rights / license: research collection in copyright - …...2016 i abstract the nano-scale world is...

165
Research Collection Doctoral Thesis Aggregation, Coalescence, and Gelation of Functional Nanoparticles Author(s): Jaquet, Baptiste P. H. Publication Date: 2016 Permanent Link: https://doi.org/10.3929/ethz-a-010793615 Rights / License: In Copyright - Non-Commercial Use Permitted This page was generated automatically upon download from the ETH Zurich Research Collection . For more information please consult the Terms of use . ETH Library

Upload: others

Post on 26-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

Research Collection

Doctoral Thesis

Aggregation, Coalescence, and Gelation of FunctionalNanoparticles

Author(s): Jaquet, Baptiste P. H.

Publication Date: 2016

Permanent Link: https://doi.org/10.3929/ethz-a-010793615

Rights / License: In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection. For moreinformation please consult the Terms of use.

ETH Library

Page 2: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

Diss. ETH Nr. 23763

Aggregation, Coalescence, and Gelation of Functional Nanoparticles

A thesis submitted to attain the degree of

DOCTOR OF SCIENCES of ETH ZURICH

(Dr. sc. ETH Zurich)

presented by

Baptiste P. H. Jaquet

MSc Chemical and Bioengineering, ETH Zurich

Born on 17.03.1989

Citizen of La Sagne, NE

Accepted on the recommendation of

Prof. Dr. Massimo Morbidelli, examiner

Prof. Dr. Chih-Jen Shih, co-examiner

2016

Page 3: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing
Page 4: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

I

Abstract

The nano-scale world is gaining more interest every day, with exciting developments potentially

allowing significant improvements in many current fields of technology. Nanoparticles, or colloids, are

hence a growing field of research, even though it has been known since close to a century. However,

the size scale in which nanoparticles comes with challenges coming from meso-scale physics together

with nano-scale physics (or chemistry). The industrial production of polymeric products often makes

use of emulsion polymerization, which naturally forms polymeric nanoparticles. These particles can

exhibit a growing number of functionalities, for example the ability of forming films. From an industrial

point of view, it is important to be able to control and understand what factors govern the aggregation

and gelation of the product.

In this work, the specificities of an industrial-type product were studied. It is composed of coalescing

acrylic polymer nanoparticles, on which surfaces poly(acrylic acid) was grown during the

polymerization process.

The effects of the polyelectrolyte layer are studied both experimentally and theoretically over two

chapters of the present work, showing clearly the underlying complexity of working with industrial

products. In order to investigate these effects, aggregation experiments and diverse characterization

techniques were used in order to devise an interaction potential relating the stability of the particles to

the pH and the synthesis recipe of the particles.

The effect of coalescence on stagnant aggregation and shear-induced gelation were studied

experimentally and theoretically over the following two chapters, using population balance equations

in order to understand the interaction between coalescence and stagnant aggregation. In the light of

Page 5: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

II

these findings, shear-induced gelation is shown to result from the competition between aggregation,

coalescence and breakup.

Finally, a rather unexpected surfactant adsorption mechanism on polyelectrolyte-covered particles

leading to their aggregation was revealed experimentally and was explained by molecular dynamics

simulations.

Page 6: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

III

Résumé

Le monde des échelles de grandeurs `nano` prend de plus en plus d’ampleur dans le temps, avec des

développements excitants pouvant rendre possible des améliorations significatives dans divers

domaines courants de la technologie. Les nanoparticules, ou colloïdes, sont donc un domaine de

recherche en pleine croissance, bien qu’il soit connu depuis près d’un siècle. Malgré cela, l’ordre de

grandeur typique des nanoparticules fait qu’elles viennent avec des problématiques spécifiques à la fois

à la physique ‘méso’ et la ‘nano-‘ physique (ou chimie). Cela fait qu’une compréhension quantitative

des facteurs gouvernant leur physique est loin d’être achevée. La production industrielle de polymères

fait souvent recours à la polymérisation en émulsion, qui par nature génère des nanoparticules de

polymère. Ces nanoparticules peuvent exhiber un nombre de fonctionnalisation constamment croissant,

par exemple la capacité de former des films lorsque la dispersion est séchée. D’un point de vue

industriel, il est essentiel de pouvoir comprendre, prévoir et contrôler quels facteurs vont mener à une

agrégation et une gélation du produit.

Dans le présent travail, les spécificités d’un produit de type industriel sont étudiées. Ce produit est

composé de nanoparticules acryliques pouvant coalescer, sur la surface desquelles des chaines de

poly(acide acrylique) ont été crûes durant le procédé de polymérisation.

Les effets de la couche de polyéléctrolyte sont tout d’abord étudiés sur deux chapitres du présent travail,

à la fois d’un côté expérimental et d’un côté théorique. Ces deux chapitres révèlent clairement la

complexité inhérente aux produits industriels. Afin d’étudier ces effets, des expériences d’agrégation et

de caractérisation variées ont étés appliquées afin de mettre au point un modèle d’interaction liant la

stabilité colloïdale de ces particule au pH et à leur recette de synthèse.

Page 7: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

IV

Les effets de la coalescence sur l’agrégation stagnante et induite par cisaillement sont ensuite étudiés

sur les deux chapitres suivants. Tout d’abord, un modèle de bilan de population a été développé afin de

prendre en compte les effets de la coalescence sur l’agrégation stagnante. A la lumière de ce

développement, la gélation induite par cisaillement est expliquée et comprise comme étant le résultat

d’une compétition entre l’agrégation, la coalescence et le cassage des agrégats formés.

Enfin, un mécanisme contre-intuitif d’adsorption inversée de surfactant ionique sur la surface de

particules couvertes de polyéléctrolyte menant à leur agrégation est démontré expérimentalement et

expliqué par simulation de dynamique moléculaire.

Page 8: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

V

Acknowledgements

First of all, I want to thank Prof. Massimo Morbidelli for giving me the opportunity to work in his

group. He allowed me to grow as a scientist and as an individual. I enjoyed the possibility of conducting

my research independently and with his confidence. I want to thank the other persons that made this

work possible. Of course, I want to thank Dr. Hua Wu for his support and suggestions, as well as his

expertise.

I want to thank Prof. Chih-Jen Shih for accepting to be my co-examiner, for his availability and

comprehension.

Many thanks to BASF SE, and in particular to Dr. Bernd Reck, Dr Volodymyr Boyko, Dr. Immanuel

Willerich, Dr. Oliver Labisch and Dr. Roelof Balk

A great thanks of course goes to Prof. Marco Lattuada and Prof. Giuseppe Storti. Their contribution to

the model development of the chapter 4 of this thesis cannot be stressed enough.

I want to thank very gratefully Dr. Stefano Lazzari for the work we did together and the support both

scientifically and morally. A great thank also goes to Dr. Tommaso Casalini for his invaluable help in

proving some intuitions are sometimes right. I want to thank Prof. Miroslav Soos for his support and

expertise in the latest stages of this work; you helped me a lot, taking some of your precious time for

me.

I want of course to thank my master, research projects, and “Hilfsassistenten” for their invaluable help

in following my experimental ideas and trusting me. In particular, I want to thank Kevin Maurer,

Antoine Klaue, Luca Colonna, Guido Zichittella, Gabriele Colombo, Daniel Balderas and Marc Siggel

for their help.

Page 9: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

VI

Thanks to Bastian Brand, to Alberto Cingolani, and Thomas Villiger for the helpful discussions and

experimental “tips” and explanations. I also want to thank all my friends; in particular David Pfister,

Rushd Khalaf, Lucrèce Nicoud, Marta Owczarz, Vincent Diederich, Bastian Brand, Marco Furlan,

Antoine Klaue, Anna Beltzung, Thomas Villiger, and Fabian Steinebach, and all of the Morbidelli group

for all the good times inside and outside the lab.

Un merci spécial va à mes parents, ainsi qu’à mes amis de Suisse Romande, pour m’avoir soutenu

pendant les bons et les mauvais jours de cette aventure. Merci d’avoir compris que mes absences

régulières ne sont pas le signe que je vous oublie.

Ein grosses Dankeschön geht auch selbstverständlich zu Caroline. Danke, dass du mich unterstützt und

auf mich gewartet hast während der letzten Periode von meinem Doktorat. Und auch Danke, dass du

mit mir die schönen Zeiten des Lebens geniesst.

Page 10: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

VII

Table of Contents

Abstract I

Résumé III

Acknowledgements ................................................................................................................................ V

Table of Contents ................................................................................................................................ VII

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

1.1 Aggregation of industrial functional nanoparticles ....................................................................... 2

1.2 Aggregation, Gelation, and Shear ................................................................................................. 2

1.3 Coalescence ................................................................................................................................... 4

1.4 Scope of the present work ............................................................................................................. 4

Chapter 2 Stabilization of Polymer Colloid Dispersions with pH-sensitive poly(Acrylic Acid) brushes7

2.1 Introduction ................................................................................................................................... 7

2.2 Experimental section ................................................................................................................. 9

2.2.1 The Colloidal Systems ........................................................................................................... 9

2.2.2 Determination of the Fuchs Stability Ratio W ..................................................................... 10

2.3 Results and discussion ................................................................................................................ 14

2.3.1 Particle size and PAA brush conformation behavior ........................................................... 14

2.3.2 Stability of P1 Latex ............................................................................................................ 19

2.3.3 Contributions of PAA Brushes to Colloidal Stability .......................................................... 20

2.4 Concluding remarks .................................................................................................................... 25

Chapter 3 Investigation of the Steric Stabilization of Dispersions with pH-sensitive poly(Acrylic Acid) brushes .......................................................................................................................................... 29

3.1 Introduction ................................................................................................................................. 29

3.2 Materials and Methods ................................................................................................................ 32

3.2.1 The latexes ........................................................................................................................... 32

3.2.2 Characterization of the particles .......................................................................................... 32

3.2.3 Aggregation Experiments .................................................................................................... 33

3.2.4 Simulation of the aggregation curves ................................................................................... 34

3.2.5 Modeling of the interaction potential ................................................................................... 34

3.3 Results and Discussion ............................................................................................................... 35

3.3.1 Characterization of the particles .......................................................................................... 35

3.3.2 Aggregation experiments and the Fuchs stability ratio ........................................................ 38

3.3.3 Steric model ......................................................................................................................... 40

Page 11: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

VIII

3.4 Concluding remarks .................................................................................................................... 42

Chapter 4 Interplay between Aggregation and Coalescence of Polymeric Particles: Experimental and Modeling Insights ................................................................................................................................ 45

4.1 Introduction ................................................................................................................................. 45

4.2 Materials and methods ................................................................................................................ 47

4.2.1 Materials .............................................................................................................................. 47

4.2.2 Particles synthesis ................................................................................................................ 47

4.2.3 Light Scattering .................................................................................................................... 48

4.2.4 Differential Scanning Calorimetry ....................................................................................... 49

4.3 Model development .................................................................................................................... 49

4.3.1 Population balance equations ............................................................................................... 49

4.3.2 Calculation of hR t , gR t and S q,t ............................................................ 53

4.4 Numerical Solution ..................................................................................................................... 55

4.5 Results and discussions ............................................................................................................... 56

4.5.1. Particle characterization and experimental conditions ........................................................ 56

4.5.2 Aggregation at room temperature: gR t , hR t and S q,t .............................. 57

4.5.3 Aggregation at higher temperatures ..................................................................................... 67

4.6 Conclusions ................................................................................................................................. 69

Chapter 5 Effects of Coalescence on the Shear-Induced Gelation of Colloids ................................... 73

5.1. Introduction ................................................................................................................................ 73

5.2. Materials and methods ............................................................................................................... 76

5.2.1 Materials .............................................................................................................................. 76

5.2.2 Latex synthesis and characterization .................................................................................... 76

5.2.3 Critical coagulation concentration determination ................................................................ 77

5.2.4 Zeta potential measurement ................................................................................................. 78

5.2.5 Colloidal titrations ............................................................................................................... 78

5.2.6 Couette-flow gelation experiments ...................................................................................... 78

5.2.7 Stirred-Tank DLCA experiments ......................................................................................... 79

5.3.1 Effect of temperature on stagnant aggregation .................................................................... 81

5.3.2 Gelation Curves ................................................................................................................... 83

5.3.3 Effect of salt concentration and temperature on gelation time ............................................. 85

5.3.4 Effect of temperature on cluster breakup ............................................................................. 89

5.4. Conclusion ................................................................................................................................. 95

Chapter 6 The Counter-Intuitive Aggregation of poly(Acrylic Acid)-Grafted Nanoparticles after

Surfactant Addition 97 6.1 Introduction ................................................................................................................................. 97

6.2 Materials and Methods ................................................................................................................ 98

Page 12: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

IX

6.2.1 Particles synthesis ................................................................................................................ 98

6.2.2 Particles characterization ..................................................................................................... 98

6.2.3 Aggregation experiments ..................................................................................................... 99

6.2.4 Computational methods ....................................................................................................... 99

6.3 Results and discussion .............................................................................................................. 103

6.3.1 Latex aggregation kinetics ................................................................................................. 103

6.3.3 Molecular dynamics simulations ....................................................................................... 107

6.4 Concluding remarks .................................................................................................................. 110

Chapter 7 Concluding remarks .......................................................................................................... 113

Chapter 8 Appendix ........................................................................................................................... 115

8.1 Experimental proof for the coalescence between particles ....................................................... 115

8.2 - time-evolution of the average fractal dimension .................................................................... 117

8.3 – Derivation of average quantities h,effR x, t and gR x,t ................................................. 120

8.3.1 Calculation of h,effR x,t ................................................................................................ 120

8.3.2 Calculation of gR x,t .................................................................................................... 127

8.4 Derivation of the discretized Equations .................................................................................... 128

8.5 Parameter Values employed in the simulations ........................................................................ 131

8.6 Measured Fractal dimensions .................................................................................................... 134

8.7 Results at higher temperatures .................................................................................................. 135

8.8 Symbol List for chapter 4 .......................................................................................................... 137

Page 13: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

X

Page 14: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

1

Chapter 1

Introduction

Colloidal and nanoparticles in general play a growingly important role in the current active fields of

research, as well a production. In fact, an important number of commercial polymeric products are used

as dispersions. These cover the fields of paints and coatings, lacquers, adhesives, paper industry, textiles

and leathers[1-3] In order to manufacture these products, emulsion polymerization is commonly used.

Colloidal suspensions are not limited to plastics production, as nanoparticles dispersions nowadays

offer a constantly growing scope of new applications, ranging from photovoltaics [4, 5] to the medical

field [6-8]. Because of the size of colloidal particle (1-1000 nm typically), they are exhibiting physical

and dynamical behaviors common with the macro- scale, as well as with the chemical size scale[9, 10].

These features offer a great deal of complexity and interest, making it uneasy to predict how a colloidal

system will react to an external stimuli, even though colloids have been studied since close to a century

[11, 12].

Great progress has been made in the understanding of colloidal systems since the initial development

of this field [13]. However, most of the theories and mathematical expressions devised in order to

describe the dynamics of dispersions are centered on ideal systems [14]. The use of typical mathematical

models to describe the manufacturing of real-world, complex colloidal dispersions usually is inefficient

at predicting and understanding quantitatively the influences of the control parameters on the dynamics

of the products.

Page 15: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

2

1.1 Aggregation of industrial functional nanoparticles

The use of emulsion polymerization to form copolymeric latexes is widely spread in industrial

applications. The quality control of latexes used for high-end applications, such as film-forming

adhesive latex [15], in which optical properties must be closely controlled, is essential. In order to ensure

a reproducibly high quality product, while maximizing productivity, a close control on the aggregation

kinetics is necessary. The use of stabilizers is commonly found in polymer production. However, in the

case of complex polymeric products, where up to ten different monomers are combined and

polymerized together, having a clear understanding of what factors will influence the colloidal stability

is dramatically complex. The main feature of the particles influencing the colloidal stability is evidently

the state of the particles’ surface. Charge density, composition, softness of the interface controls the

dynamics of nanoparticle interactions. In the present work, the case of industrial transparent adhesive

nanoparticle is studied. The industrial production of this high-end commercial product features

significant challenges, coming mainly from two features; the particles are made of a low-Tg copolymer,

which implies that the particles are melting into each other upon contact. On the other hand, in order to

ensure a reasonable commercial production, they exhibit a complex functionalized surface, composed

of a polyelectrolyte “brush-like” layer. This is pH-responsive, and its growth and morphological

properties are poorly understood, as the polymerization method does not allow for an accurate control

of its formation. The challenge here is to clarify the link between the recipe and the morphology, and

then between the morphology and the colloidal behavior.

1.2 Aggregation, Gelation, and Shear

The processes leading to aggregation of colloidal particles are relatively well-known [16]. The

lyophobic particles are stable in suspension as long as a sufficient energy barrier has to be overcome in

order for the surfaces of two particles to contact. The energy allowing to cross the energy barrier can

be supplied by Brownian motion in the case of stagnant aggregation [17], or by an external force field.

The latter case corresponds typically to shear-induced aggregation [18, 19].

When colloidal particles aggregate, whether in stagnant conditions or under the influence of an external

force field, it has been observed that they form fractal structures. Random fractal structures are self-

Page 16: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

3

similar (in size scale), and random in nature. They are related to the well-known geometrically

spectacular Mandelbrot fractal set [20], but their random nature do not imply such a degree of organized

geometric self-similarity. The fractal nature of colloidal aggregates imply that the size of an aggregate

is related to its mass by a power-law. This implies that colloidal aggregates are intrinsically porous.

Hence, the volume occupied by the aggregates increases non-linearly to the mass of the aggregates.

This means that along an aggregation process, the occupied volume fraction increases. When the whole

volume is occupied, the aggregates “percolate”, forming a network of aggregated particles, not allowing

the movement of the whole suspension anymore. This process is called gelation, and is macroscopically

recognized as a liquid-solid phase transition of the suspension[21]. Controlled gelation has been

successfully applied in order to form porous monoliths, with applications in chromatography. Gelation

can also be an undesired phenomenon; for example in the industrial production of polymeric

nanoparticles.

When polymeric particles are manufactured industrially or in lab scale, they get stirred, pumped,

filtered, heat-exchanged etc. Many of the unit operations taking place in a typical manufacturing process

imply a significant shear being applied to the latexes. Shear is an external force field allowing to bring

kinetic energy in the dispersion. If the shear rate is sufficient, it may allow nanoparticles which are

stable in stagnant conditions to aggregate. Shear-induced aggregation has been extensively studied in

the last decades, allowing to understand the mechanisms governing it. It has been revealed that the rate

of aggregation under shear is the results of the competition between the shear rate and the inter-particle

interaction potential [22].

Breakup of colloidal aggregates is one of these relevant mechanisms. While only rarely observed under

stagnant aggregation, because the adhesion energy between colloidal particles is typically much higher

than the thermal agitation present in the dispersion, breakup of aggregates caps up the size that

aggregates can physically reach under shear. The breakup rate results of the competition between the

size of the aggregate, on the shear rate (ultimately hydrodynamic stress), and on the adhesion energy of

the clusters [23].

Page 17: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

4

Under shear, the evolution of aggregate size is significantly different from what is observed in stagnant

conditions. Where stagnant aggregation typically generates a monomodal size distribution function,

shear-induced aggregation is self-accelerating, generating very large aggregates which cohabit with

primary particles or small clusters. This leads to possible “explosive gelation”, during which the

viscosity of the suspension –after an induction time- sharply rises to infinity, producing a solid gel piece

in small durations.

1.3 Coalescence

This work is focused on film-forming particles. Their physical properties are tuned in order to ensure

that the particles will fuse into each other upon drying of the dispersion. In order to obtain this result, it

is necessary that the particles’ material exhibits a finite viscosity. This means that the glass transition

temperature Tg of the material must be located below the film-forming temperature (mostly room

temperature).

Coalescence has a significant impact on both stagnant and shear-induced aggregation and gelation. In

fact, under stagnant conditions, coalescence will allow the clusters to get more compact, thus increasing

their fractal dimension, and to decrease their occupied volume fraction. This evidently is reflected in

longer gelation times. Similarly under shear conditions, coalescence allows the particles to

interpenetrate. This of course decreases the occupied volume fraction, delaying the gelation. But,

simultaneously, interpenetration of particles reinforce significantly their adhesion energy, thus

reinforcing the clusters, which are then able to grow to larger sizes before breaking apart under shear

stress. The competition between these two effects of coalescence has been reported previously [24], but

never systematically studied.

1.4 Scope of the present work

In the present work, the general case of the aggregation and gelation polyelectrolyte-stabilized, film-

forming nanoparticles is studied. In chapter 2, the initial understanding of the mechanics of poly(acrylic

acid)-covered industrial latex stagnant aggregation is studied and allows to gain qualitative

understanding of the relevant parameters controlling the stabilization of such latexes.

Page 18: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

5

In chapter 3, a steric interaction model is used in a tentative to use the aggregation properties in order

to characterize the properties of the polyelectrolyte layer of industrial systems.

In chapter 4, the effect of coalescence on diffusion-limited cluster aggregation is studied, and a

population balance equation model is devised in order to predict the size evolution of clusters formed

by ad-hoc synthesized acrylic particles with different Tg.

In chapter 5, the same particles are used in order to understand the effect of coalescence on shear-

induced aggregation and gelation using rheology.

In chapter 6, a more industrial problem is presented. The presence of surfactant, commonly used during

emulsion polymerization, is shown to have a counterintuitive effect on poly(acrylic acid)-covered

particles. In fact, head-on adsorption of SDS mediated by salt bridges is shown to occur by molecular

dynamics, and to lead to surfactant-induced aggregation in pH ranges in which the particles would not

aggregate without SDS.

Page 19: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

6

Page 20: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

7

Chapter 2

Stabilization of Polymer Colloid Dispersions with pH-sensitive poly(Acrylic Acid) Brushes

2.1 Introduction

When linear polyelectrolyte (PE) chains are densely grafted on a surface, they form so-called surface

PE brushes [25]. The strong electrostatic interactions among the highly charged chains within the

brushes lead to substantially different properties with respect to surface monolayers of uncharged

macromolecular chains. Thus, great attention has been given recently to these systems [26-33].

Weak (or so-called annealing) PEs such as poly(acrylic acid) (PAA) and poly(methacrylic acid)

(PMAA), when exposed to aqueous media, may undergo conformational changes, depending on the

system pH and salinity. Under alkaline conditions, due to deprotonation of their carboxylic groups, they

are highly charged and hydrophilic, leading to stretching of the backbone and good solubility. Instead,

under strongly acidic conditions, because of protonation of the carboxylic groups, they become weakly

charged and more hydrophobic, with the backbone in a collapsed state, thus less soluble in water.

Accordingly, weak PEs can be used to form brushes on polymeric surfaces, which respond to pH

changes in the environment, leading to potential applications in various areas such as nano-scale

actuators [34], membrane modifiers for pH-controlled permeation [28, 35], charge-driven reversible

polymer and protein adsorption [28, 36-38], nanoparticle stabilization [39], etc.

In this work, we focus on the stability behavior of colloidal particles grafted with PAA brushes. It

is known that weak PEs like PAA and PMAA when grafted on the surface change substantially their

ionization behavior with respect to their un-grafted state in water. Because of the strong electrostatic

Page 21: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

8

attraction among the mobile counterions in solution and the opposite charges on the brushes, the

concentration of the mobile counterions in the brush-occupied region increases towards the particle

surface [40, 41]. It follows that the total local charge density (polymer and mobile ions) is substantially

lower in the inner than in the outer region of the brushes. On this process is based, the so-called local

electroneutrality approximation (LEA) [40, 42], which states that the counterion concentration in the

inner region is approximately equal to that of the charged groups on the brushes. As a consequence, the

conformation of the brushes in the inner region is different from that in the outer one [43]. In addition,

experimental observations reveal that the apparent pKa of surface-grafted PMAA differs from that of

PMAA in solution and depends also on both brush thickness and density [44]. In addition to the

electrostatic repulsion arising from the particle charge, grafted hydrophilic polymer chains are known

to stabilize colloidal particles via steric interactions [45]. These steric interactions arise from the mixing

free energy of the polymer chains grafted on the surface [46]; if the dispersant is a good solvent for the

polymer composing the brushes, overlapping two grafted layers (during the aggregation process for

example) will lead to a repulsive effect. The consequence of all the above theoretical and experimental

results is that the contribution of the PAA brushes to the stability of colloidal particles is not a trivial

function of the AA concentration on the particle surface.

In this chapter we investigate experimentally the effect of surface-grafted PAA brushes on the

stability of polymer colloids at different pH values, i.e., under their different ionization and

conformational states. For this, we considered colloidal particles with different thickness of the PAA

brushes and measured their Fuchs stability ratio, W, based on the doublet formation kinetics, at different

pH values, as a function of electrolyte concentration. In the sequel, we first describe and characterize

three types of colloidal particles with different PAA brushes, and measure their W values. Next, we

compare the measured W values of the three colloids, and discuss them with respect to the expected

conformation of the PAA brushes.

Page 22: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

9

2.2 Experimental section

2.2.1 The Colloidal Systems

The polymer colloids used in this work are three types of butylacrylate-mehtylmathacrylate-acrylic acid

copolymer latexes, referred to as P1, P2 and P3 in the following, respectively, synthesized ad-hoc by

BASF SE (Ludwigshafen, Germany) through emulsion polymerization, using Na2S2O8 as initiator and

sodium dodecyl sulfate as emulsifier. The characteristics of the three latexes are reported in Table 1. It

is seen that the main difference is the amount of acrylic acid (AA): 0%, 1% and 2% for P1, P2 and P3

latexes, respectively. Thus, particles of P1 latex possess only the sulfate ( 3-OSO ) fixed charges, while

those of P2 and P3 latexes carry both sulfate and carboxylic ( -COO ) charges.

The SDS (sodium dodecyl sulfate) surfactant present in the original latexes, as well as residual

initiator, soluble oligomers and all other possible electrolytes, have been removed completely by ion

exchange resin (Dowex MR-3, Sigma-Aldrich), according to a procedure reported elsewhere [47]. After

cleaning, the conductivity of the mother liquor is <10 μS/cm and its surface tension is close to that of

deionized water, indicating that the colloidal systems are free of SDS and any other electrolyte. The

particle volume fraction for the three latexes after cleaning is about =20%, with pH around 2.7. The

cleaning also leads to replacement of the counterions of the fixed charges (e.g., 3-OSO and -COO )

from Na+ to H+ (the Na+ concentration after cleaning measured by atomic absorption spectrometry is

equal to about 3 ppm).

The particles characterization has been performed using dynamic and static light scattering (DLS

and SLS). The SLS and DLS measurements were carried out using a BI-200SM goniometer system

(Brookhaven Instruments, USA), equipped with a solid-state laser, Ventus LP532 (Laser Quantum,

U.K.) of wavelength 0=532 nm, as the light source, in the scattering angle range of 15 to 150. When

used for DLS, the angle was set to 90. The reported results are averages of at least 2 independent series

of at least 3 measurements. The nominal radius of the particles, Rp, has been evaluated by fitting the

Page 23: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

10

measured form factor from the SLS measurements, P(q), defined as ( ) ( ) / (0)P q I q I (where I is the

scattered intensity, q is the magnitude of the scattering wavevector, defined as

0 04 / sin[ / 2]q n with n0, 0 and the refractive index of water, the wavelength of the incident

light, and the scattering angle, respectively), with the Rayleigh-Debye-Gans (RDG) theory for spheres.

The nominal size of the primary particles after cleaning has been obtained by fitting the form factor,

P(q) measured under natural pH and 10 mM NaCl, with RDG theory. All the obtained Rp values are

reported in Table 2.1.

Table 2.1 Characteristics of the three latexes used in this chapter

Latex name Composition (weight ratio)a pH Rp

P1 BA (50) : MMA (50) + 0.7% SDS 5.1 84.1 nm

P2 BA (49.5) : MMA (49.5) : AA (1) + 0.7% SDS 4.9 81.0 nm

P3 BA (49) : MMA (49) : AA (2) + 0.7% SDS 4.2 90.0 nm

a—BA: n-butyl acrylate; MMA: methyl methacrylate; AA: acrylic acid; SDS: sodium dodecyl sulfate

2.2.2 Determination of the Fuchs Stability Ratio W

The contribution of different surface groups to the colloidal stability has been quantified by measuring

the Fuchs stability ratio (W) values under different conditions. To this aim, we have conducted all

experiments for the W measurements at the particle volume fraction =5.0% and different pH, using

NaCl and H2SO4 as the destabilizers.

The principle: The W measurement is based on the observation of the doublet formation kinetics

as discussed in previous work [17, 48]. In particular, at the very initial stage of the aggregation, the

main process is conversion of primary particles to doublets, and therefore the kinetics of primary

Page 24: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

11

particle consumption can be approximated as follows:

211,1 1

dNK N

dt (2.1)

where N1 and K1,1 are the primary particle number concentration and doublet formation rate constant,

respectively. Introduction of conversion of primary particles to doublets, )0 1 0( /x N N N , into Eq.

2.1, where N0 (3p3 / [4 ]R ) is the initial number concentration of primary particles, yields:

1,1 01

xK N t

x

(2.2)

Thus, from the measured x values as a function of time, t, the slope of the / (1 )x x vs t line leads to

the estimation of K1,1. Deviations from linear behavior of this line indicate that doublets or larger

clusters also contribute to the aggregation process, and the experiments have to be repeated at lower

conversion values.

From the data above, we can obtain the Fuchs stability ratio, W, based on the following

relationship:

1,1/BW K K (2.3)

where KB (=8kT/[3μl]), with k the Boltzmann constant, T the absolute temperature, and l the viscosity

of the dispersant, is the so-called Smoluchowski rate constant, defining the diffusion-controlled

collision rate between particles of equal size.

It is clear that to obtain K1,1 from Eq. 2.2, we need measure the x value as a function of t. This is

done by taking proper samples of the aggregating system and making SLS measurements. According

to the RDG theory [49], the scattered light intensity I(q) for a system containing only primary particles

Page 25: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

12

and doublets can be expressed as:

1 1 2 2

1 2

( ) ( ) 4 ( )

(0)

I q N P q N P q

I N N

(2.4)

where N2 is the number concentration of doublets, and P1(q) and P2(q) are the form factor of the primary

particles and doublets, respectively. Within the RDG theory, the expression for P1(q) is given by:

2

p p p1 3

p

sin( ) ( ) cos( )( ) 9

( )

qR qR qRP q

qR

(2.5)

The expression for P2(q) depends on whether coalescence occurs or not between the two aggregated

particles in the doublet [48]. For the present very soft acrylate copolymer particles (Tg≈15 ̊ C), we have

confirmed experimentally, as shown in the annex 8.1, that coalescence between two particles does

occur. Then, considering two primary particles coalescing and forming a new sphere, P2(q) is given by

21/3 1/3 1/3p p p

2 1/3 3p

sin(2 ) (2 )cos(2 )( ) 9

(2 )

qR qR qRP q

qR

(2.6)

By substituting Eqs 2.5 and 2.6 into Eq. 2.4 and considering that 1 0(1 )N x N= - and 2 0 / 2N xN= ,

we have:

1

( )1 ( )

(0) ( ) 1

I q xF q

I P q x= -

+ (2.7)

where

3 3 3

p p p

p p p

sin( 2 ) (( 2 )cos( 2 )1( ) 2

2 sin( ) ( )cos( )

qR qR qRF q

qR qR qR

(2.8)

Therefore, the plot, 1( ) / [ (0) ( )]I q I P q versus F(q), is a straight line, and its slope yields the value of

the conversion x.

The experimental procedure: We first mix the latex with a proper electrolyte solution so as to reach

Page 26: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

13

a desired particle volume fraction, electrolyte concentration and pH (sulfuric acid and sodium hydroxide

were used to tune the pH) to start the aggregation under stagnant conditions. Then, we take samples

from the aggregation system at different times and dilute them immediately in Milli-Q water for

quenching the aggregation and performing the SLS measurements. In addition, for each aggregating

system, pH was measured after the aggregation experiments.

Figure 2.1 Examples of the x/(1-x) vs. t plot for the estimate of K1,1 based on Eq. 2.2, in the case

of P2 latex using H2SO4 as the destabilizer at three concentrations: 0.116 mol/L (), 0.123 mol/L ()

and 0.130 mol/L ().

Fig. 2.1 shows typical plots of / (1 )x x- vs. t at the very initial stage of the aggregation process

for the P2 latex at =5%, using H2SO4 as the destabilizer at Cs =0.116, 0.123 and 0.130 mol/L. They all

are straight lines, whose slopes lead to estimate of the corresponding K1,1=3.19×10-24, 1.30×10-23 and

3.23×10-23 m3/s, and then, based on Eq. 2.3, to the Fuchs stability ratio values of W=3.86×106, 9.51×105,

and 3.82×105, respectively.

Page 27: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

14

2.3 Results and discussion

2.3.1 Particle size and PAA brush conformation behavior

To verify the presence of the PAA brushes on the particles of P2 and P3 latexes with 1% and 2%

AA, respectively, we have measured the Rh value of the particles by DLS as a function of pH at a

controlled ionic strength of 25 mM, and the results are shown in Fig. 2.2. It is seen that for P3 latex the

Rh value increases asymptotically as pH increases and the difference between the smallest and largest

Rh values is about 8 nm. These results allow us to conclude that the PAA brushes are indeed attached

to the particle surface of P3 latex. In the case of P2 latex, the Rh value increases with pH only marginally,

about 2 nm, indicating much shorter PAA brushes. Since the thickness of the brushes of P2 latex with

1% AA is only 1/4 of P3 latex with 2% AA, the brush density should be larger for P2 than for P3.

Figure 2.2 The hydrodynamic radius, Rh, of the primary particles as a function of pH, determined by

DLS, for P2 and P3 latexes after cleaning.

To further confirm the presence of the PAA brushes on the particle surface, we have measured the

form factor, P(q) by SLS, at different pH values, and the results are shown in Figs 2.3a and 2.3b for P2

Page 28: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

15

and P3 latexes, respectively. The difference in the P(q) curves obtained at different pH is very small; it

follows that the radius of gyration (Rg) of the particles at different pH is also practically the same. Note

that for P3, the Rg value does increase about 1 nm with pH in the given pH range, but this is certainly

incomparable with the Rh variations in Fig. 2.2. Therefore, the almost constant Rg and the significant

increase in Rh with pH together indicate that the growth in Rh results only from a layer of polymer on

the particle surface, instead of swelling of the whole particles. This conclusion arises because from the

definition of Rg: 2 2g i iR M r m , where M is the total mass, im the individual mass and ir is the

distance of im from the center, if the whole particle swells as the Rh variation indicates, due to increase

in ri, Rg must vary proportionally with Rh. Instead, since the mass of the particle is dominated by the

P(BA-MMA), the variation in the inertia by the 1 or 2% mass of PAA is really negligible..

Results of zeta-potential measurements showed no clear trend in the pH range, 5 to 9, where

ionization and conformation variations of the PAA brushes occur. This is due to the displacement of

the slipping plane induced by the transition from the collapsed to stretched state of the brushes [50],

making the zeta-potential measurements far from being a good indication of the surface potential.

Further characterization of the PAA brushes of P3 latex by DLS has been carried out by measuring

the Rh value of the particles as a function of NaCl concentration at each fixed value of pH in the range,

2.2 to 11.0. These experiments show typical behaviors of the annealing PE brushes. As shown in Fig.

2.4a, for pH[2.2,6.0], since most of the AA units are protonated and the brushes are in the collapsed

state, the Rh value is insensitive to the salt concentration. At pH 9.1, the initial slight increase in Rh

(CNaCl<100 mM) could indicate a transition from Osmotic Annealing brush regime (OsmA) to Salted-

brush regime (Sd) (CNaCl>100 mM). In the OsmA regime, AA ionization in the brush interior is

suppressed but increases as the salt concentration increases.

Page 29: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

16

Figure 2.3 The form factor of the primary particles, P(q) at various pH values, determined by

SLS, for (a) P2 and (b) P3 latexes after cleaning.

The increased osmotic pressure due to the increased counterion concentration in the brush causes

the brush swelling. This might explain the slight Rh increase in Fig. 2.4a. It must be noted that although

the observed Rh increase is rather small, it is larger than the experimental error. Further increase in CNaCl

leads to the Sd regime, where the behavior of the annealing PE brushes becomes similar to that of

quenched (strong) PE brushes. In this case, the brush thickness (i.e., Rh) decreases with the salt

Page 30: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

17

concentration. At pH 11.0 in Fig. 2.4a, due to the higher level of AA ionization, the Sd regime is reached

at a much lower CNaCl value, and we can observe only the decreasing part, which merges with the same

trend of the case at pH 9.1. When the data in Fig. 2.4a are plotted in the form of brush thickness, in

Fig. 2.4b, the slope in the Sd regime is about -1/3, in good agreement with theoretical and experimental

results in the literature [41, 51].

It must be noted that we have also measured Rh at extremely low pH (=1.5) for P3 latex, and we

have reported only one Rh value in Fig. 2.4a at CNaCl=10 mM. Further increase in CNaCl leads to

aggregation of the particles. The only reported Rh value drops below the plateau value observed at

intermediate pH values. This arises mostly due to the fact that apart from the PAA brushes, sulfate

groups ( 3OSO ) coming from the persulfate initiator are also present, which can be protonated at such

a low pH value. This leads to further decrease in the inter- and intra-chain electrostatic repulsion, as

well as the osmotic pressure of the counterions present in the brush, consequently, to further collapse

of the brushes.

For P2 latex, the low pH-responsiveness of its Rh value does not allow to perform the same type

of analysis. It is therefore unsafe to speculate about the difference in the absolute thickness or surface

grafting density of the PE brushes between P2 and P3. Moreover, since the particles were synthetized

through emulsion copolymerization, the acrylic acid content along the chains is unknown and can differ

from P2 to P3.

Page 31: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

18

Figure 2.4 Mean hydrodynamic radius (Rh) (a) and the estimated brush thickness () in the salted-

brush regime (b) as a function of NaCl concentration at various pH values for P3 latex. The broken

line represents the theoretically predicted relationship, 1/3C .[41]

Page 32: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

19

2.3.2 Stability of P1 Latex

Among the three latexes, P1 is the simplest one possessing only fixed sulfate charges on the particle

surface, without the PAA brushes. The corresponding measured W value at low pH are shown in Fig.

2.5a as a function of the concentration of both NaCl and H2SO4, while the corresponding measured pH

values are shown in Fig. 2.5b.

Figure 2.5 Fuchs stability ratio W (a) and the corresponding pH (b) for P1 latex as a function of the

electrolyte concentration.

As expected, the W value decreases as the NaCl or H2SO4 concentration increases. In addition, as

Page 33: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

20

shown in the figure, the electrolyte concentration value where W reaches the value of about one, i.e.,

the CCC (critical coagulant concentration) for fast aggregation, is much smaller for H2SO4 (~0.15

mol/L) than for NaCl (~0.50 mol/L). Note that this difference is not due to the fact that H2SO4 is a

bivalent salt. At such low pH, H2SO4 is only singly deprotonated, and acts as a symmetrical monovalent

electrolyte. To understand the difference in the W values measured above in the presence of NaCl and

H2SO4, one has instead to consider the difference in the association of Na+ and H+ to the surface fixed

charge groups, 3OSO . In particular, the association of H+ with sulfate groups is stronger than that of

Na+ [52]. This leads to a much stronger reduction of the surface charge density in the presence of H+

than in the presence of equal amount of Na+, thus making H2SO4 much more effective in destabilizing

the P1 latex than NaCl.

The above results clearly demonstrate the importance in accounting for the counterion association

with the surface ionizable groups when studying colloidal stability, as also demonstrated in other related

systems[53-55].

2.3.3 Contributions of PAA Brushes to Colloidal Stability

The particles of P2 and P3 latexes, apart from the fixed sulfate charge groups, contain 1% and 2% AA

monomers, respectively, which, as discussed in the previous section, are in the form of PAA brushes

on the particle surface. Clearly, their carboxylic groups, when deprotonated, i.e., COO , contribute

to the surface charges. Fig. 2.6a compares the W values for latexes, P1, P2 and P3, as a function of the

NaCl concentration at pH[1.8-9.0].

Page 34: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

21

Figure 2.6 Fuchs stability ratio W (a) for P1 (crossed symbols), P2 (hollow symbols) and P3 (full

symbols) latexes as a function of the NaCl concentration, at pH as reported in (b).

It can be observed that the stability of P1 is weakly dependent on pH. In fact raising the latter to pH=9.1

only makes the P1 latex as stable as the protonated P2 and P3. This weak dependency is clearly reflects

the low pKa of the sulfate groups, which ensures that at pH>3.0, all the chargeable groups of P1 are

ionized and raising the pH more does not affect the particle charge. It is also seen that P1 is slightly less

stable than P2 and P3 at low pH (1.9-2.5).The stabilizing effect of the acrylic acid in this low pH range

can be explained by the fact that the latter is less hydrophobic than MMA or BA, even in its protonated

Page 35: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

22

state. Therefore, in the case of P2 and P3, the overlap of their hydrophilic surfaces upon contact leads

to a repulsive interaction. This effect can easily be understood in the scope of polymer solution theories.

It must be noted that the W values for P2 and P3 are basically identical, indicating that increasing AA

(thus carboxylic groups) from 1% to 2% does not affect the colloidal stability in these conditions. The

stabilizing effect of protonated acrylic acid can be even more clearly observed in Fig. 2.7, where the

presence of pAA on the particles’ surface increases shifts the stability curves towards higher electrolyte

concentrations. P1 is slightly less stable than P2 and P3 under acidic conditions.

The situation however changes at higher pH values. The W values for P2 and P3 measured at

various pH values are shown in Figs 2.6a as a function of the concentration of NaCl. At pH around 5.5,

the W curves in Fig. 2.6a for both P2 and P3 move towards larger NaCl concentrations, indicating that

the stability of the latexes increases with respect to that at pH<3. It is therefore evident that now a

significant amount of carboxylic groups have been deprotonated and contribute to the colloidal stability.

In the case of P3 latex, the W values shift much more towards larger NaCl concentrations than those in

the case of P2 latex. This is understandable when one considers that P3 carries 2% AA while P2 only

1%. In this range of pH, located close to the pKa of acrylic acid and its polymer, it is expected that

small differences in the composition of the brushes can lead to a different repulsive effect due for

instance to change in effective pKa. Note that presence of the PE brushes on the particle surface may

contribute to the colloidal stability by steric interactions [45], this point will be covered later.

Page 36: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

23

Figure 2.7 Fuchs stability ratio W for P2 (hollow symbols) and P3 (full symbols) latexes as a function

of H2SO4 concentration (a) and the corresponding pH (b).

Finally, as shown again in Fig. 2.6a, at pH~8, i.e., under alkaline conditions, although the W values

for both P2 and P3 continue to increase with respect to the case at pH=5.5, the difference between P2

and P3 is insignificant. This result is somewhat surprising if one considers that at pH=8 and such high

salt concentrations, which ensure an ionization state in the brush close to that in the bulk [41], most of

the carboxylic groups should be deprotonated, and the surface charge should be larger for particles with

Page 37: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

24

2% AA than with 1% AA. On the other hand, it is well known from various modeling and simulations

[40, 41] that for the weak PE brushes like PAA, the concentration of the mobile counterions in the

brush-occupied region increases towards the particle surface, because of the strong electrostatic

attraction between the mobile counterions and the opposite charges on the brushes. The consequence is

that the effective charge of the groups (i.e., the AA groups in the present case) is substantially lower in

the inner than in the outer region of the brush zone. In fact, scaling-type models [40, 42] proposed that

the counterion concentration in the inner region is approximately equal to that of the charged groups on

the brushes, leading to the so-called local electroneutrality approximation (LEA).

Steric interactions, which are directly related to the brush thickness and the free energy of mixing

of the polymer chains [56], are expected to play an important role in the stabilization of these lattices.

In fact, at such high ionic strength (Debye length κ-1≈3Å), the electrostatic repulsion arising before the

brushes enter in contact is not likely to stabilize the particles, as noted by Fritz et al. on similar systems

[57]. From Fig. 2.4a, it can be seen that at all the pH and salt concentrations used for aggregation, the

Rh value (thus, the thickness of the brushes) is very similar, not to say identical. This implies that the

only parameter affecting the steric repulsion which changes with the salt concentration is the mixing

free energy of the brushes’ polymer chains. This effect has been investigated in the study of semidilute

polyelectrolyte solutions. It has been determined that it is necessary to account for the monomer-

monomer electrostatic repulsion to explain critical phenomena in polyelectrolyte chains structure[58].

Taking this repulsion into account gives rise to an effective Flory interaction parameter, which depends

on the salt concentration and the ionization state of the polyelectrolyte. This point explains why the

colloidal stabilization, which comes both from the hydrophilic nature of acrylic acid and its charge,

making the polymer chains repel each other when two brushes overlap is affected by the salt

concentration.

Accordingly, since electroneutrality prevails in the inner region, the particle-particle (or brush-

brush) interaction is mainly determined by the AA groups present in the outer region of the brushes.

Now, the difference between the P2 and P3 latexes is the length of the brushes, much longer for the

second, and if the inner part is electroneutral, then the effective repulsion is determined only by the AA

Page 38: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

25

groups in the outer region, which are very similar for the two latexes. This would justify the equal

measured W values for the two latexes shown in Fig. 2.6a. Therefore, our above results can be

considered as an experimental confirmation about non-uniform ionization of weak PE brushes, in

agreement with the local electroneutrality approximation.

2.4 Concluding remarks

Poly-acrylic acid (PAA) chains are pH-sensitive polyelectrolytes, whose charging, hydrophobicity,

polarity and conformation change with pH. In this work, we study experimentally the effect of pH on

the stability of polymer colloids with surface-grafted PAA chains (brushes). In particular, we have

prepared three acrylate copolymer colloids using the same recipe, but with 0%, 1% and 2%wt PAA

brushes, respectively. The length of the brushes under the stretched state, measured from the dynamic

light scattering, is about 2 and 13 nm for 1% and 2% PAA, respectively. In the case of 0% PAA, the

particles are stabilized only by the sulfate charges coming from the initiator, while in the latter two

cases they are stabilized by both sulfate and carboxylic charges.

The colloidal stability of each latex has been characterized by measuring the corresponding Fuchs

stability ratio (W) as a function of electrolyte concentration and pH. It is found that at low pH values

(<3), the W values of the particles with 1% and 2% PAA brushes are slightly more stable than the 0%

AA latex. This stabilization certainly comes from the hydrophilic nature of the acrylic acid.

At intermediate pH (~5), the PAA brushes are partially deprotonated, and their contribution to the

colloidal stability is substantial. Moreover, to reach the same W value, one has to use almost a doubled

NaCl concentration in the case of 2% PAA with respect to the case of 1% PAA. This means that at the

intermediate pH the ionization of the carboxylic groups (thus the particle charge) is directly linked to

the total AA groups on the surface. Accordingly, higher particle charge leads to higher stability for P3

than for P2 latex.

Under alkaline conditions (pH>8), one would expect all the carboxylic groups on the brushes to

be fully ionized and participating to stability. On the contrary, we found that the W values are basically

Page 39: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

26

the same for particles with 1% PAA and 2% PAA, which implies that the contribution of the ionized

AA in the two cases is practically the same. Such an experimental result confirms the local

electroneutrality approximation (LEA) [40, 42] in the inner region of long brushes, which leads to the

conclusion that only the ionized AA groups in the outer region of long brushes contribute to the particle

stability. Accordingly, the P2 and P3 latexes that differ in the length of the brushes exhibit the same

colloidal stability.

Page 40: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

27

Page 41: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

28

Page 42: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

29

Chapter 3

Investigation of the Steric Stabilization of dispersions with pH-sensitive

poly(acrylic acid) brushes

3.1 Introduction

As seen in the previous chapter, the use of polyelectrolyte brushes to stabilize colloidal systems is an

efficient solution to avoid particle aggregation. The case of weak polyelectrolytes is particularly

interesting, because of the pH-responsiveness of the stabilization [27, 30, 41].

The stabilizing effect arising from polyelectrolyte brushes comes from two main contributions: on one

hand, the charged character of the polyelectrolyte brush allows for an increase in surface charge density,

thus increasing the interparticle repulsion [27, 45, 57]. On the other hand, the hydrophilicity of the

polymer brush gives rise to so-called steric forces. “Steric” stabilization is a well-understood

mechanism. It arises from the free energy of mixing of two lyophilic polymer layers upon overlap, and

contains a purely steric and an elastic contribution. The elastic contribution arises from the physical

compression of the polymer chains. A typical steric interaction potential can be found in the work of

Willenbacher and coworkers [57, 59]:

0 no overlap

24 12 overlap, no brush compression2 2

1

4

USterick TB

U a HSteric LPk T

BU aSteric

k TB

1 12 2 ln compression2 2 4

1

H HL

P L L

(3.1)

Page 43: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

30

0 no compression

2 2

3 / ln

2

Elastic

ElasticP

W

U

k TB

U aL

Pk T MB

H H H L

L L

23 /

6 ln 3 1 compression2

H L H

L

(3.2)

Where (3.1) represents the purely steric contribution. (3.2) is the elastic contribution. a is the size of a

monomer unit, ν1 is the size of a solvent molecule (in the present case water), φP is the volume fraction

of the polymer chains, χ is the Flory parameter, quantifying the excess free energy of mixing of polymer

with the water, MW is the molecular weight of the chains, H is the height of the brush, and L is the

surface-to-surface separation.

The industrial synthesis of colloidal polymeric products gives rise to a number of constraints and

objectives. Polymeric products are produced in large quantities, and thus must be efficiently produced,

at high volume fractions, in the shortest reaction time possible. The high volume fraction constraint can

give rise to stability problems. It is therefore commonly decided to include additives in the emulsion

polymerization recipe, in order to modify and improve the stability of the produced particles. A typical

additive for acrylic and acrylic/styrenic adhesives and film-forming polymers are acrylic acid or

methacrylic acid [60, 61]. Because of the time constraint, and in order to obtain a stabilizing effect from

the beginning of the reaction on, the additive is mixed with the monomer mixture and fed during the

reaction time together, using a starved-feed strategy.

Such a synthesis strategy is very efficient a producing stable particles, but also gives rise to a significant

uncertainty about the actual morphology of the particles. In fact, acrylic acid being water-soluble, it will

be initiated by the water-soluble radical initiator before the more hydrophobic co-monomers. The living

chains will then enter the particle phase after having added a few hydrophobic monomeric units,

rendering them water-insoluble [62]. Once the living chain has entered the particle phase, the polymer

chain grows, mostly capturing hydrophobic monomers located in the particles. Therefore, even though

starved emulsion copolymerization is well-known to allow good chain composition control, in this case,

Page 44: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

31

the chain formed rather exhibit a kind of “block-like” copolymer structure, effectively synthesizing a

surface-active polymer. The hydrophilic acrylic acid rich extremities of the polymer chains tend to stay

on the surface of the nanoparticles. Therefore, a kind a polyelectrolyte brush is formed, as presented in

the previous chapter. The challenge of this type of particles is to understand the factors influencing the

stabilization of the particles. Namely, what is the link between the recipe (amounts of additive, initiator,

size of the particles) and the colloidal stability of the obtained particles.

The characterization of such nanoparticles is a serious issue. Some analytical techniques are suited to

characterize the size (light scattering) and the charge density (titrations) of the particles. However, no

analytical technique allows to measure directly the radial composition function within nanoparticles. It

is hence rather complicated to make the link between the recipe, the morphology, and the colloidal

stability of these products. We here propose an original method consisting of using the experimentally

obtained colloidal stability in order to gather knowledge about the particles’ structure.

The interaction potential of particles exhibiting electro-steric stabilization contains the coulombic

contribution, which is rather readily modelled and only depends on the surface charge density and the

ionic strength of the bulk, and the steric contribution, which mainly depends on the conformation of the

polymer brush. It was thus decided to investigate the steric contribution to the colloidal stability of

industrially produced poly(MMA-co-BA-co-AA) particles, synthesized with different amounts of

ammonium persulfate as initiator. In order to do so, aggregation experiments were carried out at low

volume fractions, in the presence of a high NaCl concentration, ensuring the absence of interparticle

electrostatic interactions, and varying the pH of the dispersions. The pH affects the protonation state of

the PE chains, thus modifying their effective Flory parameter, governing the steric interactions. The

aggregation kinetics was followed in situ by dynamic light scattering. These kinetic curves were then

fitted using a population balance equation model in order to obtain values of the Fuchs stability

parameter as a function of the pH. The characterization of the latexes using light scattering and colloidal

titrations allowed to gain some quantitative information about the solvent-accessible polymer. Then, a

stability model containing the structural parameters was developed and the unknown parameters were

fitted in order to gain some more structural understanding of the particles.

Page 45: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

32

3.2 Materials and Methods

3.2.1 The latexes

All the latexes used in this study were synthesized ad-hoc by BASF SE (Ludwigshafen, Germany).

They were produced by starved emulsion copolymerization of 49% w/w Methyl Methacrylate 49% w/w

Butyl Acrylate, and 2% w/w acrylic acid. The amount of ionic initiator (ammonium persulfate) was

varied between 0.0% w/w (in which case a H2O2/ascorbic acid initiation was used) and 1.2% w/w. All

the latexes were dialyzed extensively against Millipore water until their surface tensions were above 70

mN/m and their conductivities lower than 10 µS/cm, ensuring a complete removal of soluble oligomers

and surfactant used in the polymerization.

Latex name Composition (weight ratio)a

L1 BA (49) : MMA (49) : AA (2) + H2O2/Asc

L2 BA (49) : MMA (49) : AA (2) + 0.2 % APS

L3 BA (49) : MMA (49) : AA (2) + 0.8 % APS

L4 BA (49) : MMA (49) : AA (2) + 1.2 % APS

a: BA=Butyl Acrylate, MMA=Methyl Metchacrylate, AA=Acrylic Acid,

APS=Ammonium Persulfate, Asc=Ascorbic Acid

3.2.2 Characterization of the particles

The latexes were characterized by light scattering and coulometric titrations. The size of the particles

was obtained by static light scattering, using a BI-200SM goniometer system (Brookhaven Instruments,

USA), equipped with a solid-state laser, Ventus LP532 (Laser Quantum, U.K.) of wavelength 0=532

nm, as the light source, at the scattering angle range 15°-150. The particle size was obtained using the

RDG theory.

Page 46: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

33

In order to gain as much structural information as possible, the swelling of the polymer brush as a

function of the pH was investigated using the same method as reported in the previous chapter. This

allows for an estimation of the brush thickness at high salt concentrations.

In order to determine the concentration of sulfate groups coming from the persulfate radical initiator

attached to the particles, as well as the amount of solvent-available acrylic acid, titrations were

performed. A suspension of φ=1% particles was prepared, with 10 mM NaCl to provide an ionic

background. The suspension was stripped with wet N2 during 5 min, and thereafter kept under mild wet

N2 flux in the overhead to prevent gas from entering the system. As the particles were extensively ion-

exchanged, their counterions are H+. Hence the concentration of protons in the system is equal to the

concentration of sulfate groups and can be titrated with a base. The titrations were performed with a 10

mM NaOH standard solution (Fixanal®, Fluka) using a Titrino (Metrohm, Switzerland) controlled by a

computer. Before each base addition, the stability of conductivity and pH were verified not to vary more

than 0.02 pH unit and 0.1 µS/cm per minute. This point was shown to be in order to obtain reproducible

and reliable results. This is due to the dynamics of ion exchange when starting with collapsed brushes

[51]. The conductivity and pH of the suspensions were recorded, and the equivalence points were

obtained by the clear kinks in the conductivity curve.

3.2.3 Aggregation Experiments

The dialyzed latexes were diluted to a volume fraction φ = 5·10-4 in order to form stock solutions. Then,

in the vials used for the measurement of the kinetics, the suited amounts of water, H2SO4 and NaCl 4M

were added to reach the desired pH, 1M NaCl and φ = 10-5 after dilution. The stock latex solution was

then added in the prepared salt solution, argon was flushed over the top of the solution, and the vial was

placed in the DLS equipment. The preparation procedure was timed and always was equal to 1 min.

The evolution of the average hydrodynamic radius with time was then followed using a BI-200SM

goniometer system (Brookhaven Instruments, USA), equipped with a solid-state laser, Ventus LP532

(Laser Quantum, U.K.) of wavelength 0=532 nm, as the light source, at the scattering angle 90. The

solutions’ pH were measured after the aggregation using a freshly calibrated pH glass electrode. Some

Page 47: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

34

tests were carried and ensured the pH was stable over the aggregation time, varying less than 0.02 pH

units.

3.2.4 Simulation of the aggregation curves

Once the aggregation kinetics were obtained by dynamic light scattering, the Fuchs stability ratio was

obtained by using a PBE approach. The kinetic kernel was already described by Lattuada et al [17]. In

the present case, the Fuchs stability ratio, the fractal dimension and the lambda parameter were fitted

using a least-square algorithm.

3.2.5 Modeling of the interaction potential

In order to get more understanding of the morphology of the particles, the W=f(pH) function obtained

experimentally was compared to a steric interaction potential equation set. From the interaction

potential, the Fuchs stability ratio can be obtained by using eq. (3.3) [53]

22

exp /2 BU l k T

W dlGl

(3.3)

Where U is the total interaction potential, G a hydrodynamic function, and 2L a

la

a dimensionless

center-to-center separation.

Using eqs (3.1) and (3.2), it is possible to devise a steric interaction stabilization model. As observed, a

significant number of parameters are relevant for the application of these works. As discussed

previously, the amount of knowledge on the particles is rather limited. For instance, even if the swelling

of the brushes can be measured experimentally, this does not give access to the thickness of the

collapsed brush, state in which aggregation actually occurs. In the same line, the grafting density or

polymer volume fraction are unknown parameters. Some assumptions were therefore made:

Mw is assumed constant amongst the samples and taken as the molecular weight of a 15 unit

Acrylic Acid chain. This corresponds the length of a pAA chain of about 10 nm, which is the

typical order of magnitude in swelling.

φP is assumed constant amongst the samples and is set to a guessed value of 0.1

Page 48: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

35

ρP = 0.2 g/cm3 is assumed constant amongst the samples and chosen such as L = 1 nm for probe

L3 in the fully collapsed state. In fact, knowing the total density of acrylic acid on the surface

from titrations, it is possible to compute the density of the brush from its thickness.

L is computed from the AA – titration results and the chosen ρP. for all samples other than L3.

The Flory parameter χ depends on pH because of the change in hydrophilicity of poly-acrylic

acid with pH. It is assumed to follow a sigmoidal curve between two extreme values:

1 2

2

01 /p

pH pH

, where χ1 and χ2 are the extreme values of the Floy parameter, and

pH0 is the transition pH value, at which half of the acrylic acid should be protonated. It plays

here the role of an “effective pKa”. All of these values are fitted to the experimental results,

using a least-square method.

3.3 Results and Discussion

3.3.1 Characterization of the particles

3.3.1.1 Particle Size and pH-responsiveness of pAA-brush

The swelling behavior of the studied latexes was studied under the same conditions as in the previous

chapter. The use of DLS allows to gain information about the diffusion properties of the particles. The

measurement of the particles by static light scattering allows to reach to the radial mass distribution

function of the particles. This has been shown previously not to be affected by the swelling of the PE

brushes [61]. The particle size obtained by static light scattering, using the RDG theory is reported in

Table 3.2.

Page 49: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

36

Table 3.2: Characterization of the particles

Latex name APS [%w/w] Rp [nm]

L1 0.0 92.1

L2 0.2 90.0

L3 0.8 87.1

L4 1.2 93.5

Using the particle size obtained by SLS as a “shell-free” size, it is possible to devise an approximate

“brush thickness” as H PR R . The shell “thicknesses” as a function of and NaCl concentration are

reported in Fig. 3.1

Figure 3.1: Brush Thicknesses for samples L2, L3 and L4 at pH=10.08

It can be observed that the apparent shell thicknesses are very similar at low salt concentrations,

conditions under which the inter-chain electrostatic repulsion swells the PE brush [41]. The change in

swelling with increasing NaCl concentration however depends on the amount of sulfate used in the

synthesis. Namely, more sulfate in the recipe reflects in a more swollen brush at a defined NaCl

10 10000

2

4

6

8

10

12

14

16

1.2 %

0.2 %

(n

m)

NaCl Concentration (mM)

0.8 %

Page 50: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

37

concentration. This indicates already that changing the synthesis recipe affects significantly the

morphology and physics of the brush. It is here impossible to speculate about the reason for this effect.

It could come from a higher AA fraction in the brush or sulfate content at the polymer chain ends,

leading to an increased intra-chain repulsion, or more simply to changes in the grafting density of the

brushes. The L1 latex was not investigated in swelling behavior, because it was aggregating at lower

pH ranges.

3.3.1.2 Surface Charges

The surface charges of the diverse latexes were measured by colloidal titrations. This allows to quantify

the surface density of sulfate groups and acrylic acid groups on the surfaces. These results are reported

in Fig. 3.2a) and b), respectively.

Figure 3.2: Fixed sulfate (a) and Carboxy (b) groups on L1, L2, L3 and L4 latexes

0

1

2

3

4

5

6

7

8

0.0 0.5 1.0 1.50

10

20

30

40

50

60

70

80

90

100

b)

Sul

fate

con

cent

rati

on (

mm

ol/K

gP) a)

Car

boxy

l gro

ups

conc

entr

atio

n (m

mol

/KgP

)

% APS

Page 51: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

38

One can notice that increasing the APS amount in the synthesis recipe increases significantly the amount

of accessible sulfate groups on the surface of the particles. At the same time, the amount of accessible

acrylic acid decreases slightly with increasing APS amount when APS is used, and is found to be lower

when a redox initiation system is used instead of persulfate.

3.3.2 Aggregation experiments and the Fuchs stability ratio

A typical example of aggregation curve for sample L2 is reported in Fig. 3.3:

Figure 3.3: Aggregation behavior for L2 under different pH conditions. The solid line is the

corresponding PBE simulation result.

It can be observed that at the lower pH range (pH<6.0), the particles follow a fast aggregation regime,

and the aggregation rate does not depend significantly on the pH. This corresponds to the DLCA regime,

where no repulsion is present. Going to higher pH values, it can then be observed that within a limited

pH difference (from pH = 6.12 to pH = 6.17), the aggregation rate decreases significantly. Going higher

with the pH value decreases the aggregation rate even further, and reaches a new lower plateau value.

This fact indicates that a limited extent of repulsion is present, but it does not increase further higher

than the value reached at pH = 7.61. These observations can be easily explained: at low pH, the acrylic

acid is protonated, which of course limits the surface charge to the small amount of sulfate groups

coming from the persulfate initiator. The protonation of acrylic acid also increases its Flory parameter,

or namely makes it less hydrophilic. Less energy is thus needed to overlap two layers of swollen pAA

0

100

200

300

400

500

600

700

800

900

1000

0 1000 2000 3000 4000 5000 6000

Dia

met

er /n

m

Time /s

5.726.056.126.176.427.618.55

Page 52: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

39

in water than at high pH. The steric interaction is therefore rather limited. Raising the pH close to the

effective pKa of the pAA brush, some of the AA units will be deprotonated, allowing it to get more

hydrophilic. At even higher pH values, the whole pAA brush is deprotonated, and further increasing the

pH does not affect it anymore. One must be aware that even when the brush is deprotonated, it does not

get swollen, unlike the conditions under which the brush thickness is measured. This is simply due to

the screening of the electrostatic repulsion within the brush by the high NaCl concentration.

Using the PBE model, the kinetics can be used to obtain the Fuchs stability parameter. One example of

fitting is found in Fig. 3.3

Figure 3.4: Evolution of the Fuchs stability ratio W with changing pH at [NaCl] = 1M, as a function

of the pH and APS content.

Summarizing the results in Fig. 3.4, which shows the evolution of the Fuchs stability ratio with pH, a

few trends can be observed. Changing the amount of persulfate initiator in the recipe changes

significantly the pH at which the particles start to be stabilized. It is very clear that increasing the amount

of persulfate gives rise to more pH-resistant particles. Where the sample P2, containing 0.2% APS starts

to get stabilized around pH=5.7, the sample P4, containing 1.2% APS starts to get stabilized around

pH=4.8.

4.0 4.5 5.0 5.5 6.01

10

100

1000

L1 L2 L3 L4

W

pH

Page 53: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

40

3.3.3 Steric model

Using the Fuchs stability ratios obtained in the previous section, the parameters needed by the steric

model in order to reproduce the values obtained can be fitted. The first step is to obtain the values of

the Flory parameter for each sample as a function of the pH. These results are presented in Fig. 3.5 .

Figure 3.5 : Flory parameter χ as a function of pH for L1, L2, L3, L4, as fitted using the Steric

stability model.

Here, it can first of all be observed that the value range for the Flory parameter χ is rather unexpectedly

broad. In fact, the Flory parameter is a feature of the polymer composition and is not expected to vary

wildly with a simple change of initiator amount. It must also be recalled that a change of 0.1 in the Flory

parameter corresponds more to a total change in polymer nature than a conformation or a fine change

in the morphology of the chains. This fact hints strongly towards two possible explanations;

remembering that the brushes present on the surface of the particles are not pure poly(acrylic acid), but

rather an AA-rich layer, it could be that the composition of the brush is influenced by the radical

concentration during the polymerization process. Another likely explanation is that, due to the limited

amount of knowledge available about the particle surface morphology, some of the assumptions made

are not realistic. In fact, it would make sense that the polymer’s volume fraction, or molecular weight,

4.5 5.0 5.5 6.0

0.33

0.36

0.39

0.42 L1 L2 L3 L4

pH

Page 54: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

41

or density could change by changing the polymerization recipe. This illustrates clearly the limitations

of the modelling approach.

Figure 3.6: Results of the fitting for the Flory parameter at high pH (left) and “effective pKa” (right).

Let us nevertheless investigate the results of assuming that the Flory parameter is a sigmoidal function

of the pH. These results are presented in Figure 3.6. It is observed that the transition pH0 and value of

the Flory parameter at high pH depend significantly on the quantity of persulfate initiator used during

the synthesis. Namely, some clear trends are witnessed: on one hand, the L1 latex seems to be

significantly different from the three other samples. This can be explained by the fact that the whole

initiation process was different. On the other hand, the fitted values of χ1 and pH0 of the three other

latexes seem to evolve on a clear linear trend as a function of APS concentration in the latex. It can be

hypothesized here that the observed linear evolution of these parameters with APS concentration can

be correlated with some changes in the pAA brush morphology. The transition pH corresponds to an

“effective pKa” of the polyelectrolyte brush, which is known to vary both with the length and density

of the brushes [44], it is impossible to conclude whether this change is due to the sulfate presence alone

or to the morphological changes that a modified initiator concentration has on the brushes formed by

emulsion polymerization. Simultaneously, the absolute value of the Flory parameter actually depends

0.0 0.2 0.4 0.6 0.8 1.0 1.20.0

0.2

0.4

0.6

0.8

1.0 1

% APS

4

5

6

7

pH0

Page 55: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

42

on the polymer chain composition, which can be locally affected by the APS concentration surrounding

the particles during the synthesis.

3.4 Concluding remarks

In the present chapter, a tentative of quantitative understanding of the stabilization arising from steric

interactions on industrial samples was performed. The aggregation kinetics of a set of polymeric

samples, in which the initiator amount was modified were followed by light scattering at high salt

concentration. A PBE model was used to obtain the Fuchs stability ratio as a function of pH. The Fuchs

stability ratio was then used in order to fit the value of the Flory parameter required in order to

understand how the pH affects the hydrophilicity of poly(acrylic acid) in water. Interestingly, changing

the amount of initiator during the polymerization changes the effective pKa of the pAA brush. This fact

is not trivial to explain, but is probably related to the chain composition change, or a change in the

grafting density within the brush.

The Flory parameter can be accurately described by a sigmoidal function, which corresponds to its

values when fully protonated and fully deprotonated, respectively. The obtained values of Flory

parameters do not compare well with what was measured in the literature using more direct

measurements. The reason for this is probably that the assumptions made in order to simplify the model

are wrong, or that the non-ideal composition of the brushes varies when the amount of initiator used in

the polymerization is changed. Since these parameters cannot be obtained experimentally, no

improvement can be expected. It would be interesting to reproduce the same set of experiments and

apply the same model on sample synthesized in a more controlled way (grafted-on for example).

Page 56: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

43

Page 57: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

44

Page 58: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

45

Chapter 4

Interplay between Aggregation and Coalescence of Polymeric Particles:

Experimental and Modeling Insights

4.1 Introduction

Various industrial processes rely on the handling of colloidal particles, may these be in the form

of suspensions, emulsions or gels. As a matter of fact, colloidal particles applications range from food

to plastics, going through paints, coatings, paper treatment, to more recent niche-fields, such as drug-

delivery.[63]

One of the key features of colloidal particles is their kinetic stability, or the fact that the

dispersed particle phase will sooner or later organize in larger structures (i.e. clusters) and phase-

separate from the continuous phase it is suspended into. This un-avoidable process is regulated by the

interplay of different phenomena occurring, such as aggregation and coalescence. Clearly the

environmental conditions (e.g. temperature, pH, shear rate) and the particles characteristics (e.g. surface

charge, primary particle size), play a key role in determining the rate at which the destabilization process

takes place.[64] Another feature of aggregating colloidal particles is their tendency to form self-similar

(i.e. fractal) structures with fractal dimensions regulated by the conditions in which the destabilization

occurs. The fractal dimension relates the aggregates mass x , with its size R :

fdx R (4.1)

For instance, rather open clusters are formed in fully destabilized systems where the cluster

aggregation is diffusion limited (DLCA), whereas in reaction-limited cluster aggregation (RLCA) more

compact clusters are formed, as not every aggregation event is an effective one due to typically charge-

Page 59: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

46

induced particles stability.[65] Note that when aggregation and coalescence occur simultaneously, the

fractal dimension of the formed clusters is typically larger as compared to the non-coalescing cases.

This is due to the fact that coalescence leads to an interpenetration of the particles constituting the

cluster, therefore compacting the cluster itself and increasing its fractal dimension.

After the pioneering work by Ulrich and Subramanian,[66] highlighting how the simultaneous

aggregation and coalescence was indeed shaping the soot cluster growth in flames, several authors dealt

with this topic. Koch and Friedlander[67] introduced a simple, yet insightful and elegant deterministic

model extending Smoluchowski’s[12] approach to account also for coalescence. After that, Xiong and

Pratsinis solved a 2-D population balance equation (PBE) based model accounting for the time-

evolution of aerosol mass and surface area, and successfully compared it to experimental data.[68, 69]

Since then, during the last decades, the picture of aerosol coalescence and aggregation has

become clearer as the characteristic times of metal particles has been quantified and the mechanism of

coalescence unveiled.[70, 71] Moreover, the dynamics of cluster coalescence has been clarified and

expressions describing the time-dependence of the fractal dimension proposed. Another notable

progress in the field of aerosols was the model proposed by Heinson et al., which captures the key

features of the explosive generation of silica nanoparticles.[72] While aerosols have been extensively

investigated in this sense, only few studies are found on polymer particles undergoing aggregation and

coalescence.[73, 74] These works helped clarifying how rubbery (i.e. fully coalescing) particles

organize into clusters,[73] and provided insights on the restructuring of preformed clusters undergoing

coalescence upon temperature increase.[74] In these works particles were typically undergoing full

coalescence 3fd , hence existing light scattering correlations could be used to compare

experimental results (i.e. gyration and hydrodynamic radii) and PBE-based predictions. At the same

time, such correlations cannot be employed for cases where the characteristic times of aggregation and

coalescence are comparable, i.e. when partial coalescence occurs.

In this framework, the aim of this paper is to shed further light on the interplay between

aggregation and coalescence along two main lines, namely i) developing a suitable 1-D PBE based

model accounting for both processes and ii) correlating simulation results with light scattering results,

Page 60: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

47

thus extending the existing correlations for rigid clusters to partially coalesced ones. To this end,

polymeric particles with different glass transition temperatures ( gT ) have been prepared and their

aggregation behavior studied in stagnant DLCA conditions. Different temperatures and particles

concentrations have been explored in order to clarify the interplay of the coalescence and aggregation

rate. The developed model, along with the light scattering correlations, was tested against the

experimental data to verify its reliability and potential.

4.2 Materials and methods

4.2.1 Materials

Butyl Acrylate (BA), Methyl methacrylate (MMA), Potassium Peroxydisulfate (KPS) and

Sodium Chloride (NaCl) have been purchased from Sigma-Aldrich, sodium dodecyl sulfate (SDS) by

Apollo Scientific. All chemicals had purities larger than 99% and were used as received. Millipore

(MQ) water stripped for two hours with nitrogen was employed as the continuous phase in the emulsion

polymerization. Ion-exchange resin (Dowex Marathon MR-3 hydrogen and hydroxide form) purchased

from Sigma-Aldrich has been employed as received after the reaction to remove the SDS adsorbed on

the particles surface.

4.2.2 Particles synthesis

The particles were synthesized by starved emulsion polymerization[75] employing the

controlled reaction environment LABMAX©. In particular, the reaction was carried out at 70oC in a 1

L jacketed reactor. The reactor was initially charged with stripped MQ water and SDS and heated up

under nitrogen atmosphere. The initiator (KPS) solution was then added, and the monomer mixture feed

was started. The conversion and particle size were followed by thermogravimetric analysis and dynamic

light scattering (DLS), respectively. The instantaneous conversion always stayed above 99%, ensuring

that starved conditions were achieved. The monomer feed was stopped when the particles reached the

desired size. The reaction mixture was then kept at 70oC during one hour, to ensure complete conversion

Page 61: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

48

of the monomer. After the synthesis, the latexes were subjected to several cycles of cleaning using ion

exchange resins until their surface tension was above 71 mN/m (Wilhelmy plate method).

4.2.3 Light Scattering

4.2.3.1 Dynamic Light Scattering

Dynamic light scattering was employed to monitor the particles size throughout the reaction

(data not shown) until the desired particle diameter (200 nm) was reached. All the measurements were

carried out using a BI-200SM goniometer system (Brookhaven Instruments, USA), equipped with a

solid-state laser, Ventus LP532 (Laser Quantum, U.K.) of wavelength 532DLS nm , as the light

source. The temperature was controlled by an external water bath with a precision of 0.1oC. The

measurements were carried out in diluted conditions (occupied volume fraction 51 10 ) and with

10 mM NaCl.

4.2.3.2 Static Light Scattering

Static light scattering was employed to simultaneously follow the particle size (through the

gyration radius gR t ) and the fractal dimension evolution. The employed instrument was a Malvern

Mastersizer 2000, equipped with a laser having 633SALS nm . The structure factor ,S q t was

obtained by dividing the measured scattered light intensity ,I q t by the form factor ,P q t of

the primary particles, obtained by measuring the scattering intensity of the primary particles in non-

aggregating conditions. The radius of gyration gR t has been obtained by fitting the structure factor

in a Guinier plot following the method described in Harshe et al.[23]

Page 62: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

49

4.2.4 Differential Scanning Calorimetry

The glass transition temperatures of the samples were determined using a Waters Q200 differential

scanning calorimeter. The samples were heated to 120 oC, cooled down to -50oC and brought back to

120oC. The heating rate was 10oC/min in all cases.

4.3 Model development

4.3.1 Population balance equations

As previously discussed, two internal coordinates are in principle necessary to describe the time

evolution of aggregating-coalescing systems, namely the cluster mass x (i.e. the number of primary

particles in the cluster) and the fractal dimension fd . The corresponding population balance equation

is therefore two dimensional:

3

1 1 1 1 1 1

0 1

3 3

1 2 1 2 1 1 2 2 1 2 1 2

0 0 1 1

, ,, , ( )

, , ( , , , ) , , d d

1( , , , ) , , , , d d d d

2

f

f ff

A

f f f f f

B

f f f f f f

C

f x d tf x d t v x,d

t d

f x d t x x d d f x d t d x

x x d d f x d t f x d t d d x x

(4.2)

1 1 2 2 1 2 1 2( , ) , , ,D D f fx g x x y g x x d d (4.3)

1 1 2 1 2,g x x x x (4.4)

1 2

1 22 1 2 1 2 / /

1 2

ln, , ,

ln fm f fm f

fmf f d d d d

d x xg x x d d

x x

(4.5)

1, 3f f

COAL

v x d dx

(4.6)

01/3 1/3p pCOAL C

p

Rx x x

(4.7)

Page 63: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

50

The cluster mass distribution , , tff x d is defined such that , ,f ff x d t dxdd represents the

concentration of clusters consisting of x to x dx primary particles having a fractal dimension

comprised between fd to f fd dd at time t . Note that term A describes the variation in time of the

rate of coalescence , fv x d , and corresponds to a convective or Liouville term.[67] The rate of change

of the fractal dimension is defined in equation (4.6) and depends on the coalescence characteristic time

of the particles (cf. equation (4.7)) as detailed already by Koch and Friedlander.[67] Term B represents

the loss of an x -sized cluster having fractal dimension fd upon aggregation with any other cluster.

Term C instead accounts for the formation of an x -sized cluster with fractal dimension fd starting

from two smaller aggregates which need to satisfy the two constitutive laws (equation (4.4) and

4.5).[76] The integration boundaries are 0; for the cluster mass and 1;3 for the fractal dimension,

accounting for all possible cluster masses and shapes. fmd instead is the fractal dimension “imposed”

by the aggregation regime, i.e. 1.7-1.8 in DLCA and 2.0-2.1 in RLCA conditions. 0, ,p p pR represent

the surface tension, radius and viscosity of a primary particle, respectively. 1 2 1 2, , ,f fx x d d instead

is the aggregation kernel, which will be defined afterwards. All symbols are defined in the Symbol List

in Appendix 8.8.

Despite the solution of multidimensional balances is indeed possible,[77] in this context a

simplification of the balance to 1-D PBEs is desirable, as a fitting with experimental data has to be

performed, requiring the iterative solution of the PBE. To reduce the problem to a 1-D PBE, it is

necessary to multiply the original 2-D PBE (cf. equation (4.2)) with 3

1

fdd and 3

1

f fd dd , while

introducing the following two distributions, along a line already developed in the literature:[78]

3

1

, , ,f fx t f x d t dd (4.8)

3

1

, , ,f f fx t d f x d t dd (4.9)

Page 64: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

51

To properly treat the further terms present in the balance (4.2), the following assumption (cf. equation

(4.10)) is made:[78]

, , , ,f D f ff x d t x t d t d x t (4.10)

where

3

1

, 1D f fd d x t (4.11)

The physical meaning of the latter assumption (4.10) is that the clusters of a given size x have the

same fractal dimension ,fd x t . The idea is to obtain two 1-D PBEs, the first one describing the

cluster mass distribution (CMD) in time, while the second one describes their average fractal dimension

evolution in time. As a difference to typically used models, the fractal dimension is time-dependent and

a function of the (average) cluster size.

The resulting 1-D PBEs read:

1 1 1 1 1 1 1 1

0 0

, 1, , , , , ,

2

xd x tx t x x x t dx x x t x t x x x dx

dt

(4.12)

1

1 1 1 1 1 1 1 1

1 1

, 1, 3 ,

1, , , , , , ,

2

fCOAL

x

f

d x tx t d x t

dt x

x t x x x t dx d x t x x x x x t x t dx

(4.13)

where

1/3COAL Cx x (4.14)

1 2 1 21/ , 1/ , 1/ , 1/ ,

1 2 1 2 1 2

2( , )

3f f f fd x t d x t d x t d x tB

c

k Tx x x x x x

(4.15)

,,

,f

x td x t

x t

(4.16)

The Liouville term vanishes in equation (4.12), as this balance accounts solely for the cluster

concentration, disregardful of aggregates shape, whereas it is still present in equation (4.13); further

details are reported in the work by Koch and Friedlander.[67] Note that 1 2,x x is the typical DLCA

Page 65: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

52

aggregation kernel, modified in order to account for a time-dependent fractal dimension. Bk is the

Boltzmann constant, T the temperature and c the viscosity of the continuous phase.

Instead of solving the two PBEs (equation (4.12) and (4.13)) to calculate the average fractal

dimension (equation (4.16)), it is possible to differentiate equation (4.16), obtaining an ODE system

describing the time-evolution of the average fractal dimension (derivation details are reported in the

appendix 8.2):

1/3

1, 3

,0 1.75

f fC

f

dd x t d

dt x

d x x

(4.17)

The initial condition selected, ,0 1.75fd x , represents a typical fractal dimension of DLCA

aggregating clusters.[64] It is hence sufficient to solve the PBEs (4.12) coupled with the ODE system

(4.17): this way, one has directly the time evolution of the cluster concentrations while accounting for

a time-dependent fractal dimension. Notably, equation (4.17) is actually analytically solvable and

results in the following expression:

1/3, 3 3 1.75 expf

C

td x t

x

(4.18)

which is very similar to the one derived by Eggersdorfer et al., who studied the fractal dimension time-

evolution of purely coalescing fractal clusters.[79]

To appreciate the features of equation (4.18), parametric simulations with different C values

for differently-sized clusters ( 50,500,5000x ) have been performed (cf. Figure 4.1).

Page 66: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

53

Figure 4.1:

Fractal dimension vs time for differently sized clusters (x = 50, 500, 5000) and different C .

A smaller C implies a faster increase in the fractal dimension for a cluster of a given size as can be

seen in Figure 4.1 a), considering 310C s (blue curve) and 410C s (green curve). Comparing

instead the time-evolution of the fd of differently sized-clusters at the same C it is observed how

smaller clusters are able to re-arrange faster towards compact structures, exhibiting larger fd (cf.

Figure 4.1 a) -1c) for 410C s ). At the same time, if the C is large enough as compared to the

process time P considered (e.g. 6 410 300 min 1.8 10C Ps s ) , no coalescence is

observed, no matter how small the cluster size considered (cf. Figure 4. 1a)-1c) for 610C s ).

4.3.2 Calculation of hR t , gR t and S q,t

Once in possession of the time-evolution of the cluster mass distribution, average properties,

such as the hydrodynamic and gyration radii, hR t and gR t , and the average structure factor

Page 67: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

54

,S q t have to be evaluated in order to compare them with experimental results obtained from light

scattering. Such average quantities are defined as follows:

2

02

,0

, , ,

, , ,

,

h

h eff

x x t S x q t dx

R tx x t S x q t

dxR x t

(4.19)

2 2

2 0

2

0

, ,

,

g

g

x x t R x t dx

R t

x x t dx

(4.20)

2

0

2

0

, ,

,

S x q t x x dx

S q t

x x dx

(4.21)

While the cluster distribution ,x t is known once the PBEs have been solved (equation (4.12)),

suitable expressions for the hydrodynamic and gyration radii of each cluster, i.e. , ,h effR x t and

,gR x t as well as for the structure factor , ,S x q t are needed. q represents the scattering wave

vector, defined as:

4

sin2

nq

(4.22)

where is the scattering angle, is laser wavelength and n is the refractive index of the continuous

phase. Note that in the present case, clusters may aggregate and coalesce simultaneously and therefore

the usually reported formulas to calculate , ,h effR x t and ,gR x t in the case of rigid spheres cannot

be employed.[80, 81] For the sake of brevity the full derivation of the average quantities gR t ,

hR t and ,S q t , is reported in Appendix 8.3.

Page 68: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

55

4.4 Numerical Solution

In the present simplified form, the balance on ,x t would require a significant number of

ordinary differential equations (ODE) to be solved, as the internal coordinate x can go up to roughly

4 510 10 units. In order to reduce the problem size, a discretization method, based on Gaussian basis

functions is employed.[77] The main advantage of the present method is the ease with which it allows

to deal with the convolution integrals. The key idea is to approximate the actual distribution function

with a sum of Gaussian basis functions:

2

1

,G

i i

Ns x x

ii

x t t e

(4.23)

This allows to solve a finite number of ODEs (namely GN ) to obtain the time-dependent coefficients

i t , while the grid positions where the Gaussians are centered ( ix ) are fixed before the integration

start. The parameter is describes the overlapping degree of the Gaussians and is fixed once the

Gaussian centers are defined:[77]

1

1i

i i

sx x

(4.24)

To obtain the discretized balances it is sufficient to plug in the approximations (4.23) in the PBE in

(4.12). Further details are discussed in the Appendix 8.4, whereas the final balance reads:

2 1 1 2

1 1 1

2 W

N N NW W W W

j j jj j jj j j

dC C C C

dt s s s

(4.25)

The vectors and matrixes occurring in equation (4.25) are also defined in Appendix C. The calculations

in the present work have been performed with 90 Gaussians, i.e. 90GN . For numerical stability, the

fixed positions ix were placed on a linear grid (i.e. such as ix i ) up to 15 units, and then spaced

logarithmically up to 43 10 units.

Page 69: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

56

4.5 Results and discussions

4.5.1. Particle characterization and experimental conditions

The different nanoparticles used in this study, prepared as described in Section 2, are presented

in Table 4.1, where details about their composition, size, stability and glass transition temperature ( gT )

are found.

Table 4.1: Particle characterization

MMA [% wt] BA [% wt] Diameter [nm] PDI [-] CCC* [M NaCl] gT [oC]

30 70 198 0.027 0.75 -12

50 50 202 0.016 0.65 17

60 40 199 0.017 0.75 36

70 30 198 0.021 0.25 54

*The critical coagulation concentration (CCC) is measured at 25oC

Particles spanning through a quite large range of gT (-12oC to 54 C) were produced. Note that the

reported gT values represent an indication of the temperature interval at which the transition between a

glassy and a rubbery polymer matrix is observed.

The aggregation behavior of the particle systems has been studied in fully destabilized DLCA

conditions, achieved by diluting the latexes in 4 M NaCl water solutions. This specific salt concentration

was chosen as it guaranteed a density match between the continuous phase and the particle phase,

preventing cluster sedimentation to interfere with the experiments. Aggregation experiments were

performed at room temperature in static light scattering (SLS) experiments, whereas in a range between

25-45oC when using dynamic light scattering (DLS). The corresponding viscosities of the continuous

phases (given the 4M NaCl concentration) are reported in Table S2 (Appendix 8.5).[82]

Page 70: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

57

4.5.2 Aggregation at room temperature: gR t , hR t and S q,t

Static light scattering (SLS) experiments were performed at 25oC and 4 M of NaCl. Both the

average radius of gyration gR t , and the structure factor ,S q t have been measured in time at

three different occupied volume fractions, namely 5 5 51 10 ,2 10 ,3 10 . In Figure 4.2 a)-c) the

measured gR t vs. time is shown for all concentrations and particle types employed.

Figure 4.2:

gR t for the different particles at 25oC and a) 51 10 , b)

52 10 and c) 53 10 .

Black diamonds: 70% MMA, red squares: 60% MMA, green triangles: 50%MMA, blue circles: 30%

MMA particles, continuous lines: corresponding model predictions. Note that the gR t

predictions for the 70% and 60% MMA overlap. All the parameters values employed to obtain the

model predictions are reported in Table D2 Appendix 8.5.

gR t increases in time for all particle types, although the lower the gT (i.e. the lower the MMA %

and the softer the particles), the smaller is the gR t observed, both in terms of absolute value and

increase rate. This might seem surprising considering that the DLCA characteristic time of (doublet)

aggregation A is composition-independent:

11

31

8C

A DLCApart B partC k TC

(4.26)

a) b) c)

Page 71: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

58

As can be seen from equation (4.26), A depends only on the temperature of the system (25oC in this

set of experiments) and the viscosity of the continuous phase C (cf. Table S2, Appendix 8.5) and it is

defined for all systems once the particle number concentration partC is fixed. The smaller gR t

observed for the softer particles (at a given particle concentration) can be rationalized as follows. When

soft particles aggregate into clusters, neighboring particles (in the aggregate) will eventually coalesce

with one another, leading to a more compact cluster with a smaller gR t as compared to equally-

sized aggregates consisting of rigid particles, which are typically more open. This structural difference

affects also the aggregation rate: open clusters have a larger collision radius than compact ones (of the

same mass) and their aggregation probability is therefore larger. Thus, although A is per definition

composition-independent, the softness of the particles in an aggregate indirectly impacts the

aggregation rate by affecting the cluster spatial organization. Deepening whether and to which extent

particles in a cluster undergo coalescence, is therefore key.

A powerful and yet simple way to understand this is based on the comparison of the relevant

characteristic times involved. Besides the already introduced characteristic time of doublet coalescence

C (cf. equation (4.7)) and of doublet aggregation A (cf. equation (4.26)), also the characteristic time

of the entire aggregation process ( P ), defined as the total time for which the system is observed, plays

an important role in this context. Notably, P A in all cases, otherwise no clusters would be formed

if the process time was shorter than the characteristic time of doublet formation. To ease the analysis it

is convenient to introduce the ratio N of the doublet coalescence and aggregation characteristic times:

C

A

N

(4.27)

It is now possible to distinguish three limiting cases:

i) N <<1 - as soon as the primary particles aggregate, they undergo instantaneous coalescence

and the obtained clusters are fully compact all along the aggregation process.

Page 72: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

59

ii) 1N - doublet aggregation and coalescence occur at a comparable time scale. Since the

process time P A C the clusters grow while coalescing, hence rather uniformly compact

aggregates are formed. Such uniformity should be granted even though clusters born at different

times exhibit different coalescence extents, as the clustering process is mediated by cluster-

cluster aggregation, which averages out possible local non-homogeneities.

iii) N >>1 - in this latter case, two further sub-cases have to be distinguished:

a) P C - no coalescence occurs as this process is too slow (it lasts even longer than the

full aggregation process); open clusters are formed.

b) P C - coalescence occurs slowly and only after the clusters are formed. The aggregate

“history” or “life” becomes relevant: neighboring particles in a cluster formed at the

beginning of the aggregation process will exhibit a larger degree of coalescence as compared

to particles which only “recently” became neighbors. On average, the clusters will exhibit

(according to the absolute values of P and C ) a rather open structure with coalesced

braches. As compared to case ii) a smaller degree of compactness (i.e a smaller fd ) is

expected to be found, unless P .

Being based on the A and C of doublets, the latter picture is a simplistic one, nevertheless

representing an insightful view into the aggregation-coalescence process.

When fitting the model predictions against the experimental data using as only fitting parameter

C , one obtains the following results: C (i.e. 810C s ) for the 70% and 60% MMA particles,

4300C s for the 50% MMA case and 100C s for the 30% MMA particles. The simulated

gR t are reported in Figure 4.2 a)-c) (continuous lines) for the different particle concentrations; the

good agreement with the experimental data shows the model capability to capture the underlying

physics of the model. Given such C values, it is possible to compare them with

Page 73: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

60

150min 9000P s and A (cf. Table 4.2) along the previously described three limiting cases i)-

iii).

Table 4.2 A at 25oC and 45oC, for the different particle systems

A s at 25oC A s at 45oC

1 x 10-5 48 31

2 x 10-5 24 15

3 x 10-5 16 10

It is clear that being C for the 60% and 70% MMA cases, these two types of particles will never

undergo coalescence (cf. case iii a) ) given that C >> A ( N >>1) and C >> P . Note that the

predictions of gR t for these two cases are superimposed and represent a typical DLCA case of

non-coalescing particles (cf. Figure 4.2a)-c)). Such a behavior is consistent with the gT of these particles

systems, which is larger (36oC and 54oC for the 60% and 70% MMA, respectively) than the

experimental one (25oC), thus coalescence is not expected to occur. The 50 % MMA containing

particles instead belong to case iii b) because while C A 1N for any , P C , allowing the

formed clusters to undergo partial coalescence. This observation is supported by the gT of the 50%

MMA particles, which is slightly lower than the experimental one (17oC vs 25oC), thus allowing partial

coalescence to occur. The 30% MMA case on the other hand falls in case ii) as C A 1N ; being

the values very close, a significant extent of coalescence, much larger than the one of the 50% MMA

particles, is expected to occur, especially considering the long process time P A C and the

significant difference in terms of gT vs T (-12oC vs 25oC).

Page 74: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

61

To further explore the model potential, the predictions of the hydrodynamic radius hR t

(according to equation (4.19)) of the 60%, 50% and 30% MMA containing particles have been

compared with the corresponding DLS experimental results at 25oC and 51 10 in Figure 4.3.

Figure 4.3: hR t of the 60% MMA particles (red squares), the 50% MMA particles (green

triangles) and the 30% MMA particles (blue circles) and their corresponding simulations (continuous

lines) All the parameters employed to obtain the model predictions are reported in Table S3

(Appendix 8.5).

Note that the C values employed in these simulations were the ones obtained from the fitting of the

gR t data, hence the simulated curves in Figure 4.3 are purely predictive. The reasonable agreement

between simulations and experiments evidences the model reliability and its capability to capturing the

complex physics of the aggregating-coalescing system. When particles are significantly (if not almost

completely) coalescing (30% MMA case), a limited overestimation of the model prediction is observed

for both gR t and hR t . This is probably due to the fact that the equations derived for the

calculation of gR t and hR t have been obtained for the intermediate situation of partially

coalescing particles and not for full coalescing systems.

Page 75: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

62

Having explored the model effectiveness in describing size averages, it is now desirable to test

its performance in terms of cluster-structure related quantities and to deepen the interplay between

aggregation and coalescence in determining the clusters spatial organization. As a first step, the average

number of particles constituting the clusters (i.e. the cluster mass), AVEN t , defined as:

0

0

,

,AVE

x x t dx

N t

x t dx

(4.28)

is calculated at 25oC at 5 5 51 10 ,2 10 ,3 10 for the 60%, 50% and 30% MMA containing

particles, employing the previously fitted values of C . The simulation results are reported in Figure

4.4.

Figure 4.4: AVEN t at 25oC for 5 5 51 10 ,2 10 ,3 10 . Red continuous lines: 60% MMA;

Green dashed lines: 50% MMA particles (almost overlapped with the 60% MMA), dotted blue lines:

30% MMA particles. All the model parameters values employed to obtain the model predictions are

reported in Table S3 (Appendix 8.5).

As can be seen from Figure 4.4, the average mass of the clusters is smaller for aggregates consisting of

soft particles. This is due to the decreased reactivity of soft clusters as a result of their higher

compactness and reduced collision radius. Notably though, the observed differences in AVEN (for one

Page 76: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

63

set of ) among the different types of particles are relatively small: in the considered time-interval the

largest relative difference between cluster masses is of about 10% (comparing the 30% MMA and the

60% MMA particles). In other words, the difference in reactivity induced by the different softness is

present, but only mildly affects the average cluster mass. From the gR t trends (cf. Figure 4.2) on

the other hand, where the soft 30% MMA particles were showing much smaller radii (by at least 100%)

when compared to the 60% MMA particles, the significant impact of the particle softness on the clusters

morphology can be appreciated. More explicitly, the softness of the aggregating particles strongly

affects the spatial organization of the resulting clusters but only mildly impacts the overall aggregation

kinetics.

Having qualitatively discussed the impact of coalescence on the aggregates structure, it is now

interesting to attempt a more quantitative description. For this reason, the SLS experimental data on the

average structure factor ,S q t (obtained at 25oC, 4 M NaCl and three different occupied volume

fractions, 5 5 51 10 ,2 10 ,3 10 ) have been compared with the corresponding predicted

,S q t (cf. equation (4.21)) at three different time points, namely 60, 90 and 120 minutes. Note that

the model is employed here in “prediction mode”, as the employed C values in the different cases are

those previously fitted against the gR t data set. The results of the comparison are shown in Figure

4.5.

Page 77: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

64

Figure 4.5: Experimental (symbols) and predicted (continuous lines) ,S q t at three different times:

60 min (blue squares), 90 min (red triangles), and 120 min (black circles). All the parameters employed

to obtain the model predictions are reported in Table S3 (Appendix 8.5).

From Figure 4.5 it can be seen that the predictions of the model well-describe the experimental

observation in the cases of 60% and 50% MMA containing particles, whereas a poor agreement is found

in the 30 % MMA case. This implies that the model is indeed suitable for describing not only average

sizes, but gives also meaningful insights regarding the structure of clusters undergoing simultaneous

aggregation and coalescence. This holds provided that there is no or only partial coalescence: when

approaching full coalescence the model offers only a qualitative description. As a matter of fact, for the

30% MMA case, the slope of ,S q t vs. q (which represents the fd t of the clusters)[64]

increases in time, suggesting that cluster coalescence is significantly occurring. To further prove this

Page 78: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

65

point, the scattered intensity for the 30% MMA clusters has been recalculated assuming fully coalesced

clusters (i.e. spheres) using the Mie theory.[83] The radii of the spheres were calculated as

1/30S pR x R x , while the scattered light intensity from the whole distribution has been computed as:

0

0

, , ,

,

,

x t I x q t dx

I q t

x t dx

(4.29)

Note that ,I q t was then normalized by multiplication with a constant value. In the experimental

conditions considered, the polarization of the light was shown to have no significant effect on the

obtained function (data not shown). In Figure 4.6 the comparison between the experimental and the

simulated ,I q t at 60, 90 and 120 min, is reported for the 30% MMA case for the three

concentrations investigated, corresponding to 5 5 51 10 ,2 10 ,3 10 .

Figure 4.6: Experimental (symbols) and predicted (continuous lines) ,I q t at a) 51 10 , b)

52 10 and c) 53 10 for three different times: 60 min (blue squares), 90 min (red

triangles), and 120 min (black circles) for the 30% MMA particles. All the parameter values

employed to obtain the model predictions are reported in Table S3 (Appendix 8.5).

The reasonable accordance between experimental data and simulations indicates i) that the 30% MMA

particles are significantly coalesced and ii) that such coalescence is not a complete one, as otherwise a

full overlapping would have been observed.

Page 79: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

66

The average fractal dimension of the clusters, fd t , was estimated from the experimental

data at 120mint in order to provide a more direct glance at the aggregates spatial organization (cf.

Appendix 8.6, Table S5). Two different ways to calculate the fd t have been used, namely taking

the slope from the double log plot of ,S q t vs q and from 0gR vs I . In some cases it was not

possible to employ the ,S q t data in this sense, as the aggregates under investigation were too

small to get a sufficiently large fractal regime. Notably, apart from the clusters made of 30% MMA

particles, who significantly coalesce exhibiting a 2.3fd t , all other systems have the typical

fd t of non-coalescing clusters in DLCA conditions. For the 60% and 70% MMA particles this is

expected as their gT is much larger than the process temperature and therefore coalescence cannot occur.

On the other hand, the 50% MMA containing particles showed a smaller gR t (as compared to the

60% and 70% cases) and have a gT which is slightly lower than the experimental temperature (17oC vs.

25oC), therefore an effect of coalescence on fd t was expected to be present. This apparent

contradiction can be understood by recalling the different characteristic times for the 50% MMA

particles: 4300C s , 16 48A s (according to the different , cf. Table 4.2) and 9000P s .

While C >> A ( N >>1) and the initially formed clusters are indeed quite open and do not have the

time to coalesce, the total process time is large enough ( P C ) for the aggregates to undergo partial

coalescence, i.e. case iii b). These aggregates are significantly ramified at their “birth” and their

coalescence has to proceed through their branches first, before an extensive change in their structure is

appreciated. Such process is particularly slow and therefore no evident change in the fractal dimension

was appreciated. Interestingly, it turns out that while the fd t is a useful tool to assess and quantify

the extent of coalescence, in some cases measuring also the gR t and hR t is desirable as their

Page 80: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

67

time-evolution can be of great help in unravelling structural and spatial information about the

aggregating-coalescing clusters.

4.5.3 Aggregation at higher temperatures

To further test the reliability of the developed model, experiments at higher temperatures (25-

45oC) have been performed with the 60%, 50% and 30% MMA particles. The 70% MMA containing

particles were not considered as their gT is of about 55oC and these particles would therefore remain

“rigid” even at 45oC; higher temperatures were not explored in order to avoid water evaporation which

could bias the experiments. Note that to properly compare the experimental results at different

temperatures, the measured hR t have been plotted against the non-dimensional time, N :

11

8

3B

N part partA

k Ttt C t C

(4.30)

Note that by employing N , the effect of particle concentration and temperature are “filtered out”,

allowing to better appreciate the coalescence effect.

When considering the 30% MMA containing particles, it is worth mentioning that a significant

coalescence was already observed at 25oC, being their gT equal to -12oC. When increasing the

temperature by 10 degrees, the situation was found to be almost unchanged, as shown in Figure S1

(Appendix 8.7), where the hR t of the 30% MMA particles is reported for both 25oC and 35oC

against the non-dimensional time, N .This means that a substantial coalescence (very likely an almost

complete one) occurred in both cases and that for this reason no significant difference in hR t can

be appreciated; in fact in both cases 100C s was employed as fitting parameter.

When studying the 50% MMA containing particles (with a gT of 17oC), a larger coalescence

extent is expected when increasing the temperature above 25oC. Indeed, this is observed when

Page 81: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

68

measuring the hR t vs. N at 30oC, 35oC and 45oC: progressively smaller radii and radii increase

are observed (cf. Figure 4.7).

Figure 4.7: hR t vs. time of the 50% MMA particles at 25oC (black squares), 30oC (blue

diamonds), 35oC (green triangles) and 45oC (red circles) along with the corresponding simulations

(continuous lines). All the parameter values employed to obtain the model predictions are reported in

Table S4 (Appendix 8.5).

Increasing the temperature, the softness of the particles increases, causing the clusters to be more

compact and slower aggregating, an effect already discussed in the frame of the composition change

for the gR t and hR t data set at lower temperature. When fitting C against the experimental

hR t data, the following results are obtained: 1100s, 360s, and 210s, for 30oC, 35oC and 45oC,

respectively. The simulation results are reported in Figure 4.7 (continuous lines). Once more the quality

of the prediction is quite good, considering that only one fitting parameter, C , has been used. Given

the optimized values, it is possible to compare them to A (cf. Table 4.2) in the frame of the four limiting

cases identified in paragraph 5.1. At 25oC C A and partial coalescence occurs because of the long

process times, cf. case iii b). When the temperature is raised, C progressively decreases almost

Page 82: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

69

approaching a situation where C A . As a result, the coalescence extent is expected to be larger

(indeed the radii are smaller) and the aggregation kinetics slowed down due to the diminished collision

radii. Note that, beyond 35oC the coalescence seems to be almost complete as little difference is

observed between the experimental hR t at 35oC and 45oC. The model slightly overestimates such

predictions of almost fully coalesced systems, as already seen at lower temperatures with the 30% MMA

particles. When plotting the fitted C of the 50% MMA case against 1/T in a semilog plot (cf. Figure

S2, Appendix 8.7), it is possible to appreciate the temperature dependency, already observed in the

literature for both metallic and polymeric particles:[71, 74]

expC

BA

T

(4.31)

A similar analysis (at T = 25, 30, 35, and 45oC) was conducted with the 60% MMA containing

particles but for the sake of brevity it has not been shown. Moreover, as the 60% MMA particles possess

a gT of about 35oC, no real difference between the samples is observed between 25oC and 30oC. Even

beyond this temperatures (i.e. for 35oC and 45oC) only mild differences have been observed

experimentally.

4.6 Conclusions

In the present paper, a deterministic model accounting for the simultaneous aggregation and

coalescence of colloidal particles has been developed. The model is based on 1-D population balance

equations (PBE), whose solution gives access to the particle size distribution as a function of time. The

occurring coalescence is accounted for employing one parameter, the characteristic time of coalescence,

C . Literature correlations,[80] linking the PBE to experimental light-scattering information such as

hR t , gR t and qS t , have been extended in order to account for partial coalescence.

The developed model has then been successfully tested against light scattering data of DLCA

aggregating colloidal particles. In particular, by tuning the available parameter C , it was possible to

Page 83: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

70

well-reproduce the observed experimental trends employing polymeric particles exhibiting a broad

range of glass transition temperature values (-12oC - 55oC) in a range of temperatures comprised

between 25oC and 45oC. Limitations of the model were found when the coalescence extent was

significant, whereas for partial and no coalescence a good accordance was observed, both in terms of

average sizes gR t and hR t , as well as for structural parameters, such as qS t and

fd t .

Based on the characteristic times of aggregation ( A ), coalescence ( C ) and the entire

aggregation process ( P ), it was possible to identify three limiting cases, useful to appreciate the extent

of coalescence and the resulting cluster spatial organization. Recalling that P A in all cases, as

otherwise no aggregation would take place, and introducing the ratio of characteristic time of doublet

coalescence over aggregation, /C AN , it is possible to distinguish among the following situations.

i) full, instantaneous coalescence occurs when N <<1 and uniform, compact clusters are obtained all

along the aggregation process; ii) doublet aggregation and coalescence are occurring at a comparable

rate for 1N . Since P A C rather uniform, compact aggregates are formed as the clusters

coalesce while growing; iii a) no coalescence occurs and open clusters are formed when N >>1 and

P C ; iii b) coalescence occurs slowly and only after the clusters are formed when N >>1 and

P C . The initially formed clusters will be rather open; while for large enough P , the aggregates

branches will start to coalesce, hence the cluster “history” becomes of significant importance. For

P the clusters will fully coalesce and become comparable to the situation of case i).

This analysis revealed, among other things, how the fractal dimension of the clusters is not

always enough to appreciate partial coalescence (cf. case iii b)) and that by measuring gR t and

hR t a more complete picture is obtained.

Page 84: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

71

The present model might be of help in better characterizing and controlling colloidal processes

where the interplay of aggregation and coalescence regulates the clusters size and structure, which is of

great practical importance for the final applications.

Page 85: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

72

Page 86: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

73

Chapter 5

Effects of Coalescence on the Shear-Induced Gelation of Colloids

5.1. Introduction

Colloidal systems and nanoparticle suspensions are gaining continuously growing interest with time, as

new applications in the nano-scale are developed almost on daily basis[84-88]. In most of the cases,

these suspensions consist of lyophobic particles dispersed in a continuous phase. The thermodynamic

equilibrium thus consists of two separated phases; one particle-rich and one particle-poor phase. In

order to prevent the phase separation of the suspensions, the particles are commonly rendered kinetically

stable with the help of various strategies allowing to build a repulsive energetic barrier between the

particles, such as the use of charged ligands, surfactants, and steric stabilizers[89]. A typical example

of such colloidal system are polymer nanoparticles produced by emulsion polymerization, a widely

industrially used process[90]. Even though colloid science is a relatively “old” scientific topic, some of

the basic trends of these systems are still unclear and even a qualitative understanding is lacking. One

feature of dispersed colloidal particles is the ability to form fractal aggregates when the stabilization

strategy fails[91, 92]. Colloid instability arises either if the employed surfactant loses its efficacy (e.g.

high electrolyte concentration in the case of ionic surfactants)[89], or when the particle suspensions are

sheared, a condition which is very often met in industrial applications, during pumping or filtration

operations. The formation of aggregates can be either desired, if the aimed product is a powder, but it

may also compromise the quality of the whole product, as in the cases of coatings, paints and adhesives.

Page 87: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

74

In either case, aggregation should be tightly controlled. In the most drastic situations, the formation of

particle aggregates can also lead to a process called gelation, when the porous clusters interconnect,

percolate and form one “macro-aggregate” which spans through the whole container[93]. Such a

process may occur both in stagnant, as well as in shear conditions, where in the latter case it is

significantly more abrupt. The process of shear-induced gelation has been a topic of intensive research

in the last years and is rather well understood[18, 19, 22, 94]. If the applied shear rate is high enough,

the added energy can overcome the repulsion barrier keeping the particles apart. The governing rate of

doublet formation in the case of shear-induced aggregation of particles with stabilization has been

developed by Zaccone et al[22], and reads (in the simplified form):

/k 2 Pe1,1 8 Pe m BU Tm

B

Uk Dac e

k T

(5.1)

mU is the value of the interparticle potential at its maximum and 33 / BPe a k T , the Péclet

number, with α being a flow-type related constant, μ is the viscosity, is the shear rate, a is the particle

radius, kB is the Boltzmann constant, D the diffusion coefficient of the particles and T is the temperature..

The first term in the exponential represents the colloidal stabilization, slowing down the aggregation

and competing with the second term, representing the energy brought by the shearing. It can thus be

observed that above a critical size (or shear rate), the shear contribution takes over the stabilization

contribution, allowing for a drastic speeding up of the aggregation. Once small aggregates are formed

and reach the aggregation critical size, they aggregate with each other following an “explosive”

behavior because of the third-power size dependency of Pe

Once the aggregates reach a critical size, the applied stress becomes too large for them to resist, and

breakup occurs[95, 96]. The critical size depends on the topology of the aggregates and on the force

holding the particles together upon contact, which is ultimately related to the surface properties of the

particles. Breakup and re-aggregation typically lead to a compacting of the cluster, since more compact

clusters exhibit a higher average coordination number, which decreases the interface area between water

and the hydrophobic surface. At the same time, increasing the average coordination number increases

the resistance of the clusters to stress.

Page 88: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

75

If the material composing the nanoparticles exhibits a non-infinite zero-shear viscosity, coalescence can

occur once particles aggregated[97, 98]. A significant number of important polymeric products actually

undergo coalescence; such as coatings[99] and adhesives[24, 100]. This phenomenon has been

discussed in a previous article[101] for polymer particles and has been shown to have several effects

on the structure of the colloidal aggregates generated in stagnant conditions: reduction in cluster size,

increase in fractal dimension[97].

When combining coalescence and shear-induced aggregation, several effects can be expected: on one

hand, coalescence reduces the volume occupied by the fractal clusters, hence delaying the gelation. On

the other hand, partial coalescence may strengthen the bonds formed between particles, thus allowing

the clusters to grow to larger sizes before breaking, hence reducing the gelation time. The interplay of

these two competing effects makes it hard to predict what will be the effect of changing environmental

and particle parameters (e.g. temperature, particle viscosity, and shear rate) on the stability of the

suspension.

In the present work, we investigate experimentally the effect of (partial) coalescence on the kinetics of

shear-induced gelation. More accurately, we devised an experimental procedure allowing to distinguish

the contributions of aggregation, breakup and coalescence on ad-hoc synthesized polymer latexes. In

particular, we synthetized acrylic polymer colloidal particles with desired size and glass transition

temperature (Tg), and measured their gelation times in a Couette-Flow device as a function of

temperature and salt concentration. We then investigated the structural properties of the aggregates

using stirred-tank breakup experiments and the stagnant colloidal stability using dynamic light

scattering. This allowed to decouple the respective effects of coalescence on cluster breakup,

temperature effect on colloidal stability, and coalescence on the reduction of occupied volume fraction.

The present article investigates the case of hydrophobic acrylic polymeric particles in water, even

though the physical phenomena playing a role here are independent on the nature of the system and can

be found with inorganic particles as well as with other dispersants.

Page 89: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

76

5.2. Materials and methods

5.2.1 Materials

All synthesis and analysis chemicals were bought from Sigma-Aldrich, with purities higher than 99%.

The water used was filtered with a Millipore Milli-Q® (MQ) system.

5.2.2 Latex synthesis and characterization

The latexes used during this study have been synthesized by starved feed emulsion polymerization as

described in a previous article[101]. They consist of nanoparticles of random copolymer of methyl

methacrylate (MMA) and butyl acrylate (BA) dispersed in water. The initiator and surfactant used

during the polymerization were potassium persulfate (KPS) and sodium dodecyl sulphate (SDS),

respectively. The latexes were cleaned from surfactant using ion-exchange resin (Dowex® Marathon™

MR-3, Sigma-Aldrich) until their surface tension reached values close to the ones of water (>70 mN/m,

measured with a DCAT 21 and a Wilhelmy plate, Dataphysics, Germany). The ratio of MMA/BA

defines the glass transition temperature of the obtained particles. The particles were characterized in

terms of diameter and polydispersity index by dynamic light scattering using a Zetasizer Nano ZS

(Malvern, UK) at volume fractions of φ = 10-4. To measure the particles glass transition temperature

(Tg), calorimetry was employed, as already reported[101].

An overview of the produced particles in terms of their composition and properties is reported in Table

1.

Page 90: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

77

Table 5.1: Particle characterization

MMA [% wt] BA [% wt] Diameter [nm] PDI [-] gT [oC]

30 70 198 0.027 -12

50 50 202 0.016 17

60 40 199 0.017 36

70 30 198 0.021 54

5.2.3 Critical coagulation concentration determination

In order to measure the particles CCC, the stagnant aggregation process has been investigated by

dynamic light scattering (DLS) using a BI-200SM goniometer system (Brookhaven Instruments, USA),

equipped with a solid-state laser, Ventus LP532 (Laser Quantum, U.K.) of wavelength 0=532 nm, as

the light source, with scattering angle set to 90. The sample was thermostated by a water bath (Julabo,

Germany). The particles were first diluted with water to reach a polymer volume fraction φ = 10-4. The

NaCl solution was then diluted to the needed concentration, so that the mixture reaches φ = 10-5 and the

desired test salt concentration after mixing. Both solutions were then tempered in an oven set at the

desired temperature and left for equilibration for 15 minutes. When the desired temperature was

reached, the salt solution was rashly added to the polymer suspension and shortly mixed, in order to

avoid unwanted aggregation due to local higher concentrations of salt or polymer. The time needed to

mix the solutions and to set them in the DLS was measured and always equal to 1 min.

Increasing amounts of salt lead to faster aggregation. Above a certain NaCl concentration, the

aggregation stops accelerating. Such concentration corresponds to the CCC.

Page 91: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

78

5.2.4 Zeta potential measurement

In order to investigate the stagnant stability of the studied systems, their zeta potentials have been

measured at the two extreme temperatures of our range (25 and 45oC). Furthermore, to quantitatively

describe the colloidal stability of the used latexes experimentally measured values of zeta potential

(Zetasizer Nano ZS, Malvern, UK) has been compared with those calculated using generalized colloidal

stability model[102]using two different NaCl concentration. The polymer volume fraction for these

experiments was always φ = 5·10-5.

5.2.5 Colloidal titrations

In order to determine the concentration of sulfate groups coming from the persulfate radical initiator

attached to the particles, titrations were performed. A suspension of φ=1% particles was prepared, with

10 mM NaCl to provide an ionic background. The suspension was stripped with wet N2 during 5 min,

and thereafter kept under mild wet N2 flux in the overhead to prevent gas from entering the system. As

the particles were extensively ion-exchanged, their counterions are H+. Hence the concentration of

protons in the system is equal to the concentration of sulfate groups and can be titrated with a base. The

titrations were performed with a 10 mM NaOH standard solution (Fixanal®, Fluka) using a Titrino

(Metrohm, Switzerland) controlled by a computer. The conductivity and pH of the suspensions were

recorded, and the equivalence point was obtained by the clear minimum in the conductivity curve.

5.2.6 Couette-flow gelation experiments

The gelation experiments were carried out in an ARES rheometer (Rheometric Scientific, USA)

equipped with a Couette-Flow device. The radial gap was 0.35 mm, and the gap under the inner cylinder

was set to 3 mm, in order to avoid the contribution coming from the liquid below the bob. The

preparation of the suspensions was the following: first, the diluted salt solution and the polymer latex

were thermostated to the experiment temperature. Then the salt was added to the polymer rashly under

mild stirring. The final dispersion was composed of polymer φ = 5%, NaCl at the desired concentration,

Page 92: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

79

and water. The obtained solution was inserted in the thermostated rheometer cup and the bob was

lowered. A solvent trap was used to prevent water evaporation. The used steady shear rate was 4700 s-

1. The viscosity of the suspension was measured over shearing time.

As it is important to ensure that the observed gelation in the rheometer was due to shear, and not simply

to the destabilization of the particles, a part of the mixture was kept in an oven under stagnant conditions

at the same temperature as the experiment. The size of the particles kept under stagnant conditions was

measured after the end of the rheological experiment by DLS (Zetasizer Nano ZS, Malvern, UK), and

always proved the suspension to remain stable, as no change in size was observed. Each shearing

experiment was repeated at least twice to ensure reproducibility. The reported gel time corresponds to

the time-coordinate of the intersection between the viscosity plateau before the onset of gelation, and

the sharp rise in viscosity occurring during the gelation.

5.2.7 Stirred-Tank DLCA experiments

For the breakup experiments, a 2 L stirred-tank reactor was used. It was equipped with a 60 mm Rushton

turbine and four metallic cylindrical baffles. A diluted polymer dispersion φ = 10-5 was inserted in the

tank, taking great care of removing all the air bubbles present in the reactor while filling it up

completely. In fact, the presence of air would lead to adsorption of some particles or aggregates on the

surface of the bubbles and lead to undesirable effects on the aggregation and breakup behaviour. To

ensure that no air would enter the reactor, a magnetic coupling was used to drive the stirrer. For the

same purpose, part of the excess polymer suspension was pumped in a vertical tube (1.5 m high)

connected to the reactor, setting it under light overpressure.

The reactor was equipped with an electrical heating jacket to control its temperature. Once the desired

temperature of the reactor was reached, 60 mL of a 2 M MgCl2 solution was added using a syringe

through a port at the bottom of the reactor, ensuring that the particles are fully destabilized (DLCA

aggregation under shear). The system was then let to equilibrate during one hour. The aggregation and

breakup of the particles was followed by small angle static light scattering (SALS), using a Mastersizer

Page 93: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

80

2000 equipped with a flow cell (Malvern, UK). The mean radius of gyration (Rg) and fractal dimension

(Df) of the aggregates were obtained from the structure factor of the aggregates, S q . In particular,

Rg was extracted from a Guinier plot and the slope of the curve in the fractal region was used to extract

Df, using relation fDS q q . In order to measure the size of the aggregates in the reactor, a small

amount (≈10 mL) of suspension was gently diluted in 100 mL of water, and injected in the SALS

instrument. The obtained solution was continuously slowly injected in the cell, in order to prevent

sedimentation of the clusters. Each measurement has been repeated 3 time and the reported values

represent the average values.

Page 94: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

81

5.3. Results and Discussion

5.3.1 Effect of temperature on stagnant aggregation

Before discussing the impact of coalescence on shear-induced aggregation, as a reference we first

investigate the particles stability in stagnant conditions. As the shear experiments are performed at

different salt concentrations and temperatures, it is necessary to deepen the effects of these variables on

the intrinsic colloidal stability of the particles employed. Aiming at this, the critical coagulation

concentrations and zeta potentials of the studied particles were measured at the two extreme

temperatures in the studied range, namely 25 and 45oC. The results of these analyses are presented in

Figure 5.1

Figure 5.1 : Critical Coagulation Concentration of NaCl vs particle composition (MMA fraction in

percent) at different temperatures (a) and Zeta Potential vs particle composition at two different salt

concentrations and two temperatures (b)

0.0

0.2

0.4

0.6

0.8

1.0

30 40 50 60 70 80

-70

-60

-50

-40

-30

-20

-10

0(b)

25°C 45°C

CC

C /M

NaC

l

(a)

10 mM, 25°C10 mM, 45°C140 mM, 25°C140 mM, 45°C

Zet

a P

oten

tial

/mV

% MMA

Page 95: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

82

As observed, the CCC decreases significantly with increasing temperature for all particle compositions

considered. Conversely, while the zeta potential decreases (in absolute value) with increasing salt

concentration, it does not change significantly with the temperature. This indicates that the ionic

equilibrium of the sulfate groups present on the surface is not influenced by the temperature, and hence

the electrostatic repulsion is constant with temperature. In order to understand more quantitatively the

interaction giving rise to the stability of the colloidal systems, a generalized stability model was

used[102] in order to solve the Poisson-Boltzmann equation, taking into account the counterion

association[103]. The interaction potential was then computed using the DLVO theory, where the total

interaction potential equals the sum of van der Waals attraction and electrostatic repulsion. Let us

analyze as an example the case of the 30% MMA sample. The surface charge was obtained by titration

and is equal to 0.016 mol/KgP. The generalized stability model allows to compute the ionic equilibrium

and the surface potential of the particles. The obtained surface potential at 10 mM NaCl and 25°C is -

76 mV. This compares well with the obtained zeta potential in the same conditions, which is equal to -

73 mV. Increasing the ionic strength to 140 mM gives rise to a surface potential of -29 mV, according

to the model results. This agrees reasonably well with the experimentally obtained -40 mV, considering

the fact that the model is assuming ideal behavior of all involved ions. The so determined surface charge

allows to compute the interaction potential, and more specifically to evaluate the CCC, or the minimum

salt concentration at which the potential barrier value is equal to 0. The CCC computed at 25oC

employing the DLVO theory is equal to 160 mM. This contrasts strongly with the measured CCC

around 800 mM (cf. Figure 5.1), concentration of salt, at which basically no electrostatic repulsion is

present anymore. This strong discrepancy suggests the presence of non-DLVO stabilizing interactions,

which for lyophobic colloidal systems are often associated with hydration forces[104, 105]. It hence

was decided to include a non-DLVO force in the model, in the typical hydration exponential potential

form 20 0

0exphyd

dU aF l l

where d is the surface-to-surface separation The force constant 0F

was fitted in order to match the experimental value of the CCC, whereas the characteristic decay length

Page 96: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

83

0l was chosen as the size of a water molecule[89]. It was found that this force constant is equal to

3.1·106 Nm-2 at 25°C. This value is in the same range (107 Nm-2) than previously obtained results[105],

keeping in mind that the surface of the particles in the present study is significantly different.

The effect of temperature on the hydration force is found to be the governing factor influencing the

stability of the particles. In fact, it is found that 0F decreases from 3.1·106 Nm-2 at 25°C to 1.3·106 Nm-

2 at 45°C. Computing the CCC at 45°C using 3.1·106 Nm-2 as force constant results to 850 mM NaCl.

It must be here pointed out that the used stability model takes into account every effect of temperature

on the Poisson-Boltzmann distributions and on van der Waals forces, but not on counterion association.

It is however a conservative approximation, as ion association is an endergonic exothermic

process[103], it will be disfavored by an increase in temperature. This means that the computed surface

potential is lower (in absolute value) than the actual one. Therefore, the particles should become even

more stable by increasing the temperature than what was computed without taking this effect into

account, confirming once more that a non-DLVO repulsive interaction is present, and decreases in

intensity with increasing temperature.

5.3.2 Gelation Curves

The typical process of shear-induced cluster gelation is presented here. It is worth noting, that the

particulate system is stable from a colloidal point of view as long as no shear is applied. When shear is

applied, the energy brought to the system allows the particles to overcome the stabilizing energy barrier

and hence to aggregate. As shown in eq. (5.1), aggregation is strongly accelerated by the shear rate, but

also by the cube of the particle size, in the exponential term of the Arrhenius-type rate constant. This

leads to an explosive behaviour in viscosity, as show in figure 2.

Page 97: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

84

Figure 5.2 : Gelation curves for the 50% MMA sample at different NaCl concentrations

An approximate cluster size for the gel point can be derived from the condition where the effective

volume fraction of cluster reaches 0.5, which corresponds roughly to the occupied volume fraction of

randomly packed spheres[106]. The fractal scaling law is well-established and reads:

Df

gf

Ri k D

a

(5.2)

where i is the average number of particles per cluster and fk D a prefactor depending on the

fractal dimension according to empirical relation 2.084.46f fk D D [107]. Considering eq (5.2) it is

possible to obtain the relation between the size of the formed clusters at the gel point and the system

parameters, e.g. phi, size of primary particles and fd

1

3

, 2

fDfc

g gel

k DR a

(5.3)

In the present system, the critical radius of gyration is equal to ,cg gelR =87 μm. This implies that if all

the primary particles are converted into clusters, these clusters must be able to grow at least up to the

aforementioned size in order to allow gelation with primary particles of 100 nm radius and 5% of initial

volume fraction.

0 2000 4000 6000 8000 100000

1

2

3

4

/m

Pa

s

Time /s

140 mM 130 mM 120 mM

Page 98: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

85

However, as the aggregates grow the hydrodynamic stress to which they are exposed increases up to

the point when it becomes larger than the aggregate strength, leading to cluster breakup.[108]

It has been observed in this study that no shear-induced gelation is obtained with the MMA/BA

copolymers if T<Tg at φ=5%, as it is the case for the 60% MMA sample at T=25°C and the 70% MMA

sample at all studied temperatures (Tg=36°C and Tg=54°C, respectively). In these cases, it was

impossible to find a salt concentration at which gelation occurred under shear, at least in the 4h shear

experiment performed. It is possible to observe gelation of these samples in these conditions, but only

in stagnant conditions using NaCl concentration in the range of 200-500 mM, depending on the sample

and the temperature. Notably, the formed gels are remarkably weak, since tilting the container where

they are formed is sufficient to break their structures and allow the flowing out of the slurry. This reflects

clearly the effect of breakup on gelation: if these acrylic particles are not coalescing, the bonds formed

between the particles are not strong enough to allow the growth of large enough clusters to reach

gelation under shear stress.

5.3.3 Effect of salt concentration and temperature on gelation time

The gel times (i.e. the time at which the viscosity abruptly spikes) measured under constant shear in the

rheometer for the three different samples at three different salt concentrations and three different

temperatures are reported in Figure 5.3

Page 99: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

86

Figure 5.3 : Gel Times vs NaCl concentration for 30% MMA (a), 50% MMA (b), 60% MMA (c) in

the Couette-Flow device at different temperatures

First, let us comment on the effect of NaCl concentration on the gelation time (tgel). It can be observed

that for all the samples at all temperatures, increasing the salt concentration leads to a decreased tgel.

This effect is expected, as an increased salt concentration leads to a decreased electrostatic repulsion

due to the screening of the fixed charges[11, 109]. A decreased repulsion leads to a faster aggregation,

as expressed in eq. (5.1). Nevertheless, it should be observed that the dependency of the gel time on the

0

50

100

150

200

250

300

0

50

100

150

200

250

300

115 120 125 130 135 140 1450

50

100

150

200

250

300

(c)

(b)

25°C35°C45°C

Gel

Tim

e /m

in

(a)

25°C35°C45°C

Gel

Tim

e /m

in

35°C 45°C

Gel

Tim

e /m

in

CNaCl

/mM

Page 100: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

87

salt concentrations is a function of the temperature. It is observed that for the “harder” samples (60%

MMA and 50% MMA), the gelation time decreases monotonically with increasing temperature. This

effect arises because of two main reasons; on one hand, as shown in section 3.1, all the used samples

are less stable when the temperature is increased. On the other hand, as shown in our previous

paper[101], increasing the temperature of these samples increases significantly the coalescence rate. An

increase in coalescence rate corresponds to a higher extent of interpenetration between the particles,

hence giving rise to more shear-resistant (i.e. non-breaking) clusters. It is very likely that these two

effects are present and it is non-trivial to suggest when one mechanism prevails over the other one. This

point will be investigated in section 3.4.

When considering material properties, namely the particles with 30% MMA (cf. Table 1) an interesting

phenomenon occurs here (see Figure 3a). It can be seen that increasing the temperature from 25°C to

35°C, the gelation time increases while a further increase to 45°C, results in gel time decreases. This

initial increase in gelation time can be understood by taking into account the second effect of

coalescence, which is the decrease in occupied volume fraction resulting from the compaction of the

clusters. In fact, if full coalescence occurs, the obtained clusters will occupy exactly the same volume

as the particles constituting them, forming a larger sphere. Therefore, if full coalescence occurs, no

gelation is possible, because the clusters cannot percolate without an increase in occupied volume

fraction. It is likely that the initial increase in gelation time observed in the case of the 30% MMA

sample is due to a decrease in coalescence time. It has been reported previously[101] that these soft

particles seem to reach, between 25oC and 35oC under diluted DLCA conditions, an almost fully

coalesced state in comparable times. To evaluate relative importance of involved mechanisms we

perform an analysis of the characteristic times of aggregation, coalescence, and breakup. Following our

previous work,[97, 101] the characteristic coalescence time of clusters can be estimated (is defined) as:

1/30

12

Polcoal

ai i

(5.4)

where pol is the polymer particle viscosity, and 12 the interfacial tension between the phases 1

(particle) and 2 (water).

Page 101: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

88

The competition between aggregation and coalescence is not the same in stagnant and shear-induced

aggregation conditions. Namely while stagnant DLCA aggregation tends to slow down with process

time (and cluster size) because of lower diffusivity[110], shear-induced aggregation explosively speeds

up. The small clusters formed at the initial stages of aggregation can hence coalesce, but it requires a

much lower viscosity to allow coalescence on the cluster size range. It therefore makes sense to observe

the effect of decreased coalescence time on gelation time in temperature regions where the full

coalescence was already reached in more diluted conditions. The fact that gelation occurs is also a

strong proof that full coalescence is not occurring, for the reasons stated above. By increasing the

temperature further to 45°C, the gelation time decreases again. This is due to similar effects as those

observed for the other samples, namely a colloidal destabilization or a decrease in breakup rate due to

strengthened bonds between particles.

Let us now discuss the fact that the slope of the tgel vs. NaCl concentration depends on the temperature.

Namely, the influence of salt concentration decreases when the temperature is higher. It seems that the

salt is effective at speeding the gelation up to a certain point, at which the gelation time only depends

very weakly on the salt concentration. It is rather clear that this effect is not due to coalescence; in fact

it is only observed for the 50 % MMA and 60 % MMA samples in the studied temperature range. In

addition, it is not expected that coalescence changes dramatically the effect of salt on aggregation. It is

quite likely that the observed smaller salt effect at larger temperature is due to the already covered non-

DLVO force stabilizing the systems, which is a function of temperature. Since the latexes are stable

under stagnant conditions at the salt concentrations and temperature used, it is certain that some

repulsive energy barrier is present, coming from electrostatics and non-DLVO interactions. This

potential maximum is competing with the hydrodynamic stress, as shown in eq. (5.1), and one can

define a critical Péclet number over which the aggregation rate is almost independent on the interaction

potential[18], the particles aggregate every time they collide. It is hence rather straightforward to

understand that the critical Péclet number for aggregation of the primary particles (or small aggregates)

is reached when enough salt is added or/and when the temperature is high enough. The particles then

aggregate at the same rate, independently on the absolute magnitude of the interaction potential, as long

Page 102: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

89

as it is low enough. Adding more salt to the suspension has therefore no more effect. The effect is

however not observed for the softest material (30 % MMA). This is likely due to the fact that the

coalescence rate is much higher for this sample. The occupied volume reduction effect is hence

superimposed on the aggregation/breakup competition.

5.3.4 Effect of temperature on cluster breakup

In order to determine the relative effects of decreased breakup rate and decreased colloidal stability

during the gelations in the rheometer, breakup experiments were carried out in a stirred-tank reactor

under DLCA conditions and relatively low volume fractions. In these conditions, the clusters grow due

to the convection and Brownian motion, capturing all the particles they meet, until their critical breakup

size is reached. They then break, giving rise to smaller fragments or primary particles. This dynamic

equilibrium determines the average cluster size in the system[108]. Since the initial volume fraction is

set constant in all experiments and the particle size is constant, the number of particles is also constant.

Therefore, the aggregation rate is the same in all experimental conditions. The average cluster size is

hence only defined by the breakup rate, which depends on the size of the clusters, but also on the

morphology (fractal dimension), and on the binding strength between the particles. The average size

and fractal dimension of the observed clusters for the 60% and 70% MMA sample at different

temperatures is reported in Figure 5.4.

Page 103: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

90

Figure 5.4: Average radius of gyration (a) and fractal dimension for 60% MMA and 70% MMA

samples in the stirred-tank experiments, as a function of temperature.

Looking first at the effect of temperature on the size and fractal dimensions of the clusters of rigid

particles (70% MMA), it can be observed that a slight increase in size is observed between 35°C and

45°C, while the fractal dimension does not show any significant change, with values around 2.8. A

constant fractal dimension indicates that the clusters’ morphologies are not changing, and thus the only

factor governing the critical breakup size is the adhesion force. This moderate increase (1.6 µm to 6.1

µm) can have several origins coming from the effect of temperature on the adhesion of the particles.

Let us analyse two models for adhesion in the literature: according to the well-known Johnson-Kendall-

Roberts (JKR) model, the pull-off force (the force at which the particles are pulled apart) reads

12

3

2cF a [111] The increased cluster size could be explained by an increase in surface energy

25 30 35 40 450

5

10

15

20

25

30

25 30 35 40 452.0

2.2

2.4

2.6

2.8

3.0(b)

60 % MMA 70% MMA

<R

g> /

m

Temperature /°C

(a)

60 % MMA 70% MMA

Df /-

Temperature /°C

Page 104: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

91

12 . Such a variation is conceivable, considering the change in colloidal stability observed in the

previous section, which was attributed to a change in non-DLVO interactions. Another popular

adhesion model is the development by Maugis and Pollock[112], allowing for plastic and elastic

deformation. In this case, the pull-off force also depends on the Young modulus (E) of the material. The

pull-off force namely increases with 2

1

E, for both plastic and elastic deformation. It is well known that

the Young modulus decreases significantly when the temperature approaches the glass transition

temperature for amorphous polymers[113]. It is hence likely that the glassy particles become more

elastic while staying under the Tg, (=54°C) allowing the growth of larger clusters.

For the 60% MMA particles, with a lower Tg (=36°C), a more striking effect is observed; the size of

the obtained clusters while T<Tg stays in the same range as for the harder material, it increases

dramatically when the glass transition temperature is reached. The fractal dimension also stays constant

at high values around 2.8, independently on the temperature. This indicates that coalescence only occurs

on the particle scale in these conditions, without being able to significantly influence the morphology

of the clusters. This implies again that the critical breakup size of the aggregates is governed by the

adhesion of the particles only[114]. It must be noted here that all the measurements at T>Tg are

underestimating the size of the obtained aggregates. In fact, it is observed that the formed aggregates

grow to very high sizes (>100 µm) in these conditions, and tend to adhere to the surface of the reactor.

Hence, no real steady-state in size is reached, as some material is taken away from the suspension.

Measurements have been taken after 10 min, 30 min and 60 min, and a decrease in concentration was

systematically observed when T>Tg, while the cluster size stayed constant with time. This is the reason

why no data is reported for the 60% MMA particles above T=39°C as well as for the softer materials,

as most of the material then gets lost in the reactor. This effect actually only accentuates the fact that

the glass transition temperature is the point at which a critical change in binding strength is occurring.

The source of this dramatic change in shear stability can have several sources; on one hand, the surface

energy could be changing, as reported for the hard particles. However, there is no clear reason about

why crossing the glass transition temperature would change the surface energy abruptly. The change in

Page 105: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

92

CCC with temperature is also observed for the samples with lower Tg. On the other hand, the most

likely reason why the clusters can withstand the shear at larger sizes is the presence of coalescence,

occurring only at gT T . When coalescence is present, even at rather low extent, the polymer actually

flows from one particle to the other as eq. (5.4) shows. A significantly higher force is thus needed to

take apart the particles, which can now also dissipate some of the stress by viscous relaxation.

Let us now discuss more quantitatively the forces coming into play in the breakup phenomenon. First

of all, the effect of the applied shear has to be consideredError! Reference source not found.. The

maximum shear rate in this experimental system running at 500 rpm is equal to 17300 s-1, evaluated

from the scaling for the maximum dissipation rate proposed by Soos et al [115]. It is hence possible to

compute the hydrodynamic stress 25

8hydF d to which the formed clusters should be able to resist

in order to allow gelation in the rheometer experiments. This hydrodynamic force, corresponding to

,cg gelR , is equal to 283 nN. The hydrodynamic forces to which the clusters withstand in the stirred-tank

experiments are reported in Figure 5.5. As one can see, as long as the temperature stays below the glass

transition temperature of the polymer, the force to which the clusters can withstand stays at least one

order of magnitude below the force to which the critical clusters for gelation in the rheometer are

subjected. This explains very clearly why no gelation can occur below the glass transition temperature

for these clusters. It is however observed that the critical value of 283 nN is not reached in the stirred

tank experiments. This is due to the under-estimation of the cluster size due to the loss of material by

adhesion on the reactor. However, these results show clearly that the criticality in gelation ability is due

to the strengthening of the bonds between the particles constituting the clusters. Moreover, the particle-

particles bonds keep on reinforcing when the coalescence process is eased by increasing the

temperature.

Page 106: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

93

Figure 5.5 : hydrodynamic force to which the clusters are subjected as a function of temperature in

the stirred-tank experiments.

Let us now summarize the understanding of the process of gelation under shear in the presence of

coalescence. First of all, if the temperature is set below the Tg of the particles, the binding strength upon

particle contact is too low to allow large clusters to form under shear. This limits the reachable range of

occupied volume fraction, not allowing gelation to occur at the primary particle volume fraction used

in these experiments.

If the temperature is set at (or close to) the Tg, the particles start coalescing. This gives rise to an increase

in contact surface area, which reflects in an increased binding strength. This allows the particles to form

larger clusters upon contact, hence allowing gelation.

If the temperature is above the Tg, three scenarios can be distinguished, depending on the relative

aggregation and coalescence times. Here it must be pointed out that the characteristic coalescence time

is not a parameter that is trivial to define. As already discussed above, the coalescence rate depends on

the size of the coalescing object. Hence small clusters may be coalescing faster than they aggregate, but

large clusters of the same material could well be unaffected by coalescence in their whole structure, but

only in the small size scales. Therefore, their occupied volume fraction and fractal dimension could not

be affected. Keeping the above discussed phenomena in mind, one can define the following cases:

25 30 35 40 450

25

50

75

100 60 % MMA 70 % MMA

Fhy

d /nN

Temperature /°C

Page 107: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

94

If τC >> τA, the main effect will be to accelerate the aggregation rate constant and decrease the breakup

rate constant due to colloidal destabilization and increased binding strength, leading to a faster gelation.

If τC ≈ τA, the final effect will depend on the significance of the two processes at elevated temperature.

If the coalescence rate increases faster than the particles are destabilized, the gelation time will increase.

If on the contrary the coalescence rate is only weakly affected by temperature, the gelation will occur

faster due to the aforementioned reasons.

If τC << τA, the formed clusters turn into spheres faster than they grow. This leads to an impossibility for

them to percolate, hence preventing any gelation.

An interesting point is to consider the effect of the parameters that were set fixed in the present article.

Let us for example analyse the effect of changing the initial particle volume fraction. If the volume

fraction is decreased, the rate of aggregation will be decreased, due to the fact that aggregation is a

bimolecular process. This leads to a slowing down of the cluster size increase with time. It means that

each cluster spends more time at one size before aggregating further and growing larger. As eq. (5.4)

shows, the coalescence characteristic time increases with increasing cluster size. This means that at

lower volume fractions and at a defined dimensionless time / At , the clusters will have coalesced

to a higher extent. This allows the clusters to ultimately grow to larger sizes before reaching their critical

breakup rate, possibly allowing a faster gelation.

This discussion explain why it is non-trivial to predict even qualitatively the effect of temperature on

the gelation time of a sheared suspension. It highly depends on the specific interaction potential for the

studied system, on the evolution of coalescence characteristic time with temperature, on the specificities

of adhesion for hard particles, on surface tension, on the volume fraction and the shear rate.

Page 108: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

95

5.4. Conclusion

In this article, the intricate effects governing shear-induced gelation of soft colloidal particles have been

studied. To this purpose, same-sized MMA/BA copolymer latex of various glass transition temperatures

have been synthesized by starved emulsion polymerization.

An initial characterization in stagnant conditions revealed how changing the aggregation temperature

of MMA/BA copolymer particles lead to a significant decrease of non-DLVO repulsive forces. A

generalized stability model has been used in order to rationalize the relative effects of electrostatics and

non-DLVO forces in stagnant aggregation.

Using stirred-tank diluted experiments, it has been shown that crossing the Tg of polymer particles

significantly affects the aggregates strength and thus also their breakup rate, even if the temperature

change is relatively small.

Moreover, it has been shown that increasing temperature while above the Tg can lead to two opposite

effects on gelation time: i) a reinforcement of the clusters, decreasing the breakup rate and hence the

gelation time, or ii) such an higher extent of coalescence that the occupied volume fraction decreases,

therefore delaying the gelation. The outcome of a shearing experiment thus depends on the characteristic

times of aggregation and coalescence.

In summary, aggregation, coalescence and cluster breakup all have a significant influence on gelation

time under shear, and are all influenced by temperature, either independently (such as interaction

potential), or depending on one another (such as decreased breakup upon increased coalescence state).

This article identifies independently the main factors influencing these phenomena, allowing to gain a

deeper understanding of the scientifically and industrially relevant process that is colloidal gelation.

Page 109: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

96

Page 110: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

97

Chapter 6

The Counter-Intuitive Aggregation of poly(Acrylic Acid)-Grafted

Nanoparticles after Surfactant Addition

6.1 Introduction

The production of polymers regularly makes use of emulsion polymerization. This process gives rise to

particles of lyophobic polymer dispersed in a continuous phase, which is often water. The polymer

particles so generated are in a metastable state, with the equilibrium configuration being a phase-

separated system. These lyophobic particles hence have to be stabilized by introducing a repulsive force

between the particles. The nature of this repulsive force can be electrostatic, steric, or linked to the

nature of the surface, such as hydration forces [9, 45, 116, 117].

On an industrial level, the most cost-effective polymerization conditions is to maximize the polymer

volume fraction, while avoiding aggregation. To this purpose, several stabilization mechanisms can be

combined. In this study, acrylic copolymer particles synthesized in semi-batch emulsion polymerization

are studied. The stabilization strategy applied during the synthesis here was double: on one side, a

standard amount of anionic surfactant was used, and on the other side, small amounts of polymerizable,

water-soluble weak acid were added to the monomer feed. More specifically, 0%-2% acrylic acid (AA)

was added to the hydrophobic methyl methacrylate/butyl acrylate mixture. This synthesis method has

already been shown to give rise to a hydrophilic, pH-responsive shell consisting mainly of acrylic acid,

while the more hydrophobic monomers form the core of the particles[61]. The synthesized particles

Page 111: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

98

exhibit remarkable colloidal stability when the acrylic acid is deprotonated (pH > 7.0), while a decreased

stability is observed when the pH is decreased below the pKa of the functional monomer.

The goal of this study is to expend the stability study to the case where acrylic acid is used in

combination with sodium dodecyl sulfate (SDS). To this purpose, the aggregation kinetics of several

latexes synthesized ad hoc were followed at different pH values by dynamic light scattering, with

varying amounts of AA and SDS in the particles and the dispersant, respectively. It is observed that

while SDS stabilizes the particles when the pH is low, it destabilizes dramatically the particles when

the AA-shell is deprotonated. In order to understand this phenomenon at the molecular level, molecular

dynamics simulations were performed to identify a “reversed” SDS adsorption on the particle surface.

This adsorption mechanism relies on salt bridges between AA and the sulfate head of the surfactant,

mediated by the high sodium concentration in the bulk.

6.2 Materials and Methods

6.2.1 Particles synthesis

The studied latexes were supplied and synthesized ad hoc by BASF SE. They are copolymers of methyl

methacrylate and butyl acrylate in equal weight fractions, in which desired amounts of Acrylic Acid

were added for stabilizing purposes. The particles were synthesized by semi-batch emulsion

polymerization, using ammonium persulfate (APS) radical initiation system. The latexes were

synthesized in presence of sodium dodecyl sulfate (SDS) as surfactant. The surfactant has subsequently

been removed by extended period of dialysis, which was carried until the surface tension of the mother

liquor was >70 mN/m and the conductivity < 10 µS/m.

6.2.2 Particles characterization

The characterization of the latexes was carried by Static and Dynamic light scattering, and the results

are summarized in Table 1:

Page 112: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

99

Name Main Monomers Functional Monomer

[% wt AA] Diameter [nm] PDI [-]

P0 MMA/BA None 161 0.002

P1 MMA/BA 1.0 171 0.005

P2 MMA/BA 2.0 191 0.005

P3 Styrene/BA 2.0 169 0.006

Table 6.1: Characterization of the latexes used

The NaOH, NaCl, HCl, H2SO4 and SDS were all purchased to Sigma-Aldrich with the highest purities

available (>99%) and used without further purification. The water used was filtered with a Millipore

Milli-Q® (MQ) system.

6.2.3 Aggregation experiments

The aggregation process was followed by dynamic light scattering (DLS) using a BI-200SM goniometer

system (Brookhaven Instruments, USA), equipped with a solid-state laser, Ventus LP532 (Laser

Quantum, U.K.) of wavelength 0=532 nm, as the light source, with scattering angle set to 90. The

sample was thermostated by a water bath (Julabo, Germany). The preparation procedure was as follows:

Diluted latex samples were prepared, with a mass fraction φw=5·10-4 in Milli-Q water. The 20 mL vials

were cleaned and dried with nitrogen. Milli-Q water, 2 M NaCl, 2% wt. SDS solution, 0.1 M H2SO4 or

10 mM NaOH solution were added to the vial, after having being freshly filtered through a 0.45 µm

syringe filter. The overhead was then flushed with Argon, in order to avoid pH variations during the

aggregation experiments. Then, the diluted latex was added rashly to the solution, which was

subsequently stirred and set in the light scattering instrument. The overall duration of the procedure of

latex addition and starting of the measurement was measured and always equal to 1 minute.

6.2.4 Computational methods

Atomic charges

Atomic charges for polymethyl methacrylate and polyacrylic acid were computed starting from small

oligomers composed by six repeating units with a syndiotactic configuration; chain ends were saturated

Page 113: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

100

with methyl groups. Two different structures were considered for PAA, in order to take into account

the different protonation states due to the pH of the solution. A fully ionized polyacrylic acid (that is,

where all carboxyl groups are present as carboxylate) was employed to mimic high pH values, whereas

a fully undissociated PAA was considered for low pH simulations.

These structures were firstly optimized in vacuo by means of density functional theory (DFT)

calculation at B3LYP/6-31G(d,p) level of theory[118-120]. The obtained geometries were used as input

for a further optimization in implicit water, described through integral equation formalism polarizable

continuum model (IEFPCM)[121, 122] at 300 K, at the same level of theory. Atomic charges were then

fitted starting from electrostatic potentials computed at B3LYP/6-311+G(d,p)[123] level of theory in

implicit water at 300 K, by means of RESP formalism[124, 125]. Charge fitting was performed with a

two-step procedure. In primis, atomic charges were determined by assigning the proper overall charge

value to the considered oligomer, i.e., 0 for PMMA and undissociated PAA, and -6 for the fully ionized

PAA. In the second step, charge equivalence for chemically equivalent atoms was imposed. Oxygen

atoms of carboxylate groups, e.g., have the same partial atomic charge, in order to take into account

charge delocalization due to resonance. The two central units and the ending units of each oligomer

were used for building the polymer chains used in molecular dynamics simulations.

The same charge derivation protocol (structure optimization, electrostatic potential calculation and

atomic charges fitting through RESP formalism) was used for sodium dodecyl sulfate, whose overall

charge is equal to -1. Computations were performed by means of Gaussian 09 software[126].

Creation of the molecular systems

The size of the primary particle is too big to allow an atomistic description of the entire system;

moreover, such analysis would be not meaningful, because the interactions with the surfactant occur

only at the polymer/solvent interface and not in the polymer bulk phase. Since the primary particle size

(100 nm) is much higher than the polyacrylic acid chain length in the fully stretched conformation (10

nm, assumed as the characteristic length at the particle/solvent interface), the system is modeled as a

flat grafted surface of polymethyl methacrylate.

Page 114: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

101

The PMMA surface was built as a random coil of 96 polymer chains containing 20 repeating units. Four

polyacrylic acid chains were positioned at 0.1535 nm from the surface and grafted to the closest PMMA

chain by applying a harmonic restraining potential k0(x – x0)2, where k0 and x0 are equal to 2.5363 · 105

KJ mol-1 nm-2 and 0.1535 nm respectively, representative of a chemical bond between sp3 carbon atoms

according to General Amber Force Field (vide infra). Polymer brushes were placed at a distance of 4.1

nm each other, assuming a uniform hexagonal packing. Each polyacrylic acid chain is composed by 40

repeating units, which corresponds to a length of 10 nm in the fully extended conformation and a

monomer density per unit PMMA surface of 4.5 · 10-6 mol m-2, consistently with the experimental

conditions (3.6 · 10-6 mol m-2). Two different systems were created, corresponding to high pH (where

polyacrylic acid chains are fully ionized) and low pH environment (in which undissociated polymer is

grafted on the surface).

a) b)

Figure 6.1: Molecular model for polymethyl methacrylate surface (transparent VdW spheres) grafted

with polyacrylic acid chains (black residues) in explicit water molecules (blue dots) (a); hexagonal

packing of PAA chains (solvent is omitted for the sake of clarity) (b). Each image contains the

simulation box replicated once along x and y.

Each grafted surface (which lies on the xy plane) was solvated with explicit TIP3P water

molecules[127]. No water molecules were added in the xy plane around the PMMA block; in other

words, the base of simulation box coincides with the PMMA random coil. Na+ ions were also added,

Page 115: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

102

where needed, in order to assure electroneutrality. This procedure results in a periodic box with an initial

size of 8.2 x 7.1 x 20 nm (where the PMMA thickness is 6.8 nm).

The initial structure is represented in Figure 1, where the simulation box has been replicated along x

and y direction in order to highlight the resulting hexagonal packing when periodic boundary conditions

are employed.

Both systems were equilibrated through 200 ns molecular dynamics simulations in order to reach a

stable arrangement of the brushes on the surface. The attainment of an equilibrated geometry was

checked through time evolution of Solvent Accessible Surface (SAS) and maximum chain height from

the surface related to polyacrylic acid.

Sodium dodecyl sulfate was then added to the periodic box, and additional 500 ns molecular dynamics

simulations were performed for all systems (high and low pH, with and without SDS). According to

experimental conditions, SDS is in large excess with respect to polyacrylic acid; at both pH, the number

of SDS molecules added is equal to the number of PAA monomers (160).

Parameterization and simulation protocol

Polymethyl methacrylate and polyacrylic acid were parameterized according to General Amber Force

Field (GAFF)[128], which proved to be suitable for the simulation of amphiphilic and charged

polymers[129, 130]; ion parameters optimized for TIP3P water molecules were taken from Joung and

Cheatham works[131, 132].

Molecular dynamics simulations were performed by means of GROMACS package, version 5.0.2[133].

Electrostatic long-range interactions were computed through the Particle Mesh Ewald (PME)

method[134], adopting a cut-off value equal to 12 Å; the same value was employed for Lennard-Jones

interactions. Neighbor list was updated every 5 fs, and all covalent bonds involving hydrogen atoms

were restrained through LINCS algorithm[135]. A time step equal to 2 fs, along with Leap-Frog

algorithm, were employed to compute molecular trajectories.

The adopted computational protocol was the following: first of all, energy minimization was applied in

order to remove bad contacts between the solute and the randomly placed solvent molecules.

Page 116: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

103

Temperature was raised to 300 K through 200 ps in NVT ensemble; a small harmonic restraint was

applied to the solute in order to avoid wild fluctuations. Density and pressure were equilibrated through

10 ns in NPT ensemble at 300 K and 1 atm. During this phase, a harmonic restrain was applied to

polyacrylic acid chains in order to avoid artifacts. Finally, molecular dynamics simulations were

performed in NPT ensemble at 300 K and 1 atm. Temperature and pressure were controlled through

velocity rescale thermostat[136] and semi-isotropic barostat[137] respectively. Data were collected

every 20 ps.

6.3 Results and discussion

6.3.1 Latex aggregation kinetics

In order to establish a base case, let us investigate the aggregation kinetics of the P0 latex, i.e. the “bare”

particles. This latex is stable after dialysis only thanks to the relatively low charge density of its surface,

coming from the radical initiator used, which leaves an alkyl sulfate group at the end of each polymer

chain. A relatively high concentration of NaCl (0.6 M) is used in order to screen the electrostatic

repulsion between the particles, allowing the acrylic acid containing particles to aggregate. All the

aggregation measurements are performed at low volume fraction (φ=10-5), which allows to follow the

hydrodynamic radius change as a function of time in situ. The results of these experiments are shown

in Figure 6.2a.

Page 117: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

104

Figure 6.2: Kinetics of aggregation for P0 (a), P1 (b), P2 (c and d), as a function of pH (a, b, and d)

and of surfactant concentration (c). The surfactant concentration for a,b, and d is always equal to

[SDS] = 0.01 %w/w. The pH value for c) is equal to pH = 9.5

As expected, the change in pH has no effect on the kinetics of aggregation of this low-surface charge

density latex, since no group is to be protonated with changing pH. The presence of SDS allows the

surface to get sufficiently charged and hydrophilic to avoid aggregation. It must be pointed here that

the used SDS concentration, while being very high in terms of surface coverage due to the low polymer

volume fraction, stays significantly below the CMC[138, 139]. The presence of a small amount of

micelles cannot be excluded, but it is observed that this amount is not high enough to induce aggregation

due to depletion forces[140]. The presence of SDS simply stabilizes the particles by adsorption.

Let us now investigate what the effect of surfactant on aggregation kinetics is for the highest amount of

AA-shell (2% Acrylic Acid, P2) at high pH (=9.5). As well as the effect of pH at fixed surfactant

concentration (=0.01% SDS). The experimental conditions are the same as in the P0 latex case. The

results are showed in Figure 6.2c and d, respectively.

0 10 20 30 40 50 60

100

200

300

0 20 40 60 80 10075

100

125

150

175

200

0 20 40 60 80 100 120 140 160

100

200

300

400

500

0 20 40 60 80 100 120 140100

200

300

400

500

pH = 1.08, SDS pH = 8.1, SDS pH = 1.78, no SDS pH = 8.01, no SDS

Rh

/nm

Time /min

a)

d)c)

b) pH = 1.13 pH = 2.92 pH = 6.16 pH = 11.43

Rh

/nm

Time /min

0% SDS 0.002%SDS 0.01% SDS 0.02% SDS

Rh

/nm

Time /min

pH = 1.02 pH = 1.92 pH = 3.24 pH = 9.51 pH = 11.43

Rh

/nm

Time /min

Page 118: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

105

Two effects are to be observed here: on one hand, increasing the surfactant concentration at high pH

(>9.0) tends to destabilize the P2 particles. On the other hand, it is remarkable that decreasing the pH

below the pKa of polyacrylic acid [30, 61] has a stabilizing effect on these particles. It has been observed

previously with very similar particles that decreasing the pH below the pKa of acrylic acids tends to

destabilize the particles in absence of surfactant [61]. This effect is expected, since the particles are

stabilized by a hydrophilic charged polyelectrolyte shell. Decreasing the pH below the pKa of the

polyelectrolyte leads to protonation of the charged units, hence decreasing simultaneously the surface

charge (electrostatic repulsion), and the steric stabilization coming from the hydrophilicity of the

polymer. Here, the opposite trend is observed: in the presence of surfactant only, decreasing the pH re-

stabilizes the particles. Decreasing the pH to extreme values (pH=1.02) leads to a slight increase in

aggregation rate. This is due to the protonation of the sulfate groups, generated during the emulsion

polymerization[61].

The stabilizing impact of lower pH is even clearer if the amount of acrylic acid in the polymer is lower.

(Figure 6.2b). For the P1 particles, the aggregation does not occur when the pH is lowered.

These results indicate clearly that the combination of the presence of surfactant, combined with the

deprotonated poly-(acrylic acid) surface and high salt concentration leads to the aggregation. This effect

cannot be due to side effects of the presence of SDS, such as depletion forces, since the experiments

take place below the CMC of SDS. The phenomenon does not happen in the absence of acrylic acid,

and only occurs at high pH, where the solubility of SDS is increased.

In order to investigate this aggregation phenomenon in more details, and more specifically the

mechanism of SDS adsorption on the particles, a fourth latex was synthesized (P3) in which methyl

methacrylate was replaced by styrene. Styrene is significantly more hydrophobic than MMA, and is

able to adsorb a significantly higher amount of SDS than the latter[141]. It is observed in Figure a with

P3 that no destabilization is observed in the presence of SDS, independently on the pH value. This

observation hints strongly towards the fact that the adsorption of SDS on the particle surface is

responsible for the observed destabilizing effect for P1 and P2. More accurately, the amount of

surfactant adsorbed on the particle hydrophobic core modulates the repulsive interactions between the

particles, preventing the destabilizing effect coming from the acrylic acid combination.

Page 119: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

106

In order to complete the experimental investigation of the aggregation mechanism, it was decided to

make the P1 latex aggregate as reported in Figure 6.2b. during 60 min at pH=11.43, and then the pH

value was lowered by addition of H2SO4 to pH=3.02 and short mild agitation of the vial using a glass

rod. It was observed that the aggregates formed at high pH redispersed, while the suspension stayed

stable during the following hour. This indicates clearly that the aggregation mechanism of these

particles in the presence of SDS is not related to what is observed in absence of SDS, where increasing

the pH to re-stabilize the particles thanks to the acrylic acid is only able to stop the aggregation process.

Figure 6.3: Aggregation kinetics of P3 at different pH values (a) and P1 (b) with 0.01% w/w SDS

concentration. The pH value for (b) was lowered from 11.43 to 3.02 using sulfuric acid at time = 60

min.

Summarizing this article’s finding, it is observed that the combination of a highly charged

polyelectrolyte layer, a relatively high salt concentration (0.6 M NaCl), and surfactant leads to a

dramatic decrease in colloidal stability. The usual stabilizing components of the DLVO theory, together

with the added contribution of the electro-steric stabilization due to polymer brushes are not able to

explain why this combination leads to destabilization. The presence of surfactant, adsorbing on the

surface due to its hydrophobic tail, should stabilize the particles, and an increased pH value should

deprotonate the polyelectrolyte brush, thus leading to an increased stabilization. The exact opposite

trends are observed.

0 20 40 60 80 100 120

100

125

150

175

200

0 20 40 60 80 100 120 14075

100

125

150

175

200

225b) 1.16

8.95 11.45

Rh

/nm

Time /min

a)

Rh

/nm

Time /min

Page 120: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

107

The previously exposed experimental results were not reported so far, and reveal a rather

counterintuitive behavior in the interaction of poly-(acrylic acid) coated particles. In order to investigate

and understand the mechanistic reason for such trends, it was decided to simulate a similar system using

molecular dynamics.

6.3.3 Molecular dynamics simulations

Surface equilibration

First of all, 200 ns simulations have been performed at both low and high pH without SDS, in order to

obtain an equilibrated arrangement for the grafted surface.

At low pH, the undissociated polyacrylic acid chains collapse on the polymethyl methacrylate surface,

because of the attainment of hydrophobic interactions between PAA aliphatic backbone and PMMA,

as shown in Figure 4a. On the other hand, at high pH the ionized polymer chains interact each other

through Na+ bridges between the negatively charged carboxyl groups, thus avoiding the collapse on the

particle surface (Figure 6.4b).

The time evolution of PAA solvent accessible surface is represented in figure 4c and d; in particular,

the values have been normalized with respect to the corresponding values exhibited in the fully stretched

conformation of the polymeric chain, where the SAS is maximum.

At low pH, SAS decrease is mainly due to the hydrophobic fraction of the surface (Figure 6.4c), whose

exposure to the solvent is reduced by the collapse on PMMA surface driven by the hydrophobic

interactions between polymers.

On the contrary, at high pH the hydrophobic SAS fraction remains essentially unaffected, while the

hydrophilic one decreases of about 30% with respect to its initial value (Figure 6.4d). Chain interactions

through Na+ bridges, indeed, reduce the exposure to the solvent of charged carboxyl moieties.

Page 121: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

108

a) b)

c) d)

Figure 6:. Representative equilibrated strucures at low pH (a) and high pH (b); Na+ ions are

represented as blue dots, while water is omitted for the sake of clarity; normalized Solvent Accessible

Surface as a function of simulation time at low pH (c) and high pH (d).

SDS effect on stability

As mentioned, after surface equilibration SDS molecules were added in the simulation box and

additional 500 ns were performed for all systems (high and low pH, with and without surfactant) in

order to obtain a meaningful comparison and a better understanding of the surfactant effect. System

equilibration was again verified through the convergence of polyacrylic acid SAS and chain height; the

analysis of molecular trajectories was performed considering the last 100 ns.

Generally speaking, the behavior of SDS molecules can be rationalized by identifying three different

situations. The surfactant can remain free in solution, can be adsorbed on PMMA surface (because of

hydrophobic interactions between the polymer and the aliphatic tail) or can interact with PAA chains.

Page 122: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

109

Molecular trajectories showed that the interactions between the surfactant and polymer brushes depend

on PAA protonation state, i.e. on pH.

At low pH, the surfactant is mainly bound to polymer brushes through hydrophobic interactions

between the hydrophobic tail of SDS and PAA backbone. The polar group of the surfactant can interact

with PAA carboxyl moieties through hydrogen bonds, but it remains exposed to the solvent.

Notably, at high pH the binding between polyacrylic acid and SDS occurs through Na+ bridges between

the dissociated carboxyl moieties and the negatively charged sulfate groups. This electrostatic-driven

aggregation can be highlighted by means of radial distribution function between oxygen atoms of PAA

carboxyl groups and sulfur atom of SDS (Figure 6.5).

a) b)

Figure 6.5. Radial distribution function between PAA oxygen atoms and SDS sulphur atom (a);

example of salt bridge between PAA and SDS (b).

Radial distribution function shows that the interaction between the SDS sulfate group and carboxyl

moieties is stronger at high pH thanks to the attainment of salt bridges. As mentioned, at low pH

hydrogen bonds between undissociated carboxyl moieties and the polar head of SDS can occur, but

such interactions are less favored if compared to the electrostatic-driven binding at high pH.

Because of this arrangement, the hydrophobic tail of the surfactant remains exposed to the solvent, thus

increasing the hydrophobicity of the surface and promoting particle aggregation. Such phenomenon can

Page 123: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

110

be quantified by computing the overall solvent accessible surface of PAA and the surfactant bound to

the polymer brushes, as highlighted in figure 6.6.

a) b)

Figure 6.6. Fraction of hydrophobic and hydrophilic overall solvent accessible surface (PAA + bound

SDS) at low pH, before and after SDS addition (a); fraction of hydrophobic and hydrophilic overall

solvent accessible surface (PAA + bound SDS) at high pH, before and after SDS addition (b).

While at low pH the addition of the surfactant does not appreciably modify the hydrophilicity of the

exposed surface, at high pH the presence of SDS increases the hydrophobicity of the particle/solvent

interface, thus favoring colloidal destabilization of the latex.

6.4 Concluding remarks

In this chapter, the example of the sometimes counter-intuitive behavior of colloidal particles is shown.

It also demonstrates the efficiency of using colloidal aggregation as a tool in order to investigate the

properties of the surfaces. From the kinetics of aggregation of some industrial latexes, an interesting

effect was observed: adding anionic surfactant, usually used in order to improve the stability of colloidal

systems, brings instability in the system, making particles aggregate in pH ranges where the rest of the

stabilization strategy is the most efficient. This “reversed” adsorption of SDS on the surface, mediated

by salt bridges, also gives rise to the possibility of creating pH-sensitive systems, where reversible

aggregation is observed. Such systems could be of significant interest for the general nanoparticle

technological applications, as well as in simpler polymer production. In fact, if the aggregation is

Page 124: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

111

reversible, its presence in a manufacturing process is less likely to lead to a catastrophic failure of the

process.

The present chapter is equally an arguments towards the collaboration between industrial chemistry and

molecular dynamics. A counter-intuitive behavior was observed experimentally, but only simulations

could explain it on the physical level. The possibility of confirming the presence of reversed adsorption

experimentally is relatively small, since probing the surface of nanoparticles would require the use of

small-angle neutron scattering, which is not a readily accessible setup.

Page 125: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

112

Page 126: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

113

Chapter 7

Concluding remarks

The production and characterization of industrial polymeric latexes gives rise to a multitude of complex

behaviors which, even if rather well understood independently, prevent any educated guessing in their

combined effects. In the present work, the case of copolymeric film-forming particles stabilized with

polyelectrolytes was studied. First, the effect of the polyelectrolyte layer was studied in stagnant

conditions, revealing the pH-responsiveness of the weak polyelectrolyte brush and of its stabilizing

effect, which was shown to act not only on the electrostatic level, but also gives rise to steric

stabilization. In the third chapter, the effect and pH responsiveness of the steric stabilization was

examined while reducing the electrostatic contribution by addition of salt. A steric interaction model

with pH-dependent Flory parameter was devised in order to gain knowledge about the characteristics

of the particles’ surfaces. In the fourth chapter, the coalescing properties of “bare” particles synthesized

to exhibit different glass transitions were simulated using a population balance equation model for the

aggregation in stagnant conditions. In the fifth chapter, the effect of coalescence on shear-induced

aggregation and gelation was studied experimentally and revealed systematically the presence

competition between aggregation, coalescence and breakup in this process. In the last chapter, a

counterintuitive reversed surfactant adsorption mechanism was revealed by molecular dynamics,

leading to aggregation of stable particles covered with polyelectrolyte brushes.

Overall, the present work is a strong argument in favor of using even simplified mathematical models

and simulations in order to tackle complex practical industrial problems. Starting from an intricate

system where several features are expected to play a role, the best strategy is to take each of the effects

Page 127: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

114

independently, studying it systematically from the experimental and theoretical point of view, before

coming back to the full complexity of the initial problem.

Page 128: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

115

Chapter 8

Appendix

8.1 Experimental proof for the coalescence between particles

One crucial difference between coalescence and non-coalescence systems is that during the aggregation,

the total surface area of the particles in the system decreases for the former and remains constant for the

latter. Then, if an ionic surfactant is adsorbed on the particle surface, when coalescence occurs, the

reduction in the total particle surface area changes the surfactant adsorption equilibrium, and more

surfactant molecules are adsorbed on the particles, leading to increase in the colloidal stability. The

consequence is that one would observe that the doublet formation rate decreases with time for a

coalescence system, while it remains constant for a non-coalescence one.[48, 53]

Therefore, to demonstrate the occurrence of coalescence for our colloidal particles, we have

measured the rate constant for the doublet formation, K1,1 for P3 latex in the presence of SDS, based on

the technique described in the text. The results are shown in Figure 8.1. It is seen that the K1,1 value

decreases monotonically as the time increases, confirming that the total surface area decreases along

the coalescence during the doublet formation, leading to increase in the adsorbed surfactant molecules

on the particle surface, thus the stability.

However, if no surfactant is present in the system, even if coalescence occurs, the K1,1 value does

not change with time. An example is shown in Figure 8.2, for P2 latex in the absence of surfactant. This

arises because even though coalescence may change the density of the fixed charges on the doublet

Page 129: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

116

surface, it does not change the fixed charge density on the primary particles, thus without effect on K1,1.

Figure 8.1 Doublet formation rate constant, K1,1, determined from the initial aggregation kinetics at

CNaCl=0.50 mol/L for P3 latex in the presence of surfactant SDS. The solid curve is guiding the eyes.

Figure 8.2 Doublet formation rate constant, K1,1, determined from the initial aggregation kinetics for

P2 latex in the absence of surfactant.

2E‐25

2E‐24

2E‐23

0.0E+0 5.0E+3 1.0E+4 1.5E+4

K1,

1, m

3 /s

t, s

1E-24

1E-23

1E-22

0 500 1000 1500 2000

K1,

1, m

3 /s

t, s

Page 130: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

117

8.2 - time-evolution of the average fractal dimension

,,

,f

x td x t

x t

(S1)

3

13

1

, ,,

,,

, ,

f f

f

f

d f x y t ddx td d d

d x tdt dt x t dt

f x y t dd

(S2)

3

13

1

3 3 3 3

1 1 1 123

1

, ,,

, ,

, , , ,, , , ,

, ,

f f ff

f f

f f

f f f f f f f f

f f

d f x d t dddd x t d

dt dtf x d t dd

f x d t f x d td dd f x d t dd d f x d t dd dd

t t

f x d t dd

(S3)

, ,

3 3 3

1 1 13 3 3

1 1 1

, ,,

, , , ,, ,

,

, , , , , ,

, , ,1 1,

, ,

f

x t x t

t t

f f

f f f f f ff

f f f f f f

x t x td x t

ff

f x d t f x d td dd d f x d t dd dd

dd x t t t

dtf x d t dd f x d t dd f x d t dd

dd x t x t xd x t

dt x t t x t

tt

(S4)

Page 131: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

118

1 1 1

0

1 1 1 1 1

0

1 1 1

0

1 1

, , ,1 1,

, ,

1, 3 ,

, 1, , , t ,

,

1, , , , ,

2

, , ( , , ) d1

,, 1

( ,2

ff

fCOAL

f

x

f

f

dd x t x t x td x t

dt x t t x t t

x t d x tx

dd x tx t x x x t dx

dt x t

d x t x x x t x x t x t dx

x t x t x x t x

d x tx t

x x x

1 1 1

0

, ) , , dx

t x x t x t x

(S5)

1 1 1

0

1

1 1 1 1 1

0

2

1 1

0

1 1, 3 ,

,

, ,, , ,

,

,1, , , ,

2 ,

,, , ( , ,

,

fCOAL

f

T

xf

T

f

x t d x tx t x

dd x t x tx x t x t dx

dt x t

d x tx x x t x x t x t dx

x t

d x tx t x t x x t

x t

1

3

1 1 1 1 1

0

4

)d

,1( , , ) , , d

2 ,

T

xf

T

x

d x tx x x t x x t x t x

x t

(S6)

Comparing the terms T1 and T3, and T2 and T4 respectively, one gets:

Page 132: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

119

1 1 1

0

,

,,1 3 , , ,

, ,

f

f

d x t

d x tx tT T x x t x t dx

x t x t

,x t

1 1 1

0

1 1 1 1 1 1

0 0

, ( , , ) d

1 3 , , , , , , ( , , ) d 0f f

x t x x t x

T T d x t x t x x t dx d x t x t x x t x

(S7)

2 4 1 1 1 1 1

0

1 1 1 1 1

0

,1, , , ,

2 ,

,1( , , ) , , d 0

2 ,

xf

xf

d x tT T x x x t x x t x t dx

x t

d x tx x x t x x t x t x

x t

(S8)

Therefore, considering (S7) and (S8) the final form of equation (S6) reads:

, 1

,fdd x t

dt x t

1,

COAL

x tx

3 ,

, 13 ,

f

ff

COAL

d x t

dd x td x t

dt x

(S9)

Page 133: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

120

8.3 – Derivation of average quantities h,effR x, t and gR x,t

8.3.1 Calculation of h,effR x,t

In the following, the Kirkwood-Riseman (KR) theory[80, 142, 143] is extended to the case of

clusters with partially-overlapping primary particles, by using Rotne-Prager’s tensor corrected for

particles overlapping.[144, 145]

Let us start by considering the general formalism of KR theory. The basic idea is to use the

linear relationship between force and velocity, so that for each particle in a cluster one can write:

6i p i iR F u v (S10)

where Fi is the force acting on the ith particle, ui is the ith particle velocity and vi is the fluid velocity at

the center of the ith particle. The fluid velocity is written as a combination of the unperturbed fluid

velocity and of the perturbations due to all other particles in the cluster:

1

'N

i i ij jj ì

v v T F (S11)

The perturbation term is computed using the Kirkwood-Riseman theory. The meaning of Equation

(S11) is that the velocity perturbation on the ith particle due to the jth particle is proportional to the force

applied to the latter. The proportionality term is given by the Oseen tensor, which defines the velocity

profile created by a point-like force located at the center of the jth particle:

2

1

8i j

ijc ij ijR R

R RT I (S12)

where Rij is the distance between the centers of the ith and the jth particles. According to the original

formulation of KR theory, particles are considered as point forces in the calculation of the perturbation

they cause to the surrounding fluid. It is however possible to utilize a more appropriate formulation,

which extends the Oseen tensor to the case where particles size is accounted for. The corresponding

tensor is the Yamakawa (or Rotne-Prager) tensor:[144, 145]

Page 134: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

121

2

2 2 2

21 1

8 3i j p i j

ijc ij ij ij ij

R

R R R R

R R R RT I I (S13)

In the case of particles partially interpenetrated, such as those where (partial) coalescence is occurring,

the following form of Rotne-Prager tensor can be used:[146]

1 9 3

16 32 32

ij i jij

p p p ij

R

R R R R

R RT I (S14)

Equation (S14) is only valid when the interparticle distance 2ij pR R . For interparticle distances

larger than the sum of the particle diameter, Equation (S13) should instead be used. At this point, we

will proceed as usual in the case of rigid clusters. We will assume that each particle is subject to the

same average perturbation (mean field approach), and that the particle-particle correlation function can

be used to describe the distribution of particles around the average particle in the cluster. One can

observe that according to Equation (S11) the perturbation that a particle experiences depends on the

relative position of the other particles, and that such perturbation is not spherically symmetric. When

using a mean-field approach, it is instead assumed that, inside a cluster, each particle is surrounded by

all other particles uniformly distributed around it. It is therefore sufficient to compute the average of

Equations (S13) and (S14) over all possible orientations, which leads to a spherically symmetric

perturbation field having the following form:[80]

1

for 26ij ij pang

ij

R RR

T I (S15)

and:

1 1

1 for 26 4

ijij ij pang

p p

RR R

R R

T I (S16)

In this manner, we can write the following equation for the total hydrodynamic force experienced by a

cluster moving with a velocity u in a viscous fluid (obtained after a few algebraic manipulations):

2

0

6

1 4

pNR

g r r h r dr

F u (S17)

Page 135: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

122

In Equation (S17), the function h(r) is the following:

11 for 2

4

for 2

pp

pp

rh r r R

R

Rh r r R

r

(S18)

Therefore, the final equation for the cluster hydrodynamic radius is the following:

2

0

,

1 4 , ,

ph

xR tR x t

g r x t r h r dr

(S19)

One can observe that, if there is no overlapping between particles, equation (S19) is identical to the one

developed for fractal clusters, since one only needs to use the second expression for h(r) in Equation

(S18). More explicitly, the integral in Equation (S19) needs to be split in two parts, as follows:

2

2

0 2

,1

1 4 , , 1 4 , ,4

p

p

ph R

pp R

xR tR x t

rg r x t r dr R t g r x t rdr

R t

(S20)

Even the particle-particle correlation function needs to be modified. In fact for DLCA clusters it

becomes:

f

f

,

nn2

3p

3

p,

0 if r ,

if r= ,4·

( , , ) · if r 2 ,

· ·exp if r 2 ,,

D

b

b

d

x

x

d

x t

x t

x

x

t

x t

tt

Nr t

t

ag r x t rR

c rr

R

(S21)

where t is the center-to-center distance of two particles in contact for a given extent of coalescence

(if 2 pt R t , there is no coalescence, otherwise 2 pt R t ) not to be confused with D

which is instead the Dirac delta function. Note that pR t and t are calculated by solving the

following equation system:

Page 136: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

123

2

303

2

2 14

4

1 22 2

ADp

p

pp

p p

S tt R t

R t

RR t

t t

R t R t

(S22)

where 0pR is the radius of a single primary particle, while pR t is the radius of a primary particle

embedded in a doublet, undergoing coalescence (assuming that the coalescing particles maintain their

spherical shape). ADS t is the surface area of a coalescing doublet which can be estimated from the

following equation:

/min

min max minCtAD AD

ADC

dS t S t SS t S S S e

dt

(S23)

where minS is the surface area of a fully coalesced doublet, while maxS the one of a non-coalesced

doublet (to any extent). Both pR t and t change in time as a function of the coalescence extent.

t tends to zero as full coalescence implies an overlap of the geometrical centers of the two

coalescing particles. pR t instead is equal to primary particles radius 0pR if no coalescence occurs,

and reaches the value of 1/302 pR at full coalescence (i.e. the radius of a sphere obtained by the complete

fusion of two spheres of radius 0pR ). More details can be found in Jia et al.[74]

Several input parameters are required to employ the , ,g r x t function (cf. equation (S21)),

namely the cut-off length t , the cut-off exponent x and the empirical functions

, , nna x b x N x and ,c x t . As for ,x t , the cut-off length is defined as follows:[81]

1/ ,, 1.45 fd x t

px t R t x (S24)

The empirical functions , , na x b x N x and x , can be estimated instead using the following

empirical equation:[81]

Page 137: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

124

m

m

x eF x d

x e z

(S25)

where x is the cluster mass and F x represents the generic function to be estimated (

, ,a x b x x or nnN x ). The values of the empirical parameters of equation (S25) are available

in the literature and are summarized in Table S1.[81]

Table S1

Values of the parameters , ,d e z and m in DLCA condition to be used in equation (S25), in order to

calculate the functions , ,a x b x x and nnN x in equation (S21).

Parameters of the g r function d e z m

nnN 2.0342 1.1477 0.9997 1

a 0.0095 4.1292 0.1997 2

b * 0.6425 6.2352 5.1747 1

2.1976 3.8377 -0.1784 1

* Note: for 7x , 0b

The function ,c x t is obtained by imposing the following equality:

2

0

1 4 , ,x r g r x t dr

(S26)

As a result, one gets:

3 3

3

,

4 11 2

3,

, ,, 241 ,

,

f

b x b x

nn bp

d x

I

p

t

f fNC

a xx N x t t

R t b xc x t

d x t d x t

R

x t t

x t x x t x

(S27)

Page 138: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

125

Having clarified the different parameters occurring in the pair correlation function , ,g r x t , it is

possible to continue the integration of equation (S20), which can be re-written as follows:

2 22 3

0 0 2

, , ,

,

1 4 , , , , 4 , ,p P

P

A B C

Ph R R

PP R

I x t I x t I x t

xR tR x t

g r x t r dr g r x t r dr R t g r x t rdrR t

(S28)

Two different cases have to be distinguished, namely i) pt R t and ii) pt R t ,

as the integration boundaries have to be appropriately chosen.

In case i), for pt R t the three integrals become:

3 3nn3

p

,

1,

, 2 , ,, ,

24

2, ,

, ,

· 3

f

x x

d x t

f p f fINC

b

A

Np

C

b

I

bx xt t

t x

d x t R t d x t d x tx t x t t

x

N aI x t

R b

c

t x xR x t x t x

(S29)

f

4 4nn3

p

,

p

1

1, 2

4· 4

, , 1 , 1, 1 2 2, ,

,

,

,

f

b x b x

b

x t

B

df p

d

f fINC INC

I

d R t d d

N x a xx t t t t

R t b x

x t

x t x t

t x t x tc x t x t

R t x x x x

(S30)

p

1,

,

1, 121

,,, ,

,

f

f

df f

C i

x t

x t ncd

x t x tdc x t tx

dtI

x x x tR xt

t

(S31)

Page 139: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

126

While for case ii), for pt R t :

3 3nn

3p

1, 2

4· 3

b x b x

A pb x

N x a xI x t R t t

b xR t

(S32)

4 4nn

3p

1, 2

4· 4

b x b x

pbB

a xNx t t R t t

R t bI

x

(S33)

f

22nn3

p

p

1

1, 2 2

4· 2

, ,, 1 1, 21 ,

,

b xb x

pC

dff f

i

b

d nc

N x a xx t t R t

t R t b x

x t x tc x t t

I

R t x x

d dx t

t xx

(S34)

After calculating ,hR x t according to the proper case, the following equation is employed to link it

to the , ,h effR x t , accounting for the rotational diffusivity:

, 12

,,

2 ,1 0.5 1 1

3 ,

hh eff

g

f

R x tR x t

qR x t

d x t

(S35)

To calculate the average hydrodynamic radius, hR t also the scattering structure factor has to be

known. This quantity can be obtained from the following integral:

2

0

1 sin( )( , , ) 1 4 ( , , )

qrS x q t r g r x t dr

qrx

(S36)

Note that only the first term of the particle-particle correlation function, corresponding to the Dirac

delta peak of the nearest neighbors, can be evaluated analytically:

2

0

sin1 sin( )( , , ) 1 4 ( , , )nn

q qrS x q t N x r g r

xx t dr

q qr

(S37)

The rest of the integral of the pair-correlation function in equation (S37) has been obtained by numerical

quadrature.

Page 140: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

127

8.3.2 Calculation of gR x,t

The equation for the radius of gyration from the particle-particle correlation function (the same as for

the case of , ,h effR x t , equation (S21)) in the case of a non-coalescing cluster is the following:

2 2 4, 0

4( , ), ,

2g g pR R r g r x t drx

x t

(S38)

In the case of clusters undergoing coalescence, the integral remains the same, while the first term on

the right hand side, which is the radius of gyration of a primary particle, would need to be modified. It

has to become the radius of gyration of the unit element of a cluster undergoing coalescence. This is

not a well-defined object and many approximations need to be introduced. Since the relevance of this

term, especially for large clusters is almost negligible, we will keep it as the non-coalesced primary

particle radius of gyration. Therefore, the integration of equation (S38) using equation (S21) leads to

the following result:

5 52 2, 3

22

2

,

412

2 5

4 2 22· 1 ,

2

1,

, , ,,

,

f

b b

g g p nn bp

d

p f fINC

p

x t

x t t t t tt

x t t x t x tt

a xR R N x

x b x R

c R d d

x x R x x

x t

t x t

(S39)

The function ,c x t appearing in Equation (S39) is defined as for the hydrodynamic radius, cf.

equation (S27). Similarly, the functions , , , nna x b x x N x and ,x t have been previously

discussed (Table S1).

Page 141: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

128

8.4 Derivation of the discretized Equations

Let us start to treat the PBE on the cluster concentration ,x t reported in equation (12) (in the main

text). The approximation for the latter balance, stated: 2

1

,G

i i

Ns x x

ii

x t t e

. Note that the GN

Gaussians are placed in the positions ix and have heights i . In order to reduce the problem size, let

us evaluate the function ,x t in the jx positions where the Gaussians are placed:

2

1

,G

i j i

Ns x x

j ii

x t t e

(S40)

Considering that [1... ]Gj N , one can re-write equation (S40) GN times:

21 2

2 211 1 1 2 1 2

2 2 22 21 2 1 2 2 2

2

1 1 21

21 2

1

1

,

,

,

G

i i

N NG G

G

G

i i NG

G

G Gi N iG

Ns x x

i s x xs x x s x xi NN

s x x s xs x x s x xi N

i

N Ns x x

ii

t et e t e t ex t

x t t e t e t e t e

x t

t e

2

2 2 2

1 1 2 2

1 2

NG

N N N N NG G G G G

G

x

s x x s x x s x x

Nt e t e t e

(S41)

Equation (S41) can be manipulated as follows:

22 2

11 1 1 2 1 2

22 2

21 2 1 2 2 2

2 2 2

1 1 2 2

1 1

2 2

,

,

,

N NG G

N NG G

GG N N N N NG G G G G

s x xs x x s x x

s x xs x x s x x

NN s x x s x x s x x

e e ex t tx t te e e

tx te e e

(S42)

Defining then 2expi j i j ix s x x one gets:

Page 142: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

129

1 1 2 1 11 1

2 1 2 2 2 2 2

1 2

,

,

,

G

G

GG G G G G

N

N

NN N N N N

C

x x xx t tx t x x x t

tx t x x x

(S43)

A more compact writing of equation (S43) is then:

C (S44)

Where the matrix C , of size G GN N , is the so-called change of base matrix, allowing to “move”

from the cluster-mass space to the Gaussian one.

For further details on the method and to appreciate the versatility of this methodology, literature papers

should be considered,[77, 147, 148] where different applications systems (polymers, colloidal particles,

protein fibrils) have been approached with this technique. The vectors and matrices occurring in the

discretized balance are defined as follows.

The C matrix is reported in equation (S43), whereas the weighted 1 2,W W vectors read:

1

2

1 1

2 2

11

12

1/ , 1/ ,

1 1

1/ , 1/ ,

2 21 2

1/ , 1/ ,

0 0 0 0 0 0

0 0 0 0 0 0

0 0 0 0

f f

f f

f N f NG G

G G

W

W

d x t d x t

d x t d x t

d x t d x t

N N

C W C

C W C

x x

x xW W

x x

(S45)

The elements of the matrix (of size 2N N ) and of the vector size

2 1N ), are defined

as:

Page 143: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

130

,

2

, , ,

,

,

exp

i j i jj i

i j i j i j

i j i j

i ji j

i j

s s

x s x x

x x x

s ss

s s

(S46)

The elements of the W vector of size 2 1N are calculated as follows:

1 2,

W WWi j i j

j is s

(S47)

Page 144: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

131

8.5 Parameter Values employed in the simulations

Table S2

Viscosities of the continuous phases ( C ) containing 4 M NaCl at different temperatures.

oT C C Pa s

25 31.27 10

30 31.14 10

35 31.04 10

45 48.67 10

Page 145: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

132

Table S3

Parameters used for simulations reported in Figure 2-3-4-5-6

Parameter Value

Bk 231.38 10 / KJ

T 298.15 K

C 31.27 10 Pa s

0pR 71.0 10 m

5 5 51.0 10 / 2.0 10 / 3.0 10

n 1.332

90 o

DLS 9532 10 m

C

70% MMA 60% MMA 50% MMA 30% MMA

810 s 34.3 10 s 21.0 10 s

Page 146: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

133

Table S4

Parameters used for simulations reported in Figure 7, S1, S2

Parameter Value

Bk 23 11.38 10 KJ

T 298.15 / 303.15 / 308.15 / 313.15 K

C 3 3 3 41.27 10 /1.14 10 /1.04 10 /8.67 10Pa s Pa s Pa s Pa s

0pR 71.0 10 m

51.0 10

n 1.332

90 o

DLS 9532 10 m

C

298.15 K 303.15 K 308.15 K 318.15 K

50 % MMA 34.3 10 s 31.1 10 s 23.6 10 s 22.1 10 s

30 %MMA 21.0 10 s - 21.0 10 s -

Page 147: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

134

8.6 Measured Fractal dimensions

Table S5 – fd t measurement of the different particles after 120 min of DLCA aggregation from

two different sources, ,S q t vs q and 0gR vs I

51 10 52 10

53 10

MMA % ,S q t vs q 0gR vs I ,S q t vs q

0gR vs I ,S q t vs q 0gR vs I

30 - 2.45 - 2.33 - 2.98

50 1.77 1.77 1.8 1.78 1.80 1.86

60 1.73 1.86 1.74 1.73 1.78 1.87

70 1.81 1.77 1.80 1.87 1.80 1.72

Page 148: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

135

8.7 Results at higher temperatures

Figure S1 hR t vs N of the 30% MMA particles at 25oC (blue squares) and 35oC (red circles)

along with their corresponding simulations (continuous lines, which overlap as full coalescence is

reached already at 25oC). All the parameter values employed to obtain the model predictions are

reported in Table S4 (Appendix 8.5).

N[-]

0 20 40 60

Rh(t

) [n

m]

0

100

200

300

400

Page 149: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

136

Figure S2 C versus 1/T for the 50% MMA containing particles, using 343.70 10A s and

42.55 10B and equation (31) of the main text

Note that the term B is of the same order of magnitude as the ones reported in the literature.[71, 74]

Being interested in producing partially coalesced clusters, the present model could be used to calibrate

the coalescence process exploiting the temperature dependence in Figure S2 and equation (31) in the

main text; such dependence is readily available once a few selected experimental points are obtained.

Page 150: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

137

8.8 Symbol List for chapter 4

Symbol Meaning

a x Function in equation (S21)

b x Function in equation (S21)

,c x t Function in equation (S21)

C Base of change matrix to “move” from Gaussian to concentration space

d Parameter in equation (S25)

fd Internal coordinate of the distribution function representing the fractal dimension

,fd x t Average fractal dimension of an x -sized cluster at time t

fmd Fractal dimension of the aggregation regime

pD Particle diameter

e Parameter in equation (S25)

, ,ff x d t 2-D particle distribution function

F x Function to calculate , , ,na x b x N x x

jF Drag fore acting on the jth particle in a cluster

, ,g r x t Particle-particle correlation function

1 1 2

2 1 2 1 2

,

, , ,f f

g x x

g x x d d

Constitutive laws

Bk Boltzmann constant

m Parameter in equation (S25)

N Ratio between coalescence and aggregation characteristic time (equation (27) in the main text

nN Average number of nearest neighbor particles

q wave vector

r Radial distance from the center of a reference particle

,m jr Distance between the center of the mth and the jth particles

Page 151: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

138

,gR x t radius of gyration of an x-sized cluster

gR t Average radius of gyration of the whole cluster mass distribution

,heffR x t Effective hydrodynamic radius of an x-sized cluster accounting for rotational diffusivity

,hR x t Hydrodynamic radius of an x-sized cluster

hR t Average hydrodynamic radius of the whole cluster mass distribution

0pR Primary particle radius

pR t Radius of a primary particle (embedded in a doublet) undergoing coalescence

is Gaussian overlapping parameter

, ,S x q t Structure factor of an x-sized cluster, measured at a wave vector q and at time t

ADS t Surface area of a coalescing doublet

minS Minimum surface area of a doublet (i.e. fully coalesced doublet)

maxS Maximum surface area of a doublet (i.e. doublet with no coalescence extent)

t Time

T System temperature

jmT Oseen (S12) or Yamakawa (S13) tensor

u Cluster velocity

, fv x d rate of coalescence

'jv Velocity perturbation experienced by the th particle in a cluster

x Internal coordinate in the distribution functions representing the cluster mass

Cluster friction tensor

1 2,W W Weight matrixes for numerical method

z Parameter in equation (S25)

Greek Letters

i Coefficient of the Gaussians used to approximate the distribution function

Page 152: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

139

1 2 1 2, , ,f fx x d d

1 2,x x

Aggregation kernels used in the 2-D and 1-D PBEs

Cut-off exponent in the correlation function

, W Vectors employed in the discretized PBE

x Euler gamma function

INC x Euler incomplete gamma function

D Dirac Delta function

c Continuous phase viscosity

p Particle viscosity

Cut-off length in the correlation function

p Particle surface tension

A Characteristic time of doublet aggregation

C Characteristic time of doublet coalescence

COAL x Characteristic time of coalescence of an x-sized cluster

N Non-dimension time

P Characteristic time of the whole aggregation process

Occupied volume fraction

,x t 1-D distribution function obtained as 0th order moment of the 2-D distribution function , ,ff x d t

,x t 1-D distribution function obtained as 1st order moment of the 2-D distribution function , ,ff x d t

Page 153: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

140

Page 154: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

141

References

1. Elias, H.-G. and R. Mülhaupt, Plastics, General Survey, 2. Production of Polymers and Plastics, in Ullmann's Encyclopedia of Industrial Chemistry. 2000, Wiley-VCH Verlag GmbH & Co. KGaA.

2. Elisabetta, C., A Comparison of Tackified, Miniemulsion Core−Shell Acrylic Latex Films with Corresponding Particle-Blend Films: Structure−Property Relationships. Langmuir, 2009. 25(18): p. 11021-11031.

3. Jovanovic, R., Emulsion-based pressure-sensitive adhesives: A review. Polymer Reviews, 2004. C44(1): p. 1-51.

4. Dkhissi, Y., S. Meyer, D. Chen, H.C. Weerasinghe, L. Spiccia, Y.-B. Cheng, and R.A. Caruso, Stability Comparison of Perovskite Solar Cells Based on Zinc Oxide and Titania on Polymer Substrates. ChemSusChem, 2016. 9(7): p. 687-695.

5. Karg, M., T.A.F. König, M. Retsch, C. Stelling, P.M. Reichstein, T. Honold, M. Thelakkat, and A. Fery, Colloidal self-assembly concepts for light management in photovoltaics. Materials Today, 2015. 18(4): p. 185-205.

6. Kannan, J.M.F.K., Mechanically stable antimicrobial chitosan–PVA–silver nanocomposite coatings deposited on titanium implants. Carbohydrate polymers, 2015. 121: p. 37-48.

7. Trinh, N.D., Preparation and characterization of silver chloride nanoparticles as an antibacterial agent. Advances in Natural Sciences: Nanoscience and Nanotechnology, 2015. 6(4): p. 45011-45011.

8. Andrade, F.A.C., L.C.d.O. Vercik, F.J. Monteiro, and E.C.d.S. Rigo, Preparation, characterization and antibacterial properties of silver nanoparticles–hydroxyapatite composites by a simple and eco-friendly method. Ceramics International, 2016. 42(2, Part A): p. 2271-2280.

9. Israelachvili, J.N., Intermolecular and surface forces. 3rd ed. 2011, Burlington, MA: Academic Press. xxx, 674 p.

10. Chapter 1 - Introduction, in Studies in Interface Science, D. Jan K.G, Editor. 1996, Elsevier. p. 1-67.

11. Derjaguin, B.V. and L. Landau, Theory of the Stability of Strongly Charged Lyophobic Sols and of the Adhesion of Strongly Charged Particles in Solutions of Electrolytes. Acta Phys. Chim. URSS, 1941. 14: p. 633-662.

12. von Smoluchowski, M., Experiments on a mathematical theory of kinetic coagulation of coloid solutions. Zeitschrift Fur Physikalische Chemie--Stochiometrie Und Verwandtschaftslehre, 1917. 92(2): p. 129-168.

13. Chapter 9 - Phase Separation Kinetics, in Studies in Interface Science, D. Jan K.G, Editor. 1996, Elsevier. p. 559-634.

14. Chapter 8 - Critical Phenomena, in Studies in Interface Science, D. Jan K.G, Editor. 1996, Elsevier. p. 495-558.

15. Keddie, J.L., Film formation of latex. Materials Science and Engineering: R: Reports, 1997. 21(3): p. 101-170.

16. Kovalchuk, N.M. and V.M. Starov, Aggregation in colloidal suspensions: Effect of colloidal forces and hydrodynamic interactions. Advances in Colloid and Interface Science, 2012. 179–182: p. 99-106.

17. Lattuada, M., P. Sandkühler, H. Wu, J. Sefcik, and M. Morbidelli, Aggregation kinetics of polymer colloids in reaction limited regime: experiments and simulations. Adv. Colloid Interface Sci., 2003. 103(1): p. 33-56.

18. Zaccone, A., D. Gentili, H. Wu, and M. Morbidelli, Shear-induced reaction-limited aggregation kinetics of Brownian particles at arbitrary concentrations. The Journal of Chemical Physics, 2010. 132(13): p. 134903.

19. Xie, D., H. Wu, A. Zaccone, L. Braun, H. Chen, and M. Morbidelli, Criticality for shear-induced gelation of charge-stabilized colloids. Soft Matter, 2010. 6(12): p. 2692-2698.

Page 155: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

142

20. Douady, A. and J.H. Hubbard, Étude dynamique des polynômes complexes. 1984: Université de Paris-Sud, Département de Mathématique.

21. Lattuada, M., A. Zaccone, H. Wu, and M. Morbidelli, Population-balance description of shear-induced clustering, gelation and suspension viscosity in sheared DLVO colloids. Soft Matter, 2016. 12(24): p. 5313-5324.

22. Zaccone, A., H. Wu, D. Gentili, and M. Morbidelli, Theory of activated-rate processes under shear with application to shear-induced aggregation of colloids. Physical Review E, 2009. 80(5): p. 051404.

23. Harshe, Y.M., M. Lattuada, and M. Soos, Experimental and Modeling Study of Breakage and Restructuring of Open and Dense Colloidal Aggregates. Langmuir, 2011. 27(10): p. 5739-5752.

24. Lazzari, S., B. Jaquet, H. Wu, and M. Morbidelli, Shear stability of inverse latexes during their polymerization process. AIChE Journal, 2015. 61(4): p. 1380-1384.

25. Advincula, R.C., W.J. Brittain, K.C. Caster, J. Rühe, and (Editors), Polymer brushes. 2004, Weinheim: Wiley-VCH.

26. Jusufi, A., C.N. Likos, and M. Ballauff, Counterion distributions and effective interactions of spherical polyelectrolyte brushes. Colloid Polym. Sci., 2004. 282(8): p. 910-917.

27. Claesson, P.M., E. Poptoshev, E. Blomberg, and A. Dedinaite, Polyelectrolyte-mediated surface interactions. Adv. Colloid Interface Sci., 2005. 114-115: p. 173-187.

28. Ulbricht, M. and H. Yang, Porous polypropylene membranes with different carboxyl polymer brush layers for reversible protein binding via surface-initiated graft copolymerization. Chem. Mater., 2005. 17(10): p. 2622-2631.

29. Zhou, F. and W.T.S. Huck, Surface grafted polymer brushes as ideal building blocks for "smart" surfaces. Phys. Chem. Chem. Phys., 2006. 8(33): p. 3815-3823.

30. Ballauff, M., Spherical polyelectrolyte brushes. Prog. Polym. Sci., 2007. 32(10): p. 1135-1151. 31. Jain, P., G.L. Baker, and M.L. Bruening, Applications of polymer brushes in protein analysis

and purification. Annu. Rev. Anal. Chem., 2009. 2(1): p. 387-408. 32. Hinrichs, K., D. Aulich, L. Ionov, N. Esser, K.-J. Eichhorn, M. Motornov, M. Stamm, and S.

Minko, Chemical and structural changes in a pH-responsive mixed polyelectrolyte brush studied by infrared ellipsometry. Langmuir, 2009. 25(18): p. 10987-10991.

33. Polzer, F., J. Heigl, C. Schneider, M. Ballauff, and O.V. Borisov, Synthesis and analysis of zwitterionic spherical polyelectrolyte brushes in aqueous solution. Macromolecules, 2011. 44(6): p. 1654-1660.

34. Comrie, J.E. and W.T.S. Huck, Exploring actuation and mechanotransduction properties of polymer brushes. Macromol. Rapid Commun., 2008. 29(7): p. 539-546.

35. Ito, Y., M. Inaba, D.J. Chung, and Y. Imanishi, Control of water permeation by pH and ionic strength through a porous membrane having poly(carboxylic acid) surface-grafted. Macromolecules, 1992. 25(26): p. 7313-7316.

36. de Vos, W.M., G. Meijer, A. de Keizer, M.A. Cohen Stuart, and J.M. Kleijn, Charge-driven and reversible assembly of ultra-dense polymer brushes: formation and antifouling properties of a zipper brush. Soft Matter, 2010. 6(11): p. 2499-2507.

37. Evers, F., C. Reichhart, R. Steitz, M. Tolan, and C. Czeslik, Probing adsorption and aggregation of insulin at a poly(acrylic acid) brush. Phys. Chem. Chem. Phys., 2010. 12(17): p. 4375-4382.

38. Dai, J., Z. Bao, L. Sun, S.U. Hong, G.L. Baker, and M.L. Bruening, High-capacity binding of proteins by poly(acrylic acid) brushes and their derivatives. Langmuir, 2006. 22(9): p. 4274-4281.

39. Ghosh, S., W. Jiang, J.D. McClements, and B. Xing, Colloidal stability of magnetic iron oxide nanoparticles: Influence of natural organic matter and synthetic polyelectrolytes. Langmuir, 2011. 27(13): p. 8036-8043.

40. Ross, R.S. and P. Pincus, The polyelectrolyte brush: poor solvent. Macromolecules, 1992. 25(8): p. 2177-2183.

41. Zhulina, E.B. and O.V. Borisov, Poisson–Boltzmann theory of pH-sensitive (annealing) polyelectrolyte brush. Langmuir, 2011. 27(17): p. 10615-10633.

Page 156: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

143

42. Zhulina, E.B., O.V. Borisov, and T.M. Birshtein, Structure of grafted polyelectrolyte layer. J. Phys. II France, 1992. 2(1): p. 63-74.

43. Bartels, C. and D. Ronis, Competition between conformational and chemical equilibrium in suspensions of polyelectrolyte-coated particles. Macromolecules, 2011. 44(8): p. 3174-3178.

44. Schüwer, N. and H.-A. Klok, Tuning the pH sensitivity of poly(methacrylic acid) brushes. Langmuir, 2011. 27(8): p. 4789-4796.

45. Pincus, P., Colloid Stabilization with Grafted Polyelectrolytes. Macromolecules, 1991. 24(10): p. 2912-2919.

46. Jones, A. and B. Vincent, Depletion flocculation in dispersions of sterically-stabilised particles 2. Modifications to theory and further studies. Colloids and Surfaces, 1989. 42(1): p. 113-138.

47. Zaccone, A., H. Wu, M. Lattuada, and M. Morbidelli, Correlation between colloidal stability and surfactant adsorption/association phenomena studied by light scattering. J. Phys. Chem. B, 2008. 112(7): p. 1976-1986.

48. Gauer, C., H. Wu, and M. Morbidelli, Effect of surface properties of elastomer colloids on their coalescence and aggregation kinetics. Langmuir, 2009. 25(20): p. 12073-12083.

49. Bohren, C.F. and D.R. Huffman, Absorption and Scattering of Light by Small Particles. 1983, New York: Wiley.

50. Ohshima, H. and T. Kondo, Electrophoresis of large colloidal particles with surface charge layers. Position of the slipping plane and surface layer thickness. Colloid Polym. Sci., 1986. 264(12): p. 1080-1084.

51. Lozsán R, A.E. and M.S. Romero-Cano, Carboxylated core–shell particles: I. A system showing hindered swelling behavior. J. Colloid Interface Sci., 2009. 339(1): p. 36-44.

52. Martell, A.E. and R.M. Smith, Critical stability constants. 1989, New York: Plenum Press. 53. Jia, Z.C., C. Gauer, H. Wu, M. Morbidelli, A. Chittofrati, and M. Apostolo, A generalized model

for the stability of polymer colloids. J. Colloid Interface Sci., 2006. 302(1): p. 187-202. 54. Jia, Z.C., H. Wu, and M. Morbidelli, Application of the generalized stability model to polymer

colloids stabilized with both mobile and fixed charges. Ind. Eng. Chem. Res., 2007. 46(16): p. 5357-5364.

55. Ehrl, L., Z.C. Jia, H. Wu, M. Lattuada, M. Soos, and M. Morbidelli, Role of counterion association in colloidal stability. Langmuir, 2009. 25(5): p. 2696-2702.

56. Vincent, B., J. Edwards, S. Emmett, and A. Jones, Depletion flocculation in dispersions of sterically-stabilised particles ("soft spheres"). Colloid Surf., 1986. 18(2-4): p. 261-281.

57. Fritz, G., V. Schädler, N. Willenbacher, and N.J. Wagner, Electrosteric Stabilization of Colloidal Dispersions. Langmuir, 2002. 18(16): p. 6381-6390.

58. Prabhu, V.M., M. Muthukumar, G.D. Wignall, and Y.B. Melnichenko, Polyelectrolyte chain dimensions and concentration fluctuations near phase boundaries. The Journal of Chemical Physics, 2003. 119(7): p. 4085-4098.

59. Vesaratchanon, J.S., K. Takamura, and N. Willenbacher, Surface characterization of functionalized latexes with different surface functionalities using rheometry and dynamic light scattering. Journal of Colloid and Interface Science, 2010. 345(2): p. 214-221.

60. Crassous, J.J., L. Casal-Dujat, M. Medebach, M. Obiols-Rabasa, R. Vincent, F. Reinhold, V. Boyko, I. Willerich, A. Menzel, C. Moitzi, B. Reck, and P. Schurtenberger, Structure and Dynamics of Soft Repulsive Colloidal Suspensions in the Vicinity of the Glass Transition. Langmuir, 2013. 29(33): p. 10346-10359.

61. Jaquet, B., D. Wei, B. Reck, F. Reinhold, X. Zhang, H. Wu, and M. Morbidelli, Stabilization of polymer colloid dispersions with pH-sensitive poly-acrylic acid brushes. Colloid and Polymer Science, 2013. 291(7): p. 1659-1667.

62. Daswani, P. and A. van Herk, Entry in Emulsion Copolymerization: Can the Existing Models be Verified? Macromolecular Theory and Simulations, 2011. 20(8): p. 614-620.

63. Cosgrove, T., Colloid science principles, methods and applications. 2nd ed. 2010, Chichester, U.K.: Wiley. 375 S.

64. Berg, J.C., An introduction to interfaces & colloids - the bridge to nanoscience. 2010, Hackensack: World Scientific. 785.

65. Lin, M.Y., H.M. Lindsay, D.A. Weitz, R.C. Ball, R. Klein, and P. Meakin, Universality in Colloid Aggregation. Nature, 1989. 339(6223): p. 360-362.

Page 157: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

144

66. Ulrich, G.D. and N.S. Subramanian, Particle Growth in Flames .3. Coalescence as a Rate-Controlling Process. Combustion Science and Technology, 1977. 17(3-4): p. 119-126.

67. Koch, W. and S.K. Friedlander, The Effect of Particle Coalescence on the Surface-Area of a Coagulating Aerosol. Journal of Colloid and Interface Science, 1990. 140(2): p. 419-427.

68. Xiong, Y., M.K. Akhtar, and S.E. Pratsinis, Formation of Agglomerate Particles by Coagulation and Sintering .2. The Evolution of the Morphology of Aerosol-Made Titania, Silica and Silica-Doped Titania Powders. Journal of Aerosol Science, 1993. 24(3): p. 301-313.

69. Xiong, Y. and S.E. Pratsinis, Formation of Agglomerate Particles by Coagulation and Sintering .1. A 2-Dimensional Solution of the Population Balance Equation. Journal of Aerosol Science, 1993. 24(3): p. 283-300.

70. Buesser, B., A.J. Grohn, and S.E. Pratsinis, Sintering Rate and Mechanism of TiO2 Nanoparticles by Molecular Dynamics. Journal of Physical Chemistry C, 2011. 115(22): p. 11030-11035.

71. Tsantilis, S., H. Briesen, and S.E. Pratsinis, Sintering time for silica particle growth. Aerosol Science and Technology, 2001. 34(3): p. 237-246.

72. Heinson, W.R., C.M. Sorensen, and A. Chakrabarti, Computer Simulation of Aggregation with Consecutive Coalescence and Non-Coalescence Stages in Aerosols. Aerosol Science and Technology, 2010. 44(5): p. 380-387.

73. Gauer, C., Z.C. Jia, H. Wu, and M. Morbidelli, Aggregation Kinetics of Coalescing Polymer Colloids. Langmuir, 2009. 25(17): p. 9703-9713.

74. Jia, Z.C., H. Wu, and M. Morbidelli, Thermal restructuring of fractal clusters: The case of a strawberry-like core-shell polymer colloid. Langmuir, 2007. 23(10): p. 5713-5721.

75. Sajjadi, S., Particle formation under monomer-starved conditions in the semibatch emulsion polymerisation of styrene. Part II. Mathematical modelling. Polymer, 2003. 44(1): p. 223-237.

76. Kostoglou, M. and A.G. Konstandopoulos, Evolution of aggregate size and fractal dimension during Brownian coagulation. Journal of Aerosol Science, 2001. 32(12): p. 1399-1420.

77. Kryven, I. and P.D. Iedema, A Novel Approach to Population Balance Modeling of Reactive Polymer Modification Leading to Branching. Macromolecular Theory and Simulations, 2013. 22(2): p. 89-106.

78. Kostoglou, M., A.G. Konstandopoulos, and S.K. Friedlander, Bivariate population dynamics simulation of fractal aerosol aggregate coagulation and restructuring. Journal of Aerosol Science, 2006. 37(9): p. 1102-1115.

79. Eggersdorfer, M.L., D. Kadau, H.J. Herrmann, and S.E. Pratsinis, Multiparticle Sintering Dynamics: From Fractal-Like Aggregates to Compact Structures. Langmuir, 2011. 27(10): p. 6358-6367.

80. Lattuada, M., H. Wu, and M. Morbidelli, Hydrodynamic radius of fractal clusters. Journal of Colloid and Interface Science, 2003. 268(1): p. 96-105.

81. Lattuada, M., H. Wu, and M. Morbidelli, A simple model for the structure of fractal aggregates. Journal of Colloid and Interface Science, 2003. 268(1): p. 106-120.

82. Kestin, J., H.E. Khalifa, and R.J. Correia, Tables of the Dynamic and Kinematic Viscosity of Aqueous Kci Solutions in the Temperature-Range 25-150-Degrees-C and the Pressure Range 0.1-35 Mpa. Journal of Physical and Chemical Reference Data, 1981. 10(1): p. 57-70.

83. Bohren, C.F. and D.R. Huffman, Absorption and scattering of light by small particles. 2015, Weinheim: Wiley-VCH Verlag. 700 S.

84. Israelsen, N.D., D. Wooley, C. Hanson, and E. Vargis, Rational design of Raman-labeled nanoparticles for a dual-modality, light scattering immunoassay on a polystyrene substrate. Journal of Biological Engineering, 2016. 10(1): p. 1-12.

85. Alam, R., D.M. Fontaine, B.R. Branchini, and M.M. Maye, Designing quantum rods for optimized energy transfer with firefly luciferase enzymes. Nano Lett, 2012. 12.

86. Vasquez, E.S., J.M. Feugang, S.T. Willard, P.L. Ryan, and K.B. Walters, Bioluminescent magnetic nanoparticles as potential imaging agents for mammalian spermatozoa. Journal of Nanobiotechnology, 2016. 14(1): p. 1-9.

87. Kuo, C.-H., W. Li, L. Pahalagedara, A.M. El-Sawy, D. Kriz, N. Genz, C. Guild, T. Ressler, S.L. Suib, and J. He, Understanding the Role of Gold Nanoparticles in Enhancing the Catalytic

Page 158: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

145

Activity of Manganese Oxides in Water Oxidation Reactions. Angewandte Chemie International Edition, 2015. 54(8): p. 2345-2350.

88. Kong, X., J. Pakusch, D. Jansen, S. Emmerling, J. Neubauer, and F. Goetz-Neuhoeffer, Effect of polymer latexes with cleaned serum on the phase development of hydrating cement pastes. Cement and Concrete Research, 2016. 84: p. 30-40.

89. Israelachvili, J.N., Intermolecular and Surface Forces (Third Edition). 2011, Academic Press: San Diego. p. 23-51.

90. Warson, H., Emulsion polymerization, a mechanistic approach. R. G. Gilbert. Academic Press, London, 1995. Polymer International, 1996. 41(3): p. 352-352.

91. Sorensen, C.M., Light Scattering by Fractal Aggregates: A Review. Aerosol Science and Technology, 2001. 35(2): p. 648-687.

92. Berg, J.C., An Introduction to Interfaces and Colloids : The bridge to Nanoscience. 2009, Singapore: World Scientific Publishing Co Pte Ltd.

93. Lu, P.J., E. Zaccarelli, F. Ciulla, A.B. Schofield, F. Sciortino, and D.A. Weitz, Gelation of particles with short-range attraction. Nature, 2008. 453(7194): p. 499-503.

94. Zebrowski, J., V. Prasad, W. Zhang, L.M. Walker, and D.A. Weitz, Shake-gels: shear-induced gelation of laponite–PEO mixtures. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 2003. 213(2–3): p. 189-197.

95. Mohraz, A. and M.J. Solomon, Orientation and rupture of fractal colloidal gels during start-up of steady shear flow. Journal of Rheology, 2005. 49(3): p. 657-681.

96. Becker, V., E. Schlauch, M. Behr, and H. Briesen, Restructuring of colloidal aggregates in shear flows and limitations of the free-draining approximation. Journal of Colloid and Interface Science, 2009. 339(2): p. 362-372.

97. Eggersdorfer, M.L., D. Kadau, H.J. Herrmann, and S.E. Pratsinis, Aggregate Morphology Evolution by Sintering: Number & Diameter of Primary Particles. Journal of aerosol science, 2012. 46: p. 7-19.

98. Dahiya, P., M. Caggioni, and P.T. Spicer, Arrested coalescence of viscoelastic droplets: polydisperse doublets. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 2016. 374(2072).

99. van der Kooij, H.M., M. de Kool, J. van der Gucht, and J. Sprakel, Coalescence, Cracking, and Crack Healing in Drying Dispersion Droplets. Langmuir, 2015. 31(15): p. 4419-4428.

100. Karyu, N., M. Noda, S. Fujii, Y. Nakamura, and Y. Urahama, Effect of adhesive thickness on the wettability and deformability of polyacrylic pressure-sensitive adhesives during probe tack test. Journal of Applied Polymer Science, 2016. 133(27): p. n/a-n/a.

101. Lazzari, S., B. Jaquet, L. Colonna, G. Storti, M. Lattuada, and M. Morbidelli, Interplay between Aggregation and Coalescence of Polymeric Particles: Experimental and Modeling Insights. Langmuir, 2015. 31(34): p. 9296-9305.

102. Jia, Z., C. Gauer, H. Wu, M. Morbidelli, A. Chittofrati, and M. Apostolo, A generalized model for the stability of polymer colloids. Journal of Colloid and Interface Science, 2006. 302(1): p. 187-202.

103. Martell, A.E. and R.M. Smith, Inorganic Ligands, in Critical Stability Constants: First Supplement. 1982, Springer US: Boston, MA. p. 393-424.

104. Manciu, M. and E. Ruckenstein, Role of the Hydration Force in the Stability of Colloids at High Ionic Strengths. Langmuir, 2001. 17(22): p. 7061-7070.

105. Delgado-Calvo-Flores, J.M., J.M. Peula-Garcı́a, R. Martı́nez-Garcı́a, and J. Callejas-Fernández, Experimental Evidence of Hydration Forces between Polymer Colloids Obtained by Photon Correlation Spectroscopy Measurements. Journal of Colloid and Interface Science, 1997. 189(1): p. 58-65.

106. Mohanty, K.K., Porous media: Fluid transport and pore structure. By F. A. L. Dullien, Academic Press, 574 pp., 1992. AIChE Journal, 1992. 38(8): p. 1303-1304.

107. Ehrl, L., M. Soos, and M. Lattuada, Generation and Geometrical Analysis of Dense Clusters with Variable Fractal Dimension. The Journal of Physical Chemistry B, 2009. 113(31): p. 10587-10599.

Page 159: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

146

108. Soos, M., A.S. Moussa, L. Ehrl, J. Sefcik, H. Wu, and M. Morbidelli, Effect of shear rate on aggregate size and morphology investigated under turbulent conditions in stirred tank. Journal of Colloid and Interface Science, 2008. 319(2): p. 577-589.

109. Verwey, E.J.W., Theory of the Stability of Lyophobic Colloids. The Journal of Physical and Colloid Chemistry, 1947. 51(3): p. 631-636.

110. Shih, W.Y., J. Liu, W.-H. Shih, and I.A. Aksay, Aggregation of colloidal particles with a finite interparticle attraction energy. Journal of Statistical Physics, 1991. 62(5): p. 961-984.

111. Johnson, K.L., K. Kendall, and A.D. Roberts, Surface Energy and the Contact of Elastic Solids. Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, 1971. 324(1558): p. 301-313.

112. Maugis, D. and H.M. Pollock, Surface Forces, Deformation and Adherence at Metal Microcontacts. Acta Metallurgica, 1984. 32(9): p. 1323-1334.

113. Richeton, J., S. Ahzi, K.S. Vecchio, F.C. Jiang, and R.R. Adharapurapu, Influence of temperature and strain rate on the mechanical behavior of three amorphous polymers: Characterization and modeling of the compressive yield stress. International Journal of Solids and Structures, 2006. 43(7–8): p. 2318-2335.

114. Eggersdorfer, M.L. and S.E. Pratsinis, Restructuring of aggregates and their primary particle size distribution during sintering. AIChE Journal, 2013. 59(4): p. 1118-1126.

115. Soos, M., R. Kaufmann, R. Winteler, M. Kroupa, and B. Lüthi, Determination of maximum turbulent energy dissipation rate generated by a rushton impeller through large eddy simulation. AIChE Journal, 2013. 59(10): p. 3642-3658.

116. Ballauff, M., Spherical polyelectrolyte brushes. Progress in Polymer Science, 2007. 32(10): p. 1135-1151.

117. Derjaguin, B. and L. Landau, Theory of the Stability of Strongly Charged Lyophobic Sols and of the Adhesion of Strongly Charged-Particles in Solutions of Electrolytes. Progress in Surface Science, 1993. 43(1-4): p. 30-59.

118. Becke, A.D., Density functional thermochemistry. III. The role of exact exchange Journal of Chemical Physics, 1993. 98(7): p. 5648 - 5652.

119. Lee, C., W. Yang, and R.G. Parr, Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Physical Review B, 1988. 37(2): p. 785 - 789.

120. Ditchfield, R., W.J. Hehre, and J.A. Pople, Self Consistent Molecular Orbital Methods. IX. An Extended Gaussian Type Basis for Molecular Orbital Studies of Organic Molecules Journal of Chemical Physics, 1971. 54(2): p. 724 - 728.

121. Cances, E., B. Mennucci, and J. Tomasi, A new integral equation formalism for the polarizable continuum model: Theoretical background and applications to isotropic and anisotropic dielectrics. Journal of Chemical Physics, 1997. 107(8): p. 3032-3041.

122. Tomasi, J., B. Mennucci, and R. Cammi, Quantum mechanical continuum solvation models. Chemical Reviews, 2005. 105(8): p. 2999-3093.

123. Clark, T., J. Chandrasekhar, G.W. Spitznagel, and P.v.R. Schleyer, Efficient diffuse function-augmented basis sets for anion calculations. III. The 3-21+G basis set for first-row elements. Journal of computational Chemistry, 1983. 4(3): p. 294 - 301.

124. Bayly, C.I., P. Cieplak, W.D. Cornell, and P.A. Kollman, A Well-Behaved Electrostatic Potential Based Method Using Charge Restraints for Deriving Atomic Charges - the Resp Model. Journal of Physical Chemistry, 1993. 97(40): p. 10269-10280.

125. Cornell, W.D., P. Cieplak, C.I. Bayly, and P.A. Kollman, Application of Resp Charges to Calculate Conformational Energies, Hydrogen-Bond Energies, and Free-Energies of Solvation. Journal of the American Chemical Society, 1993. 115(21): p. 9620-9631.

126. Frisch, M.J., G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. Montgomery, J. A., J.E. Peralta, F. Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers, K.N. Kudin, V.N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, N.J. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J.

Page 160: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

147

Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, Ö. Farkas, J.B. Foresman, J.V. Ortiz, J. Cioslowski, and D.J. Fox, Gaussian 09, Revision A.1. 2009, Gaussian, Inc.: Wallingford CT.

127. Jorgensen, W.L., J. Chandrasekhar, J.D. Madura, R.W. Impey, and M.L. Klein, Comparison of simple potential functions for simulating liquid water Journal of Chemical Physics, 1983. 79(2): p. 926 - 936.

128. Wang, J.M., R.M. Wolf, J.W. Caldwell, P.A. Kollman, and D.A. Case, Development and testing of a general amber force field. Journal of Computational Chemistry, 2004. 25(9): p. 1157-1174.

129. Pavan, G.M., Modeling the Interaction between Dendrimers and Nucleic Acids: a Molecular Perspective through Hierarchical Scales. Chemmedchem, 2014. 9(12): p. 2623-2631.

130. Colombo, C., S. Gatti, R. Ferrari, T. Casalini, D. Cuccato, L. Morosi, M. Zucchetti, and D. Moscatelli, Self-assembling amphiphilic PEGylated block copolymers obtained through RAFT polymerization for drug-delivery applications. Journal of Applied Polymer Science, 2016. 133(11).

131. Joung, I.S. and T.E. Cheatham, Determination of alkali and halide monovalent ion parameters for use in explicitly solvated biomolecular simulations. Journal of Physical Chemistry B, 2008. 112(30): p. 9020-9041.

132. Joung, I.S. and T.E. Cheatham, Molecular Dynamics Simulations of the Dynamic and Energetic Properties of Alkali and Halide Ions Using Water-Model-Specific Ion Parameters. Journal of Physical Chemistry B, 2009. 113(40): p. 13279-13290.

133. Pronk, S., S. Pall, R. Schulz, P. Larsson, P. Bjelkmar, R. Apostolov, M.R. Shirts, J.C. Smith, P.M. Kasson, D. van der Spoel, B. Hess, and E. Lindahl, GROMACS 4.5: a high-throughput and highly parallel open source molecular simulation toolkit. Bioinformatics, 2013. 29(7): p. 845-854.

134. Essmann, U., L. Perera, M.L. Berkowitz, T. Darden, H. Lee, and L.G. Pedersen, A Smooth Particle Mesh Ewald Method. Journal of Chemical Physics, 1995. 103(19): p. 8577-8593.

135. Hess, B., H. Bekker, H.J.C. Berendsen, and J.G.E.M. Fraaije, LINCS: A linear constraint solver for molecular simulations. Journal of Computational Chemistry, 1997. 18(12): p. 1463-1472.

136. Miller, B.R., T.D. McGee, J.M. Swails, N. Homeyer, H. Gohlke, and A.E. Roitberg, MMPBSA.py: An Efficient Program for End-State Free Energy Calculations. Journal of Chemical Theory and Computation, 2012. 8(9): p. 3314-3321.

137. Berendsen, H.J.C., J.P.M. Postma, W.F. Vangunsteren, A. Dinola, and J.R. Haak, Molecular-Dynamics with Coupling to an External Bath. Journal of Chemical Physics, 1984. 81(8): p. 3684-3690.

138. Rahman, A. and C.W. Brown, Effect of Ph on the Critical Micelle Concentration of Sodium Dodecyl-Sulfate. Journal of Applied Polymer Science, 1983. 28(4): p. 1331-1334.

139. Dutkiewicz, E. and A. Jakubowska, Effect of electrolytes on the physicochemical behaviour of sodium dodecyl sulphate micelles. Colloid and Polymer Science, 2002. 280(11): p. 1009-1014.

140. Mondain-Monval, O., F. Leal-Calderon, J. Phillip, and J. Bibette, Depletion Forces in the Presence of Electrostatic Double Layer Repulsion. Physical Review Letters, 1995. 75(18): p. 3364-3367.

141. Piirma, I. and S.-R. Chen, Adsorption of ionic surfactants on latex particles. Journal of Colloid and Interface Science, 1980. 74(1): p. 90-102.

142. Kirkwood, J.G. and J. Riseman, The Intrinsic Viscosities and Diffusion Constants of Flexible Macromolecules in Solution. Journal of Chemical Physics, 1948. 16(6): p. 565-573.

143. Riseman, J. and J.G. Kirkwood, The Intrinsic Viscosity, Translational and Rotatory Diffusion Constants of Rod-Like Macromolecules in Solution. Journal of Chemical Physics, 1950. 18(4): p. 512-516.

144. Rotne, J. and S. Prager, Variational Treatment of Hydrodynamic Interaction in Polymers. Journal of Chemical Physics, 1969. 50(11): p. 4831-&.

145. Yamakawa, H. and M. Fujii, Intrinsic-Viscosity of Wormlike Chains - Determination of Shift Factor. Macromolecules, 1974. 7(1): p. 128-135.

Page 161: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

148

146. Carrasco, B. and J.G. de la Torre, Hydrodynamic properties of rigid particles: Comparison of different modeling and computational procedures. Biophysical Journal, 1999. 76(6): p. 3044-3057.

147. Kryven, I., S. Lazzari, and G. Storti, Population Balance Modeling of Aggregation and Coalescence in Colloidal Systems. Macromolecular Theory and Simulations, 2014. 23(3): p. 170-181.

148. Nicoud, L., S. Lazzari, D.B. Barragan, and M. Morbidelli, Fragmentation of Amyloid Fibrils Occurs in Preferential Positions Depending on the Environmental Conditions. Journal of Physical Chemistry B, 2015. 119(13): p. 4644-4652.

Page 162: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing
Page 163: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing
Page 164: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

Baptiste Jaquet

[email protected]

Nationality: Swiss

Born 17.03.89

Curriculum Vitae

Education

10/2011 – 9/2016 Doctor of Science in Chemical and Bioengineering , ETH Zurich

Thesis: Aggregation, Coalescence and Gelation of functional Nanoparticles

9/2010 – 10/2011 Master of Science in Chemical and Bioengineering, ETH Zurich

Thesis: Specific Salt Effect in Protein Aggregation

9/2007 – 7/2010 Bachelor of Science in Chemical and Bioengineering, EPF Lausanne

9/2004 – 7/2007 High School, Nyon

Working experience

10/2011-9/2016 Research Assistant, ETH Zurich, Morbidelli group

Page 165: Rights / License: Research Collection In Copyright - …...2016 I Abstract The nano-scale world is gaining more interest every day, with exciting developments potentially allowing

Publication List

Arosio, P., et al. (2012). "On the role of salt type and concentration on the stability behavior of a monoclonal antibody solution." Biophysical Chemistry 168–169: 19-27. Gu, S., et al. (2014). "Kinetic Analysis of the Catalytic Reduction of 4-Nitrophenol by Metallic Nanoparticles." The Journal of Physical Chemistry C 118(32): 18618-18625. Jaquet, B., et al. (2013). "Stabilization of polymer colloid dispersions with pH-sensitive poly-acrylic acid brushes." Colloid and Polymer Science 291(7): 1659-1667. Lazzari, S., et al. (2015). "Interplay between Aggregation and Coalescence of Polymeric Particles: Experimental and Modeling Insights." Langmuir 31(34): 9296-9305. Lazzari, S., et al. (2015). "Shear stability of inverse latexes during their polymerization process." AIChE Journal 61(4): 1380-1384. Jaquet, B., et al. (2016) “Effects of Coalescence on Shear-Induced Gelation of Colloids”, Langmuir, submitted