© 2008 Joshua James Taylor

130
1 ADSORPTION OF SODIUM POLYACRYLATE IN HIGH SOLIDS LOADING SLURRIES By JOSHUA JAMES TAYLOR A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2008

Transcript of © 2008 Joshua James Taylor

Page 1: © 2008 Joshua James Taylor

1

ADSORPTION OF SODIUM POLYACRYLATE IN HIGH SOLIDS LOADING SLURRIES

By

JOSHUA JAMES TAYLOR

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2008

Page 2: © 2008 Joshua James Taylor

2

© 2008 Joshua James Taylor

Page 3: © 2008 Joshua James Taylor

3

To my Father and Mother, John and Dina Taylor, who have been an encouragement throughout all my education.

Page 4: © 2008 Joshua James Taylor

4

ACKNOWLEDGMENTS

I would first like to thank my advisor Dr. Wolfgang Sigmund for his guidance and support

which has helped me through this project. I would also like to thank my committee members Dr.

Ronald Baney, Dr. Laurie Gower, Dr. Hassan El-Shall, and Dr. Charles Martin for their

intellectual discussions and advise.

I would like to thank each person in Dr. Sigmund’s group who have been a support for me

and offered constructive comments during my research. I would especially like to thank Yi-

Yang Tsai who has offered countless advice and support throughout the last four years.

Additionally, I would like to thank my friends and family who have always been there for

support and encouragement. I especially thank my parents who have always encouraged me to

further my education and have been a support during the challenging times.

Page 5: © 2008 Joshua James Taylor

5

TABLE OF CONTENTS page

ACKNOWLEDGMENTS ...............................................................................................................4

LIST OF TABLES...........................................................................................................................7

LIST OF FIGURES .........................................................................................................................8

LIST OF ABBREVIATIONS AND SYMBOLS ..........................................................................14

ABSTRACT...................................................................................................................................16

CHAPTER

1 INTRODUCTION ..................................................................................................................18

1.1 Calcium Carbonate Industry .............................................................................................18 1.2 Sodium Polyacrylate (NaPAA).........................................................................................18 1.3 High Solids Loading Slurries............................................................................................19 1.4 Objectives and Approach..................................................................................................19

2 LITERATURE BACKGROUND AND CHARACTERIZATION .......................................20

2.1 Calcium Carbonate ...........................................................................................................20 2.2 NaPAA..............................................................................................................................26 2.3 Attenuated Total Reflectance Fourier Transform Infrared Spectroscopy (ATR-

FTIR)...................................................................................................................................33 2.4 Ultraviolet-Visible Spectroscopy (UV/VIS).....................................................................40 2.5 Compression Rheology and Impedance Measurements ...................................................42

3 EXPERIMENTAL METHODS .............................................................................................44

3.1 Adsorption Isotherms........................................................................................................44 3.2 Turbidity Measurements...................................................................................................44 3.3 ATR-FTIR ........................................................................................................................45 3.4 UV/VIS Spectroscopy ......................................................................................................45 3.5 Rheology...........................................................................................................................47

4 SODIUM POLYACRYLATE ADSORPTION ONTO GCC IN HIGH SOLIDS LOADING SLURRIES ..........................................................................................................48

4.1 Adsorption Isotherms........................................................................................................48 4.2 Turbidity of NaPAA .........................................................................................................53 4.3 ATR-FTIR ........................................................................................................................57

4.3.1 Dry and Wet Samples.............................................................................................58 4.3.2 Solvent Exchange ...................................................................................................61

Page 6: © 2008 Joshua James Taylor

6

4.3.3 Water Structure in High Solids Loading Slurries...................................................63 4.3.4 NaPAA CH Band Infrared Spectra.........................................................................65 4.3.5 NaPAA Carboxylate Group Infrared Spectra.........................................................68 4.3.6 Carbonate Infrared Spectra.....................................................................................74 4.3.9 Temperature Dependence.......................................................................................78 4.3.10 Addition of Energy to Aged Slurries....................................................................79

4.4 Probe Molecules ...............................................................................................................82 4.4.1 Benzoic Acid ..........................................................................................................83 4.4.2 Gallic Acid..............................................................................................................87 4.4.3 Propionic Acid........................................................................................................92

5 WATER STRUCTURE DEPENDANCE ON SOLIDS LOADING AND AGING..............93

5.1 ATR-FTIR Water Structure..............................................................................................93 5.2 Water Confinement...........................................................................................................96 5.3 Ions as Water Structure Makers or Breakers ....................................................................98

6 WATER STRUCTURE MAKING AND BREAKING CHEMICALS...............................102

6.1 Calcium Chloride............................................................................................................102 6.2 Sodium Bicarbonate........................................................................................................102 6.3 Ethylene Glycol ..............................................................................................................105 6.4 Propylene Glycol ............................................................................................................108

7 SUMMARY AND DISCUSSION .......................................................................................112

8 CONCLUSIONS AND SUGGESTIONS ............................................................................118

8.1 Conclusions.....................................................................................................................118 8.2 Suggestions .....................................................................................................................119

LIST OF REFERENCES.............................................................................................................120

BIOGRAPHICAL SKETCH .......................................................................................................130

Page 7: © 2008 Joshua James Taylor

7

LIST OF TABLES

Table page 2-1 Several materials used for the ATR crystal. ......................................................................39

Page 8: © 2008 Joshua James Taylor

8

LIST OF FIGURES

Figure page 2-1 Unit cell of calcium carbonate with space group cR3 . A) An angled top view, B) a

side view with the dotted line through the middle being an edged....................................20

2-2 XRD of calcium carbonate powder used in this dissertation. All peaks confirm calcite structure but only the high intensity peaks are labeled with Miller indices. ..........21

2-3 Number, surface area, and volume distribution of GCC while dispersed with NaPAA in H2O as dilute samples. ...................................................................................................22

2-4 The SEM of GCC particles used in this dissertation. ........................................................23

2-5 Adsorption of polymer onto a surface. Polymer is described as containing tails, loops, and trains. ................................................................................................................27

2-6 Adsorption confirmations of a polymer. A) Pancake confirmation has an absorbed layer thickness ~ segment length, B) Confirmation similar to polymer in good solvent give adsorbs with a layer thickness ~ Rg, C) Brush confirmation has an absorbed layer thickness greater than Rg. ..........................................................................29

2-7 Repeat unit of sodium polyacrylate, NaPAA.....................................................................30

2-8 Coordination modes of a carboxylate group......................................................................31

2-9 IR scattering and adsorption in different IR techniques. ...................................................34

2-10 ATR-FTIR setup. Sample is placed onto a crystal. The IR beam reflects several times through the crystal and the evanescent wave interacts at the interface with the sample. ...............................................................................................................................35

2-11 At the interface of the crystal and sample the IR beam penetrates past the surface of the crystal and decays exponentially..................................................................................36

2-12 The electric field amplitude at the interface of the crystal and sample. The electric field decays at an exponential rate in the sample. The depth of penetration, Dp, is determined when the electric field is attenuated to 36.8% of its total intensity. ...............37

3-1 Calibration curve for gallic acid. Absorption measurements preformed at a wavelength of 210 nm with a sample thickness of 1 cm. ..................................................46

3-2 Calibration curve for benzoic acid. Absorption measurements preformed at a wavelength of 224 nm with a sample thickness of 1 cm. ..................................................47

4-1 Titration curve of NaPAA. Black curve is the titration of a NaPAA in deionized water. Red curve is the 1st derivative of the titration curve. .............................................49

Page 9: © 2008 Joshua James Taylor

9

4-2 Calibration curve for NaPAA with the titration technique. ...............................................50

4-3 The adsorbed amount of NaPAA compared to the 100% adsorption line.........................51

4-4 Langmuir adsorption isotherm of NaPAA by GCC in a 75 wt% solids loading slurry. This model is a poor fit with a R2 value of 0.8187. ...........................................................52

4-5 Temkin adsorption isotherm of NaPAA by GCC in a 75 wt% solids loading slurry. This model is a poor fit with a R2 value of 0.8529. ...........................................................52

4-6 Freundlich adsorption isotherm of NaPAA by GCC in a 75 wt% solids loading slurry. Freundlich model fits well with a R2 value of 0.96. ..............................................53

4-7 Turbidity of NaPAA in water with increasing monovalent salt concentration. High concentrations of sodium ions within slurry do not cause the NaPAA to precipitate. ......54

4-8 Turbidity of NaPAA in water with increasing divalent salt concentration. Under processing condition of GCC slurries there is no indication of NaPAA precipitation. .....55

4-9 Turbidity of NaPAA in water with increasing temperature. NaPAA does not precipitate with increasing temperature. ............................................................................56

4-10 Turbidity of NaPAA in water with 10-3M CaCl2 and 10-1M CaCl2 over a three day period. At slurry conditions there was no precipitation of NaPAA..................................57

4-11 IR spectra of the carbonate bending bands located at 1449 cm-1 of a wet and a dry 75 wt% GCC slurry. The wet slurry forms a shoulder at lower wavenumbers indicating the presents of bicarbonate species. ...................................................................................58

4-12 IR spectra of the carbonate stretching bands located at 875 cm-1 of a wet and a dry 75 wt% GCC slurry. The wet slurry has a larger FWHM. ....................................................59

4-13 Second derivative of the IR spectra of the carboxyl region for a wet and a dry sample. Band at 1581 cm-1 shifts to 1586 cm-1 indicated change in coordination............59

4-14 IR spectrum of NaPAA in H2O. The OH stretching band overlaps the CH stretching bands of NaPAA. The OH bending band overlaps part of the COO- band. .....................60

4-15 IR spectrum of NaPAA in D2O. Switching from H2O to D2O shifts the stretching and bending bands so that there is no overlap with the CH or COO- stretching bands of NaPAA...........................................................................................................................61

4-16 IR spectrum of a 75 wt% GCC slurry with NaPAA. Regions of interests are boxed and discussed. ....................................................................................................................62

4-17 IR spectra of the OH stretching bands of H2O and H2O in a 75 wt% slurry. When water is in a high solids loading slurry there is a change in the structure of the water

Page 10: © 2008 Joshua James Taylor

10

indicated with the FWHM decreasing and the 3183 cm-1 shoulder peak shift to 3239 cm-1. ..........................................................................................................................63

4-18 IR spectra of the OD stretching bands of D2O and D2O in a 75 wt% slurry. When D2O is in a high solids loading slurry there is a change in the structure of the water indicated with the FWHM decreasing and the 2502 cm-1 shoulder increasing intensity..............................................................................................................................64

4-19 IR spectrum of the CH band of NaPAA in D2O. ...............................................................66

4-20 IR spectrum of the CH bands of NaPAA in a 75 wt% GCC slurry. Two bands are formed at 2980 cm-1 and 2875 cm-1 indicating adsorption of the NaPAA onto the surface of GCC. .................................................................................................................66

4-21 IR spectra of the CH bands of NaPAA in GCC slurries with increasing concentration of NaPAA. Adsorption limit is between 1 wt% NaPAA and 9 wt% NaPAA. .................67

4-22 IR spectrum of the carboxyl stretching region for NaPAA in D2O. The IR spectrum shows that NaPAA is in the ionic form with a peak at 1570 cm-1 for the COO- and no peak at 1700 cm-1 for the C=O...........................................................................................68

4-23 Second derivative of the IR spectra of the carboxylate region of NaPAA in D2O and in a 75 wt% GCC slurry. NaPAA adsorbs onto GCC in unidentate, bridging, and bidentate modes. ................................................................................................................69

4-24 Second derivative of the IR spectra of 75 wt% GCC slurries with increasing concentration of dispersant. With an excess amount of dispersant there is no unidentate adsorption mode. ..............................................................................................70

4-25 Second derivative of the IR spectra of the carboxylate region with increasing solids loading. As the solids loading of the slurries increases there is a shift of the bands toward a bridging mode. ....................................................................................................72

4-26 Second derivative of the IR spectra of the carboxylate region with aging. The unidentate band shifts from 1585 cm-1 to 1580 cm-1 with increasing age indicating a shift towards the bridging mode. .......................................................................................73

4-27 ATR-FTIR spectrum of GCC. ...........................................................................................74

4-28 IR spectra of carbonate band of GCC in D2O and in a 75 wt% GCC slurry. Formation of a shoulder indicates formation of bicarbonates. ..........................................75

4-29 IR spectra of the carbonate band of GCC slurries showing a formation of a shoulder with increase solids loading. ..............................................................................................76

4-30 IR spectra of the carbonate band of GCC in a 75 wt% slurry with aging..........................77

Page 11: © 2008 Joshua James Taylor

11

4-31 IR spectra of the carbonate band with a slurry at different temperatures. As the temperature decreases the shoulder shifts from 1311 cm-1 to 1335 cm-1...........................78

4-32 IR spectra of a 75 wt% GCC slurry before and after a heating cycle. The only difference of the two spectra is the carbonate band. ..........................................................79

4-33 IR spectra of the OD band in a 75 wt% GCC slurry at different temperatures. Lower temperatures increase the concentration of structured water. ............................................80

4-34 IR spectra of the OD band in an aged 75 wt% GCC slurry with addition of energy to the system. Adding energy to the system partially restores the fluid like water structure..............................................................................................................................81

4-35 IR spectra of the carbonate band in an aged 75 wt% GCC slurry with addition of energy to the system. Adding energy to the system partially restores the carbonate band to a fresh slurry..........................................................................................................81

4-36 Molecular structure of benzoic acid...................................................................................82

4-37 Adsorption of benzoic acid onto GCC at varying solids loading. Benzoic acid does not adsorb onto GCC. ........................................................................................................83

4-38 Second derivative of the IR spectra of benzoic acid in D2O, 20 wt% GCC slurry, and 57 wt% GCC slurry. The bands for the benzene ring and the carboxylate do not shift, indicating no adsorption............................................................................................84

4-39 IR spectrum of the CH bands of benzoic acid in D2O and in a 57 wt% GCC slurry. Formation of new bands indicates interaction of the CH bonds with GCC.......................85

4-40 Molecular structure of gallic acid. .....................................................................................86

4-41 Adsorption of gallic acid onto GCC at varying solids loading..........................................87

4-42 Second derivative of the IR spectra of gallic acid in D2O and a 20 wt% GCC slurry. The bands for the benzene ring and the carboxylate shift, indicating adsorption..............88

4-43 IR spectrum of the CH bands of gallic acid in D2O and in a 67 wt% GCC slurry. Formation of new bands indicates interaction of the CH bonds with GCC.......................89

4-44 Second derivative of the IR spectra of gallic acid in a 20 wt% and 67 wt% GCC slurry. The bands for the benzene ring and the carboxylate shift, indicating change in the coordination of the benzoic acid with increasing solids loading. ............................90

4-45 Molecular structure of propionic acid................................................................................91

4-46 IR spectrum of the CH bands of propionic acid in D2O and in a 70 wt% GCC slurry. Formation of new bands indicates interaction of the CH bonds with GCC.......................91

Page 12: © 2008 Joshua James Taylor

12

4-47 Second derivative of the IR spectra of propionic acid in D2O, 14 wt% GCC slurry, and a 70 wt% GCC slurry. No shift in the carboxylate band indicates that propionic acid does not adsorb...........................................................................................................92

5-1 IR spectra of the OD stretching bands in GCC slurries ranging from 10 wt% to 75 wt%. As the solids loading increases there is an increase in the fluid structure (2484 cm-1) and decrease in solid structure (2390 cm-1)....................................................93

5-2 Fluid to solid water structure within different solids loading slurries. There are two different regions, the first from 10 to 50 wt% and the second above 50 wt%...................94

5-3 IR spectra of the OD band in a 75 wt% slurry with aging. Increasing age causes an increase in the concentration of structured water in the slurry. .........................................95

5-4 Fluid to solid water structure within an aging 75 wt% GCC slurry. The fluid like water structure decreases with aging becoming more like a slurry without dispersant.....96

5-5 Calculations of the distance between GCC particles in different solids loading slurries with NaPAA having a radius of gyration of 2 nm, 3 nm, and 4 nm. ....................97

5-6 Calculations of the distance between NaPAA molecules in different solids loading slurries with a radius of gyration of 2 nm, 3 nm, and 4 nm. ..............................................98

5-7 Change of pH in an aging 75 wt% GCC slurry over a six day time span. A decrease in pH is due to the concentration change of species in the slurry......................................99

5-8 IR spectra of the OD band of a 75 wt% GCC slurry with increase NaPAA. There is a decrease in the structured water indicating that NaPAA is a water structure breaker in a GCC slurry. ...................................................................................................................100

6-1 IR spectra of the OD band of slurries with and without CaCl2. Ca2+ as a water structure maker prevents the increase in fluid like water of a GCC slurry......................103

6-2 IR spectra of the carbonate band of slurries with and without CaCl2. Ca2+ prevents some of the interactions with the carbonate species. .......................................................103

6-3 IR spectra of the OD band of a 75 wt% GCC slurry with 0.19M sodium bicarbonate with aging. As a weak structure maker, sodium bicarbonate allows for an increase in the structured water with aging of a 75 wt% slurry. ........................................................104

6-4 Rheology of 75 wt% GCC slurries with and without sodium bicarbonate. Sodium bicarbonate increases the viscosity. .................................................................................105

6-5 Structure of ethylene glycol. ............................................................................................106

6-6 IR spectra of the OD band of a 75 wt% GCC slurry with 0.5M ethylene glycol with aging. Ethylene glycol as a water structure breaker prevents structure water from forming while the slurry ages. .........................................................................................106

Page 13: © 2008 Joshua James Taylor

13

6-7 IR spectra of the carbonate band of slurries with and without ethylene glycol. Ethylene glycol prevents interactions with the carbonate species as it ages. ..................107

6-8 Viscosity measurements of 75 wt% slurries with and without ethylene glycol. Ethylene glycol decreases the viscosity more than adding an equal weight amount of water to the slurry. ...........................................................................................................108

6-9 Structure of propylene glycol...........................................................................................109

6-10 IR spectra of the OD band of a 75 wt% GCC slurry with 0.43M ethylene glycol with aging. Propylene glycol as a water structure breaker prevents structured water from forming while the slurry ages. .........................................................................................109

6-11 Viscosity measurements of 75 wt% slurries with and without propylene glycol. Propylene glycol decreases the viscosity at low shear rates but increases viscosity at higher shear rates. ............................................................................................................110

7-1 Adsorption of NaPAA onto GCC due to an increase in entropy with the release of structured water. A) NaPAA in solution before adsorbing onto GCC surface, B) NaPAA adsorption releases structure water from the carboxylate groups and the calcium as demonstrated in the train of the polymer. ......................................................117

Page 14: © 2008 Joshua James Taylor

14

LIST OF ABBREVIATIONS AND SYMBOLS

A absorbance

ATR attenuated total reflectance

α intramolecular expansion factor

b Freundlich exponent

C unadsorbed NaPAA (mg/L)

c concentration of the absorbing species

Dp depth of penetration

DRIFTS diffuse reflectance infrared Fourier transform spectroscopy

E electric field amplitude

ε molar absorptivity coefficient

FTIR Fourier transform infrared spectroscopy

FWHM full width at half maximum

GCC ground calcium carbonate

IR infrared

K Freundlich constant

L the path length through the sample

l effective segment length

M molecular weight

M0 segment molecular weight

N number of segments

n refractive index

NaPAA sodium polyacrylate

PCC precipitated calcium carbonate

π pi

Page 15: © 2008 Joshua James Taylor

15

Q adsorbed NaPAA per GCC (mg/g)

SEM scanning electron microscope

T transmittance

θ angle of incidence

UV/VIS Ultraviolet-visible

υ wavenumber W wavenumber

XRD x-ray diffraction

z distance from crystal surface

Page 16: © 2008 Joshua James Taylor

16

Abstract of Dissertation Presented to the Graduate School of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

ADSORPTION OF SODIUM POLYACRYLATE IN HIGH SOLIDS LOADING SLURRIES

By

Joshua James Taylor

August 2008 Chair: Wolfgang M. Sigmund Major: Materials Science and Engineering

The world demand for calcium carbonate has been increasing by 7% per year since 2002

reaching a world capacity of 71.7 megatons in 2007. The demand continues to increase due to

the diverse applications of calcium carbonate such as building materials, medicines, additives to

food, filler for plastics and paper, and more. Calcium carbonate is often stored and transported

by dispersing it into an aqueous medium with a polyelectrolyte to achieve up to 75 wt% solids

loading. One of the most frequently applied dispersants for calcium carbonate is sodium

polyacrylate (NaPAA). Higher solids loading of the calcium carbonate slurries are desired to

increase storage capacity and decrease transportation cost. In order to achieve higher solids

loading slurries the science behind the adsorption of NaPAA onto the calcium carbonate within

high solids loading slurries must be understood. Currently all research which has been

performed on the adsorption of NaPAA onto calcium carbonate has been performed in dilute

systems. The goal of this research is to understand the adsorption of NaPAA onto calcium

carbonate in high solids loading slurries up to 75 wt% ground calcium carbonate (GCC).

The adsorption of NaPAA onto calcium carbonate was investigated utilizing several

techniques including adsorption isotherms, turbidity measurements, attenuated total reflectance

Fourier transform infrared spectroscopy (ATR-FTIR), and probe molecule adsorption. The

Page 17: © 2008 Joshua James Taylor

17

adsorption of the NaPAA was determined to be due to the chelating ability of the carboxylate

groups with calcium carbonate. The carboxylate groups were determined to adsorb through

unidentate, bidentate, and bridging modes. Also, the mode of adsorption of the carboxylate

group was dependant on the solids loading and age of a slurry system. Further analysis revealed

the CH groups of adsorbed NaPAA were interacting with the surface of the GCC.

Another novel discovery demonstrated that the water structure within a GCC slurry

dispersed with NaPAA is dependant on solids loading and age. With increasing solids loading

there is a decrease in the concentration of structured water. Also, with an increase in age there is

an increase in the concentration of structured water. Chemicals with the ability to either make or

break water structure were introduced into 75 wt% solids loading slurries. The water structure

breakers demonstrated that their inclusion in the slurry decreased viscosity and prevented an

increase in the structured water due to aging.

Page 18: © 2008 Joshua James Taylor

18

CHAPTER 1 INTRODUCTION

1.1 Calcium Carbonate Industry

The world capacity for ground calcium carbonate (GCC) reached 71.7 megatons in 2007.

The capacity has grown by 7% per year since 2002 and continues to increase world wide. GCC

is diverse in its applications which include filler in plastics and paper, building materials,

fertilizer, medicines, additive to foods, and more. The paper industry alone accounts for around

38% of the calcium carbonate demand. Transportation and storage of calcium carbonate as

paper fillers are provided in either high solids loading slurries (60-75 wt%) or as dewatered

powder.

Due to calcium carbonate’s increasing demand there has also been an increase in published

research concerning calcium carbonate. Improved understanding of calcium carbonate

dispersion in aqueous systems has been desired in order to increase the efficiency of production

and decrease costs. As will be discussed in chapter 2, literature contains many papers concerned

with dilute systems of calcium carbonate but lacks research on high solids loading GCC slurries.

1.2 Sodium Polyacrylate (NaPAA)

Polyacrylic acid and its salt sodium polyacrylate (NaPAA) are one of the most frequently

applied polyelectrolytes in industry and the household. A few examples include laundering

processes, thickening agents, and dispersion of clay and calcium carbonate. NaPAA is used in

the calcium carbonate industry as a dispersant for the mineral because the solids loading can be

increased to 75 wt% while maintaining the desired viscosity. However, the complete role of

stabilization and confirmation of NaPAA in dispersing calcium carbonate at high solids content

is not clear. Previous studies about the adsorption of NaPAA will be discussed in Section 2.2.

Page 19: © 2008 Joshua James Taylor

19

1.3 High Solids Loading Slurries

As mentioned above, GCC is dispersed with NaPAA to achieve high solids loading slurries

(75 wt%). The calcium carbonate industry is constantly working to increase solids loading while

maintaining desired slurry properties in order to decrease production costs. In order to achieve

higher solids loading slurries the science behind the adsorption of NaPAA onto the GCC within

high solids loading slurries must be understood. Several published papers have studied the

adsorption of NaPAA onto calcium carbonate but all of the research has been performed on

dilute systems (less than 5 wt%) [1-15]. This is mainly due to the difficulties in measuring

adsorption in high solids content. Therefore, there is a necessity for understanding the

interaction of dispersants with particles in high solids loading slurries.

1.4 Objectives and Approach

This dissertation will focus on understanding the interaction of NaPAA, GCC, and water

within slurries with solids loading up to 75 wt%. Therefore, this work has the following goals:

• Determine adsorption mechanism of NaPAA onto GCC in high solids loading slurries. • Determine water structure’s relationship to solids loading and aging. • Propose new chemicals for improving GCC dispersion.

There are several analysis techniques that will be utilized to accomplish the listed goals.

Attenuated total reflectance Fourier transform infrared spectroscopy (ATR-FTIR) will allow for

determination of the chemical interaction from the chemical groups containing a dipole moment.

ATR-FTIR is an excellent technique for high solids loading slurries and will be discussed in

more detail in Section 2.3. Results from the IR spectra will be supplemented with adsorption

isotherms and adsorption of probe molecules in high solids loading slurries. Additional data

from viscosity measurements will aid in the understanding of high solids loading GCC slurries.

Page 20: © 2008 Joshua James Taylor

20

CHAPTER 2 LITERATURE BACKGROUND AND CHARACTERIZATION

2.1 Calcium Carbonate

Calcium carbonate is one of the most common and abundant mineral on earth and it has

three different crystalline structures: vaterite, aragonite, and calcite. Of the three crystal

structures calcite is the most stable structure. Calcite has a trigonal crystal system with space

group cR3 , see figure 2-1. Calcite is used in industry as either precipitated calcium carbonate

(PCC) or ground calcium carbonate (GCC).

Figure 2-1. Unit cell of calcium carbonate with space group cR3 . A) An angled top view, B) a side view with the vertical dotted line through the middle being an edged.

Page 21: © 2008 Joshua James Taylor

21

PCC is made by direct carbonation of hydrated lime, known as the milk process. High

purity calcium carbonate rock is crushed and heated to form lime and carbon dioxide (equation

2-1). Next, the lime is added to water to form calcium hydroxide (equation 2-2). Finally, the

calcium hydroxide is combined with carbon dioxide and calcium carbonate precipitates out

(equation 2-3).

CaCO3 + Heat → CaO + CO2 (2-1)

CaO + H2O → Ca(OH)2 (2-2)

Ca(OH)2 + CO2 → CaCO3 + H2O (2-3)

PCC’s shape and size differ from GCC and allow it to be used in different applications. PCC has

a narrower particle size distribution and the particle shape can be tailored for the application.

During processing the moisture associated with PCC can cause problems.

Figure 2-2. XRD of calcium carbonate powder used in this dissertation. All peaks confirm calcite structure but only the high intensity peaks are labeled with Miller indices.

Page 22: © 2008 Joshua James Taylor

22

GCC is removed from the ground and does not go through all the processing steps required

for PCC; therefore, the particles’ shape and size distribution differ from PCC. GCC particles

shapes are seen to be irregularly rhombohedral and the size distribution of GCC is wider than

PCC.

0 2 4 6

0

2

4

6

8

10

Perc

ent

Diameter (µm)

Number Surface Area Volume

Figure 2-3. Number, surface area, and volume distribution of GCC while dispersed with NaPAA in H2O as dilute samples.

The GCC used in this research was provided by Imerys with a purity of greater than 99%.

X-ray diffraction (XRD) was performed on the particles to determine their structure (figure 2-2)

using the XRD Philips APD 3720. Upon analysis of the diffraction pattern the GCC was

confirmed to be calcite. As mentioned previously the size distribution of GCC is wide;

Page 23: © 2008 Joshua James Taylor

23

therefore, particle size measurements were performed with the Coulter LS13320 with a light

diffraction technique at the Particle Engineering Research Center at the University of Florida.

The size distribution is displayed with differential number, surface area, and volume versus

particles size in figure 2-3. The mean sizes for number, surface area, and volume are 0.13 µm,

0.56 µm, and 1.43 µm, respectively. Ninety percent of the sample contains particles with

diameters less than 0.22 µm but 90 % of the volume of the GCC particles is provided by particles

larger than 0.23 µm.

Figure 2-4. The SEM of GCC particles used in this dissertation.

For this research the most important measurement for particles size is the surface area

because adsorption of NaPAA occurs at the surface of the particles. The data demonstrates that

Page 24: © 2008 Joshua James Taylor

24

90% of the surface area is provided by particles with less than 1.35 µm diameters. The size

distribution of the particles is also demonstrated with a scanning electron microscope (SEM)

image of the powder, figure 2-4. Also, the specific surface area is important for calculating

adsorption isotherms; therefore, the Nova 1200 instrument in the Particles Engineering Research

Center at the University of Florida was utilized to measure the specific surface area of the GCC

using the Brunauer-Emmett-Teller method. The specific surface area was determined to be

5.36 m2/g which is within the range of published data [16-18].

Calcium carbonate is slightly soluble in water and shows a complex behavior which is due

to the complex chemical equilibrium between the mineral/water interface. The distribution of

ionic species in the system is constantly changing while moving toward equilibrium. The

following equations demonstrate the possible ionic species that form within a calcium carbonate

and water system.

CaCO3(s) Ca2+ + CO32- (2-4)

H2CO3 H+ + HCO3- (2-5)

HCO3- H+ + CO3

2- (2-6)

Ca2+ + HCO3- CaHCO3

+ (2-7)

Ca2+ + OH- CaOH+ (2-8)

H+ + OH- H2O (2-9)

Knez et al. [19] discuss the chemical equilibrium of a calcium carbonate and water system in

relation to pH in detail. They mention that high solution pH promotes the dissociation of

carbonate species (H2CO3 and HCO3-) and if the system is in equilibrium with atmospheric CO2

(GCC slurries are open to the atmosphere during processing) then the activities of all carbonate

species in the solution rise with pH. Since the system is constantly moving toward equilibrium

Page 25: © 2008 Joshua James Taylor

25

then the Ca2+ activity must align it self accordingly and the activity of Ca2+ falls with rising pH.

Finally, they conclude that the optimum stability conditions are in the pH range of about 9-11.

Geoffroy et al. [4] also go into detail about the surface chemistry of calcium carbonate in

water. They focus on the pH range from 8 to 11 which is important for this research since the

pH of GCC slurries prepared in water with NaPAA have a pH of about 10. At pH lower than 8

the calcium carbonate dissolves while at pH higher than 11 the Ca(OH)2 precipitates. Geoffroy

et al. describe the ionic surface sites of calcium carbonate as hydrated forms of –Ca+ and –CO3-.

Equations 2-10 and 2-11 show reactions of the –CaOH surface sites but the pK of the reaction

shown in equation 2-11 is far above the pH of 10.

–CaOH + H+ → –Ca(OH2)+ (2-10)

–CaOH + OH- → –CaO- + H2O (2-11)

Equations 2-12 and 2-13 show the reactions of the –CO3H surface sites but the pK of the reaction

shown in equation 2-13 is far below the pH of 10.

–CO3H + OH- → –CO3(OH2)- (2-12)

–CO3H + H+ → –CO3H2+ (2-13)

Geoffroy et al. come to the conclusion that the calcium carbonate surface sites within the pH

range of 8 to 11 consist of mainly neutral sites (–CaOH and –CO3H) and ionic sites (–Ca+ and

–CO3-). Several other published papers have also discussed these equilibrium reactions which

are present in a calcium carbonate and water system [20-27].

Several papers have been published which include zeta potential measurements of

calcium carbonate in water [6, 16, 19, 23, 24, 26, 28-33]. The measurements of the surface

charge of calcium carbonate are not consistent. Moulin et al. [23] have summarized the results

of many researchers showing that the zeta potential measurements in calco-carbonic equilibrium

Page 26: © 2008 Joshua James Taylor

26

conditions have been measured as positive, negative, and variable. The large variation of the

sign of the zeta potential is due to the measurement conditions and the nature of the potential

determining ions in the system. Authors considered the thermodynamic equilibrium within the

liquid phase but did not consider the equilibrium at the gas-liquid interface. Moulin et al. take

into consideration both equilibria when designing their experiment and come to the conclusion

that the values of the zeta potential at the calco-carbonic equilibrium are canceled but is

primarily negative on both sides of the equilibrium. Finally, they determine that the potential

determining ions in the system are not OH- and H+ but are Ca2+ and HCO3-.

Another important aspect of calcium carbonate is the interaction of calcium ions with

water. Published research has shown that calcium prefers to form six-, seven-, and eight-

coordinate structures with water [34-37]. Katz et al. [34] found that the coordination structures

are asymmetrical when the calcium ions coordinate with water in an odd number such as five or

seven. The coordination of calcium with water plays an important role in the adsorption and

coordination of carboxylates as described in Section 2.2.

2.2 NaPAA

Adsorption of polymers onto surfaces is commonly utilized to disperse particles. When

particles with adsorbed polymers approach each other there is a repulsive entropic force due to

the entropy of confining the polymers which is known as steric repulsion. If the steric repulsion

force is greater than the van der Waals attractive force then the particles remain dispersed. The

polymer may take on several different conformations within the solvent and while adsorbed onto

the surface of the particles.

Depending on the segment to segment interactions of a polymer in a solution, it may

assume different conformations. If the segment to segment interactions of the polymer are weak

then the polymer assumes a random coil shape. For a polymer with a random coil shape an

Page 27: © 2008 Joshua James Taylor

27

important length scale is the root mean square radius which is known as the radius of gyration,

Rg, as given by equation 2-14

660M

MlNlRg == (2-14)

where l is the effective segment length, N is the number of segments, M is the molecular weight,

and M0 is the segment molecular weight. Equation 2-14 is only valid as long as there are no

repulsive, attractive, or excluded volume interactions between the segments of the polymer in the

ideal solvent.

Figure 2-5. Adsorption of polymer onto a surface. Polymer is described as containing tails, loops, and trains.

In non-ideal solvents the effective size of the polymer can be smaller or larger than the radius of

gyration and is referred to as the Flory radius, RF, given by equations 2-15 and 2-16

gF RR α= (2-15)

53

ln≈FR (2-16)

where α is the intramolecular expansion factor and is unity when the polymer is in an ideal

Page 28: © 2008 Joshua James Taylor

28

solvent. α will exceed unity when the polymer is dissolved into a good solvent and the polymer

swells and expands. If the polymer is in a poor solvent the segment to segment interactions are

strong and the polymer will collapse into a compact structure, α is less than one.

The adsorption of a polymer onto a surface is shown in figure 2-5. The adsorbed part of

the polymer is called a train and the non-adsorbed part of the polymer between two trains is

called a loop. The part of the polymer that is not adsorbed and is only adjacent to one train is

called a tail. Different conformations of the polymer on the surface contain varying

concentrations of each segment. Figure 2-6 demonstrates that a polymer with a high train

concentration adopts a pancake conformation. The layer thickness of a pancake conformation is

about equal to its segment size. As the concentration of trains decreases the polymer reaches a

point at which the layer thickness is similar to the radius of gyration. At this point the surface

area covered per molecule is approximately described by equation 2-17.

22

21 NlRaSurfaceAre g ≈≈ π (2-17)

As the tail of the polymer extends further from the surface and increases length, the polymer

adopts a brush conformation. Assuming the segment width is close to the segment length, l, the

fully extended polymer has a projected area approximately described by equation 2-18.

2NlaSurfaceAre ≈ (2-18)

With a good solvent the most important factor for controlling the conformation of an adsorbed

polymer is the concentration of the polymer in the solvent. A low concentration will provide a

pancake conformation while increasing the concentration eventually leads to a brush

conformation.

Polymers whose repeat unit contains an electrolyte group are called polyelectrolytes. The

conformations of polyelectrolytes are similar to non-charged polymers except the polyelectrolyte

Page 29: © 2008 Joshua James Taylor

29

Figure 2-6. Adsorption confirmations of a polymer. A) Pancake confirmation has an absorbed layer thickness ~ segment length, B) Confirmation similar to a polymer in good solvent adsorbs with a layer thickness ~ Rg, C) Brush confirmation has an absorbed layer thickness greater than Rg.

Page 30: © 2008 Joshua James Taylor

30

conformations are also dependant on salt concentration. The charges on the polyelectrolyte

cause the segments of the polymer to repel each other and give the polymer an extended

conformation in salt free water [38, 39]. Addition of salt screens the charges on the

polyelectrolyte and eventually will collapse the polymer into a conformation similar to a non-

charged polymer. Polyelectrolytes are often chosen for dispersion instead of non-charged

polymers because the electrolyte group can be chosen to adsorb onto specific surface sites of a

particle. This research has chosen a polyelectrolyte, NaPAA (molecular weight of 5,967 g/mol

and a 2.04 degree of polydispersity), which is known to have a high affinity to calcium [1, 2, 4-

11, 13-15, 40].

Figure 2-7. Repeat unit of sodium polyacrylate, NaPAA.

NaPAA is one of the most commonly used dispersing agents in the papermaking

applications [19]. NaPAA is composed of a long linear hydrocarbon backbone which contains

one carboxyl group for each repeat unit, see figure 2-7. Several published papers have devoted

their research to the interaction of carboxyl groups and/or polyacrylates with surfaces and ions

Page 31: © 2008 Joshua James Taylor

31

[12, 36, 40-74]. There are several papers that discuss the adsorption of polyacrylates onto

calcium carbonate [1, 2, 4-15, 75]. A few of the key concepts and papers will be discussed.

Figure 2-8. Coordination modes of a carboxylate group.

The carboxylate group has been shown to interact with metal cations and surfaces in four

different modes [4, 11, 36, 49, 59, 76], namely, ionic, bridging, bidentate, and unidentate as

shown in figure 2-8. Several of the papers have investigated the different modes with IR

spectroscopy because the change in bond symmetry can be detected. The ionic, bridging, and

bidentate modes have similar group symmetry. The two oxygen atoms in the bidentate mode are

interacting with one metal cation; therefore, there will be a change in the symmetrical and

asymmetrical vibration frequencies. Since the unidentate mode has one oxygen atom

coordinated with one metal cation the symmetrical and asymmetrical vibration frequencies will

be similar to a carboxylic acid group.

Geffroy et al. [4] and Dobson et al. [49] have shown that adsorption is preferred through

chelation of dicarboxylates. A five member chelate ring (consisting of the metal cation and two

adjacent carboxylates) are the most stable followed by six and seven member chelate rings. Lu

et al. [36] further explains that carboxylate groups interact with calcium ions in

Page 32: © 2008 Joshua James Taylor

32

three-dimensional seven- or eight-fold coordination and it is common for calcium ions to

coordinate with both carboxylate groups and water. Katz et al. [34] also demonstrate that both

unidentate and bidentate coordination of carboxylate groups with calcium cations are possible

when the calcium ions bind in seven- and eight-fold coordination.

Since calcium carbonate is slightly soluble in water the dissociation of calcium ions play

an important role on the conformation and shape of the NaPAA. Schweins et al. [13, 14, 71]

have discussed the collapse of NaPAA by calcium ions in three of their papers. The bound Ca2+

ions to the NaPAA chains cause the NaPAA to become much more hydrophobic. Also, all of the

binding modes of the NaPAA with Ca2+ can allow for intramolecular bridging which decreases

the coil dimensions. The collapse of the NaPAA eventually leads to a spherical shape with

different transitional shapes depending on the concentration of NaCl. Another interesting

discovery by Schweins et al. is the expansion of the NaPAA chains with an increase in Na+ ions.

Ca2+ ions are strongly bound to the NaPAA but with increasing concentration of Na+ there is a

competition between the Ca2+ ions and the Na+ ions.

Two other papers of interest are written by Geffroy et al. [5] and Sinn et al. [15] who

discuss the heat of exchange of Ca2+ binding onto NaPAA as endothermic. They discuss that the

binding of Ca2+ to NaPAA can not be due to electrostatic forces (screened Coulomb potentials or

counterion exchange) because this would lead to either an exothermic or energetically neutral

process. Since the binding of Ca2+ ions to NaPAA is spontaneous then the free energy of binding

must be negative. In order for the free energy to be negative this process must be driven by an

increase in entropy. The increase in entropy is believed to be due to the release of water

molecules. The total amount of released water molecules can be calculated by subtracting the

rehydration of Na+ from the dehydration of Ca2+ and COO-. Sinn et al. [15] calculate that a

Page 33: © 2008 Joshua James Taylor

33

minimal number of six water molecules is required to counterbalance the endothermic binding

heat. The process of Ca2+ binding with NaPAA easily releases at least six water molecules.

Finally, two more papers are of particular interest for the role of NaPAA and ions in high

solids loading slurries. Raviv et al. [12, 77] discuss in two of their papers the role of charged

polymers and ions as providing lubrication to a system. The charged ion or carboxylate group

form water sheaths around them that are tightly bound. When the water sheaths approach each

other there is strong repulsion and the repulsive force can dominate the van der Waals attraction

force. Because the water molecules are tightly bound, removal of water molecules from the

sheath require a large amount of energy but the exchange of water molecules between two

sheaths has a much lower energy barrier. This correlates to the NaPAA and ions acting as highly

effective lubricants.

2.3 Attenuated Total Reflectance Fourier Transform Infrared Spectroscopy (ATR-FTIR)

Infrared (IR) spectroscopy detects the vibrational characteristics of chemical functional

groups having a dipole moment. When an infrared light interacts with matter then the chemical

bonds will stretch, contract, and bend. The chemical functional groups adsorb infrared radiation

in specific wavenumber ranges which allows for determination of specific chemical functional

groups within a sample. The band position and width of the band is an indication of the

interaction of the functional group with its surrounding environment. IR spectroscopy is a

technique which has been used for analyzing the interaction between dispersants and particles.

There are several different IR spectroscopy techniques but the three most common include

transmission, diffuse reflectance (DRIFTS), and attenuated total reflectance (ATR). All three

techniques are able to provide data about the interface but ATR has an advantage over

transmission and DRIFTS. The advantage of ATR can be seen in figure 2-9 which shows the IR

radiation penetration of coated particles for transmission, DRIFTS, and ATR techniques.

Page 34: © 2008 Joshua James Taylor

34

Figure 2-9. IR scattering and adsorption in different IR techniques.

For both the transmission and DRIFTS techniques the IR radiation penetrates through the whole

particle providing greater opportunity for the IR radiation to be absorbed. Also, in DRIFTS it is

difficult to control the interaction of the IR radiation with the sample because the interaction is

dependant on the particle size and sampling depth. Figure 2-9 demonstrates that the interaction

of the IR radiation with a particle using the ATR technique only penetrates into a fraction of the

particle allowing for an increase in the surface to bulk signal.

The IR technique chosen for this research will be required to analyze slurries up to 75 wt%

solids loading. Transmission IR spectroscopy requires a sample to be transparent. If too much

of the signal is adsorbed while passing through the sample then reliable data cannot be obtained.

Due to the strong scattering of 75 wt% GCC slurries, the transmission technique cannot be

utilized. The sample preparation for the DRIFTS technique requires the sample to be dried and

then mixed with KBr powder. The DRIFTS technique could be used to analyze dried 75 wt%

GCC slurries but the dried samples do not represent the slurries which are discussed in Section

4.3.1. Finally, the ATR technique is ideal for obtaining spectra of liquids, semisolids, thin films,

and solids. ATR allows for measurement of strongly adsorbing samples of any thickness which

Page 35: © 2008 Joshua James Taylor

35

is a limitation of the transmission technique. Also, ATR does not require any sample

preparation; the sample is placed directly onto the crystal.

Figure 2-10. ATR-FTIR setup. Sample is placed onto a crystal. The IR beam reflects several times through the crystal and the evanescent wave interacts at the interface with the sample.

Figure 2-10 demonstrates the ATR accessory that is mounted into the FTIR. A sample is

placed onto the crystal. The IR radiation passes through the ATR crystal with several reflections

until it exits the crystal and travels to the detector. Each time the IR radiation undergoes total

reflection at the crystal surface next to the sample the exponentially decaying evanescent wave

interacts with the sample. An understanding of this phenomenon of total reflection at the

interface of two materials is required for understanding the ATR-FTIR technique.

There are two requirements for total internal reflection to occur (figure 2-11). First, the

crystal must have a higher refractive index than the sample (n1 > n2). Second, the incident angle

in the crystal must be greater than the critical angle (θ > θc). The critical angle is a function of

the refractive indices as shown in equation 2-14

θc = sin-1 n21 (2-14)

where n21 = n2/ n1 = ratio of refractive indices of sample/crystal.

Page 36: © 2008 Joshua James Taylor

36

Figure 2-11. At the interface of the crystal and sample the IR beam penetrates past the surface of the crystal and decays exponentially.

As the IR radiation undergoes total reflection it penetrates a small distance beyond the

crystal surface. The IR radiation which penetrates beyond the surface decays exponentially and

is called an evanescent wave. Figure 2-12 shows the electric field amplitude of the evanescent

wave as it passes from the crystal into the sample with refractive indices n1 and n2, respectively.

The electric field amplitude decays exponentially with distance from the crystal surface as

described by equation 2-15

E = Eo e-z/Dp (2-15)

where E = electric field amplitude, z = distance from crystal surface, and Dp = depth of

penetration. As the evanescent wave decays it also interacts with the sample which is brought

into contact with the surface. Additional to the decay described with equation 2-15, the

Page 37: © 2008 Joshua James Taylor

37

evanescent wave is attenuated by the sample’s absorbance which is why it is called attenuated

total reflectance.

Figure 2-12. The electric field amplitude at the interface of the crystal and sample. The electric field decays at an exponential rate in the sample. The depth of penetration, Dp, is determined when the electric field is attenuated to 36.8% of its total intensity.

The depth at which the evanescent wave penetrates the samples is known as the depth of

penetration (Dp). The depth of penetration is defined as the depth at which the attenuation of the

Page 38: © 2008 Joshua James Taylor

38

evanescent wave is 36.8% of its total intensity. The depth of penetration is described with

equation 2-16

2/1221

21 )(sin2

1nWn

Dp −=

θπ (2-16)

where W = wavenumber.

There are several observations of equation 2-16 that will be discussed. The depth of

penetration is inversely proportional to the wavenumber. As the wavenumber increases there is a

decrease in the Dp; therefore, the ATR spectra will show peaks that are more intense at low

wavenumbers than at high wavenumbers. Because of this dependence on wavenumber it can be

difficult to compare ATR spectra with library spectra containing transmission or DRIFTS

spectra. In order to avoid any problems, software packages have provided calculations that can

be applied to the ATR spectra to remove the wavenumber dependence.

The depth of penetration is a function of the angle of incidence. As the angle of incidence

decreases the Dp increases. There are two possible ways to change the angle of incidence. First,

the angle of the bevel on the ATR crystal could be changed. Second, it can be changed by

changing the angle of the incoming radiation. Some ATR accessories provide the ability to

change the angle of the mirrors in order to change the angle of incidence.

The depth of penetration is a function of refractive index of the sample. The refractive

index of many organic compounds is similar; therefore, there is little change in the Dp of

different organic samples and it is considered to be independent of refractive index. Quantitative

analysis is possible with the ATR technique if the refractive indices of samples are the same or

similar.

Equation 2-16 also demonstrates that the depth of penetration is a function of the refractive

index of the ATR crystal. As the refractive index of the ATR crystal increases the Dp decreases.

Page 39: © 2008 Joshua James Taylor

39

There are several different ATR crystals that can be utilized for measurements and some of the

most common ones are listed in table 2-1. Several different properties of the crystals must be

considered when choosing the correct crystal for a system. The KRS-5 (trade name for thallium

iodide/thallium bromide) crystal has a large range but is toxic and requires handling with gloves.

The KRS-5 has a low refractive index which gives high penetration depths but it is a soft

material and can be easily scratch and bent. Diamond is not easily scratched and is insoluble in

water but is expensive. Zinc selenide (ZnSe) is the most common ATR crystal material due to its

wide range, insolubility in water, difficulty to scratch, and high depth of penetration. Finally,

Silicon (Si) and germanium (Ge) are both difficult to scratch but are usually used when small

penetration depths are desired. This research utilizes the ZnSe crystal for all ATR-FTIR

measurements.

Table 2-1. Several materials used for the ATR crystal. Z Refractive index (n) Range (cm-1) KRS-5 Diamond ZnSe Si Ge

2.35 2.42 2.42 3.42 4.0

20,000 to 250 4,200 to 200 20,000 to 600 8300 to 660 5500 to 600

After all variables of the ATR-FTIR setup have been determined then the IR spectrum of a

sample is measured. The signal obtained by the FTIR is an interferogram which is then Fourier

transformed into a spectrum. The FTIR spectrum is then analyzed by determining peak

positions, intensities, and full widths at half maximum (FWHM). Peak positions can be difficult

to determine if overlapping occurs but there is a mathematical technique that can be used to

determine the peak positions of overlapping bands. This technique calculates the second

derivative of the IR spectrum. The second derivative contains three features corresponding to

each adsorption peak. There are two upward peaks and one downward peak. The lowest point

Page 40: © 2008 Joshua James Taylor

40

of the downward peak corresponds exactly to the peak position in the original IR spectrum. The

second derivative analysis technique will be utilized in this dissertation for determination of the

peak positions of overlapping bands.

The IR spectrum can also be analyzed for quantitative analysis because the height or area

of a peak is directly proportional to the concentration of the molecule within the IR beam. If the

concentration of one component is desired to be known then several standards must be measured.

Then the concentration versus absorbance of the standards is plotted and a calibration curve is

calculated. The calibration curve is linear due to Beer’s law (discussed in Section 2.4).

However, the relative concentrations of two components in two or more spectra can be compared

without determining a calibration curve by utilizing another technique. Since the intensities of

the peaks in a spectrum are proportional to their concentration then the ratio of two peak

intensities is equal to the ratio of their concentrations. The ratio of the peak intensities of one

spectrum can be compared to the ratio of the same peaks in another spectrum in order to

determine the relative change in concentration of the components. The second method is useful

when the components do not follow Beer’s law due to interaction with other species.

ATR-FTIR is a powerful technique to measure the IR adsorption due to the total internal

reflection phenomenon. The phenomenon creates an evanescent wave which decays

exponentially with distance from the crystal. This property of the evanescent wave allows for

analysis of strongly absorbing samples. The ATR-FTIR technique also allows for in situ

measurements of a sample. Due to the advantages of this technique it has been chosen for

analysis of high solids loading slurries in this research.

2.4 Ultraviolet-Visible Spectroscopy (UV/VIS)

UV/VIS spectroscopy is a common technique used to make quantitative absorption

measurements in the UV/VIS spectral region. The technique requires ultraviolet and visible light

Page 41: © 2008 Joshua James Taylor

41

beams to pass through a reference and a sample. The light which is absorbed by the sample

promotes electronic transitions form a ground state to an excited state. Depending on the

characteristics of the sample, the energy of the light being absorbed is equal to the energy

required to excite the electron. The instrument then measures the intensity of the light passing

through the sample and compares it to the intensity of the light passing through the reference.

The ratio of the intensity of light passing through the sample to the intensity of light passing

through the reference is the percent transmittance, T. The absorbance, A, is then calculated

using equation 2-17.

)log(TA −= (2-17)

Utilizing the Beer-Lambert law (equation 2-18) provides the relationship of the concentration, c,

to absorbance

LcA ××= ε (2-18)

where ε is the wavelength-dependent molar absorptivity coefficient with units M-1 cm-1 and L is

the path length through the sample. A few limitations of the Beer-Lambert law must be noted for

accurate calculations. First, the concentration of the solute species must be below about 0.02 M

to prevent any electrostatic interactions between molecules. Second, readings at high absorbance

values are unreliable and changes in refractive index can occur with high concentrations.

Absorbance readings should be kept below 1.5. Third, solution must be homogenous during

measurement. Fourth, measurements can be erroneous if any suspended particles are in the

solution or if light from other sources enters the equipment.

The Beer-Lambert law is commonly utilized to derive a linear relationship between the

absorbance and concentration. This linear relationship is possible since ε is a constant for a

material and L can be kept constant. In order to make a calibration curve, several absorbance

Page 42: © 2008 Joshua James Taylor

42

measurements of known concentrations must be performed. The data is then plotted on a

concentration versus absorbance graph and a linear fit is applied to the data. The linear fitted

line can then allow for determination of the unknown concentration with an absorbance

measurement.

2.5 Compression Rheology and Impedance Measurements

Two additional techniques were utilized in the analysis of high solids loading slurries but

they did not provide sufficient results. The data from both techniques did not contradict any of

the results discussed in this dissertation but further analysis of the data would require extensive

modeling and many assumptions; therefore, the following two techniques are briefly discussed in

this section but are not further included in this dissertation.

Compression rheology is a technique which relates the volume fraction of a slurry to the

conformation of the dispersant. Kjeldsen et al. [78] utilized this technique to establish a

quantitative link between the molecular structure of superplascticizers and the compression

rheology behavior of MgO suspensions. A centrifugal force is applied to a high solids loading

slurry until the consolidation of the slurry reaches steady state. A stress gradient develops in the

slurry which is balanced by the strength of the particle network. Determining the solids volume

fraction profile of the centrifuged sample in relation to the stress gradient gives information

about the strength of the particle networks. Next, a relationship of the solids volume fraction to

the conformation of the dispersant could be modeled but would require several assumptions.

Impedance measurements of high solids loading slurries were performed in order to

provide more insight into the water structure within a slurry. Since the impedance of water is

different than ice, then a change in the impedance of an aging slurry was expected since the

concentration of structured water increases with aging (Section 5.1). Data collected did not

Page 43: © 2008 Joshua James Taylor

43

contradict any results in this dissertation but would require extensive modeling in order to

separate the impedance contribution from the water, ions, and calcium carbonate particles.

Page 44: © 2008 Joshua James Taylor

44

CHAPTER 3 EXPERIMENTAL METHODS

3.1 Adsorption Isotherms

Slurries were prepared by mixing a dispersant with H2O in a beaker with a stir rod until the

dispersant was uniformly dissolved. Next, GCC was added slowly and mixed with the stir rod

and a spatula. The amount of dispersant was varied between experiments from 0.1 wt% to

4 wt% of the GCC weight. After the GCC, NaPAA, and water have been mechanically stirred,

the samples are centrifuged for 45 minutes at 3,000 rpm in an Eppendorf Centrifuge 5810 using

the swing bucket rotor. After 45 minutes the samples were removed and the supernatant was

decanted into a beaker. Water was added to the supernatant and 1M HCl was added until the pH

was two. The water was boiled out of the system and the remaining dispersant was dissolved

into deionized water. This step of changing the pH to 2 and evaporating off the water is

necessary because it removes any carbonate that might be in the system (carbonate will change

the results of the adsorption isotherms). Next, 1M HCl was added until the pH was two again.

Finally, 0.1M NaOH was added in 0.25ml increments and the pH was recorded. The data was

then analyzed and fit to adsorption isotherm models in Section 4-1.

3.2 Turbidity Measurements

The HACH 2100 turbidimeter was utilized to measure the turbidity of each sample

prepared. Sample preparation includes dissolving the dispersant into deionized water at the same

concentration as within a slurry. Next, several conditions are investigated by changing the

following parameters: monovalent salt concentration; divalent salt concentration; temperature;

time. Samples are poured into a glass vial. The outside wall of the glass vial is wiped with a

cloth and oil to decrease light scattering from any defects on the surface of the glass. The glass

vial is placed into the turbidimeter and the turbidity is recorded.

Page 45: © 2008 Joshua James Taylor

45

3.3 ATR-FTIR

The Thermo Electron Magna 760 FTIR Microscope was used for the ATR-FTIR

measurements. Samples were placed onto a ZnSe ATR crystal (figure 2-10). The IR spectra

were recorded from 650 cm-1 to 4000 cm-1 with 128 scans and a resolution of 2 cm-1. The

MCT/A liquid nitrogen cooled detector was utilized for measurements. A background spectrum

was taken before each measurement and the ATR correction was applied to each sample’s FTIR

spectrum.

GCC slurry preparation included the following two steps: 1. Dissolve dispersant into H2O

or D2O. 2. Using a spatula and stir rod, slowly mix in the GCC powder. For slurries prepared in

D2O the dispersant as received was dried in a furnace at 150ºC overnight before dissolving it into

D2O. The slurries prepared with water structure makers or breakers were first prepared as

describe above with NaPAA. After the slurry is prepared then a structure maker or breaker

chemical is added to the slurry and mixed.

3.4 UV/VIS Spectroscopy

Slurries with probe molecules were prepared as described above for a slurry except one

extra step. The pH was changed to 9.9 by adding NaOH before mixing in the GCC. Controlling

the pH was to ensure that the amount of deprotonated carboxylate groups were the same as the

number of deprotonated carboxylate groups in a slurry containing NaPAA. The samples were

not disturbed for 24 hrs to allow for the adsorption of the probe molecules. After the allotted

time the samples were centrifuged at 3,000 rpm for 45 minutes. The supernatant was then

decanted into a beaker and diluted with a measured amount of water. The UV/VIS adsorption

intensities were measured at wavelengths of 210 nm and 224 nm for gallic acid and benzoic acid,

respectively, with the Beckman DU 640 spectrophotometer. Next, the measured intensities were

compared to a calibration curve and the amount of probe molecules in the supernatant was

Page 46: © 2008 Joshua James Taylor

46

determined. The amount of probe molecules adsorbed was calculated from subtracting the

measured amount in the supernatant from the amount in the slurry.

Figure 3-1. Calibration curve for gallic acid. Absorption measurements preformed at a wavelength of 210 nm with a sample thickness of 1 cm.

The calibration curves for both gallic acid and benzoic acid are given in figures 3-1 and

3-2, respectively. Each curve was determined by preparing several samples with known

concentrations of dissolved probe molecules in water and measuring their UV/VIS absorption

intensities. Earlier experiments had shown that the pH of the system did not affect the intensity

measurements.

Page 47: © 2008 Joshua James Taylor

47

Figure 3-2. Calibration curve for benzoic acid. Absorption measurements preformed at a wavelength of 224 nm with a sample thickness of 1 cm.

3.5 Rheology

The modular compact rheometer Physica MCR 300 was the instrument used to measure

the rheological properties of the slurries. Measurements were performed with the cup/cylinder

instrument setup. Each sample was pre-sheared to the maximum shear stress before recording

the data used for interpretation. The pre-shearing of the samples was necessary because the

viscosity depends on the history of the slurry just before the measurement data was obtained.

Page 48: © 2008 Joshua James Taylor

48

CHAPTER 4 SODIUM POLYACRYLATE ADSORPTION ONTO GCC IN HIGH SOLIDS LOADING

SLURRIES

4.1 Adsorption Isotherms

The depletion method is often the chosen technique in literature to measure the

concentration of a solute in order to calculate its adsorption isotherms. Adsorption isotherms

have been published for several dispersants and molecules onto calcium carbonate but all the

experiments have been completed in dilute systems [1, 5, 9, 16, 24, 31, 72, 79-84]. The

following adsorption isotherm experiments were performed with 75 wt% GCC solids loading

slurries.

Potentiometric titration was used to determine the amount of dispersant in the supernatant

after centrifugation. This technique is possible because the NaPAA dispersant contains

carboxylate groups which can be protonated or deprotonated depending on the pH (equation

4-1).

R-COOH + NaOH R-COO- + Na+ + H2O (4-1)

First, the calibration curves for NaPAA had to be determined. This included dissolving a known

amount of dispersant into deionized water, decreasing pH to 2 with 1 M HCl, and then titrating

the sample with 0.1 M NaOH. The amount of NaOH added to the system was then plotted

against the pH of the system (figure 4-1 black curve). Next, the first derivative of the titration

line was taken and then the volume of NaOH between the two peaks was calculated (figure 4-1

red curve). The volume of NaOH between the peaks is proportional to the amount of dispersant

in the water. This process was repeated for several different known amounts of dispersant. The

data was then plotted with volume of NaOH between peaks versus dispersant amount and a

linear fit was applied to the data (figure 4-2). With the calibration curves determined, the

Page 49: © 2008 Joshua James Taylor

49

0 2 4 6 8 10 12 14 160

2

4

6

8

10

12pH

Volume (mL)

Titration Derivative

Figure 4-1. Titration curve of NaPAA. Black curve is the titration of a NaPAA in deionized water. Red curve is the 1st derivative of the titration curve.

amount of dispersant in an unknown sample could be determined by measuring the volume of

NaOH that is needed to titrate the sample.

Adsorption isotherm data was measured from slurries of 75 wt% GCC containing 0.1 wt%,

0.5 wt%, 1 wt%, 2 wt%, and 4 wt% NaPAA (weight percent of GCC weight). The offered

amount of NaPAA versus adsorbed amount of NaPAA was plotted in figure 4-3 with the 100%

adsorption line.

The adsorption isotherm data was then compared to three models. First, the data was

compared to the linear form of the Langmuir model which does not fit well with a R2 value of

Page 50: © 2008 Joshua James Taylor

50

Figure 4-2. Calibration curve for NaPAA with the titration technique.

0.82 (figure 4-4). The Langmuir model assumes monolayer adsorption with a surface which

consists of uniform adsorption sites. Second, the data was compared to the linear form of the

Temkin model which also had a poor fit with a R2 of 0.85 (figure 4-5). Third, the data was

compared to the linear form of the Freundlich model which fits well with a R2 of 0.96. Next, the

Freundlich constant, K, and Freundlich exponent, b, were calculated from the intercept and slope

of the linear Freundlich model (equation 4-2), respectively.

log(Q) = log(K) + b log(C) (4-2)

where Q is the adsorbed NaPAA per GCC (mg/g) and C is the unadsorbed NaPAA (mg/L). The

non linear Freundlich model (equation 4-3) was fitted with the Freundlich constant and exponent

(figure 4-6).

Q = KCb (4-3)

Page 51: © 2008 Joshua James Taylor

51

Figure 4-3. The adsorbed amount of NaPAA compared to the 100% adsorption line.

The Freundlich model assumes a non-linear relationship between the adsorbed amount and the

concentration of the solute in the liquid. This model represents adsorption with strong solute to

solute interaction. It also indicates that the solute is adsorbing onto a heterogeneous surface. In

agreement with these results, Balaz et al. [85] also demonstrated that adsorption on mineral

surfaces are best fit with the Freundlich model.

The 75 wt% GCC slurry system does not follow the Langmuir isotherm, which is a

common model to describe most adsorption processes, but displays the Freundlich isotherm.

There are several reasons why the system doesn’t follow the Langmuir isotherm. First, the

Langmuir isotherm describes a system that is dilute in which the adsorbed molecules do not

Page 52: © 2008 Joshua James Taylor

52

Figure 4-4. Langmuir adsorption isotherm of NaPAA by GCC in a 75 wt% solids loading slurry. This model is a poor fit with a R2 value of 0.8187.

Figure 4-5. Temkin adsorption isotherm of NaPAA by GCC in a 75 wt% solids loading slurry. This model is a poor fit with a R2 value of 0.8529.

Page 53: © 2008 Joshua James Taylor

53

Figure 4-6. Freundlich adsorption isotherm of NaPAA by GCC in a 75 wt% solids loading slurry. Freundlich model fits well with a R2 value of 0.96.

interact. Second, it assumes a homogeneous surface. Third, it has a defined adsorption

maximum and assumes linear adsorption at concentrations far below the maximum. Many of

these assumptions do not fit the properties of a 75 wt% GCC slurry. The GCC slurry is a

concentrated system with calcium, carbonate, and sodium ions due to GCC’s solubility in water

and the Na+ ions from the NaPAA. Furthermore, as will be demonstrated later, the chemical

interactions within a high solids loading slurry are different than a dilute system.

4.2 Turbidity of NaPAA

Adsorption isotherms indicated that the NaPAA is on the surface of the particles but they

do not distinguish between adsorption and precipitation of the dispersant. Therefore, turbidity

measurements of the dispersant at several conditions were necessary to determine if the NaPAA

Page 54: © 2008 Joshua James Taylor

54

Figure 4-7. Turbidity of NaPAA in water with increasing monovalent salt concentration. High concentrations of sodium ions within slurry do not cause the NaPAA to precipitate.

precipitates within a 75 wt% GCC solids loading slurry. First, the impact of the monovalent salt

NaCl on the turbidity was determined because the slurry system contains Na+ from the

dispersant. Samples were prepared with 2.04 g of NaPAA dissolved into 30 ml H2O and the

NaCl concentration was varied. Na+ concentrations started at 0.4 M (slurry conditions) and were

increased above 1 M. There was no increase in turbidity at all salt concentration which indicates

no precipitation (figure 4-7). Second, the impact of the divalent salt CaCl on the turbidity was

determined because the slurry system contains Ca2+ from the CaCO3. The turbidities of the

samples containing concentrations of CaCl2 from 10-4 M to 1.9 x 10-1 M were determined.

Within the processing conditions of the slurry, 10-3 M Ca2+ [20], there was not an increase in

turbidity (no indication of precipitation). When the concentration of CaCl2 was above slurry

Page 55: © 2008 Joshua James Taylor

55

Figure 4-8. Turbidity of NaPAA in water with increasing divalent salt concentration. Under processing condition of GCC slurries there is no indication of NaPAA precipitation.

conditions at 1.9 x10-1 M there was an increase in turbidity and precipitation of the dispersant

(figure 4-8).

During slurry preparation in industry the slurry experiences thermal cycles which could

possible cause the NaPAA to precipitate; therefore, it was necessary to investigate the impact of

temperature on the turbidity of NaPAA. Samples were prepared by dissolving the NaPAA into

deionized water at room temperature. The temperature of the solution was increased up to 85ºC

while taking turbidity measurements at 25ºC, 40ºC, 55ºC, 70ºC, and 85ºC (figure 4-9). There

was no increase in turbidity with increase in temperature which indicates that within processing

temperature conditions there is no precipitation of the dispersant within the slurry.

Page 56: © 2008 Joshua James Taylor

56

Figure 4-9. Turbidity of NaPAA in water with increasing temperature. NaPAA does not precipitate with increasing temperature.

As a 75 wt% GCC slurry ages, several of its properties change with time. Turbidity

measurements of the NaPAA with respect to time were performed in order to determine if there

was precipitation of NaPAA with aging. Samples were prepared with NaPAA dissolved in water

with two concentrations of CaCl2: 10-3 M CaCl2 which represents the Ca2+ ion concentration in

slurry conditions and 10-1 M CaCl2 which represents an excess concentration of Ca2+ ions. Over

a three day period there was no impact on the turbidity of the 10-3 M CaCl2 (slurry conditions)

which is an indication of no precipitation. On the third day of the 10-1 M CaCl2 solution there

was precipitation on the bottom of the container but no increase in turbidity of the liquid (figure

4-10). On the fifth day of the 10-1 M CaCl2 there was precipitation on the bottom of the

Page 57: © 2008 Joshua James Taylor

57

Figure 4-10. Turbidity of NaPAA in water with 10-3 M CaCl2 and 10-1 M CaCl2 over a three day period. At slurry conditions there was no precipitation of NaPAA.

container and the turbidity of the liquid increased. All of the turbidity measurements indicate

that within slurry conditions there is no precipitation of the dispersant.

4.3 ATR-FTIR

ATR-FTIR is a technique that allows for analysis of high solids loading slurries due to the

phenomenon of the evanescent wave described previously in Section 2.3. This technique was

utilized for analysis of the interactions of the carboxylate group, hydrocarbon groups, carbonates,

and the solvent at a variety of conditions.

Page 58: © 2008 Joshua James Taylor

58

4.3.1 Dry and Wet Samples

As mentioned before there are several techniques to analyze the GCC slurries but many of

them require either a dilute or dry sample. IR spectra of GCC slurries after drying do not

represent the chemical interactions that have taken place in the slurry. Figures 4-11, 4-12 and

4-13 compare the IR spectrum of a 75 wt% GCC slurry that had been dried and analyzed with

DRIFTS to the spectrum of a 75 wt% GCC slurry analyzed with ATR. Figure 4-11 demonstrates

that in the wet state the carbonate bending band at 1449 cm-1 has a lower frequency shoulder that

is not present in the dry sample. This indicates that there are bicarbonate species and chemical

bonds in the slurry (discussed in Section 4.3.6) which are removed upon drying. In the wet state

1600 1500 1400 1300 12000.0

0.2

0.4

0.6

0.8

1.0 75wt% Slurry ATR

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

75wt% Slurry Dried DRIFTS

Figure 4-11. IR spectra of the carbonate bending bands located at 1449 cm-1 of a wet and a dry 75 wt% GCC slurry. The wet slurry forms a shoulder at lower wavenumbers indicating the presents of bicarbonate species.

Page 59: © 2008 Joshua James Taylor

59

950 900 850 8000.0

0.2

0.4

Log

(1/R

) (a.

u.)

Wavenumber (cm-1)

75wt% Slurry ATR 75wt% Slurry Dried DRIFTS

Figure 4-12. IR spectra of the carbonate stretching bands located at 875 cm-1 of a wet and a dry 75 wt% GCC slurry. The wet slurry has a larger FWHM.

1600 1580 1560 1540-0.0004

0.0000

0.0004

d2 A/dυ2 (a

.u.)

Wavenumber (cm-1)

Dried Slurry, DRIFTS 75 wt% Slurry, ATR

Figure 4-13. Second derivative of the IR spectra of the carboxyl region for a wet and a dry sample. Band at 1581 cm-1 shifts to 1586 cm-1 indicated change in coordination.

Page 60: © 2008 Joshua James Taylor

60

the carbonate stretching band at 875 cm-1 has a larger FWHM than the dry state (figure 4-12)

indicating that some of the chemical interactions with the carbonate species in the wet slurry are

not present in the dry state. There is also a change in the antisymmetric stretching band of the

carboxylate group in NaPAA. Figure 4-13 is the second derivative of the IR spectra which

shows a shift in the band at 1581 cm-1 in the wet state to 1586 cm-1 in the dry state. The shift to a

higher wavenumber indicates that the adsorbed NaPAA in the dry state has unidentate

coordination (coordination of the carboxylate group is described in more detail in Section 4.3.5).

Analysis of the IR spectra demonstrates that the chemical interaction of a slurry are not

represented in a dry slurry; therefore, the best IR technique for this research is the ATR.

4000 3500 3000 2500 2000 1500 10000.0

0.2

0.4

0.6

0.8

1.0

bendingOH

stretchingCOO-

Log

(1/R

) (a.

u.)

Wavenumber (cm-1)

NaPAA in H2OOH stretching

CHstretching

Figure 4-14. IR spectrum of NaPAA in H2O. The OH stretching band overlaps the CH stretching bands of NaPAA. The OH bending band overlaps part of the COO- band.

Page 61: © 2008 Joshua James Taylor

61

4000 3500 3000 2500 2000 1500 10000.0

0.2

0.4

0.6

0.8

1.0

bendingOD

COO-

OD

stretching

stretching

CHstretchingL

og (1

/R) (

a.u.

)

Wavenumber (cm-1)

NaPAA in D2O

Figure 4-15. IR spectrum of NaPAA in D2O. Switching from H2O to D2O shifts the stretching and bending bands so that there is no overlap with the CH or COO- stretching bands of NaPAA.

4.3.2 Solvent Exchange

The stretching modes of OH in water extend over a range of 2900 cm-1 to 3700 cm-1 due to

hydrogen bonding in many different energy states. The bending modes of OH in water are

located at 2130 cm-1 and 1638 cm-1 as seen in figure 4-14. Also shown in figure 4-14 is the IR

spectrum of NaPAA in H2O. As seen in the spectrum the stretching modes of water overlap the

CH stretching of the dispersant and the bending mode of water interferes with the shoulder of the

carboxylate group of the dispersant. In order to analyze these regions the water bands must be

removed without changing the slurry properties. This was accomplished by substituting D2O

Page 62: © 2008 Joshua James Taylor

62

with H2O. The OD stretching modes of heavy water are at lower wavenumbers, 2200 cm-1 to

2700 cm-1, than the OH stretching modes of water (shown in figure 4-15). Also shown in figure

4-15 is the IR spectrum of NaPAA in D2O. Within the spectrum the CH bending is visible and

there is no interference with the carboxylate band. The 75 wt% GCC slurry system was then

analyzed with the ATR-FTIR and its IR spectrum is shown in figure 4-16. The regions of focus

have been boxed and labeled which include the following: CH stretching, OD stretching,

carboxylate group, and the carbonate band.

4000 3500 3000 2500 2000 1500 10000.0

0.2

0.4

0.6

0.8

1.0

[CO3]2-

COO-

OD

CH

Log

(1/R

) (a.

u.)

Wavenumber (cm-1)

75 wt% Slurry with D2O

Figure 4-16. IR spectrum of a 75 wt% GCC slurry with NaPAA. Regions of interests are boxed and discussed.

Page 63: © 2008 Joshua James Taylor

63

4.3.3 Water Structure in High Solids Loading Slurries

A novel discovery was made upon analysis of the ATR-FTIR measurements of high solids

loading GCC slurries. This discovery demonstrated that the water structure within the high

solids loading slurries is dramatically different than bulk water. This novel discovery is

important because it demonstrates that research performed in dilute solutions does not represent

all the chemical interactions taking place in a high solids loading sample even though the

composition is the same. Figures 4-17 and 4-18 demonstrate that the stretching modes of H2O

and D2O dramatically change. For H2O the IR band decreases its FWHM and the shoulder at

3183 cm-1 shifts to a higher frequency at 3239 cm-1. A similar response with D2O and D2O in

3600 3400 3200 3000 28000.0

0.2

0.4

0.6

0.8

1.0

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

75wt% Slurry H2O

Figure 4-17. IR spectra of the OH stretching bands of H2O and H2O in a 75 wt% slurry. When water is in a high solids loading slurry there is a change in the structure of the water indicated with the FWHM decreasing and the 3183 cm-1 shoulder peak shift to 3239 cm-1.

Page 64: © 2008 Joshua James Taylor

64

a 75wt% GCC slurry is shown in figure 4-18. The IR band of D2O in a slurry decreases its

FWHM along with the 2379 cm-1 shoulder shifting to 2410 cm-1 and the 2492 cm-1 shoulder

increasing intensity and shifting to 2502 cm-1. Nickolov et al. [86] refer to the lower frequency

shoulder as ice-like with stronger hydrogen bonding and the higher frequency shoulder as fluid-

like with weaker and distorted hydrogen bonds. The ice-like water structure has hydrogen bond

coordinations of four while the fluid-like water has hydrogen bond coordinations less than four.

The ice-like structure of both H2O and D2O in a slurry decreases along with an increase in the

fluid-like structure. This will be analyzed and discussed in more detail in chapter 5.

2800 2600 2400 22000.0

0.2

0.4

0.6

0.8

1.0

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

D2O 75wt% Slurry

Figure 4-18. IR spectra of the OD stretching bands of D2O and D2O in a 75 wt% slurry. When D2O is in a high solids loading slurry there is a change in the structure of the water indicated with the FWHM decreasing and the 2502 cm-1 shoulder increasing intensity.

Page 65: © 2008 Joshua James Taylor

65

4.3.4 NaPAA CH Band Infrared Spectra

The second band to focus on is the CH band of NaPAA located at 2947 cm-1 when the

dispersant is dissolved into D2O (figure 4-19). This band is a result of the CH2 and CH

asymmetric stretching bonds within the NaPAA molecule. Focusing on the same region within a

75 wt% GCC slurry there is a decrease in the intensity of the 2947 cm-1 band and two new bands

are formed, one at 2980 cm-1 and the other at 2875 cm-1 (figure 4-20). The appearance of two

new bands was an unexpected finding. The new bands located at 2980 cm-1 and 2875 cm-1 are

not present in the IR spectrum of D2O or the IR spectrum of GCC in D2O; therefore, it is

concluded that the new bands arise from a shifting of the CH bands and/or an increase in the

symmetric stretching of the CH bands. When the NaPAA is dissolved in the solution the CH

stretching bonds are not restricted but when a carboxylate group adsorbs onto the surface of the

GCC the CH bonds near that carboxylate group are in close proximity to the surface. The close

proximity to the surface could restrict the stretching bonds and cause IR bands to shift. Schmidt

et al. [87, 88] discuss the shift in the IR for CH bands of polymer/water systems due to a change

in temperature with the largest shift being about 12 cm-1 for a 20°C temperature difference.

They also derive the theoretical frequency shifts for various numbers of water molecules in the

neighborhood of methyl groups demonstrating a wavenumber shift of ~60 cm-1 for an increase

from 8 to 12 water molecules. They conclude that the hydrophobic interactions of the methyl

groups cause a shift of the CH stretching bands. Al-Hosney et al. [89] demonstrate a shift in the

CH band of 20 cm-1 for wet versus dry conditions of HCOOH adsorbed onto calcium carbonate.

The change in the wavenumber from the 2947 cm-1 band to the 2875 cm-1 band is 72 cm-1. This

difference is equivalent to the band difference from the asymmetric to symmetric stretching of

CH2. So the appearance of the two new bands is explained to be a result of NaPAA adsorption

onto the surface of GCC.

Page 66: © 2008 Joshua James Taylor

66

Figure 4-19. IR spectrum of the CH band of NaPAA in D2O.

Figure 4-20. IR spectrum of the CH bands of NaPAA in a 75 wt% GCC slurry. Two bands are

formed at 2980 cm-1 and 2875 cm-1 indicating adsorption of the NaPAA onto the surface of GCC.

Page 67: © 2008 Joshua James Taylor

67

Further investigation of this phenomenon consisted of increasing the concentration of the

NaPAA in a 75 wt% GCC slurry. Figure 4-21 demonstrates that with one weight percent

NaPAA there is formation of two bands at 2980 cm-1 and 2875 cm-1 while the original band at

2947 cm-1 decreases, indicating adsorption of NaPAA. With an increase of NaPAA exceeding

9 wt% the CH band at 2947 cm-1 increases while the other bands remain constant in intensity.

Since the bands at 2980 cm-1 and 2875 cm-1 indicate adsorption and they do not increase in

intensity above 9 wt% NaPAA then the adsorption limit is between 1 wt% and 9 wt%. Utilizing

equations 2-18 and 2-17, N = 63 and l = 0.27 nm, the monolayer adsorption amount depending

on the NaPAA conformation would be between 2.14 mg/m2 and 4.28 mg/m2, respectively.

Figure 4-21. IR spectra of the CH bands of NaPAA in GCC slurries with increasing concentration of NaPAA. Adsorption limit is between 1 wt% NaPAA and 9 wt% NaPAA.

Page 68: © 2008 Joshua James Taylor

68

4.3.5 NaPAA Carboxylate Group Infrared Spectra

As mentioned in Section 2.2 the carboxylate group is known to adsorb on several different

surfaces and due to an increase in entropy it adsorbs onto calcium ions. The carboxylate group

of NaPAA has a resonance form when dissolved in water. This is confirmed in the IR spectra

with the absence of the C=O band which would be located around 1700 cm-1 and the presence of

a COO- band at 1570 cm-1 as seen in figure 4-22. When a 75 wt% GCC slurry is prepared there

is a change in the carboxyl band. Figure 4-23 shows that the 1570 cm-1 stretching band is split

1750 1700 1650 1600 1550 15000.00

0.05

0.10

0.15

Log

(1/R

) (a.

u.)

Wavenumber (cm-1)

NaPAA in D2O

Figure 4-22. IR spectrum of the carboxyl stretching region for NaPAA in D2O. The IR spectrum shows that NaPAA is in the ionic form with a peak at 1570 cm-1 for the COO- and no peak at 1700 cm-1 for the C=O.

Page 69: © 2008 Joshua James Taylor

69

1600 1580 1560 1540 1520 1500

-0.0010

-0.0005

0.0000

0.0005

0.0010

ionic

bridging

bidentate

unidentate

NaPAA in D2Od2 A

/dυ2 (a

.u.)

Wavenumber (cm-1)

75 wt% GCC Slurry

Figure 4-23. Second derivative of the IR spectra of the carboxylate region of NaPAA in D2O and in a 75 wt% GCC slurry. NaPAA adsorbs onto GCC in unidentate, bridging, and bidentate modes.

into a band at 1581 cm-1 and 1567 cm-1 along with a band forming at 1524 cm-1 with a shoulder.

From the research of Lu et al. [36] and Young et al. [76] they have shown that each band

represents a different mode of interaction. The four possible modes of interaction are ionic,

unidentate, bidentate, and bridging (figure 2-8). Mielczarski et al. [61, 62] assigned the bands at

1575 cm-1 and 1540 cm-1 to unidentate and bidentate adsorption, respectively. Lu et al. [36]

demonstrate that the bands arise from three-dimensional precipitated calcium dicarboxylate salts

which can either be from the bulk solution or physisorbed at a surface. Therefore, in figure 4-23

the 1581 cm-1 band is representative of unidentate coordination, the 1567 cm-1 band represents

Page 70: © 2008 Joshua James Taylor

70

bridging coordination, and the bands between 1510 cm-1 and 1545 cm-1 represent bidentate

coordination with calcium.

Additional slurries with increasing concentrations of NaPAA were prepared and analyzed

with ATR-FTIR. IR spectra of 75 wt% GCC slurries containing 1 wt% and 10 wt% NaPAA are

compared to NaPAA in D2O, see figure 4-24. Initially, the NaPAA in D2O shows one band at

1570 cm-1 which represents the carboxylate group in the ionic coordination. Next, with the

1 wt% NaPAA slurry there is a splitting of the 1570 cm-1 stretching band as described above.

1600 1580 1560 1540 1520 1500

-0.0010

-0.0005

0.0000

0.0005

0.0010

bidentate

bridging andionicunidentate

NaPAA in D2O

d2 A/dυ2 (a

.u.)

Wavenumber (cm-1)

Slurry with 1 wt% NaPAA Slurry with 10 wt% NaPAA

Figure 4-24. Second derivative of the IR spectra of 75 wt% GCC slurries with increasing concentration of dispersant. With an excess amount of dispersant there is no unidentate adsorption mode.

Page 71: © 2008 Joshua James Taylor

71

The splitting of the band confirms that the adsorption of NaPAA in a 75 wt% GCC slurry is

through unidentate, bridging, and bidentate coordination states with calcium ions on calcium

carbonate. The slurry with 10 wt% NaPAA does not show any unidentate coordination,

indicated with the removal of the 1581 cm-1 band, and there is a shift in the band peak at

1524 cm-1 to 1515 cm-1 representing bidentate coordination. Also, the band at 1566 cm-1 which

represents bridging coordination may also include some NaPAA that has not adsorbed due to

overlap of the 1570 cm-1 band from NaPAA in solution. The IR spectra demonstrate that as the

amount of dispersant increases in a 75 wt% solids loading slurry the coordination of the carboxyl

species changes with the removal of the unidentate coordination.

Similar to the results of Section 4.3.3 which showed a change in the water structure with

solids loading, there is a change in the carboxylate coordination at high solids loading. Slurry

samples were prepared with 10, 30, 50, and 70 wt% GCC. The spectra of the IR carboxyl region

are shown in figure 4-25. First, the band representative of the unidentate coordination in the

10 wt% GCC slurry at 1586 cm-1 shifts to lower wavenumbers as the solids loading is increased,

eventually to 1578 cm-1 in the 70 wt% GCC slurry. Similarly, the band representative of the

bridging coordination in the 10 wt% GCC slurry at 1567 cm-1 shifts to lower wavenumbers until

it reaches 1560 cm-1 in a 70 wt% GCC slurry. Also, as the solids loading increases the band

between 1510 cm-1 and 1545 cm-1 becomes a doublet with band peaks at 1539 cm-1 and

1521 cm-1, representing the bidentate coordination. As the solids loading increases, the

dispersant shifts towards the calcium bridging mode of adsorption indicated with the decrease in

the unidentate band position and an increase in the bidentate band position. The slight change in

band positions could possibly be due to increased interactions between dispersant molecules due

to their higher concentration in the solvent. This conclusion is important because it supports

Page 72: © 2008 Joshua James Taylor

72

1600 1580 1560 1540 1520-0.0008

-0.0004

0.0000

0.0004

0.0008

bridging

bidentate

unidentate

d2 A/dυ2 (a

.u.)

Wavenumber (cm-1)

70 wt% Slurry

10 wt% Slurry 30 wt% Slurry 50 wt% Slurry

Figure 4-25. Second derivative of the IR spectra of the carboxylate region with increasing solids loading. As the solids loading of the slurries increases there is a shift of the bands toward a bridging mode.

previous data which demonstrate that the interactions within a low solids loading slurry is

different than in a high solids loading slurry.

During the first couple of days after a slurry has been prepared there is a change in several

of its properties as mentioned in the introduction. ATR-FTIR was utilized to determine if there

is any change in the interaction of the NaPAA with GCC during the aging process. A 75 wt%

GCC slurry was prepared and analyzed with the ATR-FTIR as it aged while it was less than an

hour old, 24 hours old, and 48 hours old. The IR spectra can be seen in figure 4-26. Initially, the

band representing the unidentate coordination is located at 1585 cm-1. As the system ages the

Page 73: © 2008 Joshua James Taylor

73

band shifts to 1580 cm-1. This indicates that while the slurry ages there is a decrease in the

concentration of unidentate coordination of carboxylate groups and an increase in the bridging

and/or ionic coordination of the carboxylate groups. These results would support the idea that as

the system ages there is an increase in the amount of dissolved calcium ions which are then

available for bridging coordination, which require two calcium ions per carboxylate group,

instead of unidentate coordination, which only require one calcium ion per carboxylate group

(figure 2-8).

1600 1580 1560 1540

0.0000

0.0005

increasing age

bridgingunidentate

75 wt% Slurry Aged 48 hrs

d2 A/dυ2 (a

.u.)

Wavenumber (cm-1)

75 wt% Slurry Fresh 75 wt% Slurry Aged 24 hrs

Figure 4-26. Second derivative of the IR spectra of the carboxylate region with aging. The unidentate band shifts from 1585 cm-1 to 1580 cm-1 with increasing age indicating a shift towards the bridging mode.

Page 74: © 2008 Joshua James Taylor

74

4.3.6 Carbonate Infrared Spectra

The next bands to focus on are the carbonate bands from the GCC. This includes a

stretching band at 875 cm-1 and a bending band at 1404 cm-1 (figure 4-27) which is in agreement

with literature [90]. When GCC is introduced into a slurry the FWHM for the stretching band

3000 2500 2000 1500 10000.0

0.2

0.4

0.6

0.8

1.0

combination bands

bendingcarbonate ion

carbonate ionstretching

Log

(1/R

) (a.

u.)

Wavenumber (cm-1)

Figure 4-27. ATR-FTIR spectrum of GCC.

increases. This is an indication that the carbonates are interacting with other species in the

system. Upon analysis of the carbonate bending band there is a dramatic change in the band (see

figure 4-28). The carbonate bending band at 1404 cm-1 in figure 4-27 shifts to 1449 cm-1 for a

75 wt% GCC slurry and the band also forms a low frequency shoulder. These spectra confirm

that the carbonate is interacting with other species in the system. The shoulder that forms on the

Page 75: © 2008 Joshua James Taylor

75

1600 1500 1400 1300 1200 11000.0

0.2

0.4

0.6

75wt% SlurryL

og (1

/R)

(a.u

.)

Wavenumber (cm-1)

D2O and GCC

Figure 4-28. IR spectra of carbonate band of GCC in D2O and in a 75 wt% GCC slurry. Formation of a shoulder indicates formation of bicarbonates.

carbonate bending band indicates the formation of bicarbonates in the system [86, 91, 92]. Since

CaCO3 is slightly soluble in water then the spectra confirm the dissolution of CaCO3, adding

more ions to the system. These ions play an important role on the adsorption of NaPAA and the

water structure (discussed in Section 5.3). The FWHM and shape of the carbonate bending band

is also dependent on the solids loading. Up to 40 wt% the carbonate bending band is similar to a

mixture of D2O and GCC without NaPAA. Once the solids loading is raised above 40 wt% the

band forms a shoulder which shifts to lower wavenumbers as the solids loading is increased

(figure 4-29). This is another confirmation that high solids loading slurries have different

chemical interactions than dilute systems.

Page 76: © 2008 Joshua James Taylor

76

1600 1500 1400 1300 1200 11000.0

0.2

0.4

0.6

75wt% Slurry

40wt% Slurry 50wt% Slurry 60wt% Slurry 70wt% Slurry

increasing solids

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

D2O and GCC

Figure 4-29. IR spectra of the carbonate band of GCC slurries showing a formation of a shoulder

with increase solids loading.

There is also a noticeable change in slurry properties over a period of a few days;

therefore, a slurry of 75 wt% GCC in D2O was analyzed with ATR-FTIR over a three day

period. During the three days the carbonate band did not shift but the shoulder became less

pronounced, shifting to higher frequencies as time increased (figure 4-30). The shoulder shifting

over time is an indication of a decrease in the bicarbonate species and affects the water structure

within the slurry (more details in Section 6.2).

If ordering occurs in the system and we assume that the shoulder of the carbonate band is

an indication of this ordering then the shoulder should shift in response to thermal differences. A

slurry of 75 wt% GCC in H2O was prepared and separated into three samples. The first sample

Page 77: © 2008 Joshua James Taylor

77

1600 1500 1400 1300 1200 11000.0

0.2

0.4

0.6

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

GCC and D2O

increasing age

75wt% Slurry Aged <1hr 75wt% Slurry Aged 25hrs 75wt% Slurry Aged 51hrs

Figure 4-30. IR spectra of the carbonate band of GCC in a 75 wt% slurry with aging.

was frozen, placed on the ATR crystal, and while it was melting an IR spectrum was measured.

The second sample was poured onto the ATR crystal at room temperature and the IR spectrum

was measured. The third sample was boiled, placed onto the ATR crystal, and the IR spectrum

was measured. These three spectra show that as the temperature decreases (less relative thermal

energy) there is a shift of the shoulder to a higher frequency (figure 4-31). This supports the idea

that as the age of a slurry increases (shoulder shifts to higher frequencies) there is an increase in

the order of the system.

Page 78: © 2008 Joshua James Taylor

78

1600 1500 1400 1300 12000.0

0.1

0.2

0.3

Room Temperature SlurryL

og (1

/R) (

a.u.

)

Wavenumber (cm-1)

Almost Boiling Temperature

~0°C Slurry

1311

1319

1335

Figure 4-31. IR spectra of the carbonate band with a slurry at different temperatures. As the temperature decreases the shoulder shifts from 1311 cm-1 to 1335 cm-1.

4.3.9 Temperature Dependence

In industry a high solids loading GCC slurry will experience several different thermal

environments while being processed. An experiment was designed to determine if thermal

variation of a slurry has an effect on the IR spectrum of the final slurry. Two 75 wt% GCC

slurries in D2O were prepared at room temperature. The temperature of one sample was raised to

85ºC for 15 minutes and then cooled to room temperature before analyzing the sample with

ATR-FTIR. The spectra of both samples were the same except the shoulder of the carbonate

peak was shifted (figure 4-32).

Page 79: © 2008 Joshua James Taylor

79

4000 3500 3000 2500 2000 1500 10000.0

0.2

0.4

0.6

0.8

1.0 Previously heated Slurry

Log

(1/R

) (a.

u.)

Wavenumber (cm-1)

75 wt% Slurry

Figure 4-32. IR spectra of a 75 wt% GCC slurry before and after a heating cycle. The only difference of the two spectra is the carbonate band.

Also, the structure of the water within a slurry is dependant on the temperature of the

slurry. A 75 wt% GCC slurry was prepared and separated into two samples. The first sample

was analyzed at room temperature with the ATR-FTIR. The second sample was frozen, placed

onto the ZnSe ATR crystal and analyzed as it was melting. As expected, the shoulder at

2379 cm-1 increased in intensity and the shoulder at 2502 cm-1 decreased in intensity for the

melting sample which indicates more ice like structure (figure 4-33).

4.3.10 Addition of Energy to Aged Slurries

One technique used by industry to retard slurry aging while it is in storage includes

constant stirring of the slurries. So the effect on the water structure of adding energy to an aged

Page 80: © 2008 Joshua James Taylor

80

2700 2600 2500 2400 2300 22000.0

0.2

0.4

0.6

0.8

1.0 75 wt% Slurry Room Temperature

75 w% Melting SlurryL

og (1

/R) (

a.u.

)

Wavenumber (cm-1)

Figure 4-33. IR spectra of the OD band in a 75 wt% GCC slurry at different temperatures. Lower temperatures increase the concentration of structured water.

slurry was investigated. First, a 75 wt% slurry was aged for 5 days and then the IR spectrum was

measured. Second, a small amount of energy was added to the slurry with mechanical stirring

which represents the process in industry. As seen in figure 4-34, there is a decrease in the ice-

like structure, reversing the aging process. Third, the aging process is further reversed as more

energy is added to the system through sonication. A similar result is detected in the carbonate

band (figure 4-35). As energy is added to the system the shoulder on the carbonate band shifts to

lower frequencies which is the opposite of the aging process in figure 4-30. Even though the

water and carbonate bands are not fully restored to their fresh state, these results indicate that the

aging process includes an entropic component.

Page 81: © 2008 Joshua James Taylor

81

2800 2600 2400 22000.0

0.2

0.4

0.6

0.8

1.0L

og (1

/R)

(a.u

.)

Wavenumber (cm-1)

Same slurry handsonicated

75wt% Slurry Aged3 days

Same slurry handmixed

Figure 4-34. IR spectra of the OD band in an aged 75 wt% GCC slurry with addition of energy to the system. Adding energy to the system partially restores the fluid like water structure.

1600 1500 1400 1300 1200 11000.0

0.2

0.4

0.6

0.8

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

Same slurry sonicated

75wt% Slurry Aged3 days

Same slurry handmixed

Figure 4-35. IR spectra of the carbonate band in an aged 75 wt% GCC slurry with addition of energy to the system. Adding energy to the system partially restores the carbonate band to a fresh slurry.

Page 82: © 2008 Joshua James Taylor

82

4.4 Probe Molecules

Molecules which contain a carboxylate were chosen to probe the surface of GCC in high

solids loading slurries. The interaction of the probe molecules with the surface of GCC would

provide additional understanding of the interaction between NaPAA and GCC. Benzoic acid and

gallic acid were chosen because their concentrations in a solution could be measured with a

UV/VIS spectrometer. The benzene ring absorbs ultraviolet light causing the electrons transition

from π (bonding) to π* (anti-bonding).

Figure 4-36. Molecular structure of benzoic acid.

Page 83: © 2008 Joshua James Taylor

83

Figure 4-37. Adsorption of benzoic acid onto GCC at varying solids loading. Benzoic acid does not adsorb onto GCC.

4.4.1 Benzoic Acid

Benzoic acid was chosen as a probe molecule because it contains a benzene ring and a

carboxylic acid group (figure 4-36). Adsorption experiments for benzoic acid onto GCC were

specifically performed to demonstrate the carboxyl group adsorption and/or the hydrophobic

adsorption of the benzene ring. Adsorption of NaPAA is believed to be due to the carboxylate

groups interacting with the calcium ions [2, 4, 11, 36]. Geffroy et al. [4] go into detail describing

the complexation of carboxylates with the surface of calcite required for adsorption. They

determine that carboxylates adsorb through chelation. Since it has been demonstrated that

adsorption coordination depends on solids loading, the adsorption of benzoic acid at different

Page 84: © 2008 Joshua James Taylor

84

1640 1620 1600 1580 1560 1540 15201640 1620 1600 1580 1560 1540 15201640 1620 1600 1580 1560 1540 1520

-0.008

-0.004

0.000

0.004

C C

COO-

20 wt% GCC Slurry

Wavenumber (cm-1)

57 wt% GCC Slurryd2 A

/dυ2 (a

.u.)

Benzoic acid in D2O

Figure 4-38. Second derivative of the IR spectra of benzoic acid in D2O, 20 wt% GCC slurry, and 57 wt% GCC slurry. The bands for the benzene ring and the carboxylate do not shift, indicating no adsorption.

solids loading was investigated. Before any adsorption experiments were performed it was

determined that calcium benzoate would not precipitate in the slurry because the concentration of

calcium benzoate was below the precipitation limit of 2.72 g per 100 ml at 20°C. Figure 4-37

demonstrates that benzoic acid does not adsorb onto GCC at any solids loading. Since figure

4-37 does not indicate any adsorption, then the ATR-FTIR will be utilized to confirm that the

carboxylate and the benzene ring are not interacting with the surface of the GCC before

centrifugation.

Page 85: © 2008 Joshua James Taylor

85

3050 3000 2950 2900 2850 28000.00

0.01

0.02

0.03

Benzoic acid in D2O

Log

(1/R

) (a.

u.)

Wavenumber (cm-1)

57 wt% GCC Slurry

Figure 4-39. IR spectrum of the CH bands of benzoic acid in D2O and in a 57 wt% GCC slurry. Formation of new bands indicates interaction of the CH bonds with GCC.

Slurry samples with benzoic acid were prepared at a pH of 9.9 in order to represent the pH

condition of a NaPAA and GCC slurry. Additional confirmation of the pH of the slurry was the

absence of the C=O band at 1700 cm-1. Figure 4-38 demonstrates that benzoic acid does not

adsorb onto the GCC with increasing solids loading. The carboxylate band at 1548 cm-1 for

benzoic acid in D2O does not shift or split when GCC is added to the solution which is an

indication that the carboxylate is not adsorbing onto the GCC particles. Another possible

interaction that could cause adsorption is through hydrophobic adsorption with the benzene ring

or the CH groups. The band for a benzene ring is located at 1596 cm-1 and does not shift with

increasing solids loading so the benzoic acid is not adsorbing through benzene ring. Figure 4-39

Page 86: © 2008 Joshua James Taylor

86

indicates that the CH groups are interacting with the GCC, similar to NaPAA. The IR spectra

could be detecting a small amount of benzoic acid on the surface of the GCC which is within the

error of the adsorption measurements.

The results confirm each other and are in good agreement with Geffroy et al. [4] who

demonstrated that a molecule with one carboxylate group will not adsorb onto the surface of

GCC. Therefore, the next probe molecule chosen contained an OH group which will allow the

molecule to chelate with calcium.

Figure 4-40. Molecular structure of gallic acid.

Page 87: © 2008 Joshua James Taylor

87

Figure 4-41. Adsorption of gallic acid onto GCC at varying solids loading.

4.4.2 Gallic Acid

Gallic acid was chosen as a probe molecule for the same reason that benzoic acid was

chosen but it also includes three hydroxyl groups (figure 4-40). Hydroxyl groups were chosen in

order to determine if hydrogen bonding would play an important role in adsorption. The

adsorption of gallic acid onto GCC in several different solids loading slurries is demonstrated in

figure 4-41. The amount of gallic acid adsorbed is dependant on the solids loading of the slurry

and decreases with an increase weight percent of GCC. Calculations were performed to

determine the adsorbed monolayer concentration of gallic acid is between 1.6 mg/m2 and

8.5 mg/m2. The range in adsorbed monolayer is due to the different modes of adsorption that the

gallic acid could accompany while adsorbing. Comparing the monolayer adsorption calculations

to figure 4-41 would suggest that the adsorbed amount of gallic acid at 10 wt% would contain 3

Page 88: © 2008 Joshua James Taylor

88

1620 1600 1580 1560 1540 1520-0.0024

-0.0016

-0.0008

0.0000

0.0008

bidentatebridging

20 wt% Slurry

C C

d2 A/dυ2 (a

.u.)

Wavenumber (cm-1)

Gallic acid in D2O

Figure 4-42. Second derivative of the IR spectra of gallic acid in D2O and a 20 wt% GCC slurry. The bands for the benzene ring and the carboxylate shift, indicating adsorption.

adsorbed layers, at 30 wt% would contain 2 adsorbed layers, and at 50 wt% and 60 wt% would

contain one monolayer with different modes of adsorption.

The addition of the hydroxyl groups allows for the molecule to chelate with the GCC

surface forming a seven-bond ring through one hydroxyl group, calcium ion, and carboxylate

group (similar results from Geffroy et al. [4]). A seven-member chelate ring is not as stable as a

five-member chelate ring but the chelating ability of gallic acid promotes adsorption unlike

benzoic acid which does have the ability to form a chelate with calcium. A possible adsorption

mechanism for NaPAA is to form an eight-member chelate ring but this is much less stable and

could be a reason why a slurry ages.

Page 89: © 2008 Joshua James Taylor

89

3050 3000 2950 2900 2850 28000.030

0.035

0.040

0.045

0.050

67 wt% GCC Slurry

Wavenumber (cm-1)

Gallic acid in D2O

Log

(1/R

) (a.

u.)

Wavenumber (cm-1) Figure 4-43. IR spectrum of the CH bands of gallic acid in D2O and in a 67 wt% GCC slurry.

Formation of new bands indicates interaction of the CH bonds with GCC.

Analysis of the IR spectra indicates that gallic acid adsorbs onto the surface of GCC.

Figure 4-42 demonstrates that the carboxylate band at 1550 cm-1 shifts to 1547 cm-1 when a 20

wt% GCC slurry is prepared. The carboxylate shift is a confirmation that the gallic acid is

chelating with the surface of the GCC. Another important band includes the benzene ring band

located at 1600 cm-1 for the gallic acid and D2O system at a pH of 9. When a 20 wt% GCC

slurry with gallic acid is analyzed the benzene band shifts to 1594 cm-1 indicating that the

benzene ring also plays a role in the adsorption. Figure 4-43 indicates that the CH groups are

interacting with the GCC, similar to NaPAA.

Page 90: © 2008 Joshua James Taylor

90

1640 1620 1600 1580 1560 1540 1520

-0.0004

0.0000

0.0004

bidentate

bridging

67 wt% Slurry

C C

d2 A/dυ2 (a

.u.)

Wavenumber (cm-1)

20 wt% Slurry

Figure 4-44. Second derivative of the IR spectra of gallic acid in a 20 wt% and 67 wt% GCC slurry. The bands for the benzene ring and the carboxylate shift, indicating change in the coordination of the benzoic acid with increasing solids loading.

The IR spectra for gallic acid demonstrate that with increasing solids loading there is a

change in the adsorption of the gallic acid. Figure 4-44 shows a shift of the carboxylate band at

1547 cm-1 for a 20 wt% GCC slurry to 1560 cm-1 for a 67 wt% GCC slurry. Along with the

carboxylate shift there is a shift of the benzene band from 1594 cm-1 to 1584 cm-1 for a 20 wt%

and 67 wt% slurry, respectively. These results are further conformation that as the solids loading

of a slurry increases the chemical interaction change within the system.

Page 91: © 2008 Joshua James Taylor

91

Figure 4-45. Molecular structure of propionic acid.

3050 3000 2950 2900 2850 28000.020

0.025

0.030

0.035

0.040

0.045

70 wt% GCC Slurry

Wavenumber (cm-1)

Log

(1/R

) (a.

u.)

Propionic acid in D2O

Figure 4-46. IR spectrum of the CH bands of propionic acid in D2O and in a 70 wt% GCC

slurry. Formation of new bands indicates interaction of the CH bonds with GCC.

Page 92: © 2008 Joshua James Taylor

92

4.4.3 Propionic Acid

Propionic acid was chosen as a probe molecule because it is similar in structure to the

monomer of NaPAA (see figure 4-45 for propionic acid structure). Figure 4-46 indicates that the

CH groups are interacting with the GCC, similar to NaPAA. The carboxylate band from

propionic acid in D2O at a pH of 9.9 is located at 1552 cm-1. When 14 wt% and 70 wt% GCC

slurries are prepared, the carboxylate bands do not shift (see figure 4-47). This demonstrates

again that the adsorption of a molecule onto the surface of GCC cannot occur through a single

carboxylate group but must include chelation of the molecule with calcium.

1600 1580 1560 1540 1520-0.0006

-0.0004

-0.0002

0.0000

0.0002

14 wt% GCC Slurry

d2 A/dυ2 (a

.u.)

Wavenumber (cm-1)

Propionic acid in D2O

70 wt% GCC Slurry

Figure 4-47. Second derivative of the IR spectra of propionic acid in D2O, 14 wt% GCC slurry, and a 70 wt% GCC slurry. No shift in the carboxylate band indicates that propionic acid does not adsorb.

Page 93: © 2008 Joshua James Taylor

93

CHAPTER 5 WATER STRUCTURE DEPENDANCE ON SOLIDS LOADING AND AGING

5.1 ATR-FTIR Water Structure

A novel discovery shows that the water structure within a high solids loading slurry is

distinctly different than bulk water (first discussed in Section 4.3.3 and demonstrated in figures

4-17 and 4-18). When water is introduced into a high solids loading GCC slurry there is a

decrease in the structured water, this is indicated with a decrease in absorbance of the 2379 cm-1

OD stretching band. The structured water at 2379 cm-1 is representative of water with a

2800 2600 2400 22000.0

0.2

0.4

0.6

0.8

1.0 75wt% Slurry

10wt% Slurry 20wt% Slurry 30wt% Slurry 40wt% Slurry 50wt% Slurry 60wt% Slurry 70wt% Slurry

increasing solids

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

D2O and GCC

Figure 5-1. IR spectra of the OD stretching bands in GCC slurries ranging from 10 wt% to 75 wt%. As the solids loading increases there is an increase in the fluid structure (2484 cm-1) and decrease in solid structure (2390 cm-1).

Page 94: © 2008 Joshua James Taylor

94

Figure 5-2. Fluid to solid water structure within different solids loading slurries. There are two different regions, the first from 10 to 50 wt% and the second above 50 wt%.

hydrogen bond coordination of four while the fluid like water at 2484 cm-1 is representative of

water with a hydrogen bond coordination less than four. In order to determine if this

phenomenon occurs only in 75 wt% slurries or if it also occurs at other solids loading, several

slurries were prepared ranging from 10 wt% to 75 wt%. The samples were analyzed with the

ATR-FTIR (figure 5-1) and the ratio of the fluid like band intensity (2484 cm-1) to the solid like

band intensity (2390 cm-1) was calculated for each spectrum. The ratios for each solids loading

are compared in figure 5-2. Initially, there is a change in the water structure when a slurry is

prepared at 10 wt% compared to only a GCC and D2O mixture. From 10 wt% up to 75 wt%

there are two distinct regions. The first region has a constant negative slope which indicates a

linear change in water structure up to 50 wt% solids loading. The second region has a positive

Page 95: © 2008 Joshua James Taylor

95

slope which indicates an increase in the fluid like water structure above 50 wt% solids loading.

The change in slope from region one to region two may be due to water confinement and/or ions

in the system which will be discussed in Sections 5.2 and 5.3.

2800 2600 2400 22000.0

0.2

0.4

0.6

0.8

1.0

increasing age

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

GCC and D2O

75wt% Slurry Aged <1hr 75wt% Slurry Aged 25hrs 75wt% Slurry Aged 51hrs

Figure 5-3. IR spectra of the OD band in a 75 wt% slurry with aging. Increasing age causes an increase in the concentration of structured water in the slurry.

Aging of slurries is a common problem for industry as mentioned earlier. Investigation of

75 wt% GCC slurries with the ATR-FTIR over a period of several days demonstrated a novel

discovery indicating that the water structure within a high solids loading slurry changes with age

(figure 5-3). Increasing the age of a slurry increases the concentration of structured water

making the slurry more similar to bulk water. As seen in figure 5-4, the decrease in water

structure over a two day period leads to water structure that has a higher concentration of

Page 96: © 2008 Joshua James Taylor

96

Figure 5-4. Fluid to solid water structure within an aging 75 wt% GCC slurry. The fluid like

water structure decreases with aging becoming more like a slurry without dispersant.

structured water than a 40 wt% GCC slurry. The change in water structure could be a result of

water confinement and/or ions in the system.

5.2 Water Confinement

One of the possible reasons that the IR spectra show an increase in the fluid like water

structure is due to an increase in the confinement of water. Since water’s density is higher than

ice then when water molecules are confined within nanometers they behave like water [48, 77,

93-108]. Calculations were performed to determine the distance between calcite particles and

also the NaPAA. The following assumption were made in the calculations: the dispersant has a

radius of gyration of 2 nm to 4 nm [70], dispersant molecules do not adsorb onto calcite

particles, GCC particles have a diameter of 1 µm, and FCC packing of both calcite particles and

Page 97: © 2008 Joshua James Taylor

97

Figure 5-5. Calculations of the distance between GCC particles in different solids loading slurries with NaPAA having a radius of gyration of 2 nm, 3 nm, and 4 nm.

dispersant. All of the assumption are not a completely accurate representation of the slurry

system but are necessary for estimating distances. The distance between GCC particles

decreases from 1610 nm to 128 nm with increasing solids loading from 10 wt% to 75 wt%,

respectively (figure5-5). These distances are too large to contribute to the confinement of water

according to literature. Figure 5-6 shows the calculations for the distances between NaPAA

molecules in the dispersing medium without GCC particles. As the solids loading is increased

from 10 wt% to 75 wt% the distance between NaPAA molecules changes from 19 nm to 1.5 nm,

respectively. At 75 wt% the confinement of water within 1.5 nm could contribute to the change

in water structure observed for high solids loading samples. Also, due to the Brownian motion

Page 98: © 2008 Joshua James Taylor

98

Figure 5-6. Calculations of the distance between NaPAA molecules in different solids loading slurries with a radius of gyration of 2 nm, 3 nm, and 4 nm.

of the GCC particles and NaPAA dispersant the water has less time between collisions to form

structured water as the solids loading is increased.

5.3 Ions as Water Structure Makers or Breakers

As second possible reason that the IR spectra show an increase in the fluid like water

structure is due to the ions in the system. Ions are known to interact with water molecules and

form water sheaths around them. The ions either promote or destroy water structure depending if

the ion is a water structure maker or breaker [86, 95, 109-120]. The high solids loading slurry

systems have a high concentration of ions which include calcium, carbonate, bicarbonate,

sodium, carboxylates, and hydroxyls due to GCC’s solubility and the addition of NaPAA. The

Page 99: © 2008 Joshua James Taylor

99

concentration of several of these ions is dependent on the solids loading and certain ion

concentrations have not been determined in high solids loading slurries. However, it is possible

to calculate the concentration of sodium ions and the concentration of carboxylate groups

because a known amount of NaPAA is added to the system. Calculations determine that a

75 wt% GCC slurry with 1 wt% NaPAA (weight percent of GCC weight) will contain a

concentration of 0.34 M of sodium ions which is extremely high. With such a high concentration

of sodium ions the NaPAA will be completely collapsed with a radius of gyration of 2 nm [70].

Additionally, the GCC is slightly soluble so there is dissolution and formation of calcium ions,

carbonate ions, and bicarbonate ions into the slurry system as indicated with the formation of a

Figure 5-7. Change of pH in an aging 75 wt% GCC slurry over a six day time span. A decrease in pH is due to the concentration change of species in the slurry.

Page 100: © 2008 Joshua James Taylor

100

low frequency shoulder on the carbonate band (figure 4-29). As the age of a slurry increases

there is a change in the concentration of ions in the system along with a change in the water

structure as shown before in figures 4-30 and 5-3, respectively. This change in the concentration

of ions in the system with time is also demonstrated with a decrease in the pH with aging (figure

5-7).

0.0

0.2

0.4

0.6

0.8

1.0

2700 2600 2500 2400 2300 2200

40 wt% NaPAA Slurry

Wavenumber (cm-1)

1 wt% NaPAA Slurry

Log

(1/R

) (a.

u.)

10 wt% NaPAA Slurry

Figure 5-8. IR spectra of the OD band of a 75 wt% GCC slurry with increase NaPAA. There is a decrease in the structured water indicating that NaPAA is a water structure breaker in a GCC slurry.

NaPAA as a polyelectrolyte contains carboxylate groups which Raviv et al. [12, 77] have

shown to cause lubrication of the system due to formation of water sheaths around the

carboxylate groups. Raviv et al. explain that the water molecules which form the sheaths around

Page 101: © 2008 Joshua James Taylor

101

the ions are tightly bound to the carboxylate groups but can easily and readily exchange with

other water molecules from other water sheath with which they come in contact. The fluid like

properties of water molecules bound to NaPAA are also confirmed with ATR-FTIR

measurements in Figure 5-8. As the amount of NaPAA is increased within a 75 wt% GCC slurry

there is an increase in the fluid like to solid like water structure ratio. So NaPAA in a GCC

slurry acts as a water structure breaker.

Previously as discussed in Sections 4.3.3 and 5.1 the water structure is dependent on many

variables in the slurry system. As the variables are altered there is also a change in the

concentration and types of ions in the system. The ions in the system could be the reason for the

change in the structure of water due to their water structure making and breaking abilities. In the

following chapter the water structure making and breaking abilities of chemicals will be

analyzed in high solids loading slurries.

Page 102: © 2008 Joshua James Taylor

102

CHAPTER 6 WATER STRUCTURE MAKING AND BREAKING CHEMICALS

The previous results have demonstrated that water structure plays an important role in the

dispersion of a high solids loading slurry. Several chemicals which are known to be water

structure makers or breakers have been introduced into high solids loading slurries. The effect of

the chemicals on the slurry’s physical properties and water structure are discussed.

6.1 Calcium Chloride

Calcium ions play an important role in the dispersion of GCC as discussed in chapter two.

GCC is slightly soluble in water; therefore, calcium ions are present in the solution and are free

to interact with the NaPAA, carbonates, and dispersing medium. Since calcium is known as a

water structure maker [37], then calcium chloride was introduced to a slurry in order to

determine if calcium ions will change the water structure. Figure 6-1 shows that addition of

0.1 M calcium chloride prevents the decrease in structured water which occurs when a slurry is

prepared. Since it was previously demonstrated in Section 5.1 that the concentration of

structured water increases with aging then the results from figure 6-1 would indicate that during

the aging of a slurry there is an increase in dissolved calcium.

Addition of CaCl2 to the slurry also changes the carbonate band. As seen in figure 6-2 the

slurry with CaCl2 has a carbonate bending band with a smaller FWHM. A similar result was

seen with the carbonate stretching band. This indicates that the excess amount of calcium ions in

the system prevents dissolution of the carbonate.

6.2 Sodium Bicarbonate

As demonstrated in Section 4.3.6 there is an increase in the carbonate species indicated by the

increase in the shoulder of the carbonate band. Since the slurry system contains carbonate ions

along with the sodium ions, which are introduced with the NaPAA, other species are also

Page 103: © 2008 Joshua James Taylor

103

2800 2600 2400 22000.0

0.2

0.4

0.6

0.8

1.0L

og (1

/R)

(a.u

.)

Wavenumber (cm-1)

50wt% GCC in D2O

45wt% GCC slurry

45wt% GCC Slurry with 0.1M CaCl2

Figure 6-1. IR spectra of the OD band of slurries with and without CaCl2. Ca2+ as a water structure maker prevents the increase in fluid like water of a GCC slurry.

1600 1500 1400 1300 1200 11000.0

0.2

0.4

0.6

0.8

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

50wt% GCC in D2O

45wt% GCC slurry

45wt% GCC Slurry with 0.1M CaCl2

Figure 6-2. IR spectra of the carbonate band of slurries with and without CaCl2. Ca2+ prevents some of the interactions with the carbonate species.

Page 104: © 2008 Joshua James Taylor

104

present including sodium carbonate and/or sodium bicarbonate. Sodium carbonate is known as a

strong structure maker and sodium bicarbonate as a weak structure maker [86, 113, 114, 121].

Sodium bicarbonate was added to a 75 wt% slurry and the slurry was aged for 64 hrs. Figure 6-3

demonstrates that the OD shoulder at 2379 cm-1 increases with aging which is similar to the

results for an aged slurry. The sodium bicarbonate as a water structure maker does not prevent

the change in water structure for an aging system.

Rheology measurements were performed for a slurry with and without sodium bicarbonate.

Figure 6-4 demonstrates addition of sodium bicarbonate increases the viscosity of the slurry.

2800 2600 2400 22000.0

0.2

0.4

0.6

0.8

1.0 75wt% Slurry with

Sodium Bicarbonate Aged 64hrs

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

75wt% Slurry with Sodium Bicarbonate

Figure 6-3. IR spectra of the OD band of a 75 wt% GCC slurry with 0.19 M sodium bicarbonate with aging. As a weak structure maker, sodium bicarbonate allows for an increase in the structured water with aging of a 75 wt% slurry.

Page 105: © 2008 Joshua James Taylor

105

Figure 6-4. Rheology of 75 wt% GCC slurries with and without sodium bicarbonate. Sodium bicarbonate increases the viscosity.

Since there is an increase in the viscosity this indicates that the water structure making ability of

sodium bicarbonate could be a reason for the aging of a 75 wt% slurry.

6.3 Ethylene Glycol

Ethylene glycol is a common additive to water in order to decrease the freezing

temperature (structure shown in figure 6-5). It is also known as a water structure breaker and

each oxygen is known to be fully hydrated with ~2-3 water molecules [122-124]. Figure 6-6

shows the OD band for a 75 wt% GCC slurry containing 0.50 M of ethylene glycol. The water

structure within the slurry does not change with aging due to the addition of ethylene glycol.

Also, the addition of ethylene glycol prevents any changes in the carbonate band with aging as

Page 106: © 2008 Joshua James Taylor

106

Figure 6-5. Structure of ethylene glycol.

2800 2600 2400 22000.0

0.2

0.4

0.6

0.8

1.0

Log

(1/R

) (a

.u.)

Wavenumber (cm-1)

75wt% Slurry with Ethylene Glycol Aged 64hrs

75wt% Slurry with Ethylene Glycol

Figure 6-6. IR spectra of the OD band of a 75 wt% GCC slurry with 0.5 M ethylene glycol with aging. Ethylene glycol as a water structure breaker prevents structure water from forming while the slurry ages.

Page 107: © 2008 Joshua James Taylor

107

1600 1500 1400 1300 1200 11000.0

0.2

0.4

0.6

0.8L

og (1

/R) (

a.u.

)

Wavenumber (cm-1)

75wt% Slurry with Ethylene Glycol Aged 64hrs

75wt% Slurry with Ethylene Glycol

Figure 6-7. IR spectra of the carbonate band of slurries with and without ethylene glycol. Ethylene glycol prevents interactions with the carbonate species as it ages.

shown in figure 6-7. The spectra confirm that addition of a water structure breaker prevents the

aging of the high solids loading system.

Additionally, the rheological properties of the slurries were investigated. According to

literature when ethylene glycol is added to water the viscosity increases [125-130]. Upon

addition of 0.5M of ethylene glycol to the slurry system the max viscosity of the slurry decreased

475 Pa·s which is in the opposite direction of a water and ethylene glycol system. Also, the

decrease in the viscosity is greater than the decrease in viscosity due to addition of an equal

weight of water, demonstrated in Figure 6-8. Further, the viscosity of a 48 hour aged slurry with

Page 108: © 2008 Joshua James Taylor

108

Figure 6-8. Viscosity measurements of 75 wt% slurries with and without ethylene glycol. Ethylene glycol decreases the viscosity more than adding an equal weight amount of water to the slurry.

ethylene glycol is lower than an aged slurry without ethylene glycol. This is additional

confirmation that nonionic water structure breakers may prevent the aging of a slurry.

6.4 Propylene Glycol

Ethylene glycol is a toxic chemical; therefore, it is commonly substituted with propylene

glycol (see figure 6-9 for chemical structure) which behaves similar to ethylene glycol but is

environmentally safe. Figure 6-10 shows the OD bands of a 75 wt% GCC slurry containing

0.43 M of propylene glycol less than an hour after preparation and after 48 hours of aging.

Page 109: © 2008 Joshua James Taylor

109

Figure 6-9. Structure of propylene glycol.

2700 2600 2500 2400 2300 22000.0

0.2

0.4

0.6

0.8

1.0

75 wt% Slurry with Propylene Glycol

75 wt% Slurry with Propylene Glycol Aged 48hrs

Log

(1/R

) (a.

u.)

Wavenumber (cm-1)

Figure 6-10. IR spectra of the OD band of a 75 wt% GCC slurry with 0.43 M ethylene glycol with aging. Propylene glycol as a water structure breaker prevents structured water from forming while the slurry ages.

Page 110: © 2008 Joshua James Taylor

110

Figure 6-11. Viscosity measurements of 75 wt% slurries with and without propylene glycol. Propylene glycol decreases the viscosity at low shear rates but increases viscosity at higher shear rates.

Propylene glycol shows similar results to ethylene glycol indicating no change in water structure

with aging.

Additionally, the rheological properties of the slurries were investigated. Similar to

ethylene glycol the literature shows that addition of propylene glycol to water increases the

viscosity of water [131]. Upon addition of 0.43 M of propylene glycol to a 75 wt% GCC slurry

the max viscosity of the slurry decreased 200 Pa·s (figure 6-11). Also, the viscosity of a 48 hour

aged slurry with propylene glycol is less than the viscosity of an aged slurry without propylene

Page 111: © 2008 Joshua James Taylor

111

glycol. Propylene glycol as a water structure breaker is additional confirmation that water

structure breakers prevent the aging of 75 wt% GCC solids loading slurries.

Page 112: © 2008 Joshua James Taylor

112

CHAPTER 7 SUMMARY AND DISCUSSION

The main objective of this dissertation was to determine the mechanism of adsorption for

NaPAA onto GCC in high solids loading slurries. An overview of literature provided some

insight into the behavior of calcium carbonate and NaPAA in water at dilute concentrations.

This was followed by several experiments utilizing adsorption isotherms, turbidity

measurements, probe molecules, and rheology. The presented results focused on specific parts

of the NaPAA molecule, including the carboxylate and CH groups, and focused on the

carbonates from GCC. Also, the results in chapters 5 and 6 focus on the water structure within

high solids loading slurries. This chapter summarizes all the results in order to obtain a

comprehensible understanding of the adsorption of NaPAA onto GCC in high solids loading

slurries.

As mentioned in the literature review, calcium carbonate is slightly soluble in water and

the surface shows complex behavior due to the chemical equilibrium of the mineral/water

interface. Geffroy et al. [4] come to the conclusion that within a pH range of 8 to11 the surface

of calcium carbonate in water consists mainly of neutral sites (–CaOH and –CO3H) and ionic

sites (–Ca+ and –CO3-). Katz et al. [34] determine that the calcium ions form asymmetrical

coordination structures with water. From the literature review it is apparent that the surface of

calcium carbonate in water is heterogeneous, consisting of several different sites with different

charges and hydration states. This is in agreement with the adsorption isotherms performed in

this dissertation. The adsorption of NaPAA in 75 wt% GCC slurries follows the Freundlich

isotherm which is an indication of a heterogeneous surface. Also, several of the sites on the

GCC surface must compete with calcium and sodium ions in the solution for the adsorption of

NaPAA. The ions in solution can precipitate the dispersant and cause the dispersant to be

Page 113: © 2008 Joshua James Taylor

113

ineffective. Turbidity measurements of NaPAA in water with varying concentrations of sodium

and calcium ions indicate that NaPAA does not precipitate in slurry conditions. The turbidity

results are only an indication and not a true representation of a high solids loading GCC slurry

because the ionic condition of water in a high solids loading slurry may be different.

Results from the IR spectra of the carbonate band indicated that there is formation of

bicarbonates in 75 wt% GCC slurries. There is little change in the carbonate band up to 50 wt%

GCC but above 50 wt% GCC the band forms a shoulder which continues to extend to lower

wavenumbers as solids loading increases. The shoulder is an indication of bicarbonate formation

and these results demonstrates that a dilute system does not represent a high solids loading

system. Also, it was demonstrated that the IR spectrum of a dried 75 wt% GCC slurry does not

represent an IR spectrum of an aqueous slurry, minus the water, because the dried slurry does not

contain the bicarbonate bands. As a 75 wt% GCC slurry ages there is a shift of the bicarbonate

band to higher wavenumbers which is also accompanies with a decrease in the pH from 9.9 to

9.5. Knez et al. [19] mention that the activity of carbonate species (H2CO3 and HCO3-) increases

with rising pH and the activity of calcium ions decreases with rising in pH. This would be in

agreement with the IR spectra because there is a decrease in the band (decreasing carbonate

activity) with decreasing pH over time. This would also indicate that the activity of calcium ions

in the system increases with aging.

Further investigation of the CH groups of NaPAA would require a solvent exchange. The

OH stretching region of H2O overlaps the CH stretching region of NaPAA; therefore, D2O was

exchanged for H2O allowing for analysis of the CH bands. Surprising results were obtained from

the IR spectra indicating that the CH groups of NaPAA and the probe molecules were interacting

with the GCC. Previous literature has demonstrated that shifts in the CH band are possible due

Page 114: © 2008 Joshua James Taylor

114

to temperature changes [87, 88] and aqueous versus dry conditions [89] which were discussed

with more detail in Section 4.3.4. The following is a possible explanation for the interactions of

some CH bonds with the surface of GCC. As a carboxylate adsorbs onto the surface of GCC the

CH bond near the carboxylate is brought into proximity to the surface. The surface restricts the

bending and stretching modes of the CH bond causing a change in the IR spectra. As

demonstrated in figure 4-21 the CH bands indicate that the adsorption limit of NaPAA on GCC

is between 1 wt% and 9 wt%. Calculations of monolayer adsorption are between 2.14 mg/m2

and 4.28 mg/m2 depending on the conformation of the polymer.

The carboxylate group of NaPAA interacts with cations and surfaces in four different

modes: ionic, bridging, bidentate, and unidentate. Analysis of the IR spectrum of a 75 wt% GCC

slurry demonstrated that the carboxylate groups interact in unidentate, bidentate, and bridging

modes. Probe molecules were utilized in order to determine if the carboxylate group was

directly responsible for the adsorption of NaPAA onto the surface of GCC. Analysis of the IR

spectra for benzoic acid, containing only one carboxylate group, demonstrated that the

carboxylate group does not adsorb at low or high solids loading. Adsorption measurements of

benzoic acid onto GCC support the IR results by indicating no adsorption. The next probe

molecule included propionic acid which also contains a single carboxylate group. Analysis of

the IR spectra also demonstrated no interaction of the carboxylate group with GCC. Finally,

gallic acid which contains a carboxylate group and three OH groups was used as a probe

molecule. Analysis of the IR spectra of the carboxylate group indicated interaction with the

GCC. Adsorption of gallic acid is also confirmed with adsorption measurements. The

interaction of gallic acid with the surface of GCC is possible because the molecule is able to

chelate with the surface through the carboxylate group and an OH group. NaPAA is also

Page 115: © 2008 Joshua James Taylor

115

expected to interact through chelation of the carboxylate groups. As discussed in Section 2.2,

literature supports these results with Geffroy et al. [4] and Dobson et al. [49] who also explain

that the adsorption of carboxylates is through chelation and that adsorption does not occur with

molecules which only contain one carboxylate group.

Further investigation of the adsorption of carboxylate groups onto calcium ions could

explain why a decrease in structured water is observed for a 75 wt% slurry compared to bulk

water, Section 4.3.3. There are two papers which could explain indirectly why a high solids

loading slurry demonstrates a change in water structure. One by Geffroy et al. [5] and another

by Sinn et al. [15] which discuss the exchange of calcium ions binding onto NaPAA as

endothermic. As discussed in Section 2.2, the implication of an endothermic binding means the

adsorption process is driven by an increase in entropy. The increase in entropy is believed to be

due to the release of water molecules from the dehydration of the calcium ion and carboxylate

ion. Since the calcium ion is a water structure maker, then dehydration of the ion will release the

structured water that was bound to it.

The water structure is also dependant on the solids loading of the system. Initially when a

10 wt% GCC slurry is prepared there is a decrease in the structured water. As the solids loading

increases to 50 wt% there is a small increase in the structured water. When the solids loading is

increased above 50 wt% and up to 75 wt% there is an increase in the fluid like water structure,

figure 5-2. Also accompanying the change in water structure is a change in the adsorption mode

of the carboxylate. As the solids loading of a slurry sample is increased, the carboxylates which

are adsorbing in a unidentate mode shift their adsorption toward a bridging mode. With the

bridging mode, one carboxylate group is interacting with two calcium ions instead of the

unidentate mode in which one carboxylate group interacts with one calcium ion. The water

Page 116: © 2008 Joshua James Taylor

116

structure change could be an indication of the mode of adsorption taking place for the

carboxylate groups.

The water within an aging 75 wt% GCC slurry increases in structured water. The change

could be due to the water structure making ions increasing in concentration over time. A change

in the pH with aging signifies the slurry is not in equilibrium and the concentration of calcium

ions in the system could be increasing over time. This would support the observed increase in

structured water because the calcium ion is a water structure maker. Also, with aging the

coordination mode of the carboxylate group shifts from a unidentate mode closer to a bridging

mode. The increase in calcium ions would provide more ions which are required for a bridging

mode.

Since the previous results indicate that the ions in the system and the water structure are

related to the aging of a system, several chemicals that are known to be water structure makers or

breakers were introduced into high solids loading slurries. Calcium chloride, a water structure

maker, prevented the formation of fluid like water, limited the interactions with the carbonate

species, and increased the viscosity. Sodium bicarbonate, a water structure maker, increased the

viscosity of the slurry by a factor of 8 at a shear rate of 0.01 s-1 and did not prevent the increase

in structured water with aging of the slurry. Ethylene glycol, known as a water structure breaker,

decreased the viscosity by a factor of 3.5 at a shear rate of 0.01 s-1 and prevented the change in

water structure with aging. Propylene glycol, known as a water structure breaker, also prevented

a change in the water structure but only decreased the viscosity by 1.4 at a shear rate of 0.01 s-1

and at high shear rates ethylene glycol increased the viscosity of the slurry. Results from the

water structure makers and breakers demonstrate that the slurry system’s physical properties

depend on the water structure and ions in the system.

Page 117: © 2008 Joshua James Taylor

117

From the previous discussion, a model for the adsorption of NaPAA onto GCC is shown in

figure 7-1. Part A illustrates NaPAA in solution before adsorption onto GCC. Structured water

is shown on the surface of the GCC and surrounding the NaPAA. The driving force for the

dispersant to adsorb onto the surface has been determine to be an increase in entropy. The

increase in entropy comes from the dehydration of the carboxylate groups and the calcium ions

as they interact. Part B illustrates that when the dispersant adsorbs, its trains and the surface of

GCC release the structured water. Evidence of a decrease in structured water has been

demonstrated in the IR spectra and discussed.

Figure 7-1. Adsorption of NaPAA onto GCC due to an increase in entropy with the release of structured water. A) NaPAA in solution before adsorbing onto GCC surface, B) NaPAA adsorption releases structured water from the carboxylate groups and the calcium as demonstrated in the train of the polymer.

Page 118: © 2008 Joshua James Taylor

118

CHAPTER 8 CONCLUSIONS AND SUGGESTIONS

8.1 Conclusions

This dissertation is the first work to discuss the adsorption of NaPAA onto GCC in high

solids loading slurries. Previous published works have only focused on the analysis of dilute

systems. Several techniques were utilized to investigate slurries up to 75 wt% GCC including

turbidity measurements, adsorption isotherms, rheology, and ATR-FTIR. ATR-FTIR was

extensively used because the phenomenon of the evanescent wave in the ATR-FTIR allows for

analysis of dense systems in situ. The following are several novel discoveries and conclusions

from this dissertation.

The chemical interactions and water structure within a dilute system do not represent high

solids loading slurries. The differences in the systems are demonstrated with the carboxylate

group, carbonate species, and the water structure. As the solids loading increases the carboxylate

groups adsorbed in a unidentate mode shift toward a bidentate mode of adsorption. Also, with

increasing solid loading there is formation of bicarbonates within the slurry. A novel discovery

demonstrates that there is a decrease in the concentration of structured water with the

combination of all three components of a slurry: NaPAA, water, and GCC. Increasing solids

loading above 50 wt% increases the fluid like water structure of the slurry. These results

demonstrate the changes which take place as solids loading is increase and are an important

consideration for research of high solids loading slurries.

Additionally, the water structure, carboxylate adsorption mode, and bicarbonate

interactions were demonstrated to change with aging of a slurry. The water increases in

concentration of structured water with aging. The carboxylate adsorption mode shifts from a

unidentate more to a bridging mode along with a decrease in the activity of the bicarbonate

Page 119: © 2008 Joshua James Taylor

119

species. With these results several water structure makers and breakers were shown to either

decrease slurry performance (water structure makers) or increase slurry performance (water

structure breakers). Ethylene glycol, a water structure breaker, demonstrated the best results for

improving the properties of a 75 wt% GCC slurry by preventing the water structure change with

age and decreasing the viscosity of the slurry.

Finally, a model for the adsorption of NaPAA onto GCC was proposed and discussed. The

chelating ability of NaPAA allows for adsorption onto GCC. The adsorption is due to an

increase in entropy due to the release of water molecules from the dehydration of the interacting

carboxylate groups and calcium ions.

8.2 Suggestions

This dissertation has provided several insights into the adsorption of NaPAA onto GCC in

high solids loading slurries. By utilizing these insights a few suggestion are given to improve the

dispersion of high solids loading slurries. First, chapter 6 has already demonstrated that water

structure breaking chemicals could be introduced into a slurry system to improve the properties

slurry. This option would be useful if the dispersant must not be modified. Second, improved

anchoring of the dispersant could be accomplished by adding an OH group to the carbon

adjacent the carboxylate group on each repeat unit. Adding the OH group would allow for a

five-member chelate ring to form with calcium ion which is more stable than the eight-member

ring formed by NaPAA. Third, a water structure breaking chemical group could be added to the

alkyl chain of the dispersant. This would cause the polymer to act as both the water structure

breaker and dispersant.

Page 120: © 2008 Joshua James Taylor

120

LIST OF REFERENCES

[1] J.F. Chen, T.B. He, W. Wu, D.P. Cao, J. Yun, C.K. Tan, Adsorption of sodium salt of poly(acrylic) acid (PAANa) on nano-sized CaCO3 and dispersion of nano-sized CaCO3 in water, Colloid Surf. A-Physicochem. Eng. Asp. 232 (2004) 163-168.

[2] M. Donnet, P. Bowen, N. Jongen, J. Lemaitre, H. Hofmann, Use of seeds to control precipitation of calcium carbonate and determination of seed nature, Langmuir 21 (2005) 100-108.

[3] J.E. Gebhardt, D.W. Fuerstenau, Adsorption of Polyacrylic-Acid at Oxide Water Interfaces, Colloids and Surfaces 7 (1983) 221-231.

[4] C. Geffroy, A. Foissy, J. Persello, B. Cabane, Surface complexation of calcite by carboxylates in water, Journal of Colloid and Interface Science 211 (1999) 45-53.

[5] C. Geffroy, J. Persello, A. Foissy, B. Cabane, F. Tournilhac, The frontier between adsorption and precipitation of polyacrylic acid on calcium carbonate, Rev. Inst. Fr. Pet. 52 (1997) 183-190.

[6] C. Geffroy, J. Persello, A. Foissy, P. Lixon, F. Tournilhac, B. Cabane, Molar mass selectivity in the adsorption of polyacrylates on calcite, Colloid Surf. A-Physicochem. Eng. Asp. 162 (2000) 107-121.

[7] S. Lages, R. Schweins, K. Huber, Temperature-induced collapse of alkaline earth cation - Polyacrylate anion complexes, J Phys Chem B 111 (2007) 10431-10437.

[8] J.E. Loy, J.H. Guo, S.J. Severtson, Role of adsorption fractionation in determining the CaCO3 scale inhibition performance of polydisperse sodium polyacrylate, Ind Eng Chem Res 43 (2004) 1882-1887.

[9] R. Nystrom, K. Backfolk, J.B. Rosenholm, K. Nurmi, The effect of pretreatment of calcite dispersions with anionic sodium polyacrylate on their flocculation behavior induced by cationic starch, Journal of Colloid and Interface Science 262 (2003) 48-54.

[10] J.S. Phipps, D.R. Skuse, Role of dispersants in the production of fine particle size calcium carbonate and kaolin slurries, CIM Bull. 96 (2003) 55-60.

[11] I. Pochard, P. Couchot, C. Geffroy, A. Foissy, J. Persello, Counterions contributions in polyelectrolyte adsorption, Rev. Inst. Fr. Pet. 52 (1997) 251-253.

[12] U. Raviv, S. Giasson, N. Kampf, J.F. Gohy, R. Jerome, J. Klein, Lubrication by charged polymers, Nature 425 (2003) 163-165.

[13] R. Schweins, K. Huber, Collapse of sodium polyacrylate chains in calcium salt solutions, Eur Phys J E 5 (2001) 117-126.

Page 121: © 2008 Joshua James Taylor

121

[14] R. Schweins, P. Lindner, K. Huber, Calcium induced shrinking of NaPA chains: A SANS investigation of single chain behavior, Macromolecules 36 (2003) 9564-9573.

[15] C.G. Sinn, R. Dimova, M. Antonietti, Isothermal titration calorimetry of the polyelectrolyte/water interaction and binding of Ca2+: Effects determining the quality of polymeric scale inhibitors, Macromolecules 37 (2004) 3444-3450.

[16] M. Cechova, B. Alince, T.G.M. van de Ven, Stability of ground and precipitated CaCO3 suspensions in the presence of polyethylene oxide and kraft lignin, Colloid Surf. A-Physicochem. Eng. Asp. 141 (1998) 153-160.

[17] P. Pang, Y. Deslandes, S. Raymond, G. Pleizier, P. Englezos, Surface analysis of ground calcium carbonate filler treated with dissolution inhibitor, Ind Eng Chem Res 40 (2001) 2445-2451.

[18] A. Tabtiang, R. Venables, The performance of selected unsaturated coatings for calcium carbonate filler in polypropylene, Eur. Polym. J. 36 (2000) 137-148.

[19] S. Knez, D. Klinar, J. Golob, Stabilization of PCC dispersions prepared directly in the mother-liquid after synthesis through the carbonation of (hydrated) lime, Chem. Eng. Sci. 61 (2006) 5867-5880.

[20] J.Y. Gal, J.C. Bollinger, H. Tolosa, N. Gache, Calcium carbonate solubility: A reappraisal of scale formation and inhibition, Talanta 43 (1996) 1497-1509.

[21] J. Johnston, E.D. Williamson, The complete solubility curve of calcium carbonate, J Am Chem Soc 38 (1916) 975-983.

[22] S.I. Kuriyavar, R. Vetrivel, S.G. Hegde, A.V. Ramaswamy, D. Chakrabarty, S. Mahapatra, Insights into the formation of hydroxyl ions in calcium carbonate: temperature dependent FTIR and molecular modelling studies, J Mater Chem 10 (2000) 1835-1840.

[23] P. Moulin, H. Roques, Zeta potential measurement of calcium carbonate, Journal of Colloid and Interface Science 261 (2003) 115-126.

[24] Y.C. Huang, F.M. Fowkes, T.B. Lloyd, N.D. Sanders, Adsorption of Calcium-Ions from Calcium-Chloride Solutions onto Calcium-Carbonate Particles, Langmuir 7 (1991) 1742-1748.

[25] C. Santschi, M.J. Rossi, Uptake of CO2, SO2, HNO3 and HCl on calcite (CaCO3) at 300 K: Mechanism and the role of adsorbed water, J Phys Chem A 110 (2006) 6789-6802.

[26] S.L.S. Stipp, Toward a conceptual model of the calcite surface: Hydration, hydrolysis, and surface potential, Geochim Cosmochim Ac 63 (1999) 3121-3131.

Page 122: © 2008 Joshua James Taylor

122

[27] Y.F. Yang, G.S. Gai, S.M. Fan, X.Y. Hao, Q.R. Chen, Nanostructured modification of mineral particle surfaces in Ca(OH)(2)-H2O-CO2 system, J. Mater. Process. Technol. 170 (2005) 58-63.

[28] T. Ding, E.S. Daniels, M.S. El-Aasser, A. Klein, Surface treatment and characterization of functionalized latex particles and inorganic pigment particles used in the study of film formation from pigmented latex systems, J. Appl. Polym. Sci. 99 (2006) 398-404.

[29] N.I. Ivanova, E.D. Shchukin, Mixed Adsorption of Ionic and Nonionic Surfactants on Calcium-Carbonate, Colloid Surf. A-Physicochem. Eng. Asp. 76 (1993) 109-113.

[30] R. Nystrom, G. Hedstrom, J. Gustafsson, J.B. Rosenholm, Mixtures of cationic starch and anionic polyacrylate used for flocculation of calcium carbonate - influence of electrolytes, Colloid Surf. A-Physicochem. Eng. Asp. 234 (2004) 85-93.

[31] M.G. Song, J.Y. Kim, J.D. Kim, The dispersion properties of precipitated calcium carbonate suspensions adsorbed with alkyl polyglycoside in aqueous medium, Journal of Colloid and Interface Science 226 (2000) 83-90.

[32] M.G. Song, J.Y. Kim, J.D. Kim, Effect of sodium stearate and calcium ion on dispersion properties of precipitated calcium carbonate suspensions, Colloid Surf. A-Physicochem. Eng. Asp. 229 (2003) 75-83.

[33] S. Suty, B. Alince, T.G.M. vandeVen, Stability of ground and precipitated CaCO3 suspensions in the presence of polyethylenimine and salt, J. Pulp Pap. Sci. 22 (1996) J321-J326.

[34] A.K. Katz, J.P. Glusker, S.A. Beebe, C.W. Bock, Calcium ion coordination: A comparison with that of beryllium, magnesium, and zinc, J Am Chem Soc 118 (1996) 5752-5763.

[35] F.C. Lightstone, E. Schwegler, M. Allesch, F. Gygi, G. Galli, A first-principles molecular dynamics study of calcium in water, Chemphyschem 6 (2005) 1745-1749.

[36] Y.Q. Lu, J.D. Miller, Carboxyl stretching vibrations of spontaneously adsorbed and LB-transferred calcium carboxylates as determined by FTIR internal reflection spectroscopy, Journal of Colloid and Interface Science 256 (2002) 41-52.

[37] A.A. Zavitsas, Aqueous solutions of calcium ions: Hydration numbers and the effect of temperature, J Phys Chem B 109 (2005) 20636-20640.

[38] N. Borochov, H. Eisenberg, Stiff (DNA) and Flexible (Napss) Polyelectrolyte Chain Expansion at Very-Low Salt Concentration, Macromolecules 27 (1994) 1440-1445.

[39] S. Forster, M. Schmidt, M. Antonietti, Experimental and Theoretical Investigation of the Electrostatic Persistence Length of Flexible Polyelectrolytes at Various Ionic Strengths, J Phys Chem-Us 96 (1992) 4008-4014.

Page 123: © 2008 Joshua James Taylor

123

[40] L. Dupont, A. Foissy, R. Mercier, B. Mottet, Effect of Calcium-Ions on the Adsorption of Polyacrylic-Acid onto Alumina, Journal of Colloid and Interface Science 161 (1993) 455-464.

[41] K.K. Das, P. Somasundaran, Ultra-low dosage flocculation of alumina using polyacrylic acid, Colloid Surf. A-Physicochem. Eng. Asp. 182 (2001) 25-33.

[42] K.K. Das, P. Somasundaran, Flocculation-dispersion characteristics of alumina using a wide molecular weight range of polyacrylic acids, Colloid Surf. A-Physicochem. Eng. Asp. 223 (2003) 17-25.

[43] K.K. Das, P. Somasundaran, A kinetic investigation of the flocculation of alumina with polyacrylic acid, Journal of Colloid and Interface Science 271 (2004) 102-109.

[44] J.T. Kunjappu, P. Somasundaran, K. Sivadasan, A Unique Conformational Equilibrium of Polyacrylic-Acid at the Solid-Liquid Interface, Colloid Surf. A-Physicochem. Eng. Asp. 97 (1995) 101-107.

[45] X. Yu, P. Somasundaran, Structure of sodium dodecyl sulfate and polyacrylic acid adsorption layer using nitroxide spin labeled alumina, Langmuir 16 (2000) 3506-3508.

[46] M.R. Alexander, S. Payan, T.M. Duc, Interfacial interactions of plasma-polymerized acrylic acid and an oxidized aluminium surface investigated using XPS, FTIR and poly(acrylic acid) as a model compound, Surf Interface Anal 26 (1998) 961-973.

[47] E.A. Bekturov, V.A. Frolova, G.K. Mamytbekov, Formation-destruction conditions of an interpolymer complex between a poly(acrylic acid) gel and linear poly(ethylene glycol) in methanol, Macromol. Chem. Phys. 201 (2000) 1031-1036.

[48] Z.X. Chen, Q. Gu, H.F. Zou, T.Q. Zhao, H.Y. Wang, Molecular dynamics simulation of water diffusion inside an amorphous polyacrylate latex film, J. Polym. Sci. Pt. B-Polym. Phys. 45 (2007) 884-891.

[49] K.D. Dobson, A.J. McQuillan, In situ infrared spectroscopic analysis of the adsorption of aliphatic carboxylic acids to TiO2, ZrO2, Al2O3, and Ta2O5 from aqueous solutions, Spectrochim Acta A 55 (1999) 1395-1405.

[50] J. Dong, Y. Ozaki, K. Nakashima, FTIR studies of conformational energies of poly(acrylic acid) in cast films, J. Polym. Sci. Pt. B-Polym. Phys. 35 (1997) 507-515.

[51] J. Dong, Y. Ozaki, K. Nakashima, Infrared, Raman, and near-infrared spectroscopic evidence for the coexistence of various hydrogen-bond forms in poly(acrylic acid), Macromolecules 30 (1997) 1111-1117.

[52] V. Doseva, L.L. Yasina, Aliev, II, A.M. Vasserman, V.Y. Baranovskii, Molecular dynamics and organization of micellar complexes of poly(acrylic acid) with poly(ethylene glycol)-based surfactants, Colloid J. 65 (2003) 562-566.

Page 124: © 2008 Joshua James Taylor

124

[53] L. Dupont, A. Foissy, Evaluation of the adsorption trends of a low molecular-weight polyelectrolyte with a site-binding model, Colloid Surf. A-Physicochem. Eng. Asp. 110 (1996) 235-248.

[54] A. Foissy, A.E. Attar, J.M. Lamarche, Adsorption of Polyacrylic-Acid on Titanium-Dioxide, Journal of Colloid and Interface Science 96 (1983) 275-287.

[55] C. Friedsam, A.D. Becares, U. Jonas, M. Seitz, H.E. Gaub, Adsorption of polyacrylic acid on self-assembled monolayers investigated by single-molecule force spectroscopy, New J Phys 6 (2004).

[56] F. Genin, F. Quiles, A. Burneau, Infrared and Raman spectroscopic study of carboxylic acids in heavy water, Phys Chem Chem Phys 3 (2001) 932-942.

[57] B. Hammouda, F. Horkay, M.L. Becker, Clustering and solvation in poly(acrylic acid) polyelectrolyte solutions, Macromolecules 38 (2005) 2019-2021.

[58] T. Hofmann, R.G. Winkler, P. Reineker, Influence of salt on the structure of polyelectrolyte solutions: An integral equation theory approach, J Chem Phys 119 (2003) 2406-2413.

[59] L.J. Kirwan, P.D. Fawell, W. van Bronswijk, In situ FTIR-ATR examination of poly(acrylic acid) adsorbed onto hematite at low pH, Langmuir 19 (2003) 5802-5807.

[60] S.C. Liufu, H.N. Xiao, Y.P. Li, Thermal analysis and degradation mechanism of polyacrylate/ZnO nanocomposites, Polym Degrad Stabil 87 (2005) 103-110.

[61] E. Mielczarski, J.A. Mielczarski, J.M. Cases, Molecular recognition effect in monolayer formation of oleate on fluorite, Langmuir 14 (1998) 1739-1747.

[62] J.A. Mielczarski, E. Mielczarski, Determination of Molecular-Orientation and Thickness of Self-Assembled Monolayers of Oleate on Apatite by Ftir Reflection Spectroscopy, J Phys Chem-Us 99 (1995) 3206-3217.

[63] M.A. Moharram, L.S. Balloomal, H.M. ElGendy, Infrared study of the complexation of poly(acrylic acid) with poly(acrylamide), J. Appl. Polym. Sci. 59 (1996) 987-990.

[64] G.A. Mun, Z.S. Nurkeeva, V.V. Khutoryanskiy, Complex formation between poly(vinyl ether) of ethylene glycol and poly(acrylic acid) in aqueous and organic solutions, Macromol. Chem. Phys. 200 (1999) 2136-2138.

[65] H.T. Oyama, W.T. Tang, C.W. Frank, Conformational Change in the Complex between Poly(Acrylic Acid) and Poly(Ethylene Glycol), Abstr. Pap. Am. Chem. Soc. 192 (1986) 165-POLY.

Page 125: © 2008 Joshua James Taylor

125

[66] H.T. Oyama, W.T. Tang, C.W. Frank, Complex-Formation between Poly(Acrylic Acid) and Pyrene-Labeled Poly(Ethylene Glycol) in Aqueous-Solution, Macromolecules 20 (1987) 474-480.

[67] H.T. Oyama, W.T. Tang, C.W. Frank, Complex-Formation between Poly(Acrylic Acid) and Poly(Ethylene Glycol) in Aqueous-Solution, Acs Symposium Series 358 (1987) 422-433.

[68] Z.H. Pan, A. Campbell, P. Somasundaran, Polyacrylic acid adsorption and conformation in concentrated alumina suspensions, Colloid Surf. A-Physicochem. Eng. Asp. 191 (2001) 71-78.

[69] E. Ringenbach, G. Chauveteau, E. Pefferkorn, Adsorption of Polyelectrolytes on Soluble Oxides Induced by Polyion Complexation with Dissolution Species, Journal of Colloid and Interface Science 161 (1993) 223-231.

[70] K.R. Rogan, The Variations of the Configurational and Solvency Properties of Low-Molecular-Weight Sodium Polyacrylate with Ionic-Strength, Colloid Polym Sci 273 (1995) 364-369.

[71] R. Schweins, J. Hollmann, K. Huber, Dilute solution behaviour of sodium polyacrylate chains in aqueous NaCl solutions, Polymer 44 (2003) 7131-7141.

[72] K. Vermohlen, H. Lewandowski, H.D. Narres, E. Koglin, Adsorption of polyacrylic acid on aluminium oxide: DRIFT spectroscopy and ab initio calculations, Colloid Surf. A-Physicochem. Eng. Asp. 170 (2000) 181-189.

[73] A.A. Zaman, S. Mathur, Influence of dispersing agents and solution conditions on the solubility of crude kaolin, Journal of Colloid and Interface Science 271 (2004) 124-130.

[74] J. Zhu, Z.Y. Ren, G.B. Zhang, X.Z. Guo, D.Z. Ma, Comparative study of the H-bond and FTIR spectra between 2,2-hydroxymethyl propionic acid and 2,2-hydroxymethyl butanoic acid, Spectrochim Acta A 63 (2006) 449-453.

[75] D.M. Byler, H.M. Farrell, Infrared Spectroscopic Evidence for Calcium-Ion Interaction with Carboxylate Groups of Casein, J Dairy Sci 72 (1989) 1719-1723.

[76] C.A. Young, J.D. Miller, Effect of temperature on oleate adsorption at a calcite surface: an FT-NIR/IRS study and review, Int J Miner Process 58 (2000) 331-350.

[77] U. Raviv, J. Klein, Fluidity of bound hydration layers, Science 297 (2002) 1540-1543.

[78] A.M. Kjeldsen, R.J. Flatt, L. Bergstrom, Relating the molecular structure of comb-type superplasticizers to the compression rheology of MgO suspensions, Cem. Concr. Res. 36 (2006) 1231-1239.

[79] H.B. Abderrahmen, H. Said, S. Partyka, R. Denoyel, Ester phosphate adsorption on calcium carbonate, J. Chim. Phys.-Chim. Biol. 94 (1997) 750-764.

Page 126: © 2008 Joshua James Taylor

126

[80] K. Backfolk, S. Lagerge, J.B. Rosenholm, D. Eklund, Aspects on the interaction between sodium carboxymethylcellulose and calcium carbonate and the relationship to specific site adsorption, Journal of Colloid and Interface Science 248 (2002) 5-12.

[81] M.S. Burgess, J.S. Phipps, Flocculation of PCC induced by polymer/microparticle systems: Floc characteristics, Nord. Pulp Paper Res. J. 15 (2000) 572-578.

[82] S. El-Sherbiny, H.N. Xiao, Effect of polymeric thickeners on pigment coatings: Adsorption, rheological behaviour and surface structures, J. Mater. Sci. 39 (2004) 4487-4493.

[83] S. Suty, V. Luzakova, Role of surface charge in deposition of filler particles onto pulp fibres, Colloid Surf. A-Physicochem. Eng. Asp. 139 (1998) 271-278.

[84] S. Suzuki, Black tea adsorption on calcium carbonate: A new application to chalk powder for brown powder materials, Colloid Surf. A-Physicochem. Eng. Asp. 202 (2002) 81-91.

[85] P. Balaz, A. Alacova, J. Briancin, Sensitivity of Freundlich equation constant 1/n for zinc sorption on changes induced in calcite by mechanical activation, Chem Eng J 114 (2005) 115-121.

[86] Z.S. Nickolov, O. Ozcan, J.D. Miller, FTIR analysis of water structure and its significance in the flotation of sodium carbonate and sodium bicarbonate salts, Colloid Surf. A-Physicochem. Eng. Asp. 224 (2003) 231-239.

[87] P. Schmidt, J. Dybal, J.C. Rodriguez-Cabello, V. Reboto, Role of water in structural changes of poly(AVGVP) and poly(GVGVP) studied by FTIR and Raman spectroscopy and ab initio calculations, Biomacromolecules 6 (2005) 697-706.

[88] P. Schmidt, J. Dybal, M. Trchova, Investigations of the hydrophobic and hydrophilic interactions in polymer-water systems by AIR FTIR and Raman spectroscopy, Vib Spectrosc 42 (2006) 278-283.

[89] H.A. Al-Hosney, S. Carlos-Cuellar, J. Baltrusaitis, V.H. Grassian, Heterogeneous uptake and reactivity of formic acid on calcium carbonate particles: a Knudsen cell reactor, FTIR and SEM study, Phys Chem Chem Phys 7 (2005) 3587-3595.

[90] H.A. Al-Hosney, V.H. Grassian, Water, sulfur dioxide and nitric acid adsorption on calcium carbonate: A transmission and ATR-FTIR study, Phys Chem Chem Phys 7 (2005) 1266-1276.

[91] J.R. Bargar, J.D. Kubicki, R. Reitmeyer, J.A. Davis, ATR-FTIR spectroscopic characterization of coexisting carbonate surface complexes on hematite, Geochim Cosmochim Ac 69 (2005) 1527-1542.

Page 127: © 2008 Joshua James Taylor

127

[92] W. Hage, A. Hallbrucker, E. Mayer, Carbonic-Acid - Synthesis by Protonation of Bicarbonate and Ftir Spectroscopic Characterization Via a New Cryogenic Technique, J Am Chem Soc 115 (1993) 8427-8431.

[93] J.M. Baker, J.C. Dore, J.M. Seddon, A.K. Soper, Structural modification of water in the confined layer of a lamellar liquid crystal, Chem Phys Lett 256 (1996) 649-652.

[94] M.C. Bellissent-Funel, Structure of confined water, J Phys-Condens Mat 13 (2001) 9165-9177.

[95] W.R. Bowen, H.N.S. Yousef, Effect of salts on water viscosity in narrow membrane pores, Journal of Colloid and Interface Science 264 (2003) 452-457.

[96] V. Buch, S. Bauerecker, J.P. Devlin, U. Buck, J.K. Kazimirski, Solid water clusters in the size range of tens-thousands of H2O: a combined computational/spectroscopic outlook, Int Rev Phys Chem 23 (2004) 375-433.

[97] J.M. Di Leo, J. Maranon, Confined water in nanotube, J Mol Struc-Theochem 623 (2003) 159-166.

[98] J.M. Di Leo, J. Maranon, Confined ions and water in nanotube, J Mol Struc-Theochem 709 (2004) 163-166.

[99] Y.L. Guo, H. Yui, H. Minamikawa, M. Masuda, S. Kamiya, T. Sawada, K. Ito, T. Shimizu, FT-IR study of the interlamellar water confined in glycolipid nanotube walls, Langmuir 21 (2005) 4610-4614.

[100] J. Israelachvili, H. Wennerstrom, Role of hydration and water structure in biological and colloidal interactions, Nature 379 (1996) 219-225.

[101] S. Kerisit, D.J. Cooke, D. Spagnoli, S.C. Parker, Molecular dynamics simulations of the interactions between water and inorganic solids, J Mater Chem 15 (2005) 1454-1462.

[102] J. Klein, U. Raviv, S. Perkin, N. Kampf, L. Chai, S. Giasson, Fluidity of water and of hydrated ions confined between solid surfaces to molecularly thin films, J Phys-Condens Mat 16 (2004) S5437-S5448.

[103] A.I. Kolesnikov, J.M. Zanotti, C.K. Loong, P. Thiyagarajan, A.P. Moravsky, R.O. Loutfy, C.J. Burnham, Anomalously soft dynamics of water in a nanotube: A revelation of nanoscale confinement, Phys Rev Lett 93 (2004).

[104] A.R. Layson, D. Teeters, Polymer electrolytes confined in nanopores: using water as a means to explore the interfacial impedance at the nanoscale, Solid State Ion. 175 (2004) 773-780.

Page 128: © 2008 Joshua James Taylor

128

[105] F. Mallamace, M. Broccio, C. Corsaro, A. Faraone, U. Wanderlingh, L. Liu, C.Y. Mou, S.H. Chen, The fragile-to-strong dynamic crossover transition in confined water: nuclear magnetic resonance results, J Chem Phys 124 (2006).

[106] A. Sediki, F. Lebsir, L. Martiny, M. Dauchez, A. Krallafa, Ab initio investigation of the topology and properties of three-dimensional clusters of water (H2O)(n), Food Chem 106 (2008) 1476-1484.

[107] D. Spagnoli, D.J. Cooke, S. Kerisit, S.C. Parker, Molecular dynamics simulations of the interaction between the surfaces of polar solids and aqueous solutions, J Mater Chem 16 (2006) 1997-2006.

[108] M. Tadokoro, S. Fukui, T. Kitajima, Y. Nagao, S. Ishimaru, H. Kitagawa, K. Isobe, K. Nakasuji, Structures and phase transition of multi-layered water nanotube confined to nanochannels, Chem Commun (2006) 1274-1276.

[109] M. Hancer, M.S. Celik, J.D. Miller, The significance of interfacial water structure in soluble salt flotation systems, J Colloid Interf Sci 235 (2001) 150-161.

[110] R. Mancinelli, A. Botti, F. Bruni, M.A. Ricci, A.K. Soper, Perturbation of water structure due to monovalent ions in solution, Phys Chem Chem Phys 9 (2007) 2959-2967.

[111] R. Mancinelli, A. Botti, F. Bruni, M.A. Ricci, A.K. Soper, Hydration of sodium, potassium, and chloride ions in solution and the concept of structure maker/breaker, J Phys Chem B 111 (2007) 13570-13577.

[112] Z.S. Nickolov, J.D. Miller, Water structure in aqueous solutions of alkali halide salts: FTIR spectroscopy of the OD stretching band, J Colloid Interf Sci 287 (2005) 572-580.

[113] O. Ozdemir, M.S. Celik, Z.S. Nickolov, J.D. Miller, Water structure and its influence on the flotation of carbonate and bicarbonate salts, J Colloid Interf Sci 314 (2007) 545-551.

[114] O. Ozdemir, S.I. Karakashev, A.V. Nguyen, J.D. Miller, Adsorption of carbonate and bicarbonate salts at the air-brine interface, Int J Miner Process 81 (2006) 149-158.

[115] A.K. Soper, K. Weckstrom, Ion solvation and water structure in potassium halide aqueous solutions, Biophys Chem 124 (2006) 180-191.

[116] H.J. Bakker, M.F. Kropman, A.W. Omta, Effect of ions on the structure and dynamics of liquid water, J Phys-Condens Mat 17 (2005) S3215-S3224.

[117] O.D. Bonner, R.K. Arisman, C.F. Jumper, Effect of Ions on Water Structure .2. Reinterpretation of Experimental-Data, Infrared Phys 14 (1974) 271-275.

[118] O.D. Bonner, C.F. Jumper, Effect of Ions on Water Structure, Infrared Phys 13 (1973) 233-242.

Page 129: © 2008 Joshua James Taylor

129

[119] R.D. Mountain, Solvation structure of ions in water, Int J Thermophys 28 (2007) 536-543.

[120] R.D. Mountain, D. Thirumalai, Alterations in water structure induced by guanidinium and sodium ions, J Phys Chem B 108 (2004) 19711-19716.

[121] O. Ozcan, M.S. Celik, Z.S. Nickolov, J.D. Miller, Effect of thermal stability on the flotation response of sodium carbonate salts, Miner Eng 16 (2003) 353-358.

[122] M. Heuberger, T. Drobek, N.D. Spencer, Interaction forces and morphology of a protein-resistant poly(ethylene glycol) layer, Biophys J 88 (2005) 495-504.

[123] Z.D. Nan, B.P. Liu, Z.C. Tan, Calorimetric investigation of excess molar heat capacities for water plus ethylene glycol from T=273.15 to T=373.15 K, J Chem Thermodyn 34 (2002) 915-926.

[124] T. Sato, H. Niwa, A. Chiba, R. Nozaki, Dynamical structure of oligo(ethylene glycol)s water solutions studied by time domain reflectometry, J Chem Phys 108 (1998) 4138-4147.

[125] T.M. Aminabhavi, B. Gopalakrishna, Density, Viscosity, Refractive-Index, and Speed of Sound in Aqueous Mixtures of N,N-Dimethylformamide, Dimethyl-Sulfoxide, N,N-Dimethylacetamide, Acetonitrile, Ethylene-Glycol, Diethylene Glycol, 1,4-Dioxane, Tetrahydrofuran, 2-Methoxyethanol, and 2-Ethoxyethanol at 298.15 K, J Chem Eng Data 40 (1995) 856-861.

[126] W. Hayduk, V.K. Malik, Density, Viscosity, and Carbon Dioxide Solubility and Diffusivity in Aqueous Ethylene Glycol Solutions, J Chem Eng Data 16 (1971) 143-&.

[127] S. Kirincic, C. Klofutar, Viscosity of aqueous solutions of poly(ethylene glycol)s at 298.15 K, Fluid Phase Equilibr 155 (1999) 311-325.

[128] T.F. Sun, A.S. Teja, Density, viscosity, and thermal conductivity of aqueous ethylene, diethylene, and triethylene glycol mixtures between 290 K and 450 K, J Chem Eng Data 48 (2003) 198-202.

[129] N.G. Tsierkezos, I.E. Molinou, Thermodynamic properties of water plus ethylene glycol at 283.15, 293.15, 303.15, and 313.15 K, J Chem Eng Data 43 (1998) 989-993.

[130] J. Yamanaka, H. Matsuoka, H. Kitano, N. Ise, T. Yamaguchi, S. Saeki, M. Tsubokawa, Revisit to the Intrinsic Viscosity-Molecular Weight Relationship of Ionic Polymers .3. Viscosity Behavior of Ionic Polymer Lattices in Ethylene-Glycol Water Mixtures, Langmuir 7 (1991) 1928-1934.

[131] T.F. Sun, A.S. Teja, Density, viscosity and thermal conductivity of aqueous solutions of propylene glycol, dipropylene glycol, and tripropylene glycol between 290 K and 460 K, J Chem Eng Data 49 (2004) 1311-1317.

Page 130: © 2008 Joshua James Taylor

130

BIOGRAPHICAL SKETCH

Joshua Taylor was born in 1980 on October 19 in Santa Barbara, California. He lived in

California for 10 years, Ohio for 6 years, and then finished high school in Florida. He graduated

from Riverview High School in Sarasota, Florida in June of 1999. He then went to the

University of Florida and graduated in 2004 with a Bachelor of Science in chemical engineering.

In 2005, he graduated with a Master of Science in materials science and engineering at the

University of Florida. Joshua Taylor’s life is dedicated to Yeshua HaMashiach. Out of love for

Yeshua, he lives a life that is within the standards of the Bible.