cracking in flocculated dispersions - iit bombay · cracking in flocculated dispersions ... 2.7...

69
CRACKING IN FLOCCULATED DISPERSIONS M. TECH. PROJECT Submitted in Partial Fulfillment of the Requirements for the Degree of MASTER OF TECHNOLOGY in CHEMICAL ENGINEERING by LAXMAN BHOSALE (Roll no. 06302019) Guided by Prof. M. S. Tirumkudulu DEPARTMENT OF CHEMICAL ENGINEERING INDIAN INSTITUTE OF TECHNOLOGY BOMBAY 400 076 JULY 2008

Upload: ngoque

Post on 25-Apr-2018

220 views

Category:

Documents


1 download

TRANSCRIPT

CRACKING IN FLOCCULATED DISPERSIONS

M. TECH. PROJECT

Submitted in Partial Fulfillment of

the Requirements for the Degree of

MASTER OF TECHNOLOGY

in

CHEMICAL ENGINEERING

by

LAXMAN BHOSALE

(Roll no. 06302019)

Guided by

Prof. M. S. Tirumkudulu

DEPARTMENT OF CHEMICAL ENGINEERING INDIAN INSTITUTE OF TECHNOLOGY

BOMBAY 400 076

JULY 2008

CONTENTS

Name Page No. Approval Certificate iii

Abstract iv

Acknowledgment v

List of Figures ix

Nomenclature x

Abbreviations xi

1 Introduction 1

2 Literature Review 6

2.1 Latex Film Formation 6

2.1.1 Evaporation, Particle Concentration and Ordering 6

2.1.2 Particle Deformation 7

2.1.3 Polymer Chain Diffusion across Particle Boundaries 7

2.2 Cracking in Films formed from Stable and Flocculated

Dispersions

9

2.2.1 Stable Dispersions 9

2.2.2 Flocculated Dispersions 14

3 Experimental Section 18

3.1 Preparation of Dispersions 18

3.2 Final Close-Packed Volume Fraction Measurements 20

3.3 CCT Measurements 24

3.4 Stress Measurements 25

4 Results and Discussion 31

4.1 Final Close-Packed Volume Fraction 31

4.2 Critical Cracking Thickness (CCT) 34

4.2.1 Film Surface Patterns 34

4.2.2 Crack Patterns 35

4.2.3 Thickness Profile 38

4.2.4 Comparison of Experimental CCT with CCT Obtained

from Model

42

4.3 Critical Cracking Stress (CCS) 46

5 Conclusions and Future Work 51

References

iii

Approval Certificate

Department of Chemical Engineering

Indian Institute of Technology Bombay,

The M. Tech. project titled “Cracking in Flocculated Dispersions” submitted by

Laxman R. Bhosale (Roll No. 06302019) may be accepted for evaluation. Date: 03/07/2008

Guide……………………….......................... Co-Guide…………………………………… Internal Examiner………………………....... External Examiner………………………...... Chairman………………………....................

iv

ABSTRACT

Understanding the mechanism of cracking during drying of aqueous colloidal

dispersions has received considerable attention in the recent years. In many important

products such as paints, wet clays, coatings, etc. cracking is highly undesirable.

Hence, there is a need to understand the phenomena of cracking so that failure of

films can be controlled. Most of the work reported in literature have used stable

aqueous dispersions for this type of studies (Singh and Tirumkudulu, 2007;

Tirumkudulu and Russel, 2005; Dufresne et al., 2003). Although flocculated

dispersions are used in variety of applications, film formation in such systems is less

investigated. In this work, we have studied the cracking of colloidal films formed

from flocculated dispersions of -alumina in water. Here, the extent of flocculation

is controlled by varying the pH of the dispersion and characterized in terms of close-

packed volume fraction of the dried film. The effect of flocculation on critical

cracking thickness (CCT) and critical cracking stress (CCS) is measured by changing

the pH of the dispersion. Further, the experimental results are compared with the

model predictions of Singh and Tirumkudulu (2007) and a good agreement is found

between the theory and experiments, suggesting that the model is universal and

applies equally well to stable and flocculated systems. This report begins with an

overview of the latex dispersions followed by literature review, experimental section,

results and discussion and finally conclusions.

v

ACKNOWLEDGEMENT

I seek this opportunity to express my sincere thanks to my guide,

Prof. M. S. Tirumkudulu for his sustained support and encouragement

during the course of this project. I extend my special thanks to Mr.

Karnail B. Singh for helping me in the experimental work. I am also

thankful to all my lab-mates, especially Mr. Arijit Sarkar for their help

and criticism. Finally, I would like to thank Indian Institute of

Technology, Bombay for financial and technical support.

Laxman Bhosale

30th June 2008

vi

LIST OF FIGURES Figure

No.

Title Page

No. 1.1 Schematic of a (i) Stable and (ii) Flocculated Dispersion 3

2.1 Latex film drying process 8

2.2 Three stages of film formation 8

2.3 Critical cracking thickness variation as a function of the shear

modulus

10

2.4 Schematic diagram of an isolated crack propagating across the

film

11

2.5 The measured CCT vs the characteristic scale 2/13 )2( RMG rcp

12

2.6 Effect of NaCl concentration on the CCT of alumina suspensions 14

2.7 Variation of zeta potential as a function of AKP-30 dispersion

without addition of any electrolyte

16

2.8 The semi-empirically corrected -potential, corr , of AKP-30

alumina suspensions as a function of both volume fraction and pH

16

2.9 Variation of zeta potential ( ) of ALCOA-A16 alumina

suspensions as a function of both volume fraction and pH.

Experiments were done with 10mM NaCl as background

electrolyte

17

3.1 Experimental setup for pH measurement 19

3.2 (a) Formation of packed region and movement of particles in

circular capillary; (b) Estimation of Maximum packing fraction

20

3.3 SEM of AKP-30 particles 22

3.4 SEM of ALCOA-A16 particles 23

3.5 Dektak 150 Surface profiler (i) Front view of profilometer and (ii)

Close view of Stylus and stage

24

3.6 Schematic of the experimental setup for stress measurements 25

vii

3.7 Plot of measured stress vs time for AKP-30 dispersion with

pH=3.3, 16.0i and N = 156. The measurements were done at

26.4oC and 42% RH

27

3.8 Plot of measured stress vs time for AKP-30 dispersion with

pH=6.5, 14.0i , and N = 177. The measurements were done at

26.2oC and 42% RH

28

3.9 Plot of measured stress vs time for AKP-30 dispersion with

pH=7.8, 14.0i , and N = 303. The measurements were done at

25.9oC and 38% RH

28

3.10 Plot of measured stress vs time for AKP-30 dispersion with

pH=10.0, 13.0i and N = 262. The measurements were done at

25.8oC and 32%RH

29

3.11 Plot of measured stress vs time for AKP-30 dispersion with

pH=11.1, 17.0i , and N = 234. The measurements were done at

25.8oC and 38% RH

29

3.12 Weight of the drying film as a function of time for AKP-30

dispersion spread on silicon wafer (i) Evaporation rate calculated

using all data points and (ii) Evaporation rate calculated using

initial weight of the film and final four points

30

4.1 Variation of maximum close-packed volume fraction ( m ) as a

function of pH for AKP-30 dispersion without background

electrolyte (Temp= 26.1 oC and relative humidity =43%)

32

4.2 Variation of final close-packed volume fraction ( m ) as a

function of pH for AKP-30 dispersion with background electrolyte

concentration of 50mM NaCl (Temp= 26.6 oC and relative

humidity =53%)

32

4.3 Variation of maximum close-packed volume fraction ( m ) as a

function of pH for ALCOA-A16 dispersion (Temp= 26.1 oC and

relative humidity =43%)

33

viii

4.4 Picture of crack-free dried film obtained from drying stable

dispersions (pH < 7 and pH > 12). Picture (i) is captured by using

wide angle camera and picture (ii) is captured by using

microscope.

34

4.5 Picture of crack-free dried film obtained from drying flocculated

dispersions (7 < pH < 12). Picture (i) is captured by using wide

angle camera and picture (ii) is captured by using microscope. In

picture (ii), arrows indicate regions of bright spots

34

4.6 Pictures of different dried films of stable dispersions (pH < 7)

showing cracks

36

4.7 Pictures of different dried films of dispersions at pH ~ 7.25

showing cracks

36

4.8 Pictures of circular crack formed in different dried films of

flocculated dispersions (7.75 < pH < 11.5). Arrows indicate the

circular crack.

37

4.9 Thickness profile of a film dried (pH = 2.5) at 26.7oC and 35%RH 39

4.10 Thickness profile of a film dried (pH = 5.4) at 26.7oC and 35%RH 39

4.11 Thickness profile of a film dried (pH = 9.5) at 26.7oC and 35%RH 40

4.12 Thickness profile of a film dried (pH = 11) at 26.7oC and 35% RH 41

4.13 Mean coordination numbers (M) as a function of close-packed

volume fraction for several kinds of packings

42

4.14 Extrapolation of close-packed volume fraction below 0.5 43

4.15 Predicted Critical Cracking Thickness (CCT) as a function of pH

of AKP-30 dispersion. Filled symbols represent predictions and

empty symbols represent experimental measurements

45

4.16 The measured CCT as a function of pH for AKP-30 dispersions 45

4.17 The predicted and measured dimensionless critical cracking stress

)2/( Rc as a function of pH of AKP-30 dispersion at different

film heights. Filled symbols represent predictions and empty

symbols represent the values obtained by interpolation of

experimental stress values from line equation shown in figure 4.19

47

ix

4.18 The predicted and measured dimensionless critical cracking stress

)2/( Rc as a function of pH of AKP-30 dispersion at different

film heights. Filled symbols represent predictions by increasing

factor in equation 2.1 by 2 and empty symbols represent the values

obtained by interpolation of experimental stress values from line

equation shown in figure 4.19

47

4.19 Plot of measured dimensionless critical cracking stress )2/( Rc

vs the dimensionless film thickness )2/( Rhrcp

48

4.20 Plot of measured dimensionless critical cracking stress )2/( Rc

vs the dimensionless characteristic scale )/( 2NM m . The data

points are for films of AKP-30 dispersion with varying pH. The

solid line is a power law fit with an exponent of 1/3. C is the

constant

50

x

NOMENCLATURE

Symbol Name First used

on Page No.

elasticE Elastic energy recovery per unit length of the crack 10

Es Young’s modulus of the substrate 25

G shear modulus of the particles 10

hmax Critical cracking thickness 4

oL Initial length of dispersion in capillary 21

mL Final length of packed region in capillary (after

drying)

21

M Coordination number 11

N Dimensionless film thickness ( Rh 2/ ) 11

Pmax Maximum attainable capillary pressure 12

R Particle radius 11

st Substrate thickness 26

ft Film thickness 26

c Critical cracking stress 4

m Maximum close-packed volume fraction 4

rcp Random closed packing fraction 3

Solvent-air interfacial tension 10

Zeta potential 15

Deflection angle 24

Wetting angle 13

f Bulk stress in the film 26

xi

ABBREVIATIONS CCT Critical Cracking Thickness

CCS Critical Cracking Stress

RCP Random Closed Packing

SEM Scanning Electron Microscope

1

CHAPTER 1

INTRODUCTION

Traditionally, organic solvents have been used for the preparation of paint

dispersions. Solvents which are used in these dispersions have a high boiling point, so

that the “open time” defined as the time between the application of the paint and start

of the actual drying process, is sufficiently long for the easy application of these

products. Volatile organic solvents are used to plasticize the polymer and thereby

convert it to a homogeneous film at room temperature (Routh and Russel, 2001). In

case of traditional paints and coatings, large amounts of volatile organic compounds

(VOCs) are emitted during the film formation stage. Since the traditional solvent

based coatings pose serious environmental hazards due to the highly polluting effects

of the solvents (Keddie, 1997), the coating industries are being forced to reduce VOC

emissions via government regulations. Consequently, in several European countries,

indoor use of solvent-based architectural coatings has been banned. One of the reports

published by the Environmental Protection Agency (EPA, USA) shows that the

indoor air pollution is about 10-40 times higher than the outdoor pollution and it is

attributed to the solvent emission by paints and coatings. So it is necessary to find

alternative ways to either reduce or eliminate solvents in coating formulations.

Several alternatives such as water-based coatings, applying coatings in the polymer

melt form, increasing solids volume fraction and using powder coatings exist, but

water based latex coatings have been identified as a worthy alternative because of

ease of application and faster drying (Keddie, 1997).

When solid colloidal particles are dispersed in a liquid then it is called as a sol.

On the other hand, when liquid droplets are dispersed in another liquid, it is called an

emulsion. Latex is a colloidal dispersion of polymer particles in an aqueous solution.

2

Lattices now constitute about eight percent of the world’s total production of

polymers. Latex particles are generally prepared by emulsion polymerization. In this

reaction, monomers are loaded along with an initiator and surfactant into the reactor.

The monomer goes inside the micelle and starts polymerizing. Here, the micelle itself

acts as a small reactor. The size and shape of latex particle depends on the type of

micelle formed by addition of surfactant. Usually these particles are spherical. Typical

particle size obtained by this mechanism varies from 100 nm to 1000 nm.

Latex film formation is essential to products such as paints, adhesives (Mayer

et al., 1995), caulks and sealants, paper coatings, textiles and carpets. Various authors

have mentioned other applications of latex such as in foam mattresses, composites

(Okubo et al., 1983; Crowley et al., 1992), floor polish (Brodnyan and Konen, 1964),

gloves (Dobler et al., 1992) etc.

In most applications, the requirement is that final dried film should be

continuous. The morphology of the final dry film depends on many factors such as

properties of the polymers, method of preparation of latexes and their film formation

mechanism. But many times the properties of latex coatings are inferior to solvent-

based coatings due to incomplete film formation. Water based latex coatings differ

from solvent based latex coatings in terms of their final permeability and mechanical

properties. There are many reasons for incomplete film formation such as incomplete

particle deformation, non-uniform drying, and improper chain diffusion. The quality

of the film deteriorates because of non uniformity and presence of voids in the final

film. To get the required film properties, latex particles should come together and

aggregate to form a continuous film. In this regard, Protzman and Brown (1960)

developed the concept of the Minimum Filming Temperature (MFT) which is defined

as that temperature at which latex cast film becomes continuous and clear. Below this

3

critical temperature, the dry latex is opaque and powdery. MFT primarily indicates the

lower temperature range over which a latex application will be useful. From their first

use in the paint industry, latex coatings have been used in many applications in the

previous years. Although synthetic latex is produced in large quantity, many aspects

of its formation are still not clear.

Most studies reported in literature (Singh and Tirumkudulu 2007; Keddie et

al., 1995; Sperling et al., 1993) have focused on stable dispersions. Here, particles are

well separated from each other during drying and touch each other only when the

particle volume fraction reaches that for random close packing ( rcp ). Stability of the

dispersion is mainly determined by the balance of three forces namely electrostatic

repulsion, van der Waals attraction and steric attraction or repulsion. In case of

flocculated dispersions, the interparticle stabilizing forces are weak and are overcome

by the attractive van der Waals forces causing the particles to contact well before

reaching random close packing. It means that the particles form chain like structures

for concentrations much lower than that for random close packing (figure 1.1). Due to

the formation of flocculated structures, the cracking characteristics are expected to be

different from that for stable dispersions. Thus, it is not clear whether the mechanism

controlling cracking in stable dispersions also applies to flocculated systems.

(i) (ii)

Figure 1.1 Schematic of a (i) Stable and (ii) Flocculated Dispersion

4

Terminology:

Few terms which will be used often in this report are defined as follows:

Consider an aqueous colloidal dispersion, applied on an impermeable substrate.

Final close packed volume fraction ( m ): The volume fraction of solids in the

completely dried film is known as final close packed volume fraction.

CCT: The thickness above which the drying film would crack spontaneously is

known as critical cracking thickness.

CCS: The stress at which first crack nucleates in the film is the critical cracking

stress.

Objectives:

The objectives of this project are

1) To study the effect of flocculation (pH ) on

Final close-packed volume fraction of particles ( m )

Critical cracking thickness (CCT) of the film (hmax)

Critical cracking stress (CCS) of the film ( c )

2) To compare the measured values of critical cracking thickness and critical

cracking stress with the theoretical models available in literature.

5

In this report, different stages of film formation and the relevant studies on

stable and flocculated dispersions reported in literature are discussed in chapter 2.

Chapter 3 focuses on the experimental procedure for preparing dispersions and

measurement of final close packed volume fraction, thickness of the film and tensile

stress in the drying film. Chapter 4 presents the experimental results and their

comparison with the models available in literature. We found that flocculation affects

the particle ordering which changes the final close packed volume fraction of the

dried film. Further, we have also determined the effect of flocculation on critical

cracking thickness (CCT) and critical cracking stress (CCS). Finally, we have

compared the experimental CCT and CCS with the predictions of model developed by

Singh and Tirumkudulu (2007) and found that experiments are in good agreement

with the theory. Finally, the conclusions and scope for future work are discussed in

chapter 5.

6

CHAPTER 2

LITERATURE REVIEW

2.1 Latex Film Formation:

When a thin film of latex containing soft polymer particles is applied on an

impermeable substrate and dried, a continuous, homogeneous film can be achieved

under appropriate conditions. This process is known as film formation. During film

formation, these colloidal particles must overcome their mutual repulsion in order to

form a continuous film. So lot of effort goes into keeping the colloidal particles

separated and deflocculated in order to obtain a stable dispersion.

Though many have studied the latex film formation process (Steward et al.,

2000; Zhao et al., 1990; Lin and Meier, 1995; Brown, 1956; Henson et al., 1953), the

process is not completely understood. The process of drying and film formation in

latex coatings occurs in three stages as shown in figure 2.2. These stages are

explained in the following subsections (Steward et al., 2000).

2.1.1 Evaporation, Particle Concentration and Ordering:

When a thin film of aqueous colloidal dispersion is applied on an impermeable

substrate, water evaporates resulting in an increase in solid concentration. In stable

dispersions, the strong inter-particle repulsive forces disperse the particles against the

tendency of van der Waals force and Brownian motion to bring particles together.

Further evaporation causes particles to concentrate into a close packed array (Sheetz,

1965; Routh and Russel, 1998, 2001; Tirumkudulu and Russel, 2004). The studies

made by Eckersley et al. (1994) confirm that drying rate during the first stage is

constant. If a porous substrate is used instead, then water can leave the latex either by

evaporation into the ambient atmosphere or by absorption into the porous substrate.

7

Most of the studies have concentrated on the evaporative mechanism and considered

impermeable substrates. This is also because most applications involve nonporous

substrates. Because of the liquid meniscus formed between two particles at the

periphery of the film, a low pressure zone is created. This causes the convection of

latex particles towards the packed region. The formation of packed region and

movement of particles is shown in the figure 2.1.

2.1.2 Particle Deformation:

At the beginning of the second stage, particles are observed at the surface of

latex. At this stage, forces which overcome the electrostatic and/or steric repulsions

between particles start acting (Dobler et al., 1991). Further evaporation of water

causes the development of stresses in the wet film (Winnik, 1997). If the particles are

soft, they deform and coalesce so as to close the pores whereas in case of hard

particles, stresses are released by the nucleation and propagation of cracks

(Tirumkudulu and Russel, 2005).

The various theories proposed so far show that particle deformation occurs

due to one or more of the following forces: air/water, water/polymer, or polymer/air

interfacial tension, osmotic force and surface adhesive force (Keddie, 1997).

2.1.3 Polymer Chain Diffusion across Particle Boundaries:

During stage 3, diffusion of particles takes place across particle-particle

interface to fuse the particles. Due to diffusion controlled evaporation, the drying rate

during this stage decreases drastically. The rate of polymer inter-diffusion is mainly

affected by the temperature of the system, molecular weight of polymer, cross-linking

of polymers (Zosel and Ley, 1993), organic solvents (Juhue and Lang, 1994), co-

surfactants (Kim et al., 1994) and coalescing aids. Although the three stages of film

8

formation have been studied for a long time the mechanisms responsible for drying

and particle deformation are not fully understood (Tirumkudulu and Russel, 2004).

Figure 2.1 Latex film drying process (Tirumkudulu and Russel, 2004)

Figure 2.2 Three stages of film formation (Winnik, 1997)

9

2.2 Cracking in Films formed from Stable and Flocculated

Dispersions:

2.2.1 Stable Dispersions:

When a thin film of aqueous colloidal dispersion is applied on a nonporous

substrate, it could crack due to the stresses generated in it. One of the most important

experimental studies on cracking of colloidal films formed from stable dispersion was

reported by Chiu et al. (1993). They studied the cracking during drying of wet

ceramics and found that irrespective of particle size and rigidity there exists a critical

thickness below which films do not crack. Further, this thickness was found to

increase with increase in particle size. Singh and Tirumkudulu (2007) identify two

regimes for obtaining crack free films. For dispersions containing hard particles

(stress-limited regime), the CCT increases with particle size as well as particle rigidity

while for soft particles where particles completely deform to close the pores (strain

limited regime), the critical thickness decreases with increasing particle rigidity but is

independent of particle size (figure 2.3). Further, the way particles pack is also found

to influence their cracking behavior.

10

Figure 2.3 Critical cracking thickness variation as a function of the shear

modulus (Singh and Tirumkudulu, 2007)

Drying of films is accompanied by transverse stresses created due to decrease

in volume of the film during drying. The transverse tensile stress ( ) created in the

film scales with the square of compressive strain ( ), ~ 2G , where ‘G’ is the

shear modulus. When an isolated crack is initiated in the film, stresses over the length

of h around the crack will be released from the film (Singh and Tirumkudulu, 2007).

The elastic energy recovered per unit length of the crack scales as elasticE ~ 2h . At

the same time same, there will be increase in the surface energy per unit length

( sE ~ h ), where is the water/air interfacial tension. According to Griffith’s

energy balance concept, these two energies are equal. So by equating these two

energies, we get c ~ 3/23/1 )( hG (Singh and Tirumkudulu, 2007). In case of hard

particles, even the maximum capillary pressure is not able to completely deform them,

so that the critical cracking stress can be obtained by equating the critical stress to the

11

tensile stress at the maximum capillary pressure, 3/23/1 )( hG ~ )( maxP , which

gives the critical cracking thickness as maxh ~ 2/3max

2/1 )( PG (Singh and

Tirumkudulu, 2007).

Figure 2.4 Schematic diagram of an isolated crack propagating across the film (Tirumkudulu and Russel, 2005)

The exact expression for the critical stress ( c ) was determined by

Tirumkudulu and Russel (2005) as,

3/13/1

2 21877.0

2RG

NMR mc (2.1)

Here, R is the particle radius, m is final close-packed volume fraction, M is

the coordination number and N is dimensionless film thickness ( Rh 2/ ).

12

Singh and Tirumkudulu (2007) obtained following exact expression for the

CCT of hard particles,

2/3

max

2/13

max )(2

264.0

RPRMGh m (2.2)

Here, Pmax is the maximum attainable capillary pressure.

Singh and Tirumkudulu (2007) measured critical cracking thicknesses over a

wide range of particle sizes and particle shear modulii against the characteristic scale

for the stress limited regime. They found that the observed scaling is in remarkable

agreement with the predictions over five orders of characteristic scale (figure 2.5).

They also found that the asphericity of the particles does not significantly affect the

CCT.

Figure 2.5 The measured CCT vs the characteristic scale 2/13 )2( RMG rcp (Singh and Tirumkudulu, 2007)

13

White (1982) obtained the equation for maximum capillary pressure for

packing of identical spheres in terms of the wetting angle as,

)1(cos3

)( maxm

m

RP (2.3)

If we combine equation (2.2) and equation (2.3), we get critical cracking

thickness as a function of final close-packed volume fraction as,

2/1

23

33

max )(cos27)1(4

64.0m

mRMGh (2.4)

where is the wetting angle. The above equation shows that for hard particles, if the

shear modulus and radius of particle are constant, then the CCT is directly

proportional to 23 )(/)1( mmM . Thus with increased flocculation (or decrease

in m ), the CCT is expected to increase.

14

2.2.2 Flocculated Dispersions:

Chiu et al. (1993) have reported the effect of pH and salt concentration and

thereby the degree of flocculation on critical cracking thickness. Here, the degree of

flocculation of the alumina dispersion was changed by adding varying amount of salt

(NaCl). It was found that the CCT initially increases with increase in NaCl

concentration, passes through a maximum and then starts decreasing (Figure 2.6).

Figure 2.6 Effect of NaCl concentration on the CCT of alumina suspensions (Chiu et al., 1993)

More recently, Carreras et al. (2007) studied the effect of flocculation on the

critical cracking thickness of alumina suspensions. In this study, they found that the

critical cracking thickness for dispersed suspensions (pH ~ 1.75) is 3 times greater

than that for flocculated dispersions (pH ~ 9). They attributed this increase in CCT to

the increase in close packed volume fraction for dispersed suspensions. A higher

close-packed volume fraction has more number of particle-particle interactions per

unit area and therefore the film can withstand higher stress. They also found that with

small amounts of cross-linked polymeric binder, critical cracking thickness can be

increased by a factor greater than 10.

15

Perez et al. (2006) measured the zeta ( ) potential of high purity -alumina

(AKP-30) particles as a function of pH of the dispersion (figure 2.7). From figure, it is

clear that the surface potential of particles is high and positive at low pH and

decreases with increasing pH (equation 2.5 and equation 2.6). At pH ~ 9.0, the

potential passes through the iso-electric point and decreases further with pH to

become large and negative. Since zeta potential is the direct indication of the strength

of repulsive force, the dispersion is quite stable at low pH (< 5). However, as the pH

is increased, particles are expected to flocculate and at a pH close to 9.0, the extent of

flocculation is expected to be maximum. It means that at high zeta potential values,

the particles will be ordered in the film. A decrease in the magnitude of the zeta

potential leads to a greater disorder in particle packing. So if we change the pH of

flocculated dispersion, particles will pack at a different value of close-packed volume

fraction.

Johnson et al. (1997) also measured the zeta potential of AKP-30 particles

with pH of dispersion by maintaining 0.01M KNO3 background electrolyte

concentration (figure 2.8). They found that zeta potential of particles became zero at

pH ~ 9.4 (iso-electric point). This shows that there is small a shift in the iso-electric

point of AKP-30 dispersion by addition of electrolyte.

HOAlHOAl )( 3232 (At low pH) (2.5)

OHOAlOHOAl )( 3232 (At high pH) (2.6)

16

Figure 2.7 Variation of zeta potential as a function of AKP-30 dispersion

without addition of any electrolyte (Perez et al., 2006).

Figure 2.8 The semi-empirically corrected -potential, corr , of AKP-30 alumina suspensions as a function of both volume fraction and pH. , =0.020; , =0.060;

, =0.120; , =0.200; =0.300 (Johnson et al., 1998)

17

Greenwood (2003) measured the zeta potential of ALCOA-A16 particles as a

function of pH to determine iso-electric point (figure 2.9). They reported that

dispersion with particles having a zeta potential outside the range of -30mV to

+30mV are stable. They also studied the effect of background salt concentration on

the iso-electric point of the dispersion by using SDK160 alumina particles. In this

study, they found that iso-electric point of SDK160 alumina particles shifted from 8.1

with background concentration of 50mM NaCl to 7.4 with no salt present. With

background concentration of 10mM NaCl, they found that zeta potential of ALCOA-

A16 particles becomes zero at a pH value of 8.2 (iso-electric point).

Figure 2.9 Variation of zeta potential ( ) of ALCOA-A16 alumina suspensions as a function of both volume fraction and pH. Experiments were done with 10mM NaCl as background electrolyte. , =0.142; , =0.186; , = 0.237 (Greenwood, 2003)

18

CHAPTER 3

EXPERIMENTAL SECTION

3.1 Preparation of Dispersions:

For the current study, we chose -alumina particles as these particles have

high commercial importance. Also, the tendency of these particles to attract or repel

each other can be easily controlled by controlling the pH (Velamakanni and Lange,

1991). In other words, it is easy to stabilize or destabilize the dispersion of these

particles by just changing the pH of dispersion. High purity -alumina (AKP-30)

with average particle size as 335nm and manufactured by Sumitomo Chemical Co.,

Japan was obtained from Sigma-Aldrich. Further, alumina particles (ALCOA-A16)

with average particle size 350nm and manufactured by ALCOA-A16 SG, USA were

obtained from Tata Research Development and Design Centre (TRDDC). Dilute

dispersions with initial volume fraction of 13-15% were prepared by dispersing AKP-

30 in deionized water. Similarly, dispersions for ALCOA-A16 were also prepared. It

was observed that due to extensive ageing of the dispersion in glass, the iso-electric

point changes over time (Furlong et al., 1980). Hence, all the dispersions were

prepared in poly propylene bottles. Before preparing the dispersions, polypropylene

bottles were thoroughly cleaned by soaking successively in 1M solution of HNO3,

distilled water, 1M solution of KOH and finally in distilled water, each for 12 hours.

Standard pH tablets were obtained from Merck for the calibration of pH electrode.

Three standard solutions were prepared with pH value of 4, 7 and 9.0. Initially, the pH

meter is calibrated for any two pH values (either 4 and 7 or 7 and 9.2) and then we

checked the accuracy of the pH meter by using remaining standard solution (9.2 or 4

respectively). The pH of these dispersions was adjusted using analytical grade HNO3

19

and KOH. Before carrying out the measurements, all the dispersions were re-

dispersed using cyclo-mixer and ultra-sonicator for 4-5 minutes. The pH of the

dispersion was measured by using pH meter as shown in figure 3.1.

Figure 3.1 Experimental setup for pH measurement

pH meter

pH electrode

Flocculated dispersion

Stand

20

3.2 Final Close-Packed Volume Fraction Measurements: The final close-packed volume fraction of these dispersions was measured

using a glass capillary of size 300 m in diameter. Here, the original dispersion was

allowed to dry from one end of the capillary tube under ambient conditions. The final

close-packed volume fraction ( m ) was calculated by measuring the initial and final

length of dispersion plug.

Figure 3.2 (a) Formation of packed region and movement of particles in circular capillary; (b) Estimation of final packing fraction

Figure 3.2 (a) shows a schematic of a capillary where the dispersion fills it

partially so that one of the menisci is at the mouth of the capillary at one end while the

other meniscus is well within the capillary. As drying progresses, water evaporates

from the left and particles form a close-packed array. Due to the high curvature of the

menisci at the first layer of particles, a low pressure zone is created in the fluid. So the

low capillary pressure convects the particles and fluid leading to the formation of a

packed front. Water is drawn from the fluid region to maintain the evaporative loss

and this incoming flux carries particles with it and packed front length goes on

High curvature between the first layer of particles

Movement of particles

Packed Region

High Evaporation Rate

Negligible evaporation

Lo Lm

21

increasing. While on the other end of the capillary, as there is no capillary force

acting, evaporation is negligible.

The initial length of dispersion in capillary (Lo) and the final length of packed

region (Lm) is measured. This procedure is repeated for all the dispersions with pH

values ranging from 2 to 12. Final packing fraction is obtained by taking mass balance

for solids as,

mmoo LL (3.1)

Here, o is the initial solid fraction in capillary, m is the final solid fraction in

capillary (final close-packed volume fraction), oL is the initial length of dispersion in

capillary and mL is the final length of packed region in capillary (after complete

drying).

Figure 3.3 and figure 3.4 shows the SEM image of a dried film of AKP-30 and

ALCOA-A16 dispersions respectively. As can be seen, particles are irregular and

poly-dispersed both in shape as well as in size.

22

Figure 3.3 SEM of AKP-30 particles

23

Figure 3.4 SEM of ALCOA-A16 particles

24

3.3 CCT Measurements:

The critical cracking thickness measurements were done at a temperature of

25-28oC and a relative humidity of 35-40%. Thin films of the dispersions were cast on

glass substrates by disbursing a small amount of liquid using a spin coater (slow

rotation rate ~ 20 rpm). Here, the spin coater was used to spread the liquid uniformly

over a fixed area so as to obtain a nearly circular film. Further, the thicknesses of

these films were varied by disbursing varying amounts of dispersion on the same area.

After the films were cast, they were left to dry under ambient conditions. After a film

was completely dried, it was examined under an optical upright microscope

(Olympus, BX-60) for cracks at varying magnifications. Finally, the thickness profile

of the film was measured using a surface profilometer Dektak-150 (figure 3.5). It

should be noted that in some cases the entire film was crack free which gave the lower

limit for CCT.

(i) (ii)

Figure 3.5 Dektak 150 Surface profiler (i) Front view of profilometer and (ii) Close view of Stylus and stage (www.veeco.com)

25

3.4 Stress Measurements:

The schematic of the experimental setup for the measurement of stress in the

film is shown in the figure 3.6. Here, a thin layer of the dispersion was applied on a

thin silicon wafer (thickness ~ 128 m) clamped on one side. The dimensions of the

silicon wafer used were 13mm (length) and 5mm (width). A laser beam is reflected

from the free end of the substrate and collected by a position sensitive detector (On-

track ). The position of the detector is set in such a way that the reflected beam falls

normally on the surface of the detector. The entire detector assembly is mounted on

an X-Y traverse so that the desired positioning can be done with great precision. The

position data (i.e. the x, y coordinate of the reflected beam) from the detector is

directly fed to a computer. As the film dries, the tensile stresses in the drying film

bend the substrate which further results in the movement of reflected beam on

detector surface. Finally, the displacement of the reflected beam on the detector

surface is correlated to the amount of deflection in the substrate which is finally

related to the average tensile stress in the film.

Figure 3.6 Schematic of the experimental setup for stress measurements

X

Silicon substrate

Laser Source

Latex Film

L2 L1

Deflection angle

Position Sensitive detector

26

The classical cantilever technique was used to measure the stress in a drying

film (Petersen et al., 1999). If we apply a thin layer of film on substrate, continuous

evaporation of water produces transverse stresses in the wet film. Under the influence

of stress the substrate bends with deflection angle . The substrate deflection can be

related to the equivalent stress produced in the film by taking force balance. The bulk

stress in the film ( f ) is given as,

3

6 ( )s s

ff s f

E tt L t t (3.2)

Here, Es is the Young’s modulus of the substrate, st is the substrate thickness, ft is

the film thickness, R is the radius of curvature, L is the length of the film.

Deflection angle ( ) can be related to the displacement (X) and the length of the

reflected ray for straight substrate (L2) as

1

2

1 tan ( )2

XL

(3.3)

Finally, the bulk stress in the film ( f ) is given as,

31

2

tan ( )12 ( )

s sf

f s f

E t Xt L t t L

(3.4)

Figure 3.7 to figure 3.11 shows typical stress profiles for the films for different

pH and different thickness. Peak point in the stress profile gives the value of stress at

which the film will crack i.e. critical cracking stress (CCS). Critical cracking stress

was measured by preparing films of different thicknesses. The thicknesses of these

films were varied by disbursing varying amounts of dispersion on the same area. The

drying film is immediately removed from the clamp on cracking and the evaporation

rate is calculated by measuring the weight of the drying film. Even if we know the

initial weight of the film and the weight at three-four time points after cracking, we

27

can get an accurate value of evaporation rate. This is done to get the evaporation rate

at the same conditions at which stress experiments were carried out. Figure 3.12 (i)

and (ii) shows that evaporation calculated from these two plots is the same.

Figure 3.7 Plot of measured stress vs time for AKP-30 dispersion with pH=3.3,

16.0i and N = 156. The measurements were done at 26.4oC and 42% RH

28

Figure 3.8 Plot of measured stress vs time for AKP-30 dispersion with pH=6.5, 14.0i , and N = 177. The measurements were done at 26.2oC and 42% RH

Figure 3.9 Plot of measured stress vs time for AKP-30 dispersion with

pH=7.8, 14.0i , and N = 303. The measurements were done at 25.9oC and 38% RH

29

Figure 3.10 Plot of measured stress vs time for AKP-30 dispersion with pH=10.0, 13.0i and N = 262. The measurements were done at 25.8oC and 32%RH

Figure 3.11 Plot of measured stress vs time for AKP-30 dispersion with pH=11.1, 17.0i , and N = 234. The measurements were done at 25.8oC and 38% RH

30

y = -5.59E-06x + 3.15E-02R2 = 1.00

0.0265

0.0270

0.0275

0.0280

0.0285

0.0290

0.0295

0.0300

0.0305

0.0310

0.0315

0 100 200 300 400 500 600 700 800Time (secs)

Wei

ght (

gms)

(i)

y = -5.59E-06x + 3.15E-02R2 = 1.00

0.0265

0.0270

0.0275

0.0280

0.0285

0.0290

0.0295

0.0300

0.0305

0.0310

0.0315

0 100 200 300 400 500 600 700 800Time (secs)

Wei

ght (

gms)

(ii)

Figure 3.12 Weight of the drying film as a function of time for AKP-30 dispersion spread on silicon wafer (i) Evaporation rate calculated using all data points and

(ii) Evaporation rate calculated using initial weight of the film and final four points

31

CHAPTER 4

RESULTS AND DISCUSSION

4.1 Final Close-Packed Volume Fraction:

The extent of flocculation is quantified in terms of final close-packed volume

fraction ( m ). Figure 4.1 shows the variation of final close-packed volume fraction

with pH for the AKP-30 dispersion without addition of any electrolyte. The measured

volume fraction shows direct correspondence with the reported zeta potential in that,

at low pH, the final close-packed volume fraction is highest essentially due to the

higher surface charge (figure 2.7). As the pH is increased, the stabilizing repulsive

forces between particles decrease and the dispersion becomes more unstable resulting

in particle flocculation. This is evident from lower values of final close-packed

volume fraction as we increase pH. At a pH of 9.1 (iso-electric point), the charges at

the surface are completely neutralized and there are no stabilizing forces. As

expected, the final close-packed volume fraction is the lowest at this pH, 18.0~m .

With further increase in pH the particle surface acquires negative charge which

stabilizes the dispersion. This is again evident from the higher value of final close-

packed volume fraction at high pH.

We repeated the experiment by adding 50mM NaCl salt as background

electrolyte and found the iso-electric point to occur at pH ~ 9.5 (figure 4.2). The shift

in iso-electric point is due to the specific absorption of positively charged ions on the

surface of alumina particles. These results indicate that the iso-electric point should

not change significantly with salt concentration. Further, the final close-packed

volume fraction measurements with pH of AKP-30 dispersion with and without

background electrolyte agree well with the zeta potential vs pH study reported by

Johnson et al. (1997) and Perez et al. (2006) respectively (figure 2.8 and 2.7).

32

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.0 2.0 4.0 6.0 8.0 10.0 12.0

pH of AKP-30 dispersion

Fina

l clo

se-p

acke

d vo

lum

e fra

ctio

n (

m)

Iso-electric point ~ 9.1

Figure 4.1 Variation of final close-packed volume fraction ( m ) as a function of pH for AKP-30 dispersion without background electrolyte (Temp= 26.1 oC and

relative humidity =43%)

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0 2 4 6 8 10 12 14

pH of AKP-30 dispersiom

Fina

l clo

se-p

acke

d vo

lum

e fra

ctio

n (

m)

Iso-electric point ~ 9.5

Figure 4.2 Variation of final close-packed volume fraction ( m ) as a function of pH for AKP-30 dispersion with background electrolyte concentration of 50mM

NaCl (Temp= 26.6 oC and relative humidity =53%)

33

Figure 4.3 shows the variation of final close-packed volume fraction with pH

for the ALCOA-A16 dispersion. The final close packed volume fraction is the lowest

at pH ~ 7.2. It means that particles in ALCOA-A16 attain point of zero charge at pH

of 7.2 (i. e. iso-electric point). As we have already discussed in section 2.2.2,

ALCOA-A16 dispersion with 10mM NaCl background concentration has an iso-

electric point of 8.2. Decrease in iso-electric point from 8.2 to 7.2 could be due to the

use of background salt concentration.

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0 2 4 6 8 10 12 14

pH of ALCOA-A16 solution

Fina

l clo

se-p

acke

d vo

lum

e fra

ctio

n (

m)

Iso-electric point ~ 7.2

Figure 4.3 Variation of final close-packed volume fraction ( m ) as a function

of pH for ALCOA-A16 dispersion (Temp= 26.1 oC and relative humidity =43%) .

34

4.2 Critical Cracking Thickness (CCT):

4.2.1 Film Surface Patterns:

Figure 4.4 and figure 4.5 show the pictures of the surface of crack-free dried films.

Films obtained from drying stable dispersions (i.e. pH < 7 and pH > 12) show smooth,

glossy surface (figure 4.4) whereas the films obtained from drying of flocculated

dispersions show rough surface (figure 4.5). Many bright spots are observed in the

dried films of flocculated dispersions which is clear indication of the high porosity of

the film (i.e. low final close-packed volume fraction).

(i) (ii)

Figure 4.4 Picture of crack-free dried film obtained from drying stable dispersions (pH < 7 and pH > 12). Picture (i) is captured by using wide angle camera and picture

(ii) is captured by using microscope.

(i) (ii)

Figure 4.5 Picture of crack-free dried film obtained from drying flocculated dispersions (7 < pH < 12). Picture (i) is captured by using wide angle camera and

picture (ii) is captured by using microscope. In picture (ii), arrows indicate regions of bright spots.

35

4.2.2 Crack Patterns:

Different types of crack patterns are observed in the dried films of stable and

flocculated dispersions (figure 4.6 - figure 4.8). For dispersions at low pH values (pH

< 7), regular radial cracks are observed in the dried films above critical cracking

thickness (figure 4.6). At pH values ranging from 7 – 7.75, pattern of cracks observed

was somewhat irregular (figure 4.7). For flocculated dispersions (i.e. for 7.75 < pH <

11.5), the film shows ring type of crack (figure 4.8). It appears that two cracks are

initiated at two points in the film and they propagate and meet each other completing

the circle. The reason behind the formation of such ring type of cracks at high

flocculation is not clear and needs further investigation. In these types of films, it is

very difficult to spread the film uniformly on the glass substrate. It was observed that

a drop of flocculated dispersion did not change its shape significantly during drying.

Further, we also observed the water to drain out of the film from the edges. These

results seem to indicate that once the particle concentration reaches that for

flocculation, the shape of the drop is set by the particle network. The capillary forces

are also weak due to the high porosity of the film. So it is expected that flocculated

dispersions will have higher critical cracking thickness than stable dispersions.

Pauchard et al. (1999) studied the crack patterns formed by the drying of an

isolated drop of colloidal silica suspension on a glass slide. They also observed

similar kind of different crack patterns for different salt concentrations. They found

that dried films show regular crack pattern at low salt concentrations (0-0.2 mol/l),

disordered crack pattern at intermediate salt concentration (0.4 mol/l) and an unique

circular crack at high salt concentration (above 0.4 mol/l).

36

Figure 4.6 Pictures of different dried films of stable dispersions (pH < 7) showing cracks

Figure 4.7 Pictures of different dried films of dispersions at pH ~ 7.25 showing cracks

37

Figure 4.8 Pictures of circular crack formed in different dried films of flocculated dispersions (7.75 < pH < 11.5). Arrows indicate the circular crack.

38

4.2.3 Thickness Profile:

Figure 4.9 – 4.12 shows the thickness profiles of dried films of AKP-30

dispersions of varying pH. In case of stable dispersions, the profile shows two peaks

at two ends while the thickness in the central section is lower (figure 4.9 and figure

4.10). As we have already seen that for stable dispersions, final close packed volume

fraction is high and so the capillary pressure generated convects the particles and fluid

leading to the formation of packed front at the periphery of film. Water is drawn from

the fluid region to maintain for evaporative loss and this incoming flux carries

particles with it and packed front length goes on increasing. Transverse flow

generated by capillary forces is responsible for the observed profile.

In case of flocculated dispersion, since the final close-packed volume fraction

is low, the magnitude of capillary pressure is also low. So there is no convection of

particles from central fluid region to the periphery of the film which is clear from the

shape of the thickness profile of the dried film of flocculated dispersion (figure 4.11

and figure 4.12).

Thus, the shape of the dried film gives clear evidence of capillary pressure as

the driving force for the movement of particles in drying colloidal films.

39

Figure 4.9 Thickness profile of a film dried (pH = 2.5) at 26.7oC and 35% RH

Figure 4.10 Thickness profile of a film dried (pH = 5.4) at 26.7oC and 35% RH

Micrometer

Thic

knes

s, A

ngst

rom

Micrometer

Thic

knes

s, A

ngst

rom

Micrometer

Thic

knes

s, A

ngst

rom

Micrometer

Thic

knes

s, A

ngst

rom

40

Figure 4.11 Thickness profile of a film dried (pH = 9.5) at 26.7oC and 35%RH

Micrometer

Thic

knes

s, A

ngst

rom

Micrometer

Thic

knes

s, A

ngst

rom

Micrometer

Thic

knes

s, A

ngst

rom

Micrometer

Thic

knes

s, A

ngst

rom

41

Figure 4.12 Thickness profile of a film dried (pH = 11) at 26.7oC and 35% RH

Micrometer

Thic

knes

s, A

ngst

rom

Micrometer

Thic

knes

s, A

ngst

rom

Micrometer

Thic

knes

s, A

ngst

rom

Micrometer

Thic

knes

s, A

ngst

rom

42

4.2.4 Comparison of Experimental CCT with CCT Obtained from Model:

The calculation of CCT (equation 2.4) requires a knowledge of the shear

modulus of the particle (G) and the coordination number (M). As shear modulus for

AKP-30 particles is very high ( PaG 910156 ), we can consider them as hard

particles. Figure 4.13 shows the variation of mean coordination number as a function

of close packed volume fraction (Uri et al., 2006). The value of coordination number

was taken from extrapolation of graph as shown in figure 4.14. If equation (2.4) holds

good for hard particles, then the predicted variation of CCT values as a function of pH

of dispersion is shown in figure 4.15 (with filled symbols). This variation is same as

the variation of zeta potential with pH of AKP-30 dispersion as reported by Perez et

al. (2006) (figure 2.7).

Figure 4.13 Mean coordination numbers (M) as a function of close-packed volume fraction for several kinds of packings (Uri et al., 2006)

Coo

rdin

atio

nnu

mbe

r (M

)

Close-packed volume fraction

Coo

rdin

atio

nnu

mbe

r (M

)

Close-packed volume fraction

43

y = 11.026x - 0.4023R2 = 0.9778

0

1

2

3

4

5

6

7

8

9

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

Maximum close-packed volume fraction ( m)

Coo

rdin

atio

n nu

mbe

r <K

>

Figure 4.14 Extrapolation of close-packed volume fraction below 0.5

Figure 4.16 shows the variation of measured CCT with pH of AKP-30

dispersion. At pH values of above 6 and less than iso-electric point, CCT is found to

increase with the increase in pH. These results are in line with the theory of Singh and

Tirumkudulu (2007) who predicted an increase in CCT with decrease in final close-

packed volume fraction. In case of flocculated dispersions (8.5 < pH < 11.5), we

found it difficult to measure the exact CCT value since the results were not

reproducible. For example, in some cases the ring type of crack appeared at the

periphery of the dried film. However, if the films were cast carefully, crack free films

of a much higher thickness would be obtained. Consequently, we could not ascertain

the exact CCT for 8.5 < pH < 11.5. The thickness of the film at which circular ring

type crack is appeared is greater than 175 m . So it is clear that CCT of the films

obtained from flocculated dispersions is greater than 175 m . As we have already

discussed in section 4.2.3 that in case of stable dispersions capillary pressure effect is

prominent due to high final close-packed volume fraction. So the tensile stresses

generated in the film obtained from drying of stable dispersion are very high, which

44

causes films to crack at lower thicknesses. On the other hand, in case of flocculated

dispersions, final close-packed volume fraction is very low, so the capillary pressure

effect is less and so the CCT of flocculated dispersion is higher than that of stable

dispersions. We get the same results for the measured CCT above pH of 6 (figure

4.16). If we compare figure 4.15 and 4.16, at pH values (> 6), the variation of CCT

with pH in both the figures is identical. However, at pH values (< 6.0), CCT shows an

opposite trend. Also, the measured CCT at pH ~ 11.9 is very (CCT ~ 175) is eight

times greater than the predicted CCT (CCT ~ 22). This is unexpected and needs

further investigation. One of the reason for this trend can be the high solubility of

alumina particles at low pH values. At very low pH values (pH < 4) and very high pH

values (pH > 12), alumina particles are highly soluble in water. So at these pH values,

when we allow the dispersion to dry, alumina precipitate out as drying progresses.

45

0

20

40

60

80

100

120

140

160

180

200

0 2 4 6 8 10 12 14

pH of AKP-30 dispersion

CC

T of

AK

P-3

0, (

m)

Figure 4.15 Predicted Critical Cracking Thickness (CCT) as a function of pH of

AKP-30 dispersion. Filled symbols represent predictions and empty symbols represent experimental measurements.

Figure 4.16 The measured CCT as a function of pH for AKP-30 dispersions

0

20

40

60

80

100

120

140

160

180

200

0 2 4 6 8 10 12 14

pH of AKP-30 dispersion

CC

T of

AK

P-3

0, (

m)

CCT > 175 m

0

20

40

60

80

100

120

140

160

180

200

0 2 4 6 8 10 12 14

pH of AKP-30 dispersion

CC

T of

AK

P-3

0, (

m)

CCT > 175 mCCT > 175 m

46

4.3 Critical Cracking Stress (CCS):

Theoretical critical cracking stresses (equation 2.1) for different pH values

were calculated by taking the values of AW as 72mN/m, Shear Modulus (G) equal to

156 GPa and the m value from capillary experiments (figure 4.1).

As can be seen in figure 4.17, at low pH, the critical cracking stress is highest

essentially due to high close-packed volume fraction and high coordination number.

As the pH is increased, the dispersion becomes more unstable due to decrease in

stabilizing repulsive forces between particles and results in particle flocculation. At

iso-electric point (pH=9.1), extent of flocculation is maximum and close-packed

volume fraction is minimum, so the critical cracking stress is minimum. With increase

in pH above iso-electric point, the critical cracking stress increases. Experimental

CCS values were found to be double the theoretical CCS values. If we increase the

constant in equation 2.1 by a factor of 2, predicted and measured critical cracking

stress values match well (figure 4.18). Experimental critical cracking stress variation

as a function of pH of AKP-30 dispersion is found to be in good agreement with the

critical cracking stress variation predicted by the theory of Tirumkudulu and Russel

(2005) (figure 4.17).

47

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

0 2 4 6 8 10 12 pH of AKP-30 dispersion

cR

Film Height (N) = 150Film Height (N) = 200Film Height (N) = 250Film Height (N) = 300

Figure 4.17 The predicted and measured dimensionless critical cracking stress )2/( Rc as a function of pH of AKP-30 dispersion at different film heights. Filled

symbols represent predictions and empty symbols represent the values obtained by interpolation of experimental stress values from line equation shown in figure 4.19.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

0 2 4 6 8 10 12 pH of AKP-30 dispersion

cR

Film Height (N) = 150Film Height (N) = 200Film Height (N) = 250Film Height (N) = 300

Figure 4.18 The predicted and measured dimensionless critical cracking stress )2/( Rc as a function of pH of AKP-30 dispersion at different film heights. Filled

symbols represent predictions by increasing factor in equation 2.1 by 2 and empty symbols represent the values obtained by interpolation of experimental stress values

from line equation shown in figure 4.19.

48

Figure 4.19 show the variation of dimensionless critical cracking stress vs

dimensionless film thickness. It is clear that the critical cracking stress decreases with

increase in film thickness as predicted by the model of Tirumkudulu and Russel

(2005). When an isolated crack is initiated in the film, stresses over the length of h

around the crack will be released from the film (Singh and Tirumkudulu, 2007). The

elastic energy recovery per unit length of the crack scales as 2~ hEelastic and

increased surface energy due to formation of crack varies as hEs ~ . So if we

increase the film thickness, elastic energy stored in the film will increase with 2h

where as surface energy will increase with h . It means that thick films have sufficient

elastic energy to crack at lower stress. The experimental results also show the same

analysis as predicted (figure 4.19).

0.1

1

10

10 100 1000N = hrcp/2R

c R

pH = 3.3

pH = 6.5

pH = 7.8

pH = 10.0

pH=11.1

Figure 4.19 Plot of measured dimensionless critical cracking stress )2/( Rc vs the dimensionless film thickness )2/( Rhrcp .

49

In figure 4.20, the dimensionless critical cracking stress )2/( Rc is plotted

against the dimensionless characteristic scale )/( 2NM m for dispersions with pH

values of 3.3, 6.5, 7.8, 10.0 and 11.1 respectively. The solid line is a fit from the

model of Tirumkudulu and Russel (2005) (equation 2.1) with an exponent of 1/3. The

plot shows that the measured dimensionless critical cracking stress values matches

well with the predictions of Tirumkudulu and Russel (2005) (after increasing the

coefficient in equation 2.1 by a factor of 2). Also, if we calculate the value of power

of )/( 2NM m from the experimental dimensionless critical cracking stress

)2/( Rc vs the dimensionless characteristic scale )/( 2NM m plot as shown in

figure 4.20, we get the power of )/( 2NM m as ~ (1/3), which gives the validity of

model over wide range of packing fractions. So we can say that the experiments are in

good agreement with the model of Tirumkudulu and Russel (2005). Note that the

coefficient in equation 2.1 was increased by a factor of 2 while comparing this

equation with the experimental results.

50

y = 22.96x0.3333

R2 = 0.8894

0.1

1

10

0.00001 0.0001 0.001

M m/N2

c R

/2

pH=3.3pH=6.5pH=11.1pH=7.8pH=10.0

Figure 4.20 Plot of measured dimensionless critical cracking stress )2/( Rc vs the dimensionless characteristic scale )/( 2NM m . The data points are for films of AKP-30 dispersion with varying pH. The solid line is a power law fit with an exponent of

1/3. C is the constant.

In many of the industrial applications such as paints and coatings, we need

films which are sufficiently thick as well as at the same time having no cracks. Our

study can help to design such formulations where we can get crack free film of

desired thickness. Also, in ceramics industry, it is required to fuse the particles in the

dried film at high temperature. So it is necessary that films should not crack at large

thicknesses as crack will make it difficult to fuse the particles. So this study can be

used to get crack free surface in ceramic industry.

51

CHAPTER 5

CONCLUSIONS

Cracking in drying films of flocculated dispersions has been investigated. It

has been shown that by changing the pH of the dispersion, its flocculation can be

affected which can further be used to control its cracking behavior. As pH is

increased, final packing fraction value decreases, goes through minimum and again

starts increasing. The iso-electric point obtained by using final close packed volume

fraction variation with pH for both AKP-30 as well as ALCOA-A16 is almost equal to

the value that is given in literature. Shape of the films formed after drying AKP-30

dispersions at different pH values gives the direct evidence for the capillary pressure

as driving force for the convection of particles during drying of films. We observed

different types of crack patterns as regular crack pattern (pH < 7), disordered crack

pattern (7.25 < pH < 7.75) and an unique circular crack (7.75 < pH < 11.5).

We also found that experimental critical cracking stress values are two times

of the theoretically predicted values. Experimental studies indicate that there is clear

influence of the way in which particles pack on the cracking behavior of films. We

were able to change the packing fraction over a wide range from 0.2 to 0.7 by

changing the pH of the AKP-30 dispersion. The trend of measured critical cracking

stress with pH is similar to the trend predicted by the model of Tirumkudulu and

Russel (2005). Also, the measured critical cracking thickness variation for pH values

in between 6 and 12 is similar to the trend predicted by model of Singh and

Tirumkudulu (2007). In short, we can say that the cracking behavior of AKP-30 in

water is well described by these two models, suggesting that these models are

universal and apply equally well to both stable as well as flocculated systems.

52

In this study, we have considered that the particle asphericity do not affect the

cracking behavior of flocculated dispersions. So it is required to carry out more

experiments with perfectly spherical alumina particles and compare the results.

53

REFERENCES

1. Brodnyan, J. G. and Konen, T., 1964, “Experimental study of the mechanism

of film formation”, Journal of Applied Polymer Science, 8, 687.

2. Brown, G. L., 1956, “Formation of films from polymer dispersions”, Journal

of Polymer Science, 22, 423-434.

3. Carreras, E. S., Chabert, F., Dunstan, D. E. and Franks G. V., 2007 “Avoiding

“mud” cracks during drying of thin films from aqueous colloidal suspensions”,

Journal of Colloid and Interface Science, 313, 160-168.

4. Channell, G. M. and Zukoski, C. F., 1997, “Shear and compressive rheology

of aggregated alumina suspensions”, AIChE Journal, 43(7), 1700-1708.

5. Chevalier, Y., Graillat, P. C., Joanicot, M., Wont, K., Maquet, J., Linder, P.

and Cabane, B., 1992, “Film formation with latex particles”, Colloid Polymer

Science, 270, 806-821.

6. Chiu, R. C. and Cima, M. J, 1993, “Drying of granular ceramic films: II,

drying stress and saturation uniformity”, Journal of the American Ceramic

Society, 76(11), 2769-2777.

7. Chiu, R. C., Garino, T. J. and Cima, M. J, 1993, “Drying of granular ceramic

films: I, effect of processing variables on cracking behavior”, Journal of the

American Ceramic Society, 76(9), 2257-2264.

8. Crowley, T. L., Sanderson, A. R., Morrison, J. D., Barry, M. D., Morton-

Jones, A. J. and Rennie, A. R., “Formation of bilayers and plateau borders

during the drying of film-forming latices as investigated by small-angle

neutron scattering”, Langmuir, 8, 2110.

9. Dobler, F., Pith, T., Lambla, M. and Holl, Y., 1992, “Coalescence mechanisms

of polymer colloids”, Journal of Colloid and Interface Science, 152, 1-11.

54

10. Dobler, F., Pith, T., Lambla, M. and Holl, Y., 1992, “Coalescense mechanism

of polymer colliods ii Coalescence with evaporation of water”, Journal of

Colloid and Interface Science, 152 (1), 12-21.

11. Dufresne, E. R., Corwin, E. I., Greenblatt, N. A., Ashmore, J., Wang, D. Y.,

Dinsmore, A. D., Cheng, J. X., Xie, X. S., Hutchinson, J. W. and Weitz, D. A.,

2003, “Flow and fracture in drying nanoparticle suspensions”, Physical

Review Letters, 81 (22), 4501-4504.

12. Eckersley, S. T. and Rudin, A., 1990, “Mechanism of film formation from

latexes”, Journal of Coatings Technology, 62 (780), 89-100.

13. Eckersley, S. T. and Rudin, A., 1994, “Drying behaviour of acrylic latexes”,

Progress in Organic Coatings, 23, 387-402.

14. Furlong, D. N., Freeman, P. A. and Lau, A. C. M., 1981, “The adsorption of

soluble silica at solid-aqueous solution interfaces I. Leaching from glass-An

electro-kinetic study”, Journal of Colloid and Interface Science, 80(1), 20-31.

15. Greenwood, R., 2003, “Review of the measurement of zeta potentials in

concentrated aqueous suspensions using electroacoustics”, Advances in

Colloid and Interface Science, 106, 55-81.

16. Henson, W. A., Taber, D. A. and Bradford, E. B., 1953, “Mechanism of film

formation of latex paint”, Industrial and Engineering Chemistry Research, 45

(4), 735-739.

17. Johnson, S. B., Russel, A. S. and Scales, P. J., 1998, “Volume fraction effects

in shear rheology and electro-acoustic studies of concentrated alumina and

kaolin suspensions”, Colloids and Surfaces, 141, 119-130.

18. Juhue, D. and Lang, J., 1994, “Latex film formation in the presence of organic

solvents”, Macromolecules, 27, 695-701.

55

19. Keddie, J. L., Meredith, P., Jones, R. A. L. and Donald, A. M., 1995, “Kinetics

of film formation in acrylic latices studied with multiple-angle-of-incidence

ellipsometry and environmental SEM”, Macromolecules, 28, 2673-2682.

20. Keddie, J.L., 1997, “Film formation of latex”, Material Science and

Engineering, 21, 101-170.

21. Kim, K. D., Sperling, L. H. and Klein, A., 1994, “Reptation time, temperature,

and cosurfactant effects on the molecular interdiffusion rate during

polystyrene latex film formation”, Macromolecules, 27, 6841-6850.

22. Lee, W. P. and Routh, A. F., 2004, “Why do drying films crack?”, Langmuir,

20(23), 9885-9888.

23. Lin, F. and Meier, D. J., 1995, “A study of latex film formation by atomic

force microscopy. I. A comparison of wet and dry conditions”, Langmuir, 11,

2726-2733.

24. Mayer, A., Pith, T. and Hu, G., 1995, “Effect of the structure of latex particles

on adhesion. Part II: Analogy between peel adhesion and rheological

properties of acrylic copolymers”, Journal of Polymer Science, 33, 1793.

25. Mayer, A., Pith, T., Hu, G. and Lambla, M., 1995, “Effect of the structure of

latex particles on adhesion. Part I: Synthesis and characterization of structured

latex particles of acrylic copolymers and their peel adhesion behavior”,

Journal of Polymer Science, 33, 1781.

26. Okubo, M., Seike, M. and Matsumoto, T., 1983, “Studies on suspension and

emulsion. LX. Composite polymer emulsion film with temperature-sensitive

properties”, Journal of Applied Polymer Science, 28, 383.

56

27. Perez, P. G., Pagnoux, C., Rossignol, F. and Baumard, J. F., 2006,

“Heterocoagulation between SiO2 nanoparticles and Al2O3 submicronparticles;

influence of the background electrolyte”, Colloids and Surfaces, 281, 58-66.

28. Petersen, C., Heldmann, C. and Johannsmann, D, 1999, “Internal stresses

during film formation of polymer latices”, Langmuir, 15, 7745-7751.

29. Pokluda, O., Bellehumeur, C.T. and Vlachopoulos, J., 1997, “Modification of

Frenkel’s model for sintering”, AIChE Journal, 43 (12), 3253-3256.

30. Routh, A. F. and Russel, W. B., 2001, “Deformation mechanism during latex

film formation: Experimental evidences”, Industrial and Engineering

Chemistry Research, 40, 4302-4308.

31. Routh, A.F. and Russel, W.B., 1999, “A process for latex film formation:

limiting regimes for individual driving forces”, Langmuir, 15, 7762-7773.

32. Sheetz, D. P., 1965, “Formation of films by drying of latex”, Journal of

Applied Polymer Science, 9, 3759-3773.

33. Singh, K.B. and Tirumkudulu, M.S., 2007, “Cracking in drying colloidal

films”, Physical Review Letters, 98, 218302(1)-218302(4).

34. Steward, P.A., Hearn, J. and Wilkinson, M.C., 2000,”An overview of polymer

latex film formation and properties”, Advances in Colloid and Interface

Science, 86, 195-267.

35. Tirumkudulu, M. S. and Russel, W. B., 2004, “Role of capillary stresses in

film formation”, Langmuir, 20 (7), 2947-2961.

36. Tirumkudulu, M.S. and Russel, W.B., 2005, “Cracking in drying latex films”,

Langmuir, 21, 4938-4948.

57

37. Uri, L., Walmann, T., Alberts, L., Dysthe, D. K. and Feder, J., 2006,

“Structure of plastically compacting granular packings” Physical Review E,

73, 051301.

38. Velamakanni, B. V. and F. F. Lange, 1991, “Effect of interparticle potentials

and sedimentation on particle packing density of bimodal particle distributions

during pressure filtration,” Journal of the American Ceramic Society, 74(1),

166.

39. White, L. R., 1982, “Capillary rise in powders”, Journal of Colloid and

Interface Science, 90, 536-538.

40. Winnik, M. A., 1997, “Latex film formation”, Current Opinion in Colloid and

Interface Science, 2 (2), 192-199.

41. Xia, Z.C. and Hutchinson, J.W., 2000, “Crack patterns in thin films”, Journal

of the Mechanics and Physics of Solids, 48, 1107-1131.

42. Zhao, C. L., Wang, Y., Hruska, Z. and Winnik, M. A., 1990, “Molecular

aspects of latex film formation. An energy transfer study”, Macromolecules,

23, 4082-4087.

43. Zosel, A. and Ley, G., 1993, “Influence of cross-linking on structure,

mechanical properties, and strength of latex films”, Macromolecules, 26,

2222-2227.