Vous êtes sur la page 1sur 73

DEVELOPMENT OF A FREE LIME MONITORING

SYSTEM FOR THE KRAFT RECOVERY PROCESS


USING ZETA POTENTIAL

by

Wei Ren

A thesis submitted in conformity with the requirements


for the degree of Master of Applied Science
Department of Chemical Engineering and Applied Chemistry
University of Toronto

Copyright by Wei Ren 2014

Development of a Free Lime Monitoring System for the Kraft


Recovery Process Using Zeta Potential
Wei Ren
Master of Applied Science
Department of Chemical Engineering and Applied Chemistry
University of Toronto
2014

ABSTRACT
The presence of Ca(OH)2 (or free lime) in lime mud can cause many problems in the recovery
process of kraft pulp mills. Conventional free lime analyses require extensive laboratory work
and give inconsistent results. A systematic study was performed to determine if zeta potential, a
quickly measurable dispersion characteristic, can be used to indicate the presence of free lime in
the recovery process. Measurements were made on synthetic wet lime mud to simulate samples
collected after the white liquor clarifier in pulp mills. The results show that zeta potential
increases from negative to positive when the free lime content in the lime mud exceeds a critical
threshold. This change from negative to positive zeta potential of the lime mud slurry can be
used as a basis for developing an on-line monitoring system that effectively detects free lime in
the lime mud in order to avoid problems associated with overliming in the recovery process.

ii

ACKNOWLEDGMENTS
I wish to express my utmost gratitude to my thesis supervisor Professor Honghi Tran for his
excellent advice and guidance throughout the course of this study. His enthusiasm, motivation
and encouragement were of extreme importance for the completion of my research.
I want to thank my committee members, Professor Edgar Acosta, Professor Donald Kirk and
Professor Will Cluett for their insightful comments and valuable pieces of advice.
I also want to thank Fariba Azgomi and Sue Mao for their valuable suggestions, and providing
me with essential literatures relevant to my research. My gratitude goes to Anna Ho for her help
with administrative activities. I am grateful to all my friends, especially Eric, Liming, Anton,
Michael, Masoumeh and Naz for their support and friendship. I also want to extend my
appreciation to Carolyn Winsborough for proofreading my thesis and other technical reports.
I wish to acknowledge members of the research program on Increasing Energy and Chemical
Recovery Efficiency in the Kraft Pulping Process for their discussions and advice during
consortium meetings and their financial support. I would like to thank Domtar Espanola Mill and
DMI Peace River Mill for giving me extensive information regarding kraft pulp mill operations
and providing samples used in my research.
I want to give my sincere gratitude to my parents Jiyi Ren and Hong Liu for their technical
advices during my study, but more importantly for giving me their unconditional support and
love my whole life.
Last but not least, I want to thank my fiance Nan Ma for all she has done for me over the past
ten years. Only with her inspiration, motivation, encouragement and love I am able to finish my
research. I want to dedicate this thesis to her.

iii

TABLE OF CONTENTS
ACKNOWLEDGMENTS ........................................................................................................... iii
TABLE OF CONTENTS ............................................................................................................ iv
LIST OF TABLES ....................................................................................................................... vi
LIST OF FIGURES .................................................................................................................... vii
1 INTRODUCTION .................................................................................................................... 1
1.1 General Background ........................................................................................................... 1
1.2 Causticizing Reaction and Free Lime ................................................................................. 3
1.3 Zeta Potential Overview ..................................................................................................... 5
1.4 Objectives ........................................................................................................................... 6
2 LITERATURE REVIEW ........................................................................................................ 7
2.1 Current Methods of Free Lime Analysis ............................................................................ 7
2.2 Zeta Potential () ................................................................................................................. 8
2.2.1

Zeta Potential Theory .............................................................................................. 8

2.2.2

Previous Studies and Applications........................................................................ 10

2.3 Sound Theories and Applications ..................................................................................... 13


2.4 Factors Affecting Free Lime Concentration ..................................................................... 14
2.4.1

Lime Properties ..................................................................................................... 14

2.4.2

Operating Conditions in Causticizing Plant .......................................................... 15

3 METHODOLOGY ................................................................................................................. 18
3.1 Material ............................................................................................................................. 18
3.2 Experimental Setup ........................................................................................................... 19
3.3 Experimental Procedures .................................................................................................. 21
3.3.1

Particle Size Determination .................................................................................. 21

3.3.2

Initial Experimental Procedure ............................................................................. 24


iv

3.3.3

Procedure Modification ........................................................................................ 25

3.3.4

Modified Experimental Procedure ........................................................................ 30

4 RESULTS AND DISCUSSION ............................................................................................ 32


4.1 Effect of Free Lime on Zeta Potential for Analytical Samples ......................................... 32
4.1.1

Analytical Grade Chemicals ................................................................................. 32

4.1.2

Analytical Grade Lime and Mill Green Liquor .................................................... 36

4.2 Effect of Free Lime on Zeta Potential for Mill Samples .................................................. 39
4.2.1

Effect of Lime Properties ...................................................................................... 43

4.2.2

Effect of Causticizing Reaction Time ................................................................... 45

4.2.3

Effect of Lime Mud Storage Time ........................................................................ 46

5 CONCLUSIONS .................................................................................................................... 50
6 IMPLICATIONS.................................................................................................................... 52
REFERENCES ............................................................................................................................ 54
APPENDICES ............................................................................................................................. 58
Appendix A Ammonium Chloride Test ................................................................................... 58
Appendix B Additional Sound Theories .................................................................................. 60
Appendix C ABC Titration (with Automatic Titrator) ............................................................ 61
Appendix D Thermogravimetric Analysis (TGA) / Differential Scanning Calorimetry
(DSC) ................................................................................................................................ 62
Appendix E Zeta Potential Data for Wet and Dry Lime Mud ................................................. 63
Appendix F Zeta Potential Data for Stored Sample ................................................................. 64

LIST OF TABLES
Table 2-1. Correlation between Zeta Potential and Dispersion System ....................................... 10
Table 2-2. Properties of Quicklime .............................................................................................. 14
Table 2-3. Factors Affecting Free Lime Concentration................................................................ 17
Table 3-1. Green Liquor Composition .......................................................................................... 19

vi

LIST OF FIGURES
Figure 1-1. Typical Kraft Recovery Process .................................................................................. 2
Figure 1-2. Overall Causticizing Plant Process ............................................................................ 2
Figure 1-3. Goodwins Curve - Effect of TTA on Equilibrium CE at Different Sulphidity [4] ...... 4
Figure 1-4. Acoustic and Electroacoustic Spectrometer ................................................................ 5
Figure 2-1. Double Layer Model for Zeta Potential ....................................................................... 9
Figure 2-2. Potential Charge as a Function of Distance from Particle Surface ......................... 10
Figure 2-3. Effect of Liming Ratio on Zeta Potential ................................................................... 11
Figure 2-4. Zeta Potential as a Function of Free Lime Content .................................................. 12
Figure 3-1. Zeta Potential Measurement Probe ........................................................................... 20
Figure 3-2. Particle Size Distribution for CaCO3......................................................................... 22
Figure 3-3. Particle Size Distribution for Ca(OH)2 (from Calcination of Lime Mud) ................. 22
Figure 3-4. Average Particle Sizes for Lime and Lime Mud Samples .......................................... 23
Figure 3-5. Average Zeta Potentials for Lime and Lime Mud Samples ........................................ 24
Figure 3-6. Effect of Liming Ratio on Zeta Potential at Different Causticizing Reaction
Temperatures (Pure Chemicals, 1 Hour Reaction) ...................................................................... 25
Figure 3-7. Effect of Measurement Temperature on Zeta Potential (Pure Chemicals, 1 Hour
Reaction, 90C Reaction Temperature) ........................................................................................ 26
Figure 3-8. Effect of Liming Ratio on Zeta Potential when Measuring White Liquor Slurries
Synthesized with Pure Chemicals and Mill Green Liquor Samples (1 Hour Reaction, 90C
Reaction Temperature) ................................................................................................................. 27

vii

Figure 3-9. Effect of Liming Ratio on Zeta Potential when Measuring White Liquor Slurries
Diluted with Distilled Water at Different Concentrations (Pure Chemicals, 1 Hour Reaction,
90C Reaction Temperature) ........................................................................................................ 27
Figure 3-10. Effect of Liming Ratio on Zeta Potential when Measuring Lime Mud Diluted with
Distilled Water at Different Concentrations (Pure Chemicals, 1 Hour Reaction, 90C Reaction
Temperature)................................................................................................................................. 28
Figure 3-11. Effect of Liming Ratio on Zeta Potential when Measuring Lime Mud Diluted with
Distilled Water at 1wt% (Mill Green Liquor, 1 Hour Reaction, 90C Reaction Temperature) ... 28
Figure 3-12. Sample Compartments for Zeta Potential Measurement ......................................... 29
Figure 3-13. Effect of Liming Ratio on Zeta Potential when Measuring 10mL of Lime Mud
Diluted in Distilled Water (Pure Chemicals, 1 Hour Reaction, 90C Reaction Temperature) .... 30
Figure 3-14. Experimental Procedure Flow Chart ....................................................................... 31
Figure 4-1. Measurement Repeatability with 10mL Samples (Pure Chemicals, 1 Hour Reaction,
90C Reaction Temperature) ........................................................................................................ 33
Figure 4-2. Weight Loss Profile for Lime Mud (0.6 Liming Ratio and Pure Chemicals) ............. 34
Figure 4-3. Weight Loss Profile for Lime Mud (1.2 Liming Ratio and Pure Chemicals) ............. 35
Figure 4-4. Zeta Potential and Free Lime as a Function of Liming Ratios (Pure Chemicals, 1
Hour Reaction, 90C Reaction Temperature) .............................................................................. 35
Figure 4-5. Zeta Potential as a Function of Free Lime Content (Pure Chemicals, 1 Hour
Reaction, 90C Reaction Temperature) ........................................................................................ 36
Figure 4-6. Zeta Potential and Free Lime as a Function of Liming Ratios (Mill Green Liquor, 1
Hour Reaction, 90C Reaction Temperature) .............................................................................. 37
Figure 4-7. Zeta Potential as a Function of Free Lime Content (Mill Green Liquor, 1 Hour
Reaction, 90C Reaction Temperature) ........................................................................................ 37

viii

Figure 4-8. Effect of Liming Ratio on Zeta Potential Dry and Wet Samples (Pure Lime, 1 Hour
Reaction, 90C Reaction Temperature) ........................................................................................ 38
Figure 4-9. Effect of Liming Ratio on Zeta Potential using Pure Lime and Reburned Lime (1
Hour Reaction, 90C Reaction Temperature) .............................................................................. 40
Figure 4-10. Effect of Liming Ratio on Zeta Potential Dry and Wet Samples (Reburned Lime, 1
Hour Reaction, 90C Reaction Temperature) .............................................................................. 42
Figure 4-11. Lime Reactivity Test Setup ....................................................................................... 43
Figure 4-12. Wet Slaking Curve for Powdered and Pelleted Reburned Lime .............................. 44
Figure 4-13. Effect of Liming Ratio on Zeta Potential Powdered and Pelleted Reburned Lime
(1 Hour Reaction, 90C Reaction Temperature) .......................................................................... 45
Figure 4-14. Effect of Reaction Time on Zeta Potential (Mill Green Liquor, 90C Reaction
Temperature)................................................................................................................................. 46
Figure 4-15. Effect of Storage Time on Zeta Potential (Mill Green Liquor, 1 Hour Reaction,
90C Reaction Temperature) ........................................................................................................ 46
Figure 4-16. Effect of Extended Storage Time on Zeta Potential (Mill Green Liquor, 1 Hour
Reaction, 90C Reaction Temperature) ........................................................................................ 48
Figure 4-17. Effect of Reaction Time on Zeta Potential 1 Month Storage (Mill Green Liquor,
90C Reaction Temperature) ........................................................................................................ 49
Figure 6-1. Potential Online Free Lime Monitoring System ........................................................ 53

ix

INTRODUCTION

1.1 General Background


The kraft process is the dominant pulping process, netting approximately 130 million tons/year
of pulp globally and accounting for over 90% of the worlds chemical pulp production [1]. The
main advantages of kraft pulping are its ability to handle almost all softwood and hardwood
species, the high strength and quality of kraft pulp and its high chemical efficiency (about 97%).
In this process, wood chips (which consist of fibre and lignin) are treated in a digester at an
elevated temperature and pressure in an aqueous solution consisting mainly of sodium hydroxide
(NaOH) and sodium sulphide (Na2S), known as white liquor. The fibre in the wood chips is
separated and washed to make the pulp. The residue, weak black liquor (including lignin, spent
chemicals and water), is sent to the kraft recovery system where the organic chemicals are used
as fuels to provide steam and power for the pulp mill while the inorganic chemicals are recycled
back into the pulping process.
A schematic of the kraft recovery process is shown in Figure 1-1. Weak black liquor is
concentrated through a series of evaporators to increase its solids content. The resulting heavy
(concentrated) black liquor is burned in the recovery boiler to produce steam and power. The
inorganic portion of the black liquor forms molten smelt, consisting mainly of sodium carbonate
(Na2CO3) and Na2S. The smelt is dissolved in water in the dissolving tank to produce green
liquor, which is subsequently causticized in the causticizing plant with calcium hydroxide
(Ca(OH)2) to produce NaOH according to the slaking and causticizing reaction.
Slaking:

CaO + H2O Ca(OH)2

(1)

Causticizing:

Na2CO3 + Ca(OH)2 2NaOH + CaCO3

(2)

The resulting slurry which consist of NaOH, Na2S, unreacted Na2CO3, unreacted Ca(OH)2 and
lime mud (CaCO3) is sent through a clarifier to filter out the precipitates (lime mud). The
recovered white liquor consisting mainly of NaOH and Na2S is recycled back to the digester to
be reused in the pulping process while the washed water (weak wash) is recycled to the
dissolving tank for dissolving smelt.

Figure 1-1. Typical Kraft Recovery Process [1]


The lime mud slurry is dewatered and fed into a lime kiln where it is calcined to produce
reburned lime, which is reused in the slaker. The calcination reaction is endothermic and occurs
at temperatures over 800C. The conventional causticizing process from the dissolving tank to
the lime kiln can be seen in Figure 1-2.

Figure 1-2. Overall Causticizing Plant Process [2]


Slaker and causticizer operations are instrumental for maintaining high process efficiency and
minimizing problems in the causticizing plant. Hence it is important to understand how the
causticizing reaction proceeds and the problems associated with free lime.

1.2 Causticizing Reaction and Free Lime


In this study, free lime is defined as unreacted calcium hydroxide (Ca(OH) 2) in the lime mud. In
addition, the causticizing efficiency (CE) is defined as the percentage of NaOH concentration in
the white liquor [3] ([NaOH] and [Na2CO3] are concentrations expressed in g/L Na2O):

Causticizing Efficiency (%) =

100

(3)

The causticizing reaction has a slow reaction rate and does not go to completion but reaches
equilibrium. The slaker and several causticizers are used in series to maximize the retention time
and reaction efficiency. Theoretically, the equilibrium or maximum attainable CE is the same as
the maximum CaO/Na2CO3 ratio (liming ratio). When targeting an equilibrium CE of 80%, for
example, the amount of lime added into the slaker should be adjusted so that the liming ratio is
equal to 0.80. Exceeding this ratio would cause the system to contain free lime; and the system is
said to be overlimed. Similarly, when an insufficient amount of lime is added into the slaker,
causing the liming ratio to be much lower than 0.80, the system is said to be underlimed.
The theoretical CE is dependent on the total titratable alkali (TTA) and the sulphidity of the
white liquor as defined in Equation 4 and Equation 5, respectively.
Total Titratable Alkali = [NaOH] + [Na2S] + [Na2CO3]

(4)

Sulphidity =

(5)

100

The Na2S concentration ([Na2S]) is expressed in g/L Na2O. The relationship between CE, TTA
and sulphidity is represented in the Goodwin curve, which is used to describe the highest
achievable CE at a given condition (Figure 1-3). In the causticizing plant, a CE closely equal to
the theoretical value is desired as underlimed systems containing unreacted Na2CO3 or dead load
in recovered white liquor reduces equipment efficiencies in downstream processes. Furthermore,
additional make-up white liquor is required at the digester as NaOH yield decreases. On the other
hand, at any given TTA, there will be an upper CE limit to avoid overlimed systems. Overliming
can also cause problems including poor mud settling rate, low mud solids content, high soda
content, poor kiln thermal performance, high TRS emissions and ring formations [5].

Causticizing Efficiency (%)

White Liquor TTA (g/L)

Figure 1-3. Goodwins Curve - Effect of TTA on Equilibrium CE at Different Sulphidity [4]
Therefore, the ability to maintain an optimum CE is crucial to operations in the causticizing
plant. Conventionally, the amount of lime fed into the slaker is adjusted to achieve the targeted
CE at a certain causticizing stage, such as at one of the causticizers. This means CE is used as the
indicator for detecting overlimed systems. Even though the targeted CE at mills is usually lower
than the equilibrium CE under normal operating conditions to avoid overliming, if the lime
reactivity is low or the residence time for the causticizing reaction is too short, free lime may
exist in the white liquor and subsequently in the lime mud.
Lime addition at the slaker is based on CE, which is calculated based on TTA and sulphidity of
the liquor sample collected at sampling ports in the causticizing plant (e.g. one of the
causticizers) [6]. The conventional method used to measure liquor composition is based on
titration, commonly referred to as the ABC test. It is a standard method with TAPPI (Technical
Association of the Pulp and Paper Industry) [7] which provides accurate results but involves
lengthy analytical procedures. However, the ABC test gives inconsistent measurements when the
test is performed by different operators. On the other hand, free lime is not measured regularly,
unless problems associated with overliming occur frequently.
Contrary to the conventional titration method, this study focuses on developing an online
monitoring tool for the kraft recovery process that gives fast indication on the presence of free

lime, which can be used to adjust the liming ratio at the slaker. Based on a recent study, zeta
potential () a variable that is used to characterize stability of dispersion systems is found to
be suitable for this purpose. Possible correlation was found to exist between zeta potential and
free lime based on Azgomis research [8]. A systematic study is carried out to verify the
relationship between zeta potential and free lime using an acoustic and electroacoustic
spectrometer (Figure 1-4).

Figure 1-4. Acoustic and Electroacoustic Spectrometer [9]

1.3 Zeta Potential Overview


Zeta potential is a characterization of colloidal surface charge in solutions. The electrostatic
interaction forces between particles vary depending on the dispersion composition. The stability
of the colloidal system can be determined based on the magnitude of zeta potential. Particles
with higher absolute values of zeta potential will repel each other (stable system) while particles
with lower absolute values will cause agglomerations (unstable system) [10]. The settling rate of
lime mud particles in white liquor can be affected by free lime content, where an increase in free
lime content causes a decrease in settling rate [8]. Slow settling indicates a stable system and

should theoretically have a higher zeta potential value. Hence, a positive correlation should exist
between zeta potential and free lime content when other conditions are unchanged.
The acoustic and electroacoustic spectrometer based on sound theory has not been used
previously for measurements of zeta potential of liquor slurries as an indication of free lime in
the kraft recovery process. However, due to its capability to measure dispersion systems at
higher concentration, volume and temperature compared to the conventional light scattering
techniques, it presents more opportunities to be incorporated as an online monitoring system in
the causticizing plant.

1.4 Objectives
The objective of this work is to establish a well-defined relationship between free lime and zeta
potential. This study will serve as a foundation for the implementation of an online monitoring
system in the kraft recovery process. The work consists of the following stages:

Validating previous findings of correlation between zeta potential and liming ratio;

Comparing zeta potential measurements of liquor samples produced using analytical


chemicals and using samples collected at pulp mills;

Analyzing the effects of different parameters (e.g. reaction time) on zeta potential.

This thesis contains six chapters. Chapter 1 gave a summary of the motivations and objectives
for this study. Chapter 2 is a literature review introducing current methods of free lime analysis
and their shortcomings; the advantages of using zeta potential to measure free lime; the validity
and application of acoustic and electroacoustic principles when making these measurements; and
the effect of lime properties and causticizing plant operating conditions on the presence of free
lime. Chapter 3 discusses the methodology used in this study including materials and instruments
as well as establishing a procedure for efficient measurements using the spectrometer. Chapter 4
presents and discusses the results. Finally, Chapter 5 and Chapter 6 summarize the key findings
and indicate the implications of this study.

LITERATURE REVIEW

2.1 Current Methods of Free Lime Analysis


The free lime content in lime mud is often only analyzed when overliming problems occur in the
causticizing plant. There are no standard free lime testing method currently adapted by TAPPI
and PAPTAC (Pulp and Paper Technical Association of Canada). Two conventional free lime
measurement methods used in kraft pulp mills are the ammonium chloride (NH4Cl) method and
the muffle furnace method.
The exact procedure of the ammonium chloride method can be found in Appendix A. Even
though the procedure varies from mill to mill, the fundamental concept remains the same. The
origin of this analytical method was found to be based on ASTM Standard C-25 [11] and ASTM
Standard C-114 [12], which are standard methods used in the limestone and cement industries.
However, the transition to the pulp and paper industry remains unclear. The principle of the
NH4Cl analysis is to dissolve white liquor in a weak acid solution, where the free lime (Ca(OH)2)
will form soluble calcium chloride (CaCl2). CaCl2 is used in a series of reactions until titration is
performed at the end with potassium permanganate (KMnO4).
Previous studies have indicated that the NH4Cl analysis might not be able to reliably measure the
free lime content [7] since the reaction time between Ca(OH)2 and NH4Cl took much longer to
reach completion than the suggested reaction time in the procedures (30 minutes). In addition,
the residual sodium carbonate (Na2CO3) in the sample could neutralize the weak acid solution
required to produce CaCl2 which disrupts the overall reaction.
Another method for free lime determination is based on sample weight loss due to chemical
decomposition at different temperatures, commonly used in thermogravimetric analysis (TGA).
Since Ca(OH)2 decomposes at 380C, a muffle furnace is used to heat up the sample and the
weight is recorded at 300C and 500C [13]. The difference in weight between these two
temperatures is used to determine the amount of free lime in the sample.
A problem with using this method is that other chemicals with decomposition temperatures in the
range between 300C and 500C would contribute to the total amount of free lime. For example,
magnesium hydroxide (Mg(OH)2) decomposes at approximately 350C. Thus, the amount of free

lime calculated using the muffle furnace analysis would often be higher than the actual amount,
as the total weight loss is also caused by the decomposition of Mg(OH)2.
Measuring the free lime content using either the NH4Cl method or the thermal decomposition
method has some disadvantages that make them unsuitable for integration into an online
monitoring system. The most important disadvantage is that the required laboratory work is
extensive (e.g. long reaction or heating time). There is a need for a new method that can be used
to accurately and reliably determine the free lime content. Zeta potential is a characteristic of
slurry systems that is affected by the free lime content and can be measured efficiently.

2.2 Zeta Potential ()


Many theories were developed to explain interactions in aqueous systems between particles and
the surrounding solution. Zeta potential is mainly based on the magnitude of electrical potential
difference that exists between a particles surface and the continuous medium. The potential is
affected by chemical composition and concentration in the dispersion, such as the amount of free
lime content. Previous studies were able to identify how some chemicals (e.g. Na2CO3) that exist
in the slurry systems will affect the zeta potential.

2.2.1

Zeta Potential Theory

Zeta potential is a measurement of the colloidal surface charge or the electrochemical


equilibrium on particle interfaces in an aqueous solution. It depends on the properties of the
particle surface and the solution. The electrostatic interaction forces among particles have an
important impact on the properties of the solution. The stability of the colloidal system can be
determined based on the magnitude of zeta potential. In general, dispersion systems with higher
absolute zeta potential values will cause particles to repel one another while systems with lower
absolute zeta potential values will lead to particle agglomeration resulting in faster settling [10].
Zeta potential is part of the DLVO theory, named after Derjaguin, Landau, Verwey and
Overbeek [14, 15]. It is used to quantitatively explain the collection of fine dispersions in an
aqueous solution and the interactions between charged particles in a liquid medium. DLVO
theory is based on both the effect of van der Waals and double layer forces [16]. Van der Waals
force is the sum of attractive and repulsive forces between molecules excluding covalent bonds,
hydrogen bonds and electrostatic interactions. The double layer force is based on electrostatic

charges accumulated at the surface of the particle when it is in a dispersed phase [17]. Zeta
potential is used to characterize the double layer force.
Zeta potential can be explained using the distribution of electrical charge in the double layer as
shown in Figure 2-1. Charges build up at the surface of the colloid, forming an electrostatic field
that affects other ions in the solution. Two forces are experienced by the negatively charged
colloid: the attraction forces from the oppositely charged ions (counter-ions) and the repulsion
forces from the same charged ions (co-ions) in the solution [17].

+
-

+ +

- +

+
- +

Slipping Plane
+ + + + + +
Stern Layer
+
+
+
+
+
Colloid Surface
+
+
+ - +
+
+
- + +
+
+
Diffuse Layer
+
+
+
+
+
+
+
+
- + + + +
+
Figure 2-1. Double Layer Model for Zeta Potential
When counter-ions are absorbed to the surface of the colloid, an inner layer is formed called the
Stern layer. The counter-ions on the outer surface are also attracted to the colloid but encounter
the repulsive force from the Stern layer. This dynamic equilibrium results in the formation of the
diffuse layer. The slipping plane is the imaginary boundary that sets apart the stationary layer
of fluid attached to the colloid and the free flowing dispersion medium [10].
The potential difference between the particle surface and the continuous phase is a function of
the distance from the particle surface (Figure 2-2). The absolute value of the potential charge
decreases linearly within the Stern layer and then exponentially within the diffuse layer, where it

10

approaches zero at the imaginary boundary of the double layer. Theoretically, zeta potential is
the potential difference at the slipping or shear plane.

Figure 2-2. Potential Charge as a Function of Distance from Particle Surface [18]
The type of ions in the double layer structure has a strong correlation to the sign and magnitude
of zeta potential. Table 2-1 summarizes the correlation between typical values of zeta potential
and the dispersion system [19].
Table 2-1. Correlation between Zeta Potential and Dispersion System
Zeta Potential (mV)
0 to 10
from 10 to 50
greater than 50

Dispersion System
Unstable, rapid agglomeration and fastest settling rates
Moderate stability, decreased agglomeration and slower settling rates
Excellent stability, minimum agglomeration and slowest settling rates

Azgomis research indicated a decrease in settling rate of precipitated lime mud and an increase
in zeta potential when liming ratio exceeds a critical level [8].

2.2.2

Previous Studies and Applications

Since lime mud and white liquor slurries contain mainly CaCO3, Ca(OH)2, Na2CO3, NaOH and
Na2S, a literature review was conducted to summarize previous findings of correlation between
zeta potential and these chemicals. It was found that the results varied greatly between different
research groups. The main discrepancy was on the potential-determining ions (PDI) or chemicals
that had the most effect on the zeta potential of the system. Some research groups suggested that

11

in pure aqueous suspensions, Ca2+ and CO32- were acting as the PDI [20, 21]. Other group
suggested that OH-, HCO3- and H+ were acting as the PDI [22, 23]. Aside from these chemicals,
other impurities that might exist in the slurry system could also change the surface charge based
on the intensity of their concentration [24].
In Azgomis study, zeta potential values were measured using a micro-electrophoretic apparatus
based on laser light-scattering theories [8]. White liquor was prepared using both pure lime and
reburned limes from various pulp mills and pure Na2CO3 as a substitute for green liquor.
Measurements were made on 0.1g of lime mud (separated from the white liquor) dissolved in
500mL of distilled water. At liming ratios below 1.0, zeta potentials were relatively constant and
negative between -15mV to -10mV. However, when liming ratios exceeded 1.0, zeta potentials
became positive and reached a value above 20mV (Figure 2-3). In addition, free lime (Ca(OH)2)
concentration was found to be a function of liming ratio using the thermal decomposition
method. More specifically, free lime content increased as liming ratios increased. When zeta
potential was plotted against the free lime content, a linear relationship could be observed, as
shown in Figure 2-4. Another study also demonstrated that increasing Ca2+ ions in suspension
increased the zeta potential values until Ca2+ ions saturated the particle surface [25].

Zeta Potential (mV)

60
Lime A
Lime C
Pure Lime

40

Lime B
Lime D

20

-20

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Liming Ratio ([CaO]/[Na2CO3])


Figure 2-3. Effect of Liming Ratio on Zeta Potential [8]

1.4

12

Zeta Potential (mV)

60
Lime A
Lime C
Pure Lime

40

Lime B
Lime D

20

-20

10

15

20

25

Free Lime (%)


Figure 2-4. Zeta Potential as a Function of Free Lime Content [8]
As discussed in the previous section, due to the disadvantages associated with current methods
for measuring free lime content, the possibility for their implementation as an online monitoring
system in the causticizing plant is limited. Based on Azgomis research, measuring the zeta
potential of dispersion systems and correlating the results to the amount of free lime in the
system were much faster than the traditional laboratory approach. The results were also
consistent within the scope of her study. This provides great motivation for further investigation
into the relationship between free lime and zeta potential, as well as the exploration of more
efficient control strategies in the causticizing plant. There are several aspects of Azgomis work
that require additional research and understanding:
-

White liquor used for zeta potential measurement should be synthesized with green liquor
samples obtained from kraft pulp mills instead of Na2CO3 solution; and

Zeta potential measurement should be made on more concentrated lime mud solutions to
simulate normal conditions in causticizing plants.

In order to measure zeta potential of slurry samples at higher concentrations and with higher
degrees of inconsistency in chemical composition due to the presence of impurities in actual
green liquor, a spectrometer based on sound theories is used which differs from the laser
diffraction instrument used in Azgomis study that imposed the above restrictions.

13

2.3 Sound Theories and Applications


The spectrometer for particle size and zeta potential measurement used in this study relies on the
concept of acoustic and electroacoustic theories, respectively. A basic understanding of these
theories will give insights into the reasoning and advantages behind the use of sound technology
for zeta potential measurements as oppose to the more conventional light scattering instruments.
A more detailed explanation of these theories can be found in Appendix B.
The acoustic theory for dispersion systems is based on a combined relationship between different
acoustic properties (attenuation, sound speed and angle of sound scattering) and system
properties (composition, particle size distribution and surface charge) [26].
The main acoustic interaction within a colloidal system can be divided into two groups:
attenuation caused by either absorption (conversion of acoustic energy into thermal energy) or
scattering (re-direction of acoustic energy). The absorption of ultrasound for dispersion systems
can be calculated based on density using acoustic measurement systems [26]. On the other hand,
scattering of ultrasound is harder to measure due to its nonlinear effects. Measuring both
absorption and scattering mechanisms are difficult for actual dispersion systems due to the
necessity of an extended set of input parameters, which are often unavailable. For the
spectrometer used in this study, only the absorption effect intensity of the incident sound beam
after transmission is measured, which is not affected by the non-linear scattering effects [27].
In addition to acoustic theories, electroacoustic mechanisms are utilized by this spectrometer in
making zeta potential measurements. Interactions between sound beams and electric fields in
colloidal systems result in electroacoustic effects. In the presence of sound waves, anions and
cations with different mass or friction coefficients will generate an alternating electric potential
within the solution [28]. It is often referred to as Ion Vibration Potential (IVP) and
measurement techniques based on electroacoustic theories are developed for porous bodies [29],
colloidal systems [30, 31] and colloid vibration potential [32, 33].
Since zeta potential cannot be directly measured, it is calculated from other experimentally
measurable characteristics of the system such as the IVP. The fundamental capability for all zeta
potential analyzers is that they require a means of moving liquid relative to the particle surface
that causes disturbance within the double layer; and a means of monitoring and recording the

14

generated signals. Motion induction could be accomplished by an external electric field or a


mechanical pressure field while monitoring could be accomplished by using electrical,
mechanical or optical technologies. The spectrometer used in this study accomplishes both
motion induction and signal monitoring through electrical means.
In addition to understanding why a spectrometer based on sound theories is suitable for zeta
potential measurements of concentrated slurry samples, it is also important to acknowledge the
factors that affect the amount of free lime in the slurry. The correlation between zeta potential
and free lime can then be used to monitor and/or control these factors to minimize free lime
problems in the causticizing plant.

2.4 Factors Affecting Free Lime Concentration


Zeta potential is affected by the type of ions and their concentration in the dispersion. Many
factors can affect the free lime concentration during the chemical recovery process including
properties of lime used during the causticizing reaction and the operating conditions in the
causticizing plant.

2.4.1

Lime Properties

In the lime industry, calcium oxide (CaO) is referred to as quicklime, while calcium hydroxide
(Ca(OH)2) is referred to as hydrated lime or slaked lime. Lime with high magnesium content is
termed dolomitic lime.

Pure quicklime is produced by calcination of natural limestone (CaCO3) which is similar to


calcining lime mud to produce reburned lime in the causticizing plant. During the calcination
process, carbon dioxide is released from CaCO3 producing CaO with a porous structure. This
allows water to penetrate the CaO particles during the slaking reaction. The properties of the
main calcium-containing compounds in the kraft recovery process are shown in Table 2-2.
Table 2-2. Properties of Quicklime [34]
Calcium Oxide
Molecular Weight (g/mol)
Specific Gravity
Bulk Density (g/cm3)
Specific Energy (kJ/kg)

56.1
3.2 3.4
0.88 0.96
0.44

Calcium
Hydroxide
74.1
2.3 2.4
0.40 0.56
0.67

Calcium
Carbonate
100
2.3 2.7
0.96 1.1
0.92

15

Olsen and Direnga [35] reported that the properties of lime (CaO) and lime mud (CaCO3)
influenced the properties of white liquor and the overall efficiency in the causticizing plant. Lime
kiln produces reburned lime by calcining lime mud at high temperatures. The quality of lime is
often based on its residual calcium carbonate (CaCO3) content, reactivity and availability
(fraction of CaO in the reburned lime available for reaction). Slaking and causticizing reactions
at the slaker mainly use reburned lime fed from the lime kiln, and only use fresh lime for makeup
due to losses during the causticizing cycle. Therefore when the quality of the reburned lime is
poor, the reactivity of lime will also be lower which reduces the causticizing efficiency. This will
cause the produced white liquor slurry to contain an excess amount of free lime even when the
liming ratio is lower than the equilibrium ratio.
The particle size of lime could adversely affect the kinetics of the causticizing reaction [36].
Ideal reburned lime should consist of soft pebbles with a diameter of approximately 2cm [37].
Pure lime was found to be much more reactive than reburned lime due to having lime availability
of nearly 100% and a large specific surface area of approximately 2m2/g [37]. On the other hand,
reburned lime had a lime availability in the range of 87% 92% and a specific surface area of
less than 0.5m2/g [37].
The overall causticizing efficiency is often found to be correlated to lime reactivity [38, 39]. Due
to conditions in the causticizing plant, the residence time of lime in the slaker is strictly
controlled and a highly reactive lime shall slake within 5 minutes upon contact with water [37].
Many studies concluded that the rate of the causticizing reaction is affected by the properties of
the lime [40, 41]. Therefore, in addition to using a high liming ratio, the presence of free lime in
the causticizing plant may be caused by low lime quality due to non-optimized operating
conditions at the lime kiln.

2.4.2

Operating Conditions in Causticizing Plant

Due to the importance of reburned lime properties and its continued use in the lime cycle, the
calcination process needs to be controlled carefully. At low calcining temperatures, lime
structure was found to consist of microscopic particles with a high specific surface area (high
porosity) and are termed soft-burned limes. When temperature and/or duration of the
calcinations were increased, the particles grew and lime porosity decreased, lowering the specific
surface area and reactivity of the lime [42]. These limes are termed hard-burned or sintered

16

lime. It was found that when the calcination temperature exceeded 1100C, a decrease in
causticizing efficiency occurred due to the use of sintered reburned lime at the slaker [43]. When
testing and comparing pure lime and reburned lime collected from 10 Canadian kraft pulp mills,
Dorris [42] found that lime sintering in the kiln contributed to the poor reactivity of reburned
lime. Long residence time (i.e. greater than 1 hour) in the lime kiln also causes sintering of
calcium oxide particles while short residence time results in incomplete calcination process [44].
Chemical content in the slaking water also affects the lime property. Potgieter [45] found that
slaking water containing a chloride content higher than 250mg/L enhances the solubility of lime
which leads to an increase in lime reactivity. On the other hand, slaking water containing
carbonate and sulphate contents higher than 250mg/L reduces slaking rate as CaCO3 and CaSO4
can partially or completely coat the lime particles [45]. Carbonate content has a more severe
impact on hydration rates as it is less soluble than CaSO4.
Causticizing efficiency (CE), as discussed in Section 1.2, is dependent on the amount of CaO
added in the slaker to convert Na2CO3 in the green liquor to NaOH. The ratio between CaO and
Na2CO3 is also known as the liming ratio. When trying to maximize the CE at the slaker and the
causticizers, using a liming ratio equal or higher than the stoichiometric liming ratio can induce a
higher conversion rate of NaCO3 to NaOH in a shorter period of time [43, 46]. However, due to
the adverse effect of free lime content in the causticizing cycle, targeted CE in the causticizing
plant is often 3% to 11% below the equilibrium CE [47]. In order to maintain the optimum CE,
liming ratio should be adjusted as soon as possible after changes occur in green liquor
composition (i.e. changes in Na2CO3 concentration) [48].
Non-process elements (NPE) in pulp mills are defined as elements that do not have any function
during the production of pulp or during the chemical and energy recovery process. NPE presents
challenges when monitoring and minimizing the free lime content in the causticizing plant.
Typical NPE in the causticizing cycle can be characterized by their accumulation tendencies
[49]. Some NPE, such as calcium (Ca), magnesium (Mg) and manganese (Mn) could be easily
removed from the system due to their low-solubility in alkaline solution. Other NPE were found
to have a higher tendency to remain in the system such as iron (Fe), aluminum (Al) and silicon
(Si) [49]. Potassium (K) and chloride (Cl) were found to be the most difficult to remove due to

17

their high solubility. The variability in NPE concentration in liquor systems was found to be
related to the difference in chemicals used in kraft pulp mills [50].
Among other NPE, Mg can accumulate in the causticizing cycle. Due to its similarity to CaCO3,
magnesium carbonate (MgCO3) can be calcined to magnesium oxide (MgO) in the lime kiln at
temperatures lower than lime calcination [51]. Excess MgCO3 can cause lime to become hardburned at normal operating conditions [52]. In addition, MgO is slaked by water to magnesium
hydroxide (Mg(OH)2), similar to Ca(OH)2, but it does not react with Na2CO3 in the green liquor.
High Mg content can cause increased grit production [45]. This is a typical example of how NPE
can be carried and recycled through the causticizing plant, acting as dead weight in various
chemical processes. More importantly, NPE can alter the value of zeta potential non-linearly, due
to changes in particle surface potential caused by polyvalent ions within the dispersion system.
Even though the majority of sodium content is removed during the white liquor clarifier, an
unusual form of sodium termed guarded sodium is present in the lime mud which cannot be
reduced by washing [53]. Guarded sodium will always form in CaCO3 and is not significantly
affected by the liming ratio. Since zeta potential is highly correlated to pH, an excess amount of
guarded sodium can reduce the effect of free lime on zeta potential. Table 2-3 summarizes how
lime quality or the slaking reaction is affected by different causticizing plant operating
conditions.
Table 2-3. Factors Affecting Free Lime Concentration
Operating Condition
Kiln temperature > 1100C
Kiln residence time > 1 hour
Kiln magnesium content increases
Slaking water [Cl] increases
Slaking water [SO4] and [CO3] increase

Effect on Lime Quality or Slaking Reaction


Production of sintered lime
Production of sintered lime
Production of sintered lime
Slaking rate increases
Slaking rate decreases

18

METHODOLOGY

The relationship between liming ratio and zeta potential was studied using a combination of
materials including analytical chemicals and samples collected from kraft pulp mills. White
liquor was synthesized in the laboratory using the following combinations.
1) Analytical grade sodium carbonate (Na2CO3) and calcium oxide (CaO);
2) Mill green liquor (mainly Na2CO3 and Na2S) and analytical grade CaO;
3) Mill green liquor and reburned lime (mainly CaO).
The zeta potential of white liquor and lime mud produced using different liming ratios and other
variable experimental or measurement conditions (e.g. sample concentration) were measured
using the acoustic and electroacoustic spectrometer, model number DT-1202 from Dispersion
Technology. The instrument is referred to as DT-1202 in this thesis. Thermogravimetric analysis
(TGA) was used to measure the free lime content of white liquor or lime mud. The experimental
parameters and procedures were initially based on Azgomis research [8], and then modified to
match the specific requirements of the DT-1202 and improve the overall efficiency of the
method for free lime determination.

3.1 Material
Analytical grade chemicals with a purity of >99.9% were used. For mill green liquor samples,
the standard TAPPI ABC test discussed in Section 1.2 was used to determine their properties. In
addition to TTA, effective alkali (EA) and active alkali (AA), defined below respectively in
Equation 6 and Equation 7 ([NaOH] and [Na2S] are expressed in g/L Na2O), are used to calculate
the Na2CO3, NaOH and Na2S concentration in the green liquor samples. Detailed procedures for
the ABC test can be found in Appendix C. The result of the ABC test is shown in Table 3-1.
Effective Alkali = [NaOH] + [Na2S]

(6)

Active Alkali = [NaOH] + [Na2S]

(7)

Other materials used in this study included chemicals used for calibrating DT-1202s zeta
potential measurement probe. Since the DT-1202 makes measurements by inducing electric

19

current through an electrode, instrument coating and aging would require frequent calibration
under extensive use. This was done daily to ensure the instrument gave accurate results within
the instruments limits. The calibration solution for zeta potential is 10wt% silica in 0.01M
potassium chloride (KCl) which gives a standard zeta potential value of -38.0mV. This solution
is obtained by diluting the 50wt% colloidal silica provided by Dispersion Technology with
0.01M KCl solution prepared using analytical chemicals.
Table 3-1. Green Liquor Composition
Effective Alkali (EA)
Active Alkali (AA)
Total Titratable Alkali (TTA)

g/L Na2O
29.1
47.7
121

3.2 Experimental Setup


Zeta potential measurements were measured with the DT-1202. The instrument was also used to
determine the particle size for CaO and CaCO3, which was used as initial input parameters for
calculating zeta potential. The DT-1202 determines zeta potential of dispersions based on
measurements of colloid vibration current (CVI) [54 56]. In addition, the DT-1202 uses the
macroscopic fitting method for particle size distribution (PSD) measurements, similar to light
scattering, x-ray scattering and neutron scattering [57].
Particle size measurements are mainly based on sample weight. Discussed in Section 2.3, the
attenuation of ultrasound caused by individual particles is proportional to the particles weight or
volume. The DT-1202 and the acoustic spectroscopy technique are conventionally used by other
research groups for development and quality control in a wide range of industries such as cement
and ceramics industries [58, 59]. The sampling chamber in the DT-1202 has an ultrasound
generation end and a receiving end. At the generation end, a probe sends ultrasound at different
frequencies while the probe at the receiving end catches the incident beam at various positions.
The combination of frequencies and measurement positions are used to develop an experimental
PSD model, which is fitted to either a unimodal (single peak) or a bimodal (double peak) curve.
The theoretical model that matches the experimental curve with minimum deviation was used to
calculate the particle size of the sample.

20

The DT-1202 uses electroacoustic technology to measure zeta potential. The piezo-crystal at the
bottom of the zeta potential probe receives an input pulse, producing ultrasound waves that
eventually reach the dispersion in the sample chamber through the gold electrode (Figure 3-1).

Figure 3-1. Zeta Potential Measurement Probe


The transducer and amplifier help with the adjustment and transmission of ultrasound waves
based on the user inputted concentration and composition of the dispersion. The ultrasound
pulses, when transmitted within the dispersion, generate particle motions relative to the liquid
when the densities between the two phases are different. These movements generate dipole
moments which cause oscillating potential on the gold electrode only. The stainless steel shell on
the outside of the zeta potential probe, separated by a non-conducting ceramic spacer, has an
electric potential of zero. The difference in potential between the two reference points equals the
colloid vibration current (CVI) which is used in combination with the concentration and the
particle size to calculate the samples zeta potential.
Lime reactivity tests were performed in an isotherm cylindrical Dewar container with a digital
laboratory stirrer. The SDT Q600 from TA Instruments was used for thermogravimetric analysis.

21

A detailed procedure for TGA can be found in Appendix D. The Ca(OH)2 and CaCO3
concentration are determined based on weight losses caused by their decomposition to CaO and
water or carbon dioxide, respectively. The decomposition reactions proceed as shown below:
Ca(OH)2 CaO + H2O

(8)

CaCO3 CaO + CO2

(9)

3.3 Experimental Procedures


The initial procedure was based on Azgomis research [8], but due to the differences between the
DT-1202 and the laser diffraction instrument used in Azgomis study, modifications to sample
preparation method (e.g. concentration) were made.

3.3.1

Particle Size Determination

Zeta potential measurements using electroacoustic techniques are primarily based on knowing
the concentration, density and particle size of the sample, using theories outlined in Section 2.3.
More specifically, the DT-1202 used in this study gave the most accurate and consistent results
when measuring dispersion systems with colloids having a similar particle size. Hence, the
particle size distributions (PSD) of Ca(OH)2 and CaCO3 were compared.
Lime calcinations were performed with 15g of either pure CaO or lime mud. The sample was
heated to 850C in an oven for 50 minutes to avoid soft-burned or hard-burned lime. Particle size
and zeta potential measurements were made after samples cooled down to room temperature.
For particle size measurements, calibrations were not required for the DT-1202. Dispersion
samples consisted of Ca(OH)2 (CaO dissolved in water) and/or CaCO3 and were diluted in
distilled water at 8wt% concentration. Lime mud samples were used as the basis for CaCO3
particle size measurements. For determining the particle size of Ca(OH)2, three different samples
were used: pure CaO, reburned lime from calcination of pure CaCO3 and reburned lime from
calcination of lime mud. The PSD of lime mud and reburned lime (from calcination of lime mud)
can be seen in Figure 3-2 and Figure 3-3 as produced by the DT-1202, respectively.
Each point on the particle size distribution curve is equal to the total weight percentage of all
particles with that same diameter. The average particle size of the sample is at the peak of the

22

distribution curve. When more than one peak exists for a particular sample, either two chemicals
with distinct size are present, or a high degree of agglomeration exists in the sample.

PSD (Weight Basis)

4.0

3.0

2.0

1.0

0.0
0.1

10

100

1000

Diameter (m)
Figure 3-2. Particle Size Distribution for CaCO3

PSD (Weight Basis)

2.0

1.5

1.0

0.5

0.0
0.1

10

100

1000

Diameter (m)
Figure 3-3. Particle Size Distribution for Ca(OH)2 (from Calcination of Lime Mud)
It can be observed that for lime mud, two peaks exist and a bimodal distribution is used to
characterize the particle size distribution. However, for reburned lime, a single peak exists and a
unimodal distribution is sufficient to characterize the particle size distribution. For lime mud, the

23

average particle size is around 12m at the first peak and approximately 200m at the second
peak. On the other hand, the average particle size for reburned lime is approximately 10m.
The necessity for using a bimodal distribution to model the PSD of lime mud can be caused by
the presence of both fine CaCO3 particles and more coarse particles due to agglomeration
between CaCO3, free lime and other impurities in the lime mud. The calcination process for
converting CaCO3 into CaO caused the reburned lime to have a more uniform chemical
composition and reduced the amount of agglomeration when compared to the original lime mud.
Therefore, only a unimodal distribution was required to characterize the PSD for reburned lime.
Comparison is made between the average particle sizes of the four samples, as shown in Figure
3-4. It can be concluded that the average value for each sample is within the same order of
magnitude. Lime mud has the largest diameter of 12.7m while the three Ca(OH)2 samples have
an average diameter of 10.7m. These findings are consistent with previous results where
Ca(OH)2 and CaCO3 particles were found to be in the range of 5m to 20m in diameter [8].
Since the main colloids (Ca(OH)2 and CaCO3) in the sample measured in this study have similar
sizes, the DT-1202 can be used with confidence for producing accurate and consistent results.
15

Particle Size (um)

12
9
6
3
0
Calcined CaCO3

Calcined Mud

Pure Ca(OH)2

Pure Mud

Sample Type
Figure 3-4. Average Particle Sizes for Lime and Lime Mud Samples
Zeta potentials for these samples were measured and the results are shown in Figure 3-5. The
zeta potential for all Ca(OH)2 samples is similar at an average value of 44mV. On the other hand,

24

lime mud has a zeta potential of approximately 12mV. These measurements indicate that free
lime does have an impact on zeta potential.
50

Zeta Potential (mV)

40
30
20
10
0
Calcined CaCO3

Calcined Mud

Pure Ca(OH)2

Pure Mud

Sample Type
Figure 3-5. Average Zeta Potentials for Lime and Lime Mud Samples

3.3.2

Initial Experimental Procedure

The following procedure was adapted from Azgomis research [8] to be used as a guideline for
making zeta potential measurements. Validation and modifications for the following procedural
steps will be discussed in the next section.
A sealed plastic bottle containing 100mL of pure Na2CO3 solution or mill green liquor was
heated up to 90C in a hot water bath with constant mixing. Analytical CaO or mill reburned
lime was transferred quickly into the bottle, while maintaining constant agitation. The reaction
was carried out for 1 hour, allowing the slaking and causticizing reactions to reach equilibrium.
The resulting white liquor slurry was filtered with a vacuum filtration unit at a constant vacuum
of 0.5bar, monitored with a pressure gauge. This was done to separate the white liquor and the
precipitated lime mud cake. The produced white liquor contained dissolved sodium content and
other impurities, which might adversely affect zeta potential measurements as the overall charge
in the system could be changed, as discussed in Section 2.2.2. The wet lime mud was dried in an
oven at 120C over a period of 90 minutes. The dried lime mud was diluted to 1wt%
concentration in 150mL distilled water and zeta potential was measured at room temperature.

25

3.3.3

Procedure Modification

Due to the differences between electroacoustic and light-scattering techniques and efforts to
minimize the time required for making sample measurements, the initial procedure was modified
to better suit the specific scope of this study. A series of experiments were performed to
determine the validity of each parameter used in the initial procedure (e.g. dilution
concentration) by changing its value within a reasonable range. In addition, considerations were
given to other variables that could potentially change during normal operations in causticizing
plants (e.g. lime particle size). Ultimately, a value for each parameter was chosen based on its
tendency to give consistent results with the least amount of measurement time.
The causticizing reaction temperature was changed between room temperature (approximately
25C) and 90C. The result shown in Figure 3-6 indicates that when carrying out the causticizing
reaction at room temperature, there is minimum variation on zeta potential for liming ratios
between the range of 0.8 and 1.2. This is expected as causticizing reactions at low temperature
will have slow kinetics, and the degree of completion for the causticizing reaction will be
relatively low regardless of the liming ratio. On the other hand, the causticizing reaction
completed at 90C shows a strong correlation between zeta potential and liming ratio.
50

Zeta Potential (mV)

9
S

40

90C

30
20

25C
10
0
0.8

0.85

0.9

0.95

1.05

1.1

1.15

1.2

Liming Ratio ([CaO]/[Na2CO3])


Figure 3-6. Effect of Liming Ratio on Zeta Potential at Different Causticizing Reaction
Temperatures (Pure Chemicals, 1 Hour Reaction)

26

The measurement temperature was investigated using samples produced at a liming ratio equal to
1.1. After mixing the lime mud sample with distilled water at 1wt%, the lime mud was heated to
50C the upper limit for the DT-1202 without damaging the zeta potential measurement probe.
The sample was cooled by air while the temperature was monitored using the DT-1202s
temperature probe. A measurement was made with every 5C decrease in sample temperature
until the lime mud reached room temperature (approximately 25C). The results are shown in
Figure 3-7. It can be concluded that the measurement temperature has an insignificant impact on
zeta potential, and that measurements can be made at room temperature to reduce sample
preparation time in this study.

Zeta Potential (mV)

50
40
30
20
10
0
25

30

35

40

45

50

Temperature (C)
Figure 3-7. Effect of Measurement Temperature on Zeta Potential (Pure Chemicals, 1 Hour
Reaction, 90C Reaction Temperature)
Instead of measuring zeta potential of lime mud, measurements could be made directly on the
synthesized white liquor. If correlations existed between zeta potential of white liquor and liming
ratio, it would decrease the overall analysis time. Figure 3-8 and Figure 3-9 show the effect of
liming ratio on zeta potential of white liquor with immediate measurement or after sample
dilution, respectively. It can be concluded that liming ratio has no effect on zeta potential when
measuring white liquor samples. This may be due to impurities and high sodium content in white
liquor obscuring the effect of Ca(OH)2, causing free lime to become a less determinant factor on
the zeta potential of the sample, even at very low concentrations (e.g. 0.7wt%). The direct
dilution of white liquor is therefore not desirable for accurate measurement of zeta potential.

27

100

Zeta Potential (mV)

Pure Chemicals
Mill Samples

80

Pure Chemicals

60
40

Mill Samples
20
0
0.2

0.4

0.6

0.8

1.2

Liming Ratio ([CaO]/[Na2CO3])


Figure 3-8. Effect of Liming Ratio on Zeta Potential when Measuring White Liquor Slurries
Synthesized with Pure Chemicals and Mill Green Liquor Samples (1 Hour Reaction, 90C
Reaction Temperature)
80
0.7wt%
1.5wt%

Zeta Potential (mV)

70

1.0wt%
3.0wt%

0.7wt%

60

1.0wt%

50
40

1.5wt%

30

3.0wt%

20
10
0
0.6

0.7

0.8

0.9

1.1

1.2

Liming Ratio ([CaO]/[Na2CO3])


Figure 3-9. Effect of Liming Ratio on Zeta Potential when Measuring White Liquor Slurries
Diluted with Distilled Water at Different Concentrations (Pure Chemicals, 1 Hour Reaction,
90C Reaction Temperature)
Since the spectrometer measures zeta potential by monitoring particle movements under an
electric field, the sample concentration plays an important role. Highly concentrated samples
might hinder particle movement due to increased chemical bonds between particles and

28

decreased free space in the dispersion. Figure 3-10 and Figure 3-11 show the zeta potential of
lime mud produced using pure Na2CO3 and mill green liquor samples.
60
1wt%

3wt%

Zeta Potential (mV)

50

1wt%

40
30

3wt%

20
10
0
0.4

0.6

0.8

1.2

1.4

Liming Ratio ([CaO]/[Na2CO3])

Zeta Potential (mV)

Figure 3-10. Effect of Liming Ratio on Zeta Potential when Measuring Lime Mud Diluted with
Distilled Water at Different Concentrations (Pure Chemicals, 1 Hour Reaction, 90C Reaction
Temperature)
75
Run 1
Run 3

60

Run 2
Run 4

45
30
15
0
0.4

0.6

0.8

1.2

1.4

Liming Ratio ([CaO]/[Na2CO3])


Figure 3-11. Effect of Liming Ratio on Zeta Potential when Measuring Lime Mud Diluted with
Distilled Water at 1wt% (Mill Green Liquor, 1 Hour Reaction, 90C Reaction Temperature)

29

Using pure chemicals, at 3wt% concentration, the zeta potential only increases slightly with an
increase in liming ratio. At 1wt% concentration, there is a much higher increase in zeta potential
with an increase in liming ratio. With a lower concentration, the particle movement is more
readily measurable by the zeta potential probe. When substituting Na2CO3 with mill green
liquors, similar results are obtained. However, the results between different experimental runs
are not consistent with each other. More importantly, there are no negative zeta potentials for
underlimed systems as shown in Azgomis study [8] (Figure 2-3).
In order to obtain more consistent results, the sample volume was varied by switching from the
large sample chamber with an attached magnetic stirrer (holds up to 130mL of sample) to the
small sample cup (holds up to 20mL of sample), both shown in Figure 3-12. The zeta potential
probe is inserted into the sample chamber or sits under the sample cup. The magnetic stirrer
cannot be attached to the sample cup, so when making zeta potential measurements in this
configuration, the sample will settle on top of the zeta potential probe.

Location of Zeta
Potential Probe

Sample Cup
Zeta Potential
Probe
Sample
Chamber

Figure 3-12. Sample Compartments for Zeta Potential Measurement


To reduce the sample volume for the sample cup, only 20mL of synthesized white liquor was
prepared. After filtration and drying, lime mud was added to 20mL of distilled water to achieve
1wt% in concentration. 10mL of this solution was transferred into the sample cup for zeta
potential measurements and the result is summarized in Figure 3-13. It can be seen that the
correlation between zeta potential and liming ratio is much closer to Azgomis research [8]. The
reason for the differences in results may be due to the effect of constant mixing associated with
using the large sample chamber. Due to the low concentration of lime mud samples, the current
induced by mixing may disrupt particle movements within the dispersion. The angle of sound
reflection, the main variable used to determine the sign of zeta potential, is also affected by

30

mixing. By allowing the sample to settle onto the surface of the probe, a more accurate
representation of sound induced particle movements and sound reflection angle can be captured.
80

Zeta Potential (mV)

60
40
20
0

-20
-40
-60
0.2

0.4

0.6

0.8

1.2

1.4

Liming Ratio ([CaO]/[Na2CO3])


Figure 3-13. Effect of Liming Ratio on Zeta Potential when Measuring 10mL of Lime Mud
Diluted in Distilled Water (Pure Chemicals, 1 Hour Reaction, 90C Reaction Temperature)

3.3.4

Modified Experimental Procedure

Based on the investigations in the previous section, a modified experimental procedure was used
for zeta potential measurements of lime mud that was separated from white liquor produced
using different liming ratios.
Tightly sealed plastic bottles containing 20mL of pure Na2CO3 or mill green liquor were heated
in a hot water bath at 90C with continued agitation. Different amounts of pure CaO or mill
reburned lime were transferred into the bottles to achieve a range of liming ratios. The slaking
and causticizing reactions were allowed to be carried out for 1 hour so the system had enough
time to reach equilibrium.
The resulting white liquor slurry was filtered in a vacuum filtration unit under a 0.5bar vacuum,
and the separated wet lime mud slurry was then dried for 90 minutes in an oven at 120C. The
lime mud was diluted in 20mL of distilled water to 1wt% concentration. Zeta potential
measurements were made with 10mL of diluted lime mud sample at room temperature. The
sample chamber sat on top of the zeta potential probe and the sample was allowed to settle

31

before making measurements. A simplified process diagram for the modified experimental
procedure is shown below in Figure 3-14.
Green Liquor

Lime

Causticizing Reaction in
Water Bath

White
Liquor Slurry

Vacuum
Filtration

Solids (Lime Mud Slurry)


Zeta Potential
Measurement

Dilution with
Distilled Water

Oven Drying

Figure 3-14. Experimental Procedure Flow Chart

White Liquor

32

RESULTS AND DISCUSSION

Zeta potential () of lime mud was found to be dependent on the amount of free lime in the lime
mud, based on Azgomis research [8]. To validate this relationship, zeta potential of samples
using analytical chemicals and mill samples will be compared and discussed. To further
investigate this relationship, zeta potential measurements were correlated to free lime
measurements for samples prepared under different experimental conditions (e.g. reaction time)
and measurement conditions (e.g. storage time).
Zeta potential values will be expressed in millivolts (mV), liming ratios in concentration ratio of
calcium oxide to sodium carbonate or [CaO]/[Na2CO3], and free lime content as weight
percentage in either white liquor or lime mud samples.

4.1 Effect of Free Lime on Zeta Potential for Analytical Samples


The experimental results from modifications of the initial procedure in Section 3.3.3 gave great
insights on how the acoustic and electroacoustic spectrometer can be used effectively to measure
zeta potential. It was found that the measurements were sensitive to different parameters due to
both chemical properties such as increased molecular interactions at high concentrations and
physical properties such as mixing induced currents causing chaotic particle movements. This
section will focus on the discussion of differences and similarities between using pure CaO and
either analytical Na2CO3 or mill green liquor to produce white liquor. Zeta potential of diluted
lime mud (at 1wt% concentration) will be measured with the 10mL sample cup at room
temperature (25C).

4.1.1

Analytical Grade Chemicals

Using analytical chemicals, causticizing reactions were carried out at 90C over a period of 1
hour. Experiments were repeated up to 20 times (and each experiment was measured 3 times) to
determine the reproducibility of results when using the modified experimental procedures as
established in the previous section. The result is shown in Figure 4-1, where it can be seen that
when the liming ratio is equal to or below 0.8, the zeta potential average between repeated
experiments is negative with a very small error bar. In addition, when liming ratio increases
above 1.0, the zeta potential value becomes less consistent, with slightly larger error bars, but
results of repeated experiments are all positive values.

33

40

Zeta Potential (mV)

30
20
10
0

-10
-20
-30
-40
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

Liming Ratio ([CaO]/[Na2CO3])


Figure 4-1. Measurement Repeatability with 10mL Samples (Pure Chemicals, 1 Hour Reaction,
90C Reaction Temperature)
According to Azgomis work, underliming and overliming conditions should have negative and
positive zeta potential values, respectively [8]. The result in Figure 4-1 shows a similar trend,
except when liming ratio is equal to 1.0, where the measured zeta potentials are in both positive
and negative regions, represented by the larger error bar. This is expected and will be referred to
as the transition zone where the system is either slightly underlimed or overlimed. Using pure
chemicals, the lime mud samples prepared at a liming ratio of 1.0 only have a minimum amount
of excess free lime. Therefore, each sample used for dilution may not contain enough free lime
for it to become a dominant factor when determining the zeta potential, in which case the zeta
potential will remain negative.
Using zeta potential theory outlined in Section 2.2.1, the negatively charged colloid in the lime
mud slurry can be the dissociated carbonate (CO32-) ions. Since the Stern layer consists of
calcium (Ca2+) and sodium (Na+) ions, the surface charge at the particle interface is caused by
adsorption of these charged ions. In general, dispersions at high pH will have negative zeta
potentials. Theoretically, in the lime mud slurry, the presence of sodium hydroxide (NaOH) will
cause the dispersion to have a negative zeta potential value. However, at overliming conditions,
positive zeta potentials are observed.

34

The reason for consistent measurements of positive zeta potentials when the sample contains an
excess amount of free lime might be due to changes in the double layer thickness. Since the Stern
layer contains positively charged ions, additional Ca2+ ions (dissociated from excess amount of
Ca(OH)2) can further compress the double layer when compared to having only Na+ ions. In
addition, zeta potential measurements are obtained based on electric charge generated by particle
movements. Hence, the ability of particles to move through the dispersion (known as electrical
mobility) plays an important role in determining the magnitude and the sign of zeta potential.
Combination of a thinner double layer thickness and the change in electrical mobility caused by
excess free lime can potentially change the sign of zeta potential from negative to positive.
For samples from the above experiments, the free lime content was measured using TGA. The
result for weight losses at different temperatures is shown in Figure 4-2 and Figure 4-3, for
liming ratios of 0.6 and 1.4, respectively. The initial decrease in weight at 100C corresponds to
evaporation of water. Decrease in weight at approximately 400C is due to the decomposition of
Ca(OH)2 to CaO and H2O. The final decrease in weight at approximately 800C is due to the
decomposition of CaCO3 to CaO and CO2. Comparing the two plots, the sample weight loss
caused by Ca(OH)2 decomposition is insignificant when liming ratio is 0.6. On the other hand,
there is a clear decrease in sample weight when the liming ratio is 1.2. This is expected as the
amount of free lime in underlimed systems should be much lower than in overlimed systems.

Weight (%)

110

90

100
99
98

70

97
96

350

375

400

425

450

50
0

150

300

450

600

750

900

Temperature (C)
Figure 4-2. Weight Loss Profile for Lime Mud (0.6 Liming Ratio and Pure Chemicals)

35

Weight (%)

110

90
97
96

95

70

94

93
92
350

375

400

425

450

50
0

150

300

450

600

750

900

Temperature (C)
Figure 4-3. Weight Loss Profile for Lime Mud (1.2 Liming Ratio and Pure Chemicals)
The zeta potential and free lime content for lime mud prepared at different liming ratios is shown
in Figure 4-4. Since pure chemicals are used, the free lime contents are expected to be
consistently low for liming ratios below 1.0. When overliming occurs, the free lime content
increases to almost 6wt%. The result also shows that zeta potential increases with liming ratios.
20

8
Zeta Potential
Free Lime

10

5
0

-5

-10

Free Lime (%)

Zeta Potential (mV)

15

-15
-20
-25

0
0

0.4

0.8

1.2

1.6

Liming Ratio ([CaO]/[Na2CO3])


Figure 4-4. Zeta Potential and Free Lime as a Function of Liming Ratios (Pure Chemicals, 1
Hour Reaction, 90C Reaction Temperature)

36

When zeta potential values are plotted directly against the free lime content in Figure 4-5, it can
be noted that a direct correlation exists between the two variables.
20

Zeta Potential (mV)

15
10
5

Overliming

Underliming
-5
-10
-15
-20
0

Free Lime (%)


Figure 4-5. Zeta Potential as a Function of Free Lime Content (Pure Chemicals, 1 Hour
Reaction, 90C Reaction Temperature)
Combined with the results shown in Figure 4-4, when the liming ratio is below 1.0, free lime
content is below 2wt% and the average zeta potential is negative. On the contrary, when liming
ratio is at or higher than 1.0, free lime content is above 2wt% and the average zeta potential is
positive. Since this relationship exists between the two variables, one can use zeta potential to
represent the Ca(OH)2 concentration in lime mud within a typical range of liming ratios in the
kraft recovery process. For example, based on the above results, if the system is said to be
overlimed when free lime content is higher than 2wt%, then a positive zeta potential can be used
to indicate a system with an excess amount of free lime.

4.1.2

Analytical Grade Lime and Mill Green Liquor

While still using analytical CaO, mill green liquor samples were used instead of pure Na2CO3.
The results for zeta potential and free lime measurements of filtered lime mud samples are
shown in Figure 4-6. Unlike using analytical chemicals, even for underlimed systems, free lime
content still exists in the sample (2wt%). This is due to the presence of impurities in the mill
green liquor samples, meaning the 20mL of green liquor used in the causticizing reaction does
not necessarily contain the exact amount of Na2CO3 as analyzed by the ABC test in Section 3.1.

37

The free lime content can reach as high as almost 19wt% at extreme overliming conditions.
However, similar to using pure chemicals, zeta potential is positively correlated to free lime.
30

20
Zeta Potential
Free Lime

15
10
0

10

-10

Free Lime (%)

Zeta Potential (mV)

20

5
-20
-30

0
0

0.4

0.8

1.2

1.6

Liming Ratio ([CaO]/[Na2CO3])


Figure 4-6. Zeta Potential and Free Lime as a Function of Liming Ratios (Mill Green Liquor, 1
Hour Reaction, 90C Reaction Temperature)
The zeta potential value of overlimed systems reaches a limit and does not increase with further
increase in liming ratio (Figure 4-7).
25

Zeta Potential (mV)

20
15
10
5
0

-5
-10
-15
-20
0

10

15

20

Free Lime (%)


Figure 4-7. Zeta Potential as a Function of Free Lime Content (Mill Green Liquor, 1 Hour
Reaction, 90C Reaction Temperature)

38

Zeta potential remains negative for underlimed systems. For overlimed systems, even when free
lime content is increased from 6wt% to 19wt%, the zeta potential is consistently measured to be
approximately 13mV to 15mV. Constant zeta potential at overliming conditions can also be seen
when using pure chemicals (Figure 4-5). Since the small sample cup was used, samples were
loaded directly onto the zeta potential probe. Measurements were made on particle movements as
they were induced by sound waves. When the system contained either mostly Ca(OH)2
(overlimed) or CaCO3 (underlimed), zeta potentials corresponded mainly to these chemicals and
the effects of other components were effectively hidden. Hence, zeta potential values level off at
both overliming and underliming conditions.
Causticizing plants often operate at liming ratios within the transition zone (when liming ratio is
between the range of 0.8 to 1.0), thus it is important to investigate the sensitivity of zeta potential
measurements in this region. Samples were prepared and measured at 0.01 liming ratio
increments. In addition, as a continuous effort to reduce the overall analysis time, some samples
were separated into two portions after vacuum filtration of white liquor. One portion continued
to go through the drying process (dry lime mud) while the other was measured directly (wet lime
mud). The result is summarized in Figure 4-8.
20
Dry Lime Mud
Wet Lime Mud

Zeta Potential (mV)

15
10
5
0

-5
-10
-15
-20
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

Liming Ratio ([CaO]/[Na2CO3])


Figure 4-8. Effect of Liming Ratio on Zeta Potential Dry and Wet Samples (Pure Lime, 1 Hour
Reaction, 90C Reaction Temperature)

39

When measurements are made with the dried lime mud sample; the zeta potential values levels
off at both ends of the curve similar to the previous results (Figure 4-4 and Figure 4-6).
However, the smaller increments in liming ratio give a more detailed representation of how zeta
potential changes at the transition region between liming ratios of 0.8 and 1.0. The results are
expected as zeta potential shows a gradual increase within this range and changes from negative
to positive values, before leveling off at liming ratios greater than 1.0.
By examining the differences between measurements of dry and wet lime mud samples, the
results at higher liming ratios are very close to each other. On the other hand, measurements at
lower liming ratios show a larger discrepancy between the two methods. This may be due to the
effect of impurities dissolved in water (which are removed from the lime mud during the drying
process) and their interactions with CaCO3 and Ca(OH)2 in the wet lime mud sample.
At higher liming ratios, the presence of an excess amount of Ca(OH)2 has a more dominant
effect on zeta potential when compared to the effect of the impurities. On the other hand, at
lower liming ratios, the CaCO3 does not have the same degree of influence on zeta potential as
exerted by the free lime. Zeta potential of wet lime mud samples in the transition zone will be
further studied in the next section.

4.2 Effect of Free Lime on Zeta Potential for Mill Samples


Using the same green liquor samples from the previous section, pure CaO was replaced with mill
reburned lime. The causticizing reaction continued to be carried out at 90C for 1 hour. The zeta
potential of samples obtained using reburned lime and the previous results based on pure CaO
are shown in Figure 4-9.
There are both similarities and differences between using pure CaO and reburned lime. Between
both sets of experiments, zeta potential values at high liming ratios are very close to each other.
An increase in zeta potential values at the transition region is present in both cases, with the
samples using reburned lime showing larger discrepancies between positive and negative values
as liming ratio increases. This may be due to the inconsistent quality of reburned lime used in the
causticizing reaction. Since 20mL of white liquor slurry is required, only a small amount of
reburned lime is added, causing variability in terms of lime quality when compared to pure CaO.

40

Hence, at the sensitive transition zone, the lime mud sample can be either slightly underlimed or
slightly overlimed due to the small increments of liming ratio.
20
Pure CaO
Reburned Lime

Zeta Potential (mV)

15
10
5
0

-5
-10
-15
-20
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

Liming Ratio ([CaO]/[Na2CO3])


Figure 4-9. Effect of Liming Ratio on Zeta Potential using Pure Lime and Reburned Lime (1
Hour Reaction, 90C Reaction Temperature)
In addition, differences exist in the range of the transition zone between the two sets of
experiments. For experiments using pure CaO, the zeta potential values begin to increase rapidly
at a liming ratio of approximately 0.7, while for reburned lime, this occurs at liming ratios just
before 0.8. The reason for the delayed start into the transition zone for reburned lime is due to the
effects of lime availability as outlined in Section 2.4.1. Reburned lime should be less reactive
than pure CaO due to the presence of un-calcined CaCO3, as well as having lowered
porosity/specific surface area within the lime particles. Hence, when adding the same amount of
lime in weight, reburned lime will have a lower amount of CaO available for reactions when
compared to pure chemicals. It is then expected that at a liming ratio of 0.7, samples made with
pure chemicals are already starting to become overlimed while samples made with reburned lime
remain underlimed.
Contrary to the more constant zeta potential values at higher liming ratios between the two
results, using reburned lime gives lower values at underliming conditions when compared to
using pure chemicals. The lime mud will contain a higher content of CaCO3 for samples made
with reburned lime, due to the presence of un-calcined CaCO3 in the reburned lime. In addition,

41

using reburned lime introduces additional impurities that will impact the measurement of zeta
potential. It is shown repeatedly that for measurements made in this study, zeta potential values
of dispersion systems with an excess amount of Ca(OH)2 (overlimed systems) are much less
affected by other colloids within the system, when compared to dispersion systems mainly
consisting of CaCO3 (underlimed systems).
The zeta potentials of lime mud samples prepared with reburned lime were also measured prior
to drying the samples (wet lime mud). The results are compared to measurements made after
drying the samples (dry lime mud) and are plotted in Figure 4-10. Zeta potential data is also
included in Appendix E.
The overall trend of zeta potential for the wet lime mud samples is similar to the trend for dry
lime mud samples. The zeta potential values are relatively constant at both end of the curve. The
region where zeta potential starts to increase is when the liming ratio is slightly lower than 0.9.
The main contrast between wet and dry lime mud measurements is the lack of a smooth
transition phase for the wet lime mud. In addition, zeta potential values are lower in the negative
region and higher in the positive region, when compared to the dry lime mud measurements.
These characteristics may be caused by the presence of impurities that are dissolved in water for
the wet lime mud samples, where polyvalent ions can impact zeta potential values as discussed in
Section 2.2.2.

20
Dry Lime Mud
Wet Lime Mud

Zeta Potential (mV)

15
10
5
0

-5
-10
-15
-20
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

Liming Ratio ([CaO]/[Na2CO3])


Figure 4-10. Effect of Liming Ratio on Zeta Potential Dry and Wet Samples (Reburned Lime, 1 Hour Reaction, 90C Reaction
Temperature)

42

43

Based on the differences between the two sample preparation methods, measurements using wet
lime mud samples are more advantageous for the development of a free lime monitoring system
in the causticizing plant. The most important factor is the decrease in sample preparation time,
where drying takes at least an hour to complete. This means a shorter delay time for detecting the
presence of free lime when the operating conditions (e.g. green liquor composition) change. The
lack of a smooth transition zone is also helpful when the results are used to determine whether a
system contains free lime. The sign of zeta potential can be used as a fast indication of an
overlimed system (positive values) or underlimed system (negative values) without looking at
the specific value. The large gap between the two ends of the plot is useful from a process
control standpoint. A threshold range can be set such that zeta potential within this range can
give an indication that the system is very close to becoming overlimed (i.e. between -10mV and
10mV for results presented in Figure 4-10).

4.2.1

Effect of Lime Properties

Reburned lime with different particle sizes may affect the kinetics of the causticizing reaction as
discussed in Section 2.4.1. Equipment setup for the lime reactivity test is shown in Figure 4-11.

Mixer

Thermocouple

Dewar
Flask

Figure 4-11. Lime Reactivity Test Setup

44

A lime reactivity test was performed with 20g of mill reburned lime mixed in 900mL of distilled
water. Temperature readings were recorded at 10s intervals until temperatures were constant for
at least 5 minutes. The difference between powdered and pelleted lime (ranging from 1mm to
2mm in diameter) was investigated. In addition, white liquor was synthesized at different liming
ratios using powdered or pelleted lime, and their zeta potentials were compared.
Based on the lime reactivity test, differences can be seen between lime powder and lime pellets
(Figure 4-12). Powdered lime reaches a higher temperature at steady state and reaches that
temperature at a faster rate than pelleted lime. This suggests that using powdered lime has a
higher potential to reach equilibrium CE compared to pelleted lime when the same reaction time
is maintained. In addition, the rate of temperature increase for powdered lime is not as consistent
when compared to pelleted time, indicating that the larger surface area of finely pulverized lime
will cause the slaking reaction to occur more violently and rapidly. These results indicate that
when using the same weighted lime, finely pulverized lime is more reactive than pelleted time,
which will give higher reaction efficiencies during the causticizing process.
80

Temperature (C)

Powder
Pellet

60

40

20

0
0

20

40

60

80

100

Reaction Time (min)


Figure 4-12. Wet Slaking Curve for Powdered and Pelleted Reburned Lime
The zeta potential measurements for the causticizing reaction carried out using either pulverized
lime or lime pellets are shown in Figure 4-13.

45

20
Powder
Pellet

Zeta Potential (mV)

15
10
5
0

-5
-10
-15
-20
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

Liming Ratio ([CaO]/[Na2CO3])


Figure 4-13. Effect of Liming Ratio on Zeta Potential Powdered and Pelleted Reburned Lime
(1 Hour Reaction, 90C Reaction Temperature)
These results show that when the liming ratio is equal to 0.6 (i.e. the system is underlimed),
samples prepared using pelleted lime already give positive zeta potential measurements. This
means only the outer surface of the lime pellet reacted with the Na2CO3 in the green liquor,
leaving the Ca(OH)2 core of the lime pellets in the measured lime mud sample. These results
suggest that the lime particle size and lime reactivity influences the causticizing reaction and in
turn the zeta potential measurements greatly. Excess free lime could be present for a theoretically
underlimed system when the lime quality is poor.

4.2.2

Effect of Causticizing Reaction Time

The causticizing reaction time was varied to determine the sensitivity of zeta potential
measurement using the DT-1202 as retention time in the causticizers might vary between
different kraft pulp mills. The reaction time was monitored between 1 hour and 6 hours and the
results are shown in Figure 4-14. For each specific liming ratio, small variations in zeta potential
exist at different reaction times, but the overall trend is the same. Underlimed systems will give
negative zeta potential measurements and overlimed systems will give positive zeta potential
measurements. The small variations are insignificant for the overall use of zeta potential as a tool
to detect the presence of free lime.

46

20

Zeta Potential (mV)

15
10
5
1.20LR
0.90LR

1.10LR
0.80LR

1.00LR

-5
-10
-15
-20
0

Reaction Time (h)


Figure 4-14. Effect of Reaction Time on Zeta Potential (Mill Green Liquor, 90C Reaction
Temperature)

4.2.3

Effect of Lime Mud Storage Time

The effect of storage time was also considered. Measurements were made on the wet lime mud
after storing the samples for different amounts of time (Figure 4-15).
20

Zeta Potential (mV)

15
10
5
1.15LR
0.95LR

1.10LR
0.90LR

1.05LR
0.85LR

1.00LR
0.80LR

1
0.75LR

-5
-10
-15
-20

0.5

1.5

2.5

3.5

Storage Time (h)


Figure 4-15. Effect of Storage Time on Zeta Potential (Mill Green Liquor, 1 Hour Reaction,
90C Reaction Temperature)

47

If kraft pulp mills do not have instruments for zeta potential measurements readily available, it
may be possible to send the samples for off-site analysis if overliming problems continue to
occur. In Figure 4-15, the result of short term storage (1 hour to 4 hours) is plotted. The zeta
potential trend is identical to results in the previous sections and almost no variation exists when
samples were stored up to 4 hours before measurements.
When the storage time was extended up to 1 month, the result can be seen in Figure 4-16 (data
included in Appendix F). In this case, for overlimed systems, the sign of zeta potential values
remains positive, regardless of the storage time. However, the values for underlimed systems
become positive when the samples are stored for as little as one day (i.e. when liming ratio
equals 0.9). All zeta potentials for underlimed systems become positive after one week of
storage. This suggests wet lime mud cake stored for more than a day will likely not result in a
difference between overlimed and underlimed system. This means that zeta potential
measurement needs to be made as soon as possible in order to obtain the most representative
results and indicate the possibility of an overlimed system. The reasons for the change in zeta
potentials may be due to how the causticizing reaction took place. During the reaction, Ca(OH)2
will form a coating of CaCO3 on the surface of the particles at the microscopic level. However,
during extended storage periods, this coating becomes less coherent and the Ca(OH)2 in the inner
layer becomes detectable. Since overlimed systems are already saturated with Ca(OH)2, there
will be no change in the sign of zeta potentials.
Another possible cause is the oxidation of sodium sulphide (Na2S) in wet lime mud cake to
sodium sulphate (Na2SO4) after long storage periods. Polyvalent ions, as discussed in Section
2.2.2, will cause zeta potential to shift between positive and negative values. Hence, Na2SO4 may
have an impact during zeta potential measurements when compared to Na2S. This also indicates
that Ca(OH)2 is a more dominant factor in zeta potential measurements when compared to
CaCO3 and is less likely to be affected by other chemicals, as first discussed in Section 4.1.2.

20

Zeta Potential (mV)

15
10
5

0
-5
-10
Instant
4 hours
2 weeks

-15

2 hours
1 day
3 weeks

3 hours
1 week
4 weeks

-20
0.75

0.8

0.85

0.9

0.95

1.05

1.1

1.15

Liming Ratio ([CaO]/[Na2CO3])


Figure 4-16. Effect of Extended Storage Time on Zeta Potential (Mill Green Liquor, 1 Hour Reaction, 90C Reaction Temperature)

48

49

When combining both the effect of reaction time and storage time, more insight can be obtained
about the lime mud system. Figure 4-17 includes measurements of wet lime mud cake, prepared
at different reaction times and measured after 1 month of storage. Similar to previous results,
zeta potential values become positive for underlimed systems when the reaction time is short (up
to 3 hours). However, at longer reaction times, the values stay negative even after 1 month of
storage. This is likely due to the higher causticizing efficiency as a result of longer reaction
times. This means that Ca(OH)2 is completely reacted to form CaCO3 at the microscopic level, as
opposed to having a Ca(OH)2 core with an outer layer of CaCO3. Since no free lime is hidden
within a CaCO3 coating, the value of zeta potential will not change regardless of the storage
time.
20

Zeta Potential (mV)

15
10
1.20LR
1.10LR
1.00LR
0.90LR
0.80LR

5
0

-5
-10
-15
-20
0

Reaction Time (h)


Figure 4-17. Effect of Reaction Time on Zeta Potential 1 Month Storage (Mill Green Liquor,
90C Reaction Temperature)

50

CONCLUSIONS

A systematic study was conducted to investigate the feasibility of using the zeta potential of lime
mud slurry to detect free lime content in the causticizing plant. Synthetic samples were produced
by using analytical chemicals (i.e. pure CaO and Na2CO3) or mill samples (i.e. reburned lime and
green liquor). Zeta potential measurements were made using an acoustic and electroacoustic
spectrometer (DT-1202) on wet lime mud filtered from the white liquor.
Experimental procedures were initially based on Azgomis work [8]. Due to differences in
measurement technology, a modified procedure was established by adjusting the sample dilution
concentration, sample volume and forgoing the drying process.
A range of liming ratios were used to obtain lime mud samples that varied in free lime content,
and the effect of free lime on zeta potential was determined. In addition, the sensitivity of zeta
potential measurements to other parameters were also investigated, namely lime particle size,
reaction time and storage time. The main conclusions are summarized below:
Establishment of Experimental Procedures for the Spectrometer

The particle size for reburned lime and lime mud samples obtained from pulp mills were
about 10m in diameter, similar to analytical chemicals (CaO and CaCO3). The approximate
particle size is required as one of the input parameters for zeta potential measurements.

The positive correlation between zeta potential and free lime was found only with wet lime
mud filtered from the white liquor. Measurements performed directly on the white liquor
showed no change in zeta potential when varying the free lime content regardless if the
samples are diluted.

The free lime content correlates well with zeta potential measurements when the lime mud
samples were diluted to 1wt% concentration in distilled water, compared to dilutions at
higher concentrations (e.g. 3wt%).

Zeta potential measurements need to be made with a small sampling cup (i.e. 10mL),
allowing the lime mud sample to settle on the zeta potential probe instead of constantly
mixing the sample.

51

Effect of Liming Ratios

Repeated measurements of lime mud samples produced with analytical chemicals gave
consistent zeta potential values. Positive values were obtained for overlimed systems and
negative values were obtained for underlimed systems. At liming ratios between the range of
0.8 and 1.0, both positive and negative values were measured, indicating the presence of a
transition zone.

At the same liming ratio, TGA showed a higher amount of free lime in lime mud samples
prepared using mill green liquor when compared to samples prepared using pure Na2CO3.

Zeta potentials were relatively constant for underlimed systems between the range of -15mV
and -20mV but increased as the liming ratio increased. Zeta potentials reached the range of
+10mV and +15mV for overlimed systems. The change in zeta potential values from
negative to positive was observed for lime mud prepared using either mill samples or
analytical chemicals.

Measurements performed on dried lime mud samples showed a gradual increase in zeta
potential within the transition zone from negative to positive values. On the other hand,
measurements on wet lime mud samples showed an initial slow increase in negative values
followed by an acute change in the sign of zeta potential when the system became overlimed.

Compared to the reactivity of pelleted lime, finely pulverized lime had a higher potential to
reach equilibrium causticizing efficiency in a shorter period of time. Samples prepared using
pelleted lime gave positive zeta potentials (i.e. indication of an overlimed system) even when
the liming ratio is low (e.g. at a ratio of 0.6).

The overall trend of zeta potential as a function of liming ratio was not affected by varying
the causticizing reaction time between 1 hour and 6 hours, when the lime mud samples were
measured immediately after preparation.

Storing the lime mud samples for longer than 1 day before making measurements would
cause the sign of zeta potential for underlimed systems to change from negative to positive if
the retention time for the causticizing reaction was shorter than 4 hours.

52

IMPLICATIONS

This study confirmed that correlations exist between zeta potential and free lime. In addition, this
study explored the possibility of using zeta potential as the basis for a free lime monitoring
system in the causticizing plant by analyzing the effect on zeta potential when changes occur in
measurement conditions (e.g. sample volume) and operating parameters (e.g. reaction time). The
results bring several practical implications.
Within the liming ratio range between 0.4 and 1.4, the zeta potential was found to be directly
correlated to the amount of free lime. Systems with low free lime contents had negative zeta
potential values while systems with high free lime contents had positive zeta potential values.
There was also an acute change from negative to positive zeta potential values when an excess
amount of free lime is present, if measurements were made on wet lime mud slurries (i.e.
samples collected after the white liquor clarifier).
A monitoring system can be designed to maintain an optimum liming ratio within the
causticizing plant based on zeta potential. This will replace current laboratory tests for free lime
determination such as the ammonium chloride method and the muffle furnace method. Detection
of free lime using zeta potential will be more efficient since measurements can be made directly
on the lime mud slurries and the measurement time is approximately 30 seconds, reducing the
extensive amount of laboratory work required in conventional tests. In addition, the polarized
result will clearly indicate an overlimed or an underlimed system, based on positive or negative
zeta potential, respectively.
Such a monitoring system can also be adjusted in terms of how strictly the amount of free lime
needs to be controlled. An underlimed system will have a zeta potential of approximately 20mV. As free lime in the system increase, the zeta potential will gradually increase to -10mV.
Further increase in free lime will cause the zeta potential to jump to +15mV. Hence, a strict
system setting will give a warning for excess free lime as soon as zeta potential deviates from 20mV. On the other hand, since higher liming ratios result in higher conversion rates of green
liquor to white liquor within a short period of time, a lenient system setting can be used, which
will only give a warning for excess free lime when a positive zeta potential is measured.

53

If lime mud slurry samples are collected after the white liquor clarifier in the causticizing plant,
the only laboratory procedure will be diluting the sample to 1wt% concentration before
measuring its zeta potential. However, if lime mud slurry samples prior to the white liquor
clarifier need to be measured, then a filtration step is required before diluting the collected
sample. This will ensure that dissolved NaOH content in the white liquor does not hinder the
effects of free lime on zeta potential.
In addition, a monitoring system can be directly incorporated into the causticizing plant. A
potential system installed after the white liquor clarifier is proposed in Figure 6-1. For each zeta
potential measurement, V-101 opens to allow a small amount of lime mud slurry to divert from
the main flow. V-101 then closes to allow water from the weak wash stream to dilute the lime
mud slurry sample. Once the zeta potential measurement has been taken, V-102 opens, allowing
the sample to return to the main flow. Since the volume of the diluted lime mud slurry sample is
small, it will not alter the composition in the main flow. After V-102 closes, V-101 opens again
for the next zeta potential measurement.

WL Clarifier

V-101

V-102

Mud
Slurry
Storage
Tank

H2O

Zeta Potential
Probe

Figure 6-1. Potential Online Free Lime Monitoring System

54

REFERENCES
1. Tran, H.N. and Vakkilainnen E. K., The Kraft Chemical Recovery Process, Tappi Kraft
Recovery Short Course, Tappi Press, St. Petersburg, Florida, January 7-10, 2013
2. Tran, H.N., Energy and Chemical Recovery, Pulp and Paper Course Notes (CHE564),
University of Toronto, Spring 2011
3. Sanchez, D.R., Recausticizing Principles and Practice, Tappi Kraft Recovery Short
Course, Tappi Press, St. Petersburg, Florida, January 7-10, 2013
4. Costa, D.S., Pimenta, E.M., and Figueriredo, L.S., Study Case: Optimization of
Recausticizing Plants in a Brazillian Pulp Mill, International Chemical Recovery
Conference, Williamsburg, VA, 2010
5. Riberio, J.C.T., Santos, S.M., and Tran, H.N., Experience of Low Lime Mud Solids
Problems at a Kraft Pulp Mill, O Papel, 69(6), 2008: 69 -79
6. Mao, S., Gafarova, L., and Tran, H.N., Analytical Methods for Free Lime Analysis,
Research Consortium Meeting, Toronto, Ontario, November 11-13, 2008
7. TAPPI Standard T624, Analysis of Soda and Sulfate White and Green Liquors, Technical
Association of the Pulp and Paper Industry, September 27, 2011
8. Azgomi, F., Impact of Liming Ratio on Lime Mud Settling and Filterability in the Kraft
Recovery Process, Ph.D. Thesis, University of Toronto, June 2013
9. Dukhin S.S., and Derjaguin B.V., Electrokinetic Phenomena, John Willey & Sons, New
York, 1974
10. Weiner, B.B., Tscharnuter, W.W., and Fairhurst, D., Zeta Potential: a New Approach
Canadian Mineral Analysts Meeting, September 1993
11. ASTM Standard C25 Standard Test Methods for Chemical Analysis of Limestone,
Quicklime, and Hydrated Lime, ASTM International, West Conshohocken, PA, 2006, DOI:
10.1520/C0025-06
12. ASTM Standard C114 Standard Test Methods for Chemical Analysis of Hydraulic
Cement, ASTM International, West Conshohocken, PA, 2004, DOI: 10.1520/C0114-04a
13. Bond, J., Free Lime in Lime Mud (TGA Method % CaO), DMI Peace River Pulp Division,
Peace River, Alberta, April 6, 2009
14. Derjaguin , B.V. and Landau, L., Theory of the Stability of Strongly Charged Lyophobic
Sols and the Adhesion of Strongly Charged Particles in Solution of Electrolytes, Acta Phys.
Chim, USSR, 14, 1941: 633-662
15. Verwey, E.J.W. and Overbeek, J.Th.G., "Theory of the Stability of Lyophobic Colloids",
Elsevier, 51(3), 1948: 631-636
16. Elimelech, M., Gregory, J., Jia, X., and Williams, R.A., Particle Deposition and
Aggregation Measurement: Modelling and Simulation, Butterworth-Heinemann, Boston,
1995

55

17. Lyklema, J., "Fundamentals of Interface and Colloid Science", Volumes 1, Academic Press,
1993
18. Hunter, R.J., Zeta Potential in Colloidal Science, Academic Press, London, England, 1981
19. Wakman, R., The Influence of Particle Properties on Filtration, Separation Purification
Technology, 58(2), 2007: 234-241
20. Foxall, T., Peterson, G.C, Rendall H.M., and Smith A.L., Charge Determination of Calcium
Salt/Aqueous Solution Interface, Journal of the Chemical Society, Faraday Transactions,
75(I), 1970: 1034-1039
21. Thompson, W.D., and Pownall, G.P., Surface Electrical Properties of Calcite, Journal of
Colloids and Interface Science, 131, 1989: 74-82
22. Somasundaran, P., and Agar, G.E., The Zero Charge of Calcite, Journal of Colloid
Interface Science, 24, 1967: 433-440
23. Somasundaran, P., Amankonah, J.O., and Ananthapadmabhan, K.P., Mineral Solution
Equilibria in Sparingly Soluble Mineral System, Colloid and Surfaces, 15, 1985: 309-333
24. Smallwood, P.V., Some Aspects of the Surface Chemistry of Calcite and Aragonite, Par I:
An Electrokinetic Study, Colloid and Surfaces, 255, 1977: 881-886
25. Huang, Y.C., Fowkes, F.M., Lylod, T.B., and Sanders, N.D., Adsorption of Calcium Ions
from Calcium Chloride Solution onto Calcium Carbonate Particles, Langmuir, 7, 1991:
1742-1748
26. Dukhin A.S., and Goetz, P.J., Ultrasound for Characterizing Colloids, Elsevier, 2002
27. Morse, P.M., and Uno Ingard, K., Theoretical Acoustics, Princeton University Press, NJ,
January 1, 1987
28. Debye, P., "A Method for the Determination of the Mass of Electrolyte Ions", Journal of
Chemical Physics, 1, 1933: 13-16
29. Frenkel J., On the Theory of Seismic and Seismoelectric Phenomena in a Moist Soil,
Journal Engineering Mechanics, 131(9), 2005: 879-887
30. Hermans, J., Charged Colloid Particles in an Ultrasonic Field, Philosophical Magazine,
25(168), 1935: 426-438
31. Rutgers, A.J., and Rigole, W., "Ultrasonic Vibration Potentials in Colloid Solutions, in
Solutions of Electrolytes and Pure Liquids", Transactions of the Faraday Society, 54,
1958:139-143
32. Enderby, J.A., "On Electrical Effects Due to Sound Waves in Colloidal Suspensions",
Proceedings of Royal Society of London, A207(1090), 1951: 329-342
33. Booth, F., and Enderby, J., "On Electrical Effects due to Sound Waves in Colloidal
Suspensions", Proceedings of Physical Society, A65(321), 1952
34. Lime Terminology, Standards & Properties, National Lime Association, 2007
35. Olsen, J.C., and Direnga, O.G., Settling Rate of Calcium Carbonate in the Causticizing of
Soda Ash, American Chemical Society: Industrial & Engineering Chemistry, 33(2), 1941:
204-218

56

36. Wernqvist, A., and Theliander, H., On the Kinetics of the Causticizing Reaction, Journal
of Pulp and Paper Science, 20(10), 1994: 301-315
37. Venkatesh, V., Lime Reburning: The Rotary Lime Kiln, Chemical Recovery in the Alkaline
Pulping Processes, Tappi Press, Atlanta, 1992: 153-179
38. Rothrock, C.W., The Effect of Certain Variables on the Causticizing Process, Tappi
Journal, 41(6), 1958: 241A-244A
39. Rydin, S., Haglund, P., and Mattson, E., Causticizing of the Technical Green Liquors with
Various Lime Qualities, Svensk Papperstidning, 80(2), 1977: 54-58
40. Rydin, S., The Kinetics of the Causticizing Reaction, Svensk Papperstidning, 81(2), 1978:
43-48
41. Sylwon, O., Theory and Practice in Causticizing Green Liquor, Paper Trade Journal,
143(2), 1959: 42-46
42. Dorris, G.M., and Allen, L.H., The Effect of Reburned Lime Structure on the Rates of
Slaking, Causticizing and Lime Mud Settling, Journal of Pulp and Paper Science, 11(4),
1985: J89-J98
43. Kinzner, K., Investigations of the Causticizing of Green Liquors, Proceedings of
IUPAC/EUCEPA Symposium on Recovery of Pulping Chemicals, Helsinki, May 13-17,
1968: 279-302
44. Potgieter, J.H., Potgieter, S.S., Moja, S.J., and Mulaba-Bafubiandi, A., An Empirical Study
of Factors Influencing Lime Slaking. Part I: Production and Storage Conditions, Elsevier
Science Ltd., 15(2002), December 2001: 201 - 203
45. Potgieter, J.H., Potgieter, S.S., Moja, S.J., and Mulaba-Bafubiandi, A., An Empirical Study
of Factors Influencing Lime Slaking. Part II: Lime Constituents and Water Composition,
South African Water Research Commission, 29(2) April 2003
46. Campbell, A.J., Factors Affecting White Liquor Quality: Green Liquor Concentration,
Dregs Concentration and Lime Dosage, Pulp & Paper Canada, 82(4), 1981: T121-T126
47. Campbell, A.J., The Effects of Lime Quality and Dosage on Causticizing and Lime Mud
Settling Properties, Pulp & Paper Canada, 86(11), 1985: 67-71
48. Axelsson, O., Gustafsson, E., and Wiklander, G., Improper Operating Conditions Sources
of Unnecessary Trouble and Energy Cost in the Kraft Recovery Cycle, Pulp & Paper
Canada, 84(2), 1983: 55-57
49. Keitaanniemi O., and Virkola N.E., Amounts and Behaviours of Certain Chemical Elements
in Kraft Pulp Manufacture: Results of a Mill Scale Study, Paper ja Puu, 60(0), 1978: 507522
50. Jemaa, N., Thompson, R., Paleologou, M., and Berry, R.M., Non-process Elements in the
Kraft Cycle, Part I: Sources, Levels and Process Effects, Pulp & Paper Canada, 100(9),
1999: 47-51
51. Keitaanniemi O., and Virkola N.E., Undesirable Elements in Causticizing Systems, Tappi
Journal, 65(Jul.), 1982: 89-92

57

52. Shin, H.G., Kim, H., Kim, Y.N., and Lee, H.S., Effect of Reactivity of Quick Lime on the
Properties of Hydrated Lime Sorbent for SO2 Removal, Journal of Materials Sciences and
Technology, 25(3), 2009
53. Notidis, E., The Formation of Guarded Sodium in Lime Mud, M.A.Sc Thesis, University of
Toronto, 1994
54. Dukhin, A.S., and Goetz, P.J., Method and Device for Characterizing Particle Size
Distribution and Zeta Potential in Concentrated System by Means of Acoustic and
Electroacoustic Spectroscopy, Patent USA, 6109098, 2000
55. Dukhin, A.S., and Goetz, P.J., Method and Device for Determining Particle Size
Distribution and Zeta Potential in Concentrated Dispersions, Patent USA, 6449563, 2002
56. Dukhin, A.S., and Goetz, P.J., Method for Determining Electric Properties of Particles in
Liquids by means of Combined Electroacoustic and Complex Conductivity Measurement.
Patent USA, 6915214 B2, 2005
57. ISO 20998-1:2006, Measurement and Characterization of Particles by Acoustic Methods Part 1: Concepts and Procedures in Ultrasonic Attenuation Spectroscopy, International
Organization for Standardization, February 13, 2011
58. Dukhin A.S., and Goetz, P.J., Peculiarities of Live Cells' Interaction with Micro- and
Nanoparticles, Dispersion Technology Inc., July 2010
59. Dukhin A.S., and Goetz, P.J., Acoustics and Electroacoustics for Ceramics, Dispersion
Technology Inc., March 1999
60. SDP-QC.70-160, Excess CaO% in Filtered Lime Mud, Peace River Pulp Division
61. Epstein, P.S., and Carhart R.R., The Absorption of Sound in Suspensions and Emulsions,
Journal of the Acoustical Society of America, 25(3), 1953: 553-565
62. Allegra, J.R., and Hawley, S.A., Attenuation of Sound in Suspensions and Emulsions:
Theory and Experiments, Journal of the Acoustical Society of America, 51, 1972: 15451564

58

APPENDICES
Appendix A Ammonium Chloride Test [60]
REACTIONS:
Ca(OH)2 + 2NH4Cl CaCl2 + 2NH4OH
CaCl2 + (NH4)2C2O4 CaC2O4 + 2NH4Cl
CaC2O4 + H2SO4 H2C2O4 + CaSO4
5H2C2O4 + 2KMnO4 + 3H2SO4 10CO2 + K2SO4 + 2MnSO4 + 8H2O

REAGENTS:

Ammonium Chloride (NH4Cl)

Ammonium Oxalate ((NH4)2C2O4)

4N Sulfuric Acid (H2SO4)

0.1N Potassium Permanganate (KMnO4)

EQUIPMENT:

Two 400mL Beakers

Balance

9cm Buchner Funnel and Whatman 541 Filter Paper

50mL Automatic Burette

Filtering Flask

Hot Plate

Stirring Rod

59

PROCEDURE:
1. Weigh out 10g of oven dried mud sampled directly from the mud filter. Weight = WS g.
2. Dilute with 100mL distilled water in a 400mL beaker.
3. Add 10g of ammonium chloride and stir well four times over 30min.
4. Prepare a filter flask with a 9cm Buchner funnel containing a wetted 541 filter paper.
5. Filter the sample and rinse the beaker into the funnel with 4 25mL portions of distilled water.
6. Pour the filtrate into another 400mL beaker and heat to boiling on the hot plate.
7. Remove from the hot plate and add 2g of ammonium oxalate.
8. Heat to boiling again.
9. Remove from the hot plate and filter on a 541 filter paper and wash with hot distilled water.
10. Wash the precipitate from the filter into a 400mL beaker.
11. Add enough 4N sulfuric acid to dissolve the precipitate.
12. Heat the mixture to about 90C and titrate to a faint pink with 0.1N potassium permanganate.
13. Push the filter paper into the solution and break it up with a stirring rod.
14. Continue titrating to a faint permanent pink with 0.1N potassium permanganate without rezeroing the burette. Burette reading = VS mL.

CALCULATIONS:
( )
( )

(
(

)(

)(

)
(
)

60

Appendix B Additional Sound Theories


There are six mechanisms associated with sound interactions in colloidal systems [48].

Viscous mechanism explains shear waves generated by particle oscillating under acoustic
pressure, it causes energy loss due to shear friction, dominate for particles under 3 microns;

Thermal mechanism explains temperature gradients on particle surfaces, which is the


dominate dissipation of acoustic energy for soft particles;

Scattering mechanism explains the redirection of acoustic energy flow similar to light
scattering, and it is most important to attenuation for particles larger than 3 microns;

Intrinsic mechanism refer to acoustic energy loss between homogenous phases, and it is
important when overall attenuation is low and when volume fraction of dispersion is low;

Structural mechanism explain additional energy dissipation caused by inter-particle bonds;

Electrokinetic mechanism explains interactions within the double layer of the particle and
is negligible to acoustic attenuation measurement.

A well-defined theory need to characterize these relationships for both wide frequency ranges
and dispersed system of high volume fraction. However, previous studies had proven it was
difficult to accomplish this task as the commonly used ECAH theory [61, 62] from Epstein,
Carhart, Allegra and Hawley only meet the first requirement but not the second. The ECAH
theory was constructed in two stages, starting with observations at a single particle level and then
expanding to the macroscopic level. The main disadvantage of ECAH was its neglect for
interactions between particles. Since even at macroscopic level, the theory considered the
particles to be isolated. To expand on the ECAH theory, it was necessary to simplify the main
mechanisms for colloid sound interactions as outlined in Section 2.3.

61

Appendix C ABC Titration (with Automatic Titrator)


REAGENTS:

Barium Chloride (BaCl2)

Formaldehyde (CH2O)

PROCEDURE:
1. Measure 10mL of green liquor sample into the beaker.
2. Fill up to 200mL with distilled water.
3. Add 3g of barium chloride to precipitate Na2CO3 and start stirrer.
4. Record endpoint (EP1) when pH is approximate 11.
5. Add 10mL of formaldehyde to precipitate Na2S.
6. Record endpoint (EP2 and EP3) when pH is approximately 9 and 4, respectively.

CALCULATION:
( )(
( )(
Note: NaOH and Na2CO3 expressed in g Na2O basis.

)
)

(
(

)
)

62

Appendix D Thermogravimetric Analysis (TGA) / Differential


Scanning Calorimetry (DSC)
EQUIPMENT:

TA Instrument SDT Q600

PROCEDURE:
1. Select temperature ramp test with starting temperature of 50C, final temperature of 900C
and heating rate of 20C/min.
2. Use air purge stream at flow rate of 100mL/min.
3. Calibrate the equipment with an empty alumina crucible.
4. Fill crucible with approximately 10mg of sample and start the analysis.

63

Appendix E Zeta Potential Data for Wet and Dry Lime Mud
Note: Zeta potentials are expressed in mV and liming ratios are expressed in CaO/Na2CO3
Liming
Ratio
0.40
0.41
0.42
0.43
0.44
0.45
0.46
0.47
0.48
0.49
0.50
0.51
0.52
0.53
0.54
0.55
0.56
0.57
0.58
0.59
0.60
0.61
0.62
0.63
0.64
0.65
0.66
0.67
0.68
0.69
0.70
0.71
0.72
0.73

ZP (Wet ZP (Dry
Mud)
Mud)
-13.1
-20.9
-11.0
-17.4
-11.6
-18.6
-11.8
-16.7
-11.9
-18.9
-10.4
-19.0
-12.6
-19.3
-11.3
-17.1
-11.6
-18.8
-11.2
-19.0
-13.0
-17.9
-12.1
-18.0
-11.3
-17.0
-11.7
-19.8
-12.1
-19.3
-11.9
-19.8
-13.7
-17.8
-11.4
-17.2
-12.0
-18.5
-11.8
-14.8
-11.4
-16.4
-10.3
-14.8
-12.5
-16.8
-11.1
-16.3
-9.85
-14.6
-10.5
-14.7
-11.5
-17.0
-8.19
-16.1
-9.34
-15.4
-9.34
-15.9
-9.51
-16.6
-9.19
-16.7
-11.5
-17.7
-8.87
-16.3

Liming
Ratio
0.74
0.75
0.76
0.77
0.78
0.79
0.80
0.81
0.82
0.83
0.84
0.85
0.86
0.87
0.88
0.89
0.90
0.91
0.92
0.93
0.94
0.95
0.96
0.97
0.98
0.99
1.00
1.01
1.02
1.03
1.04
1.05
1.06
1.07

ZP (Wet ZP (Dry
Mud)
Mud)
-9.63
-15.5
-8.63
-14.8
-8.70
-15.6
-8.41
-14.1
-6.15
-13.9
-9.50
-16.1
-6.05
-13.9
-2.56
-16.8
-6.72
-14.0
6.97
-14.8
-8.37
-14.6
5.83
-14.3
4.53
-11.6
-4.90
-13.0
12.3
-13.0
5.37
-12.6
7.51
-14.0
-6.00
-11.9
-1.30
-13.1
2.62
-15.5
-3.18
-9.80
11.1
-7.94
11.8
-7.88
11.1
-9.04
11.2
-10.2
9.56
14.0
12.8
14.7
10.5
-6.63
11.3
13.9
9.26
14.3
12.4
16.2
10.2
16.4
10.3
16.0
15.7
17.1

Liming
Ratio
1.08
1.09
1.10
1.11
1.12
1.13
1.14
1.15
1.16
1.17
1.18
1.19
1.20
1.21
1.22
1.23
1.24
1.25
1.26
1.27
1.28
1.29
1.30
1.31
1.32
1.33
1.34
1.35
1.36
1.37
1.38
1.39

ZP (Wet ZP (Dry
Mud)
Mud)
11.0
15.5
12.0
16.2
10.7
15.7
11.6
16.8
11.4
17.0
12.5
12.0
10.3
16.0
8.55
16.0
9.72
17.4
9.16
14.7
8.96
14.3
8.62
15.6
6.87
14.5
8.66
12.6
8.20
13.9
7.46
11.2
8.99
14.2
7.15
13.4
10.3
13.2
7.22
14.1
8.79
13.2
7.25
13.7
9.26
11.0
8.15
12.0
9.95
12.3
7.72
11.6
9.52
14.1
8.68
11.4
11.8
13.1
7.09
13.1
9.33
11.3
11.8
14.7

64

Appendix F Zeta Potential Data for Stored Sample


Note: Zeta potentials are expressed in mV and liming ratios are expressed in CaO/Na2CO3
Liming Ratio /
Storage Time
Instant
2 hours
3 hours
4 hours
1 day
1 week
2 weeks
3 weeks
4 weeks

0.75

0.80

0.85

0.90

0.95

1.00

1.05

1.10

1.15

-16.3
-16.2
-16.3
-13.0
-11.7
-6.32
3.87
7.99
3.70

-12.9
-13.6
-13.8
-11.2
-10.7
-6.29
4.12
9.03
7.86

-13.3
-12.5
-12.9
-12.1
-7.51
3.39
11.4
11.7
11.9

-11.1
-9.53
-10.4
-10.4
0.830
6.72
14.2
12.6
13.4

-8.74
-7.42
-6.03
-6.32
-2.19
6.29
10.2
9.56
9.95

11.8
13.9
14.5
13.8
11.7
4.71
5.08
5.21
12.2

15.9
14.2
15.5
13.7
13.0
11.3
15.4
12.4
16.0

16.1
15.7
15.0
14.6
13.8
11.3
16.6
13.7
15.9

15.3
14.5
13.4
14.3
12.3
10.9
15.2
12.6
16.8

Vous aimerez peut-être aussi