as random variables, in the
interval [0,1], for the purpose of our simulations.

In our simulations the dynamics for the possible collision events of the macromolecular
ellipsoidal particles, as in cases A, B, C and D, are introduced through a special algorithm code.
This enables us to follow the behavior of the macromolecular particle over N = 107 elementary
events due to stochastic Brownian and causal hydrodynamic movements for a single simulation
run.

II.4.2 For two-dimensional spatial frame (2D)

In a 2D-spatial frame, we just take the value of angle φ equal to π/2. Hence, the case A stay the
same, the case C, the Eq. (12b) and Eq. (14b) are eliminated, the Eq. (17a) become

z c (N )
θcollition = sin−1 [ ] (20a)
R

6
CHAPTER II

And the Eq. (19c) become

L
z c(N+1) = | sinθ(N+1) | (20b)
2

In our simulation, we take a necessary condition for zc aims to exclude all zc less than R:
If zc < R, we should add R.

6
CHAPTER II

Chapter III: Simulation results for spherical forms of


colloidal particles
The simulation results are displayed around the point nearest to the wall, over the interval 0.5≤ ξ
=zc/L≤ 2 and 0.5≤ ξ ≤ 4. As stated the simulations are carried out for an atomically flat surface
boundary at ξ= 0, which is used as the reference plane relating to the lowest material levels that a
particle extremity can touch in diffusive collisions.
We will consider simulating for particles in bulk solutions, distant from the surface boundary,
with the aspect ratios: re = 1.02, which corresponds to the interesting case of nearly spherical
particles. This estimation of re renders this particle hence helpful as a reference for our numerical
simulations. To circumscribe the present uncertainty as regards the topography of prepared solid
surfaces in these experiments, and the nature of the collision between the extremity of a
macromolecule and the surface, the simulations are done for different values of e and e'. They
are also presented for diverse values of α. The purpose is to calculate close to a solid surface the
stationary PDF functions P (ξ), a number of measurable quantities that depend on these PDF and
to discern their tendencies for low and high flow conditions.

III.1 Simulation results for restitution coefficients e = 1


In Fig. 8 (a) the Brownian simulations of the Jeffery PDF in 2D is given for a wide range of α
where ξ ≤ 2. The results are obtained from 107 hydrodynamic and stochastic events per simulation
run in a statistical ensemble of various runs. We assume that the restitution coefficients are e =1
and e' random. We noticed that as the peclet number α increases the PDF distribution increases
rapidly reaching higher values. However, for ξ ≥ 1.36, P (ξ) takes a nearly constant behavior for
all taken values of α.

(a) (b)

6
CHAPTER II

Fig. 8. The simulation results for the normalized spatial PDF near the solid surfaces, calculated for ∆ θ rot =0.03,
for different peclet number possibilities, as a function of ξ: (a) for a normalized distance 0.5 ≤ ξ ≤ 2; (b) for 0.5 ≤ ξ
≤ 4. The figures correspond to α = 0, 1, 10, 50, 100, for a typical Brownian restitution, e = 1 and random
hydrodynamic restitution.
In Fig.8 (b) we repeated the same simulation, but we extended the range of ξ to reach 4.
Comparing between the Fig. 8 (a) and (b), we observed that for ξ ≥ 2 the P (ξ) doesn’t show any
significant variation. To avoid the simulation time, we will limit our simulation till ξ = 2.

III.2 Simulation results for random restitution coefficients e

In Fig.9 we just changed the Brownian restitution coefficient e to become random. Then we
restarted the simulation under the same conditions. The P (ξ), for ξ ≤ 1.45, is approximately
linear for all α. For ξ ¿1.45, as α decreases P (ξ) becomes more constant. This behavior is
translated by the increase in the variation of the probability distribution due to the increase of the
hydrodynamic force. Which means that for higher hydrodynamic magnitudes (higher α) the
spherical particle is more likely to be found at higher altitudes.

Fig. 9. The simulated PDF, as in Fig. 8, calculated for the restitution coefficients e and e'random.

III.3 Simulation results for restitution coefficients e = 0

In Fig. 10, we observed an unexpected oscillation in the PDF distribution when Brownian
restitution coefficient e = 0. The PDF distribution defines the probability of finding the particle at
a specific altitude. Then any oscillation defies our expectations regarding the particle’s behavior.
It can be concluded that a peak at a certain altitude suggests that the particle is more preferably to

6
CHAPTER II

be found there. After more than one simulation, we noticed that the first descending peak
happened at ξ ≈1.28. Moreover, as α decreases the variation of this peak decreases.

Fig. 10. Typical simulation results, for the normalized PDF P (ξ), as a function of the normalized distance, ξ, of the
center of mass of the macromolecular spherical particles, measured from the reference plane. The P (ξ) are
calculated near the flat surface, for a fixed choice of the Brownian restitution coefficient e = 0 and for random e',
further for variable α.

In Fig. 11 (a)-(b)-(c), the simulation results for the normalized spatial PDF are calculated for null
Brownian restitution coefficient, for different peclet number possibilities. The figures correspond
to: (a) α = 0, (b) α = 10, (c) α = 100, where in each the hydrodynamic restitution is varied
through e' = 0, 0.5, 1. Although we varied α and e', the peak is still present in the PDF. Two
possibilities interfere in this event; the first possibility is that the particle elevated in altitude
whereas the second possibility is that this particle decreased in altitude. The presence of this peak
took place for e = 0, which means that the PDF’s oscillation is related to the null Brownian
restitution coefficient e. This null value denotes that the particle loses its energy upon its
collision with the surface; when a particle hits the surface, it stays and waits for the next event.
Based on the above, we can speculate that the spherical particle would have descended
downward that is we go for the second possibility.
We remark that around ξ = 0.5, there was a drastic increase upward in the PDF values depending
on the increase in the values of α.

In Fig. 11 (a) α is null and the hydrodynamic force doesn’t exist. In Fig. 11 (b) α is relatively small.
In these two situations, the Brownian force is dominant and hence controls the motion of the
particle giving the peak. However, in Fig. 11 (c) α is large and so is the hydrodynamic force
making it comparable to the Brownian. This equilibrium decreases the peak in the PDF.

6
CHAPTER II

(a) (b)

(c)

Fig. 11. Typical simulation results, for the normalized PDF P (ξ), as a function of the normalized distance, ξ, of the
center of mass of the macromolecular spherical particles, measured from the reference plane. The P (ξ) are
calculated near the flat surface, for a fixed choice of the Brownian restitution coefficient e = 0, and for variable e', at
conditions of low flow, α = 0, in Fig. 11 a; of relatively medium flow, α = 10, in Fig. 11 b; and of relatively high
flow, α = 100, in Fig. 11 c for comparison, where in each the hydrodynamic restitution is varied through e' = 0, 0.5,
1.

6
CHAPTER II

III.4 Simulation results for restitution coefficients e' = 0

In Fig. 12 the Brownian simulations of the Jeffery PDF in 2D is given for a wide range of α where
ξ ≤ 2. We assume that the restitution coefficients are e random and e' = 0. We noticed that as the
peclet number α increases the PDF distribution increases rapidly reaching higher values.
However, for ξ ≥ 1.6, P (ξ) takes a nearly constant behavior for all taken values of α. For the
preceding values of restitution coefficients, the PDF did not attain a significant shift such as that
when the Brownian restitution coefficient was zero. We regard that P (ξ) for α = 0 and that for α
= 1 behaved in the same rate different from those of the remaining peclet numbers whose
behaviors were also similar.

Fig. 12. The simulated PDF, as in Fig. 10, calculated for the hydrodynamic restitution coefficients e' = 0 and e
random.

In Fig. 13 (a)-(b)-(c), it is obvious that the PDF did not undergo any dramatic change to be noticed.
In these three simulations we fixed the Brownian restitution coefficient at one. Each simulation
corresponds to a specific value of α while manipulating the hydrodynamic restitution coefficient
e'. The significance of these simulations is the existence of a relationship between the PDF and
e'. This relationship was expressed as the following: when the hydrodynamic restitution
coefficient increases, the PDF also increases. In comparison with Fig. 13 (a) where α = 0 and (b)
where α = 10, the PDF ran almost similarly taking into consideration that P (ξ) for α = 0 did not
reach the same values for α = 10. This difference is clear due to the comparatively higher value
of α. The values of the PDF when α = 100 were almost identical and maximal for e' = 0.5 and 1
unlike that for e' = 0 which was saliently lower. The maximal values of the PDF recorded are
due to the relatively high values of the peclet number α.

6
CHAPTER II

(a) (b)

(c)

Fig. 13. Typical simulation results, for the normalized PDF P (ξ), as a function of the normalized distance, ξ, of the
center of mass of the macromolecular spherical particles, measured from the reference plane. The P (ξ) are
calculated near the flat surface, for a fixed choice of the Brownian restitution coefficient e = 1, and for variable e', at
conditions of low flow, α = 0, in Fig. 10 a; of relatively medium flow, α = 10, in Fig. 13 b; and of relatively high
flow, α = 100, in Fig. 13 c for comparison, where in each the hydrodynamic restitution is varied through e' = 0, 0.5,
1.

6
CHAPTER II

Chapter IV Modelling adhesion at pore boundaries


under diffusive
collisions
The possible adhesion (stickiness) of metallic and macromolecular nano-particles during a
collision with a solid surface depends on the loss of energy of the incident particle owing to the
collision. Adhesion should take place if this loss is so great that the particle has not enough
kinetic energy left to free itself away from the adhesive forces which act to keep it adhered to the
solid surface boundary. The adhesion or non-adhesion may be summarized by a basic relation
over the Newtonian restitution coefficient, e determined by

K . E f =e 2 K . E i (21)

K . E i and K . E f represent the kinetic energy of the particle before and after collision with the solid
surface boundary, so that

Adhesion occurs if K . Ef < EP


No Adhesion if K . Ef > EP

E P is the energy cost to liberate the particle during its collision rebound from the attractive Hamaker
potential V Pbinding the particle to solid surface boundary. V P, which has a strong local character,
depends on diverse factors, such as the material nature of the particle and the topography of the surface
boundary. This loss of energy depends on a number of factors, in particular the material nature of
the particle and the solid surface, the form of the particle, the rough surface topography on the
scale of the incident particle, and the Hamaker forces [29] that intervene between the particles
and the solid surfaces. The energy loss mechanisms include the transfer of the particle incident
kinetic energy into other modes such as in rotation, elastic deformation, and heat dissipation to
the impacted surface [30].

Since the notion of Newtonian restitution is of significant interest to our calculations, [31] built
up a theoretical model to evaluate this effect in terms of the diverse energy components (kinetic,
potential, deformation and adhesive), that contribute to the determination of the energy loss in a
collision. The reason for existing is to have the capacity to calculate the sticking probability and
consequent mobility of a given particle shape during its passage through a pore.

6
CHAPTER II

[31] derived a general relation for the Newtonian restitution coefficient,e, to be able to quantify
the possibility of mechanical adhesion of macromolecular and metallic nano particles on the
surface boundaries of pores through which the colloidal suspensions are flowing. This would
take place under the non-equilibrium conditions of the flow and supplements the known
thermodynamic adhesion processes. Moreover, Malak Ashkar master 2 student Derive an
equation for restitution coefficient,e, for metallic spherical particles near different types of
surface boundaries in different medium.

AR
e 2 ( 1−e 2 ) K . Ei− e ( e−1 ) =e2 ( Es , t −Es , p ) (22)
6 Zi

Where A is the Hamaker constant, R the radius of the particle, Zi the initial position of the
particle before the collision, Es , t = Vp, 0 – Vp, i is the surface energy on the way in for the particle
where Vp, 0 and Vp, i represent respectively the potential energies at equilibrium and final position,
Es , p = Vp, 0 – Vp, f is the surface energy on the way out for the particle where Vp, f is the classical
turning point; and ¿) represent the energy loss.

The possible adhesion of metallic and macromolecular nano-particles during a collision with a
solid surface depends on the loss of energy of the incident particle owing to the collision.

If and when the particle comes near the solid surface boundaries, due to hydrodynamic and
Brownian events preceding this latest position, the simulation algorithm can be switched on to
analyse whether or not the particle is in a position to submit to the consequences of the Hamaker
attraction. In this case the program algorithm between the Nth and the (N+1) th events will
decide whether or not the particle will collide with the surface boundary and calculate its
restitution coefficient using an equation such as Eq.22, and consequently whether or not it loses
sufficient energy to stick – adhere at the surface.

Adhesion should take place if the final kinetic energy of the particle is not enough to free it from
the adhesive forces which act to keep it at the solid surface boundary. E P=|V P ( N )| is the
energy necessary to liberate the particle during its collision rebound from the attractive Hamaker
potential V Pbinding the particle to solid surface boundary.

6
CHAPTER II

Using this procedure it is possible to chart the probabilities for the adhesion of different forms of
particles at surface boundaries in mesopores, and to design an algorithm to calculate by
numerical simulations the effective mobility for such particles during their flow inside
mesopores under non-equilibrium conditions.

Fig. 14. The simulation results for the restitution coefficient e, calculated for the gold spherical nanoparticle and
gold surface as a function of the initial position Zi.

In Fig. 14, done by the master 2 student Malak Ashkar, the simulations that were recorded taking
values of the restitution coefficient in the neighborhood of zero, show that the range of the initial
position varied between 3×10−4 and 4.75×10−4. As for our simulations in Fig. 11, we notice that
when the values of ξ were approaching 0.5 there can be a possibility for adhesion to take place.
Therefore, the particle will be very close to the surface. This means that Zi takes a very minimal
value in the neighborhood of e equal zero, this coincides with the findings of Fig. 14 done by M2
students.

6
CHAPTER II

Conclusion

Numerical simulations and algorithms are produced to study the dynamics of the spherical
nanoparticles in a two-dimensional spatial frame. Specifically, we consider for this situation of
an open channel pore for which we calculate the probability distribution functions (PDF) for the
spatial positions of these particles, in a streaming dilute solution close to the atomically flat solid
surfaces boundaries. We build up a theoretical model based in this case on the equations of
Jeffrey for the dynamics of solid particles in fluids and the molecular dynamics by mechanical
restitution for the diffusive collisions of the particles at the solid boundaries. Simulations are
accomplished to calculate the equilibrium PDF for spherical molecular particles in suspension in
a fluid under hydrodynamic flow. This is done for a wide range of α =γ˙/Drot, where γ˙ is the
constant shear rate of the linear hydrodynamic flow and Drot is the diffusion coefficient for the
molecular Brownian rotational motion; and for the possible range of restitution coefficients that
model interactions between molecular species and surface topographies, moreover a theoretical
model was used to calculate the restitution coefficient from basic physical principles and
quantify the energy balance during the process of a diffusive collision of a nano particle under
the influence of the repulsive forces due on one hand, and the attractive Hamaker forces acting
on the nano particle on the other, since determine under given physical conditions the outcome
for adhesion or not.

6
CHAPTER II

References

1. Atwi, A., A. Khater, and A. Hijazi, Three-dimensional Monte Carlo simulations of the dynamics of
macromolecular particles in solutions flowing in mesopores. Central European Journal of
Chemistry, 2010. 8(5): p. 1009.
2. Pozrikidis, C., Numerical Simulation of Cell Motion in Tube Flow. Annals of Biomedical
Engineering, 2005. 33(2): p. 165-178.
3. Stover, C.A. and C. Cohen, The motion of rodlike particles in the pressure-driven flow between
two flat plates. Rheologica Acta, 1990. 29(3): p. 192-203.
4. Rose, K., et al., Hydrodynamic interactions in metal rodlike-particle suspensions due to induced
charge electroosmosis. Vol. 79. 2009. 011402.
5. Subramanian, G. and D. L.  Koch Inertial effects on fibre motion in simple shear flow. Vol. 535.
2005. 383-414.
6. Zhang, Y., et al., Tethered DNA dynamics in shear flow. Vol. 130. 2009. 234902.
7. Park, J., et al., A cloud of rigid fibres sedimenting in a viscous fluid. Vol. 648. 2010. 351-362.
8. Maynard, A.D. and D.Y.H. Pui, Nanotechnology and occupational health: New technologies –
new challenges. Journal of Nanoparticle Research, 2007. 9(1): p. 1-3.
9. Kendall, B., et al., 8b - Kendall et al. (2010) SI. 2015.
10. T. Eckersley, S. and A. Rudin, Mechanism of Film Formation From Polymer Latexes. Vol. 61. 1990.
89-100.
11. Dobler, D.W. and D.N. Burt, Purchasing and Supply Management: Text and Cases. 1996:
McGraw-Hill.
12. Horn, R.G., J.N. Israelachvili, and F. Pribac, Measurement of the Deformation and Adhesion of
Solids in Contact. Vol. 115. 1987. 480-492.
13. Choi, G.Y., S. Kim, and A. Ulman, Adhesion Hysteresis Studies of Extracted Poly(dimethylsiloxane)
Using Contact Mechanics. Langmuir, 1997. 13(23): p. 6333-6338.
14. M., G.R., Emotion Labelling and Cognition1. Journal for the Theory of Social Behaviour, 1978.
8(2): p. 125-135.
15. Grzelczak, M., et al., Directed Self-Assembly of Nanoparticles. ACS Nano, 2010. 4(7): p. 3591-
3605.
16. Nie, Z., A. Petukhova, and E. Kumacheva, Properties and emerging applications of self-
assembled structures made from inorganic nanoparticles. Vol. 5. 2010. 15-25.
17. Yang, S.-M., et al., Synthesis and assembly of structured colloidal particles. Vol. 18. 2008. 2177-
2190.
18. Singh, N. and S. Bagchi, Applied ecology in India: Scope of science and policy to meet
contemporary environmental and socio-ecological challenges. Vol. 50. 2013. 4-14.
19. Chen, X. and S.S. Mao, Titanium Dioxide Nanomaterials:  Synthesis, Properties, Modifications,
and Applications. Chemical Reviews, 2007. 107(7): p. 2891-2959.

6
CHAPTER II

20. Bisquert, J., E. Palomares, and C.A. Quiñones, Effect of Energy Disorder in Interfacial Kinetics of
Dye-Sensitized Solar Cells with Organic Hole Transport Material. The Journal of Physical
Chemistry B, 2006. 110(39): p. 19406-19411.
21. Hsu , R. and P. Ganatos Gravitational and zero-drag motion of a spheroid adjacent to an inclined
plane at low Reynolds number. Vol. 268. 1994. 267-292.
22. Boeder, P., Über Strömungsdoppelbrechung. Zeitschrift für Physik A Hadrons and Nuclei, 1932.
75(3): p. 258.
23. Atwi, A., A. Khater, and A. Hijazi, Three dimensional simulation dynamics for the dilute colloidal
suspensions of rod-like polymer particles flowing in the bulk and near solid boundaries.
Macromolecular Research, 2013. 21: p. 1-11.
24. Hijazi, A. and A. Khater, Simulations of distribution functions for rod-like macromolecules in
linear flow near solid surfaces. Computational Materials Science, 2001. 20(2): p. 213.
25. Hijazi, A. and A. Khater, Brownian dynamics simulations of rigid rod-like macromolecular
particles flowing in bounded channels. Computational Materials Science, 2001. 22(3-4): p. 279.
26. Hijazi, A. and A. Khater, Boëder PDF Brownian simulations for macromolecular rod-like particles
near uneven solid surfaces. European Polymer Journal, 2008. 44(11): p. 3409.
27. Jeffery, G.B., The Motion of Ellipsoidal Particles Immersed in a Viscous Fluid. Proceedings of the
Royal Society of London. Series A, 1922. 102(715): p. 161-179.
28. Khater, A., Intensity reflection coefficients of atomically irregular surfaces in phonon scattering.
Europhys. Letters 1989. 2(7): p. 539.
29. Israelachvili, J.N., Intermolecular and Surface Forces, Third Edition. 3rd ed. 2011, San Diego,
United States of America: Academic Press Elsevier.
30. Sato, S., D.-R. Chen, and D.Y.H. Pui, Molecular Dynamics Study of Nanoparticle Collision with a
Surface – Implication to Nanoparticle Filtration. Aerosol and Air Quality Research, 2007. 7(3): p.
278-303.
31. Atwi, A., Theoretical and Numerical Calculations for the Dynamics of Colloidal Suspensions of
Molecular Particles in flowing solution inside Mesopores, in Physics2012, Université du Maine et
Université Libanaise. p. 320.

6
CHAPTER II

Table of figures

Figure 1 : Ellipsoidal particles........................................................................................................................3


Figure 2 : Spherical particles..........................................................................................................................3
Figure 3 : Rod-like particles...........................................................................................................................3
Figure 4 : A scanning electron microscopy (SEM) image of the titanium dioxide particles; with an aspect
ratio of 4.19...................................................................................................................................................3
Figure 5 : Coordinate system for ellipsoidal particle centered at the origin..........................................................7
Figure 6 : Calculated example of Jeffery’s different orbit constants C for re→ ∞................................................8
Figure 7 : Schematic representation of the spherical macromolecular particles near to the
solid surface boundary of one of the two plates which confine the hydrodynamic flow of the fluid…………………… 11
Figure 8 : The simulation results for the normalized spatial PDF near the solid surfaces, calculated for ∆ θ rot =0.03,
for different peclet number possibilities, as a function of ξ: (a) for a normalized distance 0.5 ≤ ξ ≤ 2; (b) for 0.5 ≤ ξ
≤ 4. The figures correspond to α = 0, 1, 10, 50, 100, for a typical Brownian restitution, e = 1 and random
hydrodynamic restitution...............................................................................................................................14
Figure 9 : The simulated PDF, as in Fig. 8, calculated for the restitution coefficients e and e'random.................15
Figure 10 : Typical simulation results, for the normalized PDF P (ξ), as a function of the normalized distance, ξ, of
the center of mass of the macromolecular spherical particles, measured from the reference plane. The P (ξ) are
calculated near the flat surface, for a fixed choice of the Brownian restitution coefficient e = 0 and for random e',
further for variable α.....................................................................................................................................16
Figure 11 : Typical simulation results, for the normalized PDF P (ξ), as a function of the normalized distance, ξ, of
the center of mass of the macromolecular spherical particles, measured from the reference plane. The P (ξ) are
calculated near the flat surface, for a fixed choice of the Brownian restitution coefficient e = 0, and for variable e', at
conditions of low flow, α = 0, in Fig. 11 a; of relatively medium flow, α = 10, in Fig. 11 b; and of relatively high
flow, α = 100, in Fig. 11 c for comparison, where in each the hydrodynamic restitution is varied through e' = 0, 0.5,
1.................................................................................................................................................................17
Figure 12: the simulated PDF, as in Fig. 10, calculated for the hydrodynamic restitution coefficients e' = 0 and e
random........................................................................................................................................................18
Figure 13 : Typical simulation results, for the normalized PDF P (ξ), as a function of the normalized distance, ξ, of
the center of mass of the macromolecular spherical particles, measured from the reference plane. The P (ξ) are
calculated near the flat surface, for a fixed choice of the Brownian restitution coefficient e = 1, and for variable e', at
conditions of low flow, α = 0, in Fig. 10 a; of relatively medium flow, α = 10, in Fig. 13 b; and of relatively high
flow, α = 100, in Fig. 13 c for comparison, where in each the hydrodynamic restitution is varied through e' = 0, 0.5,
1.................................................................................................................................................................19
Figure 14 : The simulation results for the restitution coefficient e, calculated for the gold spherical nanoparticle and
gold surface as a function of the initial position Zi...........................................................................................22

You might also like