alkyl ketene dimer and precipitated calcium carbonate...

112
Alkyl ketene dimer and precipitated calcium carbonate interactions in wet-end papermaking by Agatha Poraj-Kozminski Department of Chemical Engineering McGill University, Montreal March 2006 A thesis submitted to McGiII University in partial fulfillment of the requirements of the degree of Master of Engineering ©Agatha Poraj-Kozminski 2006

Upload: vanliem

Post on 20-Mar-2018

231 views

Category:

Documents


5 download

TRANSCRIPT

Alkyl ketene dimer and precipitated calcium carbonate interactions in wet-end papermaking

by

Agatha Poraj-Kozminski

Department of Chemical Engineering McGill University, Montreal

March 2006

A thesis submitted to McGiII University in partial fulfillment of the

requirements of the degree of Master of Engineering

©Agatha Poraj-Kozminski

2006

1+1 Library and Archives Canada

Bibliothèque et Archives Canada

Published Heritage Branch

Direction du Patrimoine de l'édition

395 Wellington Street Ottawa ON K1A ON4 Canada

395, rue Wellington Ottawa ON K1A ON4 Canada

NOTICE: The author has granted a non­exclusive license allowing Library and Archives Canada to reproduce, publish, archive, preserve, conserve, communicate to the public by telecommunication or on the Internet, loan, distribute and sell th es es worldwide, for commercial or non­commercial purposes, in microform, paper, electronic and/or any other formats.

The author retains copyright ownership and moral rights in this thesis. Neither the thesis nor substantial extracts from it may be printed or otherwise reproduced without the author's permission.

ln compliance with the Canadian Privacy Act some supporting forms may have been removed from this thesis.

While these forms may be included in the document page count, their removal does not represent any loss of content from the thesis.

• •• Canada

AVIS:

Your file Votre référence ISBN: 978-0-494-25006-8 Our file Notre référence ISBN: 978-0-494-25006-8

L'auteur a accordé une licence non exclusive permettant à la Bibliothèque et Archives Canada de reproduire, publier, archiver, sauvegarder, conserver, transmettre au public par télécommunication ou par l'Internet, prêter, distribuer et vendre des thèses partout dans le monde, à des fins commerciales ou autres, sur support microforme, papier, électronique et/ou autres formats.

L'auteur conserve la propriété du droit d'auteur et des droits moraux qui protège cette thèse. Ni la thèse ni des extraits substantiels de celle-ci ne doivent être imprimés ou autrement reproduits sans son autorisation.

Conformément à la loi canadienne sur la protection de la vie privée, quelques formulaires secondaires ont été enlevés de cette thèse.

Bien que ces formulaires aient inclus dans la pagination, il n'y aura aucun contenu manquant.

f/-'"

~ ..

ABSTRACT

This thesis investigates the interactions between alkyl ketene dimer (AKD)

and precipitated calcium carbonate (PCC). Although the mechanisms behind AKD

sizing and reactions with cellulose have been studied in-depth, methods describing

AKD retenti on are poorly understood. The aim of this research was to determine the

conditions and time-scale under which AKD and PCC heteroflocculate, and to

determine the influence of PCC on non-retained AKD. We also wanted to understand

the mechanisms behind AKD interactions with cP AM, and perform experiments on

the twin-wire sheet former. We outline a procedure for creating an AKD emulsion

coated by cationic starch and free of extraneous substances. We find that AKD and

PCC each homoflocculate, but no heteroflocculation occurs between the two

chemicals. This suggests that PCC and AKD do not directly interact. Instead, starch

removal from the AKD partic1e surface aids the homoflocculation of PCC (via

polymer bridging). This indicates that the alkaline environment is the most significant

factor contributing to AKD hydrolysis. More importantly, we find that AKD and PCC

are not likely to interact in the whitewater cycle. This refutes the general idea that

PCC lowers AKD retention. Adsorption kinetic experiments revealed that although

cationic AKD and cP AM do not flocculate, cP AM does increase the retention of

AKD by assisting its deposition onto fibers. Asymmetrical polymer bridging explains

both the increase in AKD retention and the behavior of AKD flocculation kinetics

with cP AM. Although the initial kinetics are nearly independent of cP AM dosage,

excess cP AM delays the achievement of maximum possible AKD retention due to

starch and cP AM re-conformation. Lastly, it was found that the addition of cP AM

results in an increase in the bond strength between AKD and fibers, thereby

eliminating AKD detachment. Studies on the twin-wire sheet former found that

cP AM increases the first-pass retenti on of AKD three-fold. The twin-wire former was

proven to be a useful tool for studying AKD retention.

ii

RESUMÉ

Le sujet de cette thèse porte sur l'étude des interactions entre le dimère de

cétène alkylé (AKD) et le carbonate de calcium précipité (PCC). Bien que les

mécanismes de l'encollage à l'AKD et de sa réaction avec la cellulose aient été

étudiés en profondeur, ceux qui gouvernent la rétention de l'AKD sont encore mal

définis. Nos recherches visaient à déterminer les conditions et l'échelle de temps sous

lesquelles l' AKD et le PCC hétéro-floculent et à déterminer l'effet du PCC sur

l'AKD non retenu. Nous voulions aussi comprendre les mécanismes de l'interaction

de l'AKD avec le polyacrylamide cationique (cP AM) et réaliser des expériences sur

la machine à p_apier à double toile. Nous décrivons un procédé pour créer une

émulsion d'AKD englobé par de l'amidon cationique sans aucune autre substance.

Nous avons observé que l'AKD et le PCC homo-floculent, mais qu'il n'y a pas

d'hétéro-floculation entre les deux produits. Ceci indique que le PCC et l'AKD

n'interagissent pas directement. Ce qui se produit plutôt est que l'amidon détaché de

la surface des particules d'AKD contribue à l'homo-floculation du PCC (par le biais

de ponts polymériques). Ceci indique qu'un environnement alcalin est le facteur

dominant qui contribue à l'hydrolyse de l'AKD. De plus grande importance est notre

découverte que l'AKD et le PCC sont peu enclin à interagir dans le cycle de l'eau

blanche. Ceci réfute la théorie que le PCC réduit la rétention de l'AKD. Des

expériences de cinétique d'adsorption ont révélé que même si l' AKD cationique et le

cP AM ne floculent pas, le cP AM augmente la rétention de l' AKD en contribuant à sa

déposition sur les fibres. La formation de ponts de polymères asymétriques explique à

la fois la hausse de la rétention de l' AKD et le comportement de la cinétique de

floculation de l'AKD avec le cPAM. Même si la cinétique initiale est indépendante

du dosage de cPAM, un excès de cPAM retarde l'atteinte du niveau maximal de

rétention à cause de la re-conformation de l'amidon et du cPAM. Enfin, nous avons

découvert que l'ajout de cPAM augmente la force des liens entre l'AKD et les fibres,

ce qui empêche le détachement de l'AKD. Nos expériences sur la machine à papier à

double toile démontrent que le cP AM multiplie par trois le niveau de rétention de

première passe de l'AKD. La machine à papier à double toile s'avère être un outil

utile pour l'étude de la rétention de l' AKD.

111

Acknowledgements

1 would like to thank Dr. Theo van de Ven and Dr. Reghan Hill for their continued

guidance and encouragement, and for their dedication to the project. They have made

me a better researcher, and 1 am etemally grateful to them for their support.

My sincere thanks to my colleagues at the Pulp and Paper Center, and particularly to

the members ofmy research group: Lojza, Marcius, Jimmy, Prasad, and Meng.

Many thanks to Louis Godbout, Dr. Alince, Dr. Petlicki, Lou Cusmich, Helen

Campbell and Ed Siliauskas, for their invaluable contributions to this project. A

special thanks to Christopher Hammock, creator of the laboratory twin-wire former.

My sincere gratitude to B. Sithole, J. Pimentel, and A. Gagne at Paprican, Pointe

Claire, for their help with AKD analysis.

And finally, thank you to my parents and friends, and to my husband Christopher, for

all your love and support, and without whom, none of this would have been possible.

IV

TABLE OF CONTENTS

Abstract

Resumé

Acknowledgements

Contents

List of Figures

List of Tables

Chapter 1: Introduction

1.1 Background

1.2 Objectives of the Thesis

1.3 Literature Review

1.3.1 Major Chemical Components

1.3.1.1 Alkyl ketene dimer (AKD)

1.3.1.2 Precipitated Calcium Carbonate (PCC)

1.3 .1.3 Retention Aids

1.3.1.4 Fibers

1.3.2 Retention Mechanisms

1.4 References

Chapter 2: AKD and PCC Interactions

2.1 Abstract

2.2 Introduction

v

ii

111

iv

v

ix

XlI

2

5

7

7

8

10

11

12

13

14

17

18

r--, 2.3 Emulsification of AKD Stabilized with Cationic Starch 19

2.3.1 Materials and Methods 20

2.4 Emulsion Analysis 21

2.4.1 Particle Size Analysis 21

2.4.2 Electrophoretic Mobility Analysis 23

2.4.3 Effect of Starch Concentration and Ratio 25

2.4.4 Comparison to Commercial Emulsion 27

2.5 Photometrie Dispersion Ana1yzer Experiments 27

2.5.1 Materials and Methods 28

2.5.2 PDA Parameters 30

2.5.3 Results and Discussion 32

2.5.3.1 PCC Homoflocculation 32

2.5.3.2 PCC and AKD F10cculation 32

2.5.3.3 PCC and Starch Flocculation 34

2.5.3.4 Effects of Starch filtration 36

2.6 Scanning Electron Microscopy Analysis 39

2.7 Conclusions 42

2.8 Acknowledgements 43

2.9 References 44

VI

Chapter 3: AKD Interactions with cP AM and Fibers

3.1 Abstract

3.2 Introduction

3.3 Materials and Methods

3.4 Materials

3.4.1 Methods: Photometric Dispersion Analyzer

3.4.2 Methods: Adsorption Kinetics

3.5 Results

3.5.1 AKD Flocculation

3.5.2 Cationic AKD Deposition on Fibers

3.5.3 Anionic AKD Deposition on Fibers

3.6 Conclusions

3.7 Acknowledgements

3.8 References

Chapter 4: Interactions of AKD and Other Papermaking Additives on a Laboratory Twin-wire Sheet Former Machine

4.1 Abstract

4.2 Introduction

4.3 Machine Description

4.4 Materials and Methods

4.5 Analysis of Results

4.6 Outline of Experiments

vu

47

48

50

50

51

52

53

53

54

60

62

62

64

66

67

67

72

74

75

4.6.1 Reproducibility Experiments 75

4.6.2 PCC Retention with Cationic Retention Aid 77

4.6.3 First-Pass Retention vs. Second-Pass Retention with AKD 78

4.6.4 AKD Retention at Various Additions 79

4.6.5 AKD Retention with cP AM 79

4.7 Conclusions 80

4.8 Acknowledgements 81

4.9 References 82

Chapter 5: Conclusions

5.1 Overview 84

5.2 Recommendations for Further Work 87

Appendix A. Principles of Flocculation 89

Appendix B. Starch Clustering and Effects of Shear and Sonication 93

Appendix C. Twin-Wire Sheet Former Machine (TWF) Instructions 96

V111

LIST OF FIGURES

CHAPTERI

Figure 1.1: Typical Approach System of a Paper Machine 3

Figure 1.2: AKD p-esterification with cellulose 8

Figure 1.3: Hydrolysis of AKD to palmitone 9

CHAPTERII

Figure 2.1: Partic1e-size distribution of AKD emulsion 22

Figure 2.2: SEM image of AKD emulsion at 1O,000X magnification 23

Figure 2.3: Theoretical and experimental values of AKD partic1e diameter 25

Figure 2.4: PDA arrangement with circulation and pump 29

Figure 2.5: RATIO of dispersed versus aggregated partic1es 30

Figure 2.6 a: Effect of varying flowrate on the floc formation in PDA for a stirring rate of 150 rpm 31

Figure 2.6 b: Effect ofvarying stirring speed in beaker on the floc formation in PDA for a flowrate of 150 mL/min. 32

Figure 2.7: PCC flocculation with AKD emulsion 34

Figure 2.8: PCC flocculation with sonicated starch (low starch additions) 35

Figure 2.9: PCC flocculation with sonicated starch (high starch additions) 35

Figure 2.10: Schematics of energy of interaction between two PCC partic1es: a) partially coated by starch- van der Waals and electrostatic attraction b) fully coated- van der Waals attractions and steric repu1sions 36

Figure 2.11: Successive filtration schematic of AKD drop lets and starch stabilizer 37

Figure 2.12: PCC flocculation with "AKD2" (AKD filtered twice) 39

Figure 2.13: SEM image of PCC partic1es 39

IX

Figure 2.14: SEM image of AKD emulsion 40

Figure 2.15(a,b,c): PCC and AKD at a 1:1 Ratio 41

Figure 2.16: SEM image of AKD emulsion after filtering away free starch; AKD drop lets aggregate 43

CHAPTERIII

Figure 3.1: Photograph of vials with various quantities ofcPAM; Amounts are given as mg cPAM/g anionic AKD. 53

Figure 3.2: AKD deposition onto fibers, at 10 mg AKD/g fiber addition, and varying cPAM addition. Inset: [mg cPAMlg fibers] 54

Figure 3.3: AKD deposition to fiber surface 55

Figure 3.4: AKD deposition on fibers a) low cPAM dosages; b) high cPAM dosages 56

Figure 3.5: Asymmetric Polymer Bridging Mechanism 57

Figure 3.6: Increased Bond Strength by Starch Re-conformation a) Initial adsorption of starch to cP AM b) Spreading of starch 58

Figure 3.7: Adsorption Kinetics of Anionic AKD emulsion to fibers, in the presence of cP AM.. 61

CHAPTERIV

Figure 4.1: Laboratory Twin-Wire Sheet Former 68

Figure 4.2: Headbox Approach System of Twin-Wire Sheet Fomier 69

Figure 4.3: Drainage Section of Twin-Wire Sheet Former 69

Figure 4.4: Photograph ofTwin-Wire Sheet Former 71

Figure 4.5: Photograph of a paper sheet produced 71

x

Figure 4.6: Additive locations and residence times on the TWF 73

Figure 4.7: Water drainage along the TWF for three duplicate runs 76

Figure 4.8: Filler retention in the wet-web for three duplicate runs on the TWF 77

Figure 4.9: Influence of cP AM on PCC retenti on on the TWF 78

Figure 4.10: The effect of cP AM on AKD retenti on 80

Xl

LIST OF TABLES

Table 1.1: Typical Concentrations and Sizes of Papermaking Chemicals for the Production of Fine Papers 7

Table 2.1: Emulsion Characterization 21

Table 2.2: Electrophoretic Mobilities and Partic1e Diameters of AKD Emulsions 38

.~ ..

XlI

CHAPTER 1

INTRODUCTION

1.1 Background

Paper producers are continually searching for new and more effective ways to

utilize chemicals and reduce operating costs without compromising paper quality.

Since the operating speeds and manufacturing volumes of paper machines are

constantly increasing, the challenge to produce paper under such conditions is highly

reliant on additives. In the wet-end section of a paper machine, many chemicals are

needed to enhance the properties ofthe paper and to improve its quality. Among these

chemicals are sizing agents, which produce hydrophobicity in paper. To study the

behavior of alkyl ketene dimer (AKD), a common sizing agent, we examine its

chemical interactions in the forming section and short-circulation whitewater loop of

a paper machine. In this study, we focus on the processes and chemicals used to

manufacture fine office paper from kraft pulp fibers.

The papermaking chemicals studied here are added to the approach system of

the paper machine (Figure 1.1). The approach system consists of a fan-pump loop

where the di lute pulp suspension is metered, diluted, and additives are added. Flow

enters the machine chest and moves to the "stuff box", which regulates the feed by

ensuring a constant head. A valve before the fan pump controls the basis weight

entering the paper machine. Both the flow from the machine chest, and whitewater

retuming from the wire pit, enter the fan pump. The fan pump is the Iargest pump in

the paper machine system, and dictates much of the shear delivered onto the paper

machine. Next, the suspension is carried to the centrifugaI c1eaners and pressure

screens, before it flows into the pressurized headbox and onto the paper machine [1].

2

CENTRIFUGAL CLEANERS

MACHINE CHEST

PRESSURE SCREENS

STUFF BOX

WlRE HEAD BOX

WIREPIT

BASIS WEIGHT VALVE

Figure 1.1: Typical Approach System of a Paper Machine (adapted from Smook [1] )

Since paper mills are being urged to close their whitewater loops, more of the

dissolved and colloidal substances (DeS) or "anionic trash", such as fines and other

debris, are retuming to the headbox in the short-circulation whitewater loop. As a

result, the properties of the water being fed into the paper machine are altered due to

the elevated levels of electrolytes. This has a severe impact on the efficiency of other

chemicals added to enhance the quality and properties of the paper. Several studies

have shown that the behavior of additives is determined, in part, by the properties of

the water. Since much of this water input into the paper machine now cornes from the

whitewater, it is important to know how these extra substances affect the water

chemistry [2, 3].

3

Closure of the whitewater cycle releases less water effluent to the

environment, consequently reducing the costs associated with the treatment of such

effluent. However, reusing this water leads to problems within the paper machine,

such as corrosion of the machinery due to high salt concentrations [4]. Another reason

to close the whitewater system is to reduce the loss of fibers, which are valuable due

to the costs associated with processing them. The whitewater will always contain a

fraction of fibers, so the volume lost should be minimized. It has also been shown that

a build-up of electrolytes results in increased conductivity, anionic trash content,

sticky deposition, and cationic polymer demand, and has an adverse effect on paper

quality [5]. For example, a build-up of negatively charged DCS can degrade the

performance of cationic retention-aid systems [6].

The decrease in first-pass retenti on of sizing agents results in paper that is less

sized. This has led to an increased use of sizing agents and retention aids to offset the

effects of DCS. Though this may be a temporary remedy, the long-term solution

requires an understanding of how to increase the first-pass retention of these

chemicals to optimize their use. The addition sequence and residence times of

additives employed in paper mills have been selected from previous experience and

knowledge. However, comprehension of the chemical reactions remains crucial in a

complete understanding of optimization.

Single-pass retenti on may be one of the most important factors influencing

paper quality and paper machine operation [1]. Many chemicals become less effective

after recirculation in the whitewater cycle. In fact, it has been shown that in

subsequent passes through the paper machine, AKD contributes significantly less to

4

sizing than in the first pass. As mentioned later, the decrease in sizing is a direct result

of AKD reactions with water, which renders it non-reactive towards cellulose [7].

1.2 Objectives of the thesis

Since AKD has a low first-pass retention (~40%), a large portion of it enters

the whitewater cycle and has ample time to interact with many types of particles,

including dissolved and colloidal substances (DCS) and precipitated calcium

carbonate (PCC), a common filler. It is hypothesized that in the recirculation loop

AKD may reverse its charge by adsorbing anionic material. When passing through

the point of zero charge, AKD particles have the ability to aggregate with other AKD

particles. Other studies show that in the presence of water, AKD reacts to form a non­

reactive ketone that does not contribute to sizing [8]. In addition, excessive usage of

AKD contaminates the whitewater system and may cause operating problems by

forming deposits in the machinery [9]. However, it has not been shown whether there

exist any significant interactions between AKD and PCC, nor has any of this work

been performed on a larger scale than in the laboratory. The experimental methods

employed by all previous researchers involved the dynamic drainage jar or other

small-scale methods.

Several theories exist to explain how AKD reacts with cellulose to form a

hydrophobie monolayer in the paper structure. Extensive literature exists examining if

a covalent bond with cellulose is necessary for sizing [8, 10-13]. However, few

outline the mechanism of AKD retention itself. This process occurs in the forming

section of papermaking and is critical to the subsequent AKD reaction with cellulose.

5

The objectives ofthis work are three-fold:

(1) To detennine the conditions and time-scale under which AKD and PCC will

heteroflocculate, and relate these findings to the behavior of AKD and PCC in

the whitewater cycle;

(2) To study the behavior of AKD under wet-end conditions; in particular, to

detennine the influence of PCC on non-retained AKD, and to understand the

mechanisms behind AKD's retention and its interactions with PCC and cationic

polyacrylamide (cP AM);

(3) To study retention of AKD on a laboratory twin'-wire sheet fonner to acquire a

more accurate interpretation of real interactions on a paper machine and to

detennine the differences in a first-pass versus second-pass retenti on value.

A discussion of these objectives comprises the contents of the three subsequent

chapters of this thesis. The remainder of this chapter presents a relevant literature

review. An overview of wet-end operations in a papennaking machine is discussed,

followed by a survey of AKD properties and interactions. In addition, PCC and

cP AM data are provided to help understand the wet-end system as a whole. Chapter 2

presents the interactions between AKD and PCC investigated on a small scale.

Chapter 3 covers the interactions of AKD with cationic polyacrylamide. Chapter 4

discusses the experiments perfonned on the laboratory twin-wire sheet fonner

machine. Finally, Chapter 5 concludes with a recapitulation of results and

recommendations for future work.

6

1.3 Literature Review

1.3.1 Major Chemical Components

In the wet-end section of a paper machine, many chemicals are added to

control properties and improve the quality of the paper. Table 1.1 provides a

summary of several chemicals used in the production of fine kraft paper, listed with

their typical concentrations and average sizes.

Table 1.1: Typical Concentrations and Sizes ofPapermaking Chemicals for the Production of Fine Pa ers

Component

Bleached Softwood Kraft (BSK) Pulp

starch

w/w% ofmass

0.5-1.0 **

0.03-1.00

Particle Diameter

30flm; (up to 3 mm length)

~80-100 nm '--'-'~'~'~"~'--'-"'--"~~~~~~-"~-"--~----I

5-10 <76flm

5-25 ~0.2-3 flm

0.05-0.2 ~0.2-2 flm

~0.03 ~100nm

* * wt. fibers/wt. water suspension

The aperture of the forming wire is typically 100 flm. Therefore, the material

that passes through this fabric and enters the whitewater loop is fillers, fines, AKD,

and retention aids. This is disadvantageous for the papermaking process, since by

entering the short circulation loop, the first-pass retention (FPR) of these materials

decreases. In the whitewater, they have time to interact with many substances that

may decrease their effectiveness. A discussion of the problems that may arise from

these reactions is found further in this chapter.

7

1.3.1.1 Alkyl ketene dimer (AKD)

AKD has been used as a sizing agent (to produce hydrophobicity in paper)

since the 1950's [14]. It is composed oftwo long carbohydrate chains, ranging from

14 to 18 carbon molecules each, and a lactone ring in the center. AKD binds to

cellulose fibers via a p-esterification reaction (Figure 1.2) [10]. This p-ester covalent

bond was found to be vital for effective sizing [8]. The two hydrophobic tails form a

hydrophobic layer on top of the sheet, retarding water penetration.

H 1

~-CH=H-Rz o-c

Il o

~ R -CH:z-c -CH-C-O

t Il Il o 0

Figure 1.2: AKD p-esterification with cellulose

AKD forms a stable colloidal emulsion in the presence of stabilizing

polymers, such as cationic starch. This gives the emulsion an overall positive charge,

enabling AKD particles to adsorb onto the negatively charged fibers. Often, another

polymer with the opposite charge is added to give the emulsion amphoteric

properties. AKD emulsions are stored with a dry solids content of 6-15%. AKD

particle sizes are in the range 0.2-2 microns, and addition levels are equivalent to

0.05-0.2% ofpure AKD based on fiber [14].

The emulsion is typically kept at low pH and cooler temperature (+3°C) to

prevent hydrolysis. Under these conditions, it can remain stable for several months.

As the pH increases, AKD begins to react with water (Figure 1.3). At a neutral pH,

AKD is hydrolyzed quickly (4%AKD/hr) and at a basic pH of 10.4, it is completely

8

hydrolyzed in less than a day [5]. AKD hydrolysis is affected by pH, as well as

temperature. Using NMR analysis, AKD was found to hydrolyze to its ketone form

(palmitone) in a short period. More importantly, it was found that this ketone does not

contribute to sizing, since its ketone bond makes it un-reactive [8]. This ketone has

also been found to cause deposition problems in paper machines [5].

Figure 1.3: Hydrolysis of AKD to palmitone

The acceleration in AKD hydrolysis can be attributed to CO{ and HC03-

ions. One the ory (not fully developed) states that PCC accelerates hydrolysis by

destabilizing the AKD. Other theories describe AKD spreading on the PCC surface,

or thirdly, sorne claim that the higher pH inside the PCC (due to Ca(OH)z ) catalyzes

the hydrolysis of AKD [5]. One study shows that AKD molecules adsorbed on

precipitated calcium carbonate (PCC) filler convert to palmitone [15]. According to a

second researcher, AKD adsorbs to filler due to its high surface area [9]. Another

states that it is the high affinity of AKD to filler responsible for AKD degradation [8].

Other tests indicate there is little or no interaction between the filler and AKD

particles at aIl [16]. With these contradictory results, it is difficult to draw any real

conclusions.

To counteract the negative effects, increasing the AKD/PCC ratio was found

to reduce the rate and extent of AKD hydrolysis. For this reason, it has become the

trend to increase the quantity of AKD in mill operations.

9

It is critical to maximize the first-pass retenti on of AKD in order to optimize

sizing. Intermediate reactions interfere with AKD depositing onto the fibers, in a

manner that reduces hydrophobicity. Solvent extraction tests indicate that sizing

occurs only when there is a covalent bond between the size molecule and cellulose

[8]. This link is necessary to ensure the proper alignment of the molecule so the

hydrophobic tails are exposed to the air interface [17]. Furthermore, it may be

possible that the bond serves as an anchor to prevent evaporation and detachment of

the molecule at higher temperatures. In general, AKD retention is a dynamic process

that involves attachment, detachment, and charge neutralization and reversaI [16].

1.3 .1.2 Precipitated Calcium Carbonate (PCC)

Precipitated calcium carbonate has become a popular filler since the change to

neutral or alkaline papermaking system was implemented. PCC is an example of a

mineraI filler that is added to the fumish to improve the optical and physical

properties of the sheet. The particles fill the spaces between the fibers, resulting in a

denser and smoother sheet. Since fillers are less expensive than fibers, they also

decrease the overall cost of paper manufacturing [1]. However, fillers weaken the

paper strength by interfering with fiber-fiber bonding and by transferring polymer

from fiber to filler surfaces [18]. Therefore, the amount of filler added to the sheet is

limited by the resulting reduction in strength and sizing quality. Most fine papers

contain anywhere between 5-25% wt filler/wt fibers.

Since PCC is soluble at lower pH, it can only be used in a neutral or alkaline

environment. CaC03 is partially soluble in water and the Ca2+ ions preferentially

10

adsorb to the crystal surface. In distilled water PCC carries a positive charge, but in

tap and mill water PCC has a negative charge due to the presence of impurities in the

water [19]. To improve retenti on by adsorption to the negatively charged fibers, a

cationic retenti on aid is added to promote PCC retention.

There are several explanations as to how PCC partic1es can be retained in the

sheet. First, PCC may deposit onto fibers by electrostatic attraction, or by charge

neutralization from polymer addition. Polymer bridging between PCC partic1es and

pulp fibers may also occur. Lastly, PCC partic1es may coagulate to form aggregates

that remain in the forming web via mechanical entrapment [20]. The maximum

amount adsorbed has been measured previously and is reported to be 250 mg PCC per

gram of fibers. This is the amount that produces a mono layer on the fiber surface,

based on a Im2 /g specific surface area offibers [21].

The surface charge of PCC is very dependent on PCC concentration, pH, and

impurities in the medium [19]. Upon addition of PCC to water (distilled or tap),

partic1es begin to aggregate, suggesting the system is unstable. These aggregates are

weak and can be broken up by shear, which under papermaking conditions, can reach

up to 104 S-1 [22]. The stability of PCC suspensions, as well as rates of flocculation,

can be measured by turbidity tests using the Photometric Dispersion Analyzer. More

on this testing procedure is presented in Chapter 2.

1.3.1.3 Retention Aids

The use of retention aids in a paper operation has many advantages. They act

to improve fines retenti on, increase drainage, improve filler retention, control

11

deposits and improve operations [6]. In the production of fine kraft paper using PCC,

cationic polyacrylamide (cP AM) is the most common retenti on aid. Cationic

polyacrylamide is an organic compound with a linear structure and a molecular

weight in the range of3-6xl06 Da.

Highly charged cationic polyelectrolytes promote deposition of PCC on fibers

by charge modification, while ones with a high molecular weight and low charge

density act by a bridging mechanism [23]. In this study, we use cPAM with a degree

of substitution of 20% and a high molecular weight, indicating a bridging polymer.

The addition of a retenti on aid has a dramatic effect on the retention of

precipitated calcium carbonate. Since both contaminated PCC 1 and fibers have a net

negative charge, they tend to repel each other. Cationic polyacrylamide imparts an

attraction (via bridging), thereby improving PCC retention. It is important to

determine the optimal quantity of cP AM to add since quantities that are too high

hinder sheet formation [6].

1.3.1.4 Fibers

Cellulose is negative1y charged. The negative charge originates from the

carboxyl groups that dissociate to various extents at different pHs. Under normal

papermaking conditions (~pH 7), the carboxyl and sulphonic acid groups are the

major contributors to the fiber charge [24].

A fraction (~1O-20%) of the fiber mass is comprised of fines. By definition,

these are the smaller segments of fibers whose diameters are less than 76 microns.

l "Contaminated" PCC refers to the particles in process or tap water whose charge has been influenced by impurities in the water.

12

Fines have a higher specific surface area (~10 m2jg) [25] than fibers (1 m2jg), which

gives them the capacity to adsorb much more material per unit surface area. Because

they are smaller than the fabric aperture, fines commonly drain through without being

retained in the sheet, taking sorne of the additives with them.

1.3.2 Retention Mechanisms

AKD retenti on can be improved by either physical or chemical methods. For

example, Mattsson [26] suggests that AKD retention can be improved physically by

adding a chemical to induce AKD molecules to aggregate prior to addition into the

approach system. This method of "preflocculation" produces aggregates that can

grow to a size up to 30 microns in diameter. In such a way, the AKD flocs are

retained in the sheet since they are less likely to pass through the wire. Tests

demonstrated that on a laboratory scale (employing the Britt Dynamic Drainage Jar),

AKD molecules that formed aggregates had a significant improvement in single-pass

retention. However, the question arose whether these aggregates would not cause an

uneven distribution in sizing on the paper sheet [26].

Instead, we consider the chemical interactions that occur between the

additives and fibers that allow us to optimize the retention. By understandingthe

interactions that occur, we can predict the effects of changing the parameters. The

intent of this study is to find the fundamental mechanisms responsible for the

interactions.

13

1.4 References

[1] Smook, G.A., "Handbook for Pulp & Paper Technologists", Angus Wilde Publications, Vancouver, 2nd Ed., (1992).

[2] Donat, V., van de Ven, T.G.M. and Paris, J., "Distribution of dissolved and colloidal substances in the forming and press sections of a paper machine", Journal of Pulp and Paper Science, 29(9) 294-298 (2003).

[3] Huber, P., Carre, B., Mauret, E. and Roux, J.-C., "The influence of fine elements build-up in the short-circulation on fibre flocculation." Preprints -International Paper and Coating Chemistry Symposium, 5th, Montreal, QC, Canada, (2003).

[4] Gavelin, G., "Paper Machine Design and Operation", Angus Wilde Publications, Vancouver, (1998).

[5] Jiang, H. and Deng, Y., "The effects ofinorganic salts and precipitated calcium carbonate filler on the hydrolysis kinetics of alkyl ketene dimer", Journal ofPulp and Paper Science, 26(6) 208-213 (2000).

[6] Allen, L.H., Polverari, M., Levesque, B. and Francis, W., "Effects of system c10sure on retention- and drainage-aid performance in TMP newsprint manufacture", Tappi Journal, 188-195 (1999).

[7] Mattsson, R., Sterte, J. and Odberg, L."Colloidal Stability of AKD dispersions." in The Science ofPapermaking: Transactions of the 12th Fundamental Research Symposium Ed.Baker.Oxford, UK, (2001). 393-415.

[8] Bottorff, K.J., "AKD sizing mechanism: A more definitive description", Tappi Journal, 77(4) 105-16 (1994).

[9] Esser, A. and Ettl, R."On the mechanism ofsizing with alkyl ketene dimer (AKD): physico-chemical aspects of AKD retenti on and sizing efficiency." in Fundamentals ofPapermaking Materials, Transactions of the 11th Fundamental Research Symposium Cambridge, UK, (1997). 997-1020.

[10] Bottorff, K.J. and Sullivan, M.J., "New insights into the alkylketene dimer (AKD) sizing mechanism", Nordic Pulp & Paper Research Journal, 8(1) 86-95 (1993).

[11] Isogai, A., "Effect of cationic polymer addition on retenti on of alkylketene dimer", Journal ofPulp and Paper Science, 23(6) 276-281 (1997).

[12] Isogai, A., "Stability of AKD-Cellulose B-Ketoester Bonds to Various Treatments", Journal of Pulp and Paper Science, 26(9) 330-334 (2000).

[13] Lindstrom, T. and Soderberg, G., "On the mechanism ofsizing with alkyl­ketene dimers, part 1. Studies on the amount of alkyl-ketene dimer required forsizing different pulps", Nordic Pulp and Paper Research Journal, 1(1) 26-33 (1986).

[14] Roberts, J.C."Neutral and Alkaline Sizing" in Paper Chemistry Ed.Roberts, J.C. New York: Blackie Chapman & Hall, (1991).

[15] Scott, W.E., "Wet End Chemistry", Tappi Press, Atlanta, (1996). [16] Champ, S. and Ettl, R., "The dynamics of Alkyl Ketene Dimer (AKD)

retention." Preprints -5th International Paper and Coating Chemistry Symposium, Montreal, QC, Canada, 285-291 (2003).

14

[17] Yu, L. and Garnier, G., "The role ofvapour deposition during internaI sizing: a comparative study between ASA and AKD", Journal of Puip and Paper Science, 28(10) 327-331 (2002).

[18] Cho, B.-U., Garnier, G. and van de Ven, T.G.M., "Parameters affecting paper formation on a pilot fourdrinier using cP AMlbentonite retenti on aids", Preprints -5th International Paper and Coating Chemistry Symposium, Montreal, QC, Canada, 193-200 (2003).

[19] Vanerek, A., Alince, B. and Van De Ven, T.G.M., "Colloidal behavior of ground and precipitated calcium carbonate fillers: effects of cationic polyelectrolytes and water quality", Journal of Pulp and Paper Science, 26(4) 135-139 (2000).

[20] Mitsui, K., "Mechanism of Fines Retention and Drainage with a Polyacrilamide/Bentonite Retention Aid", thesis, Chemical Engineering, McGill University, (2000).

[21] Alince, B., "Time factor in pigment retention", Tappi Journal, 79(3) 291-294 (1996).

[22] Tarn Doo, P.A., Kerekes, R.J. and Pelton, R., "Estimates of Maximum Hydrodynamic Shear Stresses on Fibre Surfaces in Paper machine Wet End Flows and in Laboratory Drainage Testers", Journal of Pulp and Paper Science, 10(4) 80-88 (1984).

[23] Vanerek, A., Alince, B. and van De Ven, T.G.M., "Interaction of calcium carbonate fillers with pulp fibers: effect of surface charge and cationic polyelectrolytes", Journal of Puip and Paper Science, 26(9) 317-322 (2000).

[24] Lindstrom, T. "Electrokinetics of the Papermaking Industry" in Paper Chemistry Ed.Roberts, J.C.New York: Blackie Chapman & Hall, (1991).

[25] Porubska, J., "Microstructure and properties ofTMP papers", M.Eng. thesis, Chemical Engineering, Mc Gill University, (2000).

[26] Mattsson, R., Sterte, J. and Odberg, L., "Sizing with pre-flocculated alkyl ketene dimer (AKD) dispersions", Nordic Pulp & Paper Research Journal, 17(3) 240-245 (2002).

15

CHAPTER2

AKD AND PCC INTERACTIONS

2.1 Abstract

Experiments with AKD and PCC were performed on a laboratory scale using

the Photometrie Dispersion Analyzer (PDA). What was originally thought tobe

heteroflocculation between PCC and AKD was found to be PCC homoflocculation by

starch, the AKD stabilizer. SEM photographs show that PCC and AKD each

homoflocculate; little to no heteroflocculation occurs. With the results from PDA

tests, this suggests that PCC and AKD do not directly interact. Instead, starch removal

from the AKD particle surface aids the homoflocculation of PCC (via polymer

bridging). Moreover, this starch transfer results in AKD particle instability, causing

the AKD particles to aggregate since they have been depleted of their stabilizing

polymer. Therefore, it can also be concluded that cationic starch has a higher affinity

for PCC than for AKD.

More importantly, we find that AKD and PCC are not likely to interact in the

whitewater cycle. Since no heteroflocculation occurs, this dismisses the idea that PCC

wi1110wer AKD single-pass retention.

17

2.2 Introduction

Extensive work has been done to investigate AKD and PCC interactions in

wet-end papermaking, though there are few conclusive results. It is stated in several

papers that PCC causes AKD hydrolysis, leading to a great loss of sizing, and causing

size reversion after drying [1-3]. However, the mechanism for these interactions and

the conditions under which they occur is unknown. Furthermore, no research has been

done to determine whether PCC affects the first-pass/second-pass retenti on of AKD.

Recirculation of AKD in the whitewater cycle may be detrimental to its sizing

efficiency. In this chapter, we explore ifthis is a direct effect of PCC.

The chemical interactions that occur in the wet-end depend on the order in

which chemicals are added, as well as their residence time prior to the headbox.

Typically, PCC is added near the first mix box overflow, and AKD is added next in

the second mix box [4]. Thus, AKD is likely to interact with PCC and fibers. The

interactions depend on the state of flocculation and the state of the suspension; that is,

whether PCC has adsorbed to the fibers, or formed flocs with the polymeric retention

aid. Several theories suggest what will occur upon addition of AKD. Many articles

discuss the interactions between AKD and fibers, AKD and PCC or cP AM, or how

one affects the other, but none of this research has been done on a wet-end paper

machine environment or an industrial time scale [5-8]. In this chapter, we present

results from the small-scale trials conducted to investigate the behavior of PCC and

AKD in similar laboratory environments. In Chapter 4, trials with AKD and PCC are

performed on a larger-scale using the laboratory twin-wire sheet former.

18

2.3 Emulsification of AKD Stabilized with Cationic Starch

Most large-scale AKD emulsion preparations consist of additional

components, such as surfactants, polymers, stabilizers, alcohols, etc. However, details

of the commercial formulations are usually not disclosed. Therefore, the researcher

cannot know if interactions of an emulsion are due to AKD or other chemicals within

the emulsion. To solve this problem, we developed a method to create a simple but

stable AKD emulsion in our laboratory. The emulsion consisted of only AKD and a

stabilizer (cationic starch). To design our method for emulsification, we referenced

several other researchers who produced their own AKD emulsions and developed

methods to properly analyze its properties [9, 10].

As mentioned in the previous chapter, alkyl ketene dimer (AKD) is a waxy

ester with two fatty chains, typically ranging in length from 14 to 18 carbon

molecules joined by a lactone group. It is extremely hydrophobic and insoluble in

water. Upon addition to fibers, the AKD can react with cellulose to form a f3-ester

covalent bond. However, this is a slow reaction and sizing only begins in the drying

section of a paper machine, continuing throughout the drying and curing process for

about a week afterwards [11].

For AKD to remain in emulsion, a stabilizer must be added to prevent the

AKD drop lets from coalescing during emulsification. Since AKD has a melting point

of about 500 e (depending on chain length), the emulsification is done at a higher

temperature so that the AKD is in its molten phase. Immediately after emulsification,

the solution is quenched below room temperature for the AKD to revert to a more

stable solid.

19

2.3.1 Materials and Methods

Alkyl ketene dimer (from Raisio Roe Lee) was obtained in pellet form. A 10 g

sample was melted to 60°C in preparation for sonication. A cationic (quatemary

amine-substituted) starch CATO-237 was obtained from National Starch in a powder

form. According to the manufacturer, this chemical has a nitrogen content (%N) of

0.36-0.44. This value can be converted using the following formula (2.1), which

equates to a degree of substitution of 0.043-0.053%. The degree of substitution

indicates the percentage of amine groups on the starch with a positive charge. This

translates into the overall cationic charge.

%DS = 162x%N 1400 - (152x %N)

(2.1)

The starch was dispersed in de-ionized water to either 1%, 2% or 4% starch

concentration. It was cooked for 40 minutes in a 93 oC water bath, as suggested by the

manufacturer. This hot starch solution was added to a 100 mL beaker containing the

meltedAKD.

A sonicator from Vibracell (VCF-1500, Sonics & Materials, Inc) was used for

aIl emulsifications. This apparatus operates at a frequency of 20 kHz. The tip of the

sonicator, which is one inch in diameter, was inserted into the liquid. The sample was

sonicated for three cycles of three minutes each to ensure proper homogenization.

Following sonication, the solution was immediately quenched in a metal beaker

containing 100 mL of cold de-ionized water. Hydrochloric acid (lM) was added to

reduce the pH to approximately 4. Samples were kept refrigerated until further

testing. The composition of this emulsion is presented in Table 2.1, showing a

dispersion that has 53.4 mg AKD/mL solution.

20

Table 2.1 Emulsion Characterization

Com~onent Quantit~

AKD (g) 10

Starch solution (g) 1.5

Starch solution volume (ml) 75

De-ionized water (ml) 100

1M HCI (ml) 0.5

Total Volume (ml) 187.2

[AKD] (mg AKD/ml) 53.4

2.4 Emulsion Analysis

2.4.1 Particle Size Analysis

The Malvem Mastersizer2000 (Malvem Instruments, UK) was used to test the

AKD particle size as well as particle-size distribution. This technology relies on laser

diffraction, using the Mie theory to determine the particle-size distribution. Particles

are passed through a focused laser beam, which causes them to scatter light. The

intensity of this light is measured as a function of the scattering angle by multiple

photosensitive detectors. The refractive index of AKD was taken as 1.485 (a typical

value of a wax) in water with a refractive index of 1.33. Plots such as the one seen in

Figure 2.1 were obtained. The particle diameter ranges from approximately 0.2 /lm to

5 /lm. It is normal to expect a variation in particle sizes. For the precision required

here, the emulsion is relatively monodisperse. The calculated median value (by %

volume) ofthe samples was used as the average particle diameter.

21

12

,.-.. 10 ~ 0 ....... 8 al E

6 ::l '0 > 4

2

~.01 0.1 10 100 Partiele diameter (~m)

Figure 2.1: Particle-size distribution of AKD emulsion (Malvem Mastersizer)

Periodic testing revealed that the refrigerated emulsions were stable for at

least two weeks. After four weeks, the two phases (AKD and water) began to

separate. Since AKD has a lower density than water, it rises to the top, causing

creaming. This instability can be due to creaming by coalescence, aggregation, or

both 1 [9]. AKD stabilized with the highest DS% starch was found to be most stable.

The particle size was confirmed with the Acoustic & Electroacoustic

Spectrometer (Dispersion Technology, Inc., NY). This apparatus calculates the

particle size distribution by minimizing the deviation between the measured and

calculated acoustic attenuation spectra. This method also gave a slightly higher value

of 1.27±O.59 Ilm.

In addition to particle size analysis, photographs were taken by scannmg

electron microscopy (SEM). Samples were prepared by diluting the emulsion and

placing a drop on an SEM pin. When the drop let had dried, it was sputter coated with

1 Coalescence: fusion of individual particles into one greater particle Aggregation: coming together of individual particles to form a cluster, particles not evenly dispersed

22

a gold film. Pictures were taken at 5,OOOX and 1O,OOOX magnification. As seen in

Figure 2.2, these photographs reveal a homogeneous sampling of AKD drop lets that

are less than a micron in diameter. Although most partic1es are similar in size, several

outlying drops that have likely coalesced also appear. It is interesting to note that the

partic1es are not perfectly spherical. We can only presume that this is due to the SEM

coating process, when the sample is exposed to extreme vacuum conditions.

Figure 2.2: SEM image of AKD emulsion at 1O,OOOX magnification (scale = 1 J.lm)

2.4.2 Electrophoretic Mobility Analysis

Electrophoretic mobility (EM) is a measure of the partic1e charge. This

mobility value indicates the strength of electrostatic attractions (or repulsions) a

partic1e experiences with another charged partic1e in proximity. AlI tests were done in

both de-ionized (DI) and tap water. A di lute sample ofthe AKD emulsion was poured

into a flat cell equipped with an electrode at either end. A voltage was applied and the

velocity of partic1es at a specifie distance from the tube wall could be measured on a

magnified screen divided into grids. By knowing the time, t, for a partic1e to travel

23

across a grid, and a constant K, (detennined during calibration) we can apply

Equation 2.2 to calculate the electrophoretic mobility.

EM = K = 8.717x10-Sm

2 IV

t time (s) (2.2)

Since our AKD compound is an uncharged waxy molecule, the charge it carries in

emulsion cornes solely from the stabilizing agent, cationic starch. This is in

agreement with other researchers, who have also found that the properties of an AKD

emulsion depend heavily on the stabilizer used.

Esser [12] has shown that the retention value of an AKD emulsion can be

correlated with its electrophoretic mobility. An EM value of +0.5 [(Jlm/s)/(V/cm)]

can be associated with a 10% retenti on value, while a value of +2.0 [(Jlm/s)/(V/cm)]

gives 100% AKD retention. Cationic starch typicaIly has a low mobility of +0.5

[(Jlm/s)/(V/cm)] at aIl pHs, whereas synthetic stabilizers have a value of

+3.2[(Jlm/s)/(V/cm)] at low pH, and which drops to around +2.5[(Jlm/s)/(V/cm)] at

pH 8, a typical environment for papennaking. Nevertheless, synthetic polymers carry

a mobility that is six times higher than that of starch. Since our application of AKD

occurs in a neutrallalkaline environment, it was important to choose a suitable

stabilizing agent that perfonns in this range. Our mobility measurements in tap water

revealed that the AKD emulsion carried a weak positive charge corresponding to a

mobility of +1.33 [(Jlm/s)/(V/cm)]. The mobility was higher in DI water, since

anionic impurities in the water could adsorb onto AKD.

24

2.4.3 Effect ofStarch Concentration and Ratio

Several variations of the emulsification procedure were tested. One variable

that was modified was the ratio between AKD and starch. From the literature, a

suggested range for this ratio is 0.1-0.3 (1 g AKD for each 0.1-0.3 g starch) [9].

Several trials were done with various ratios; it was found that 19 AKD: 0.15 g starch

consistently produced the same partic1e diameter, and therefore this ratio was used in

all subsequent emulsions.

Various starch concentrations were tried as well. Using starch cooked at a4%

concentration, the AKD particle sizes tend towards 2.3Jlm diameter. As the starch

concentration decreased, the partic1e size decreased as well. Using a 1% starch

solution, we obtained AKD droplets that were nearly identical to the theoretical

values (Figure 2.3).

4,------------------------------------,

• 0

• 4 % starch o 2 % starch .., 1 % starch

- Theoretical Value

oL-~==:===~~~~~--J - .., 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40

C (Starch:AKO Ratio)

Figure 2.3: Theoretical and experimental values of AKD partic1e diameter

Theoretically, we can predict the smallest stable AKD partic1e size that can occur for

various ratios using the following formula, where C is the ratio, a is the radius of

25

AKD droplets, p is the density of AKD wax, and r max is the maximum surface

capacity of starch, which is taken from the literature as lOg/m2 [13].

(2.3)

Solving for a and converting to a diameter, we obtain the line shown in Figure 2.3. As

the starch concentration approaches 1 %, the experimental particle size approaches the

theoretical size.

It was determined that concentrated solutions of starch produced larger AKD

droplets. As the starch concentration was decreased, the AKD particles became

smaller after receiving the same sonication treatment. These size variations were

attributed to starch clustering. This topic has been covered in depth by several

researchers in the area of pulp and paper, including Shirazi [13, 14]. After sorne

experiments, we found that starch clustering is sensitive to sonication. At higher

starch concentrations, sonication is not able to disperse the starch completely;

however, it is much simpler with low concentration solutions. Thus, in the

emulsification with 4% starch solutions, the starch was not able to completely un-

cluster, so fewer starch clusters were available to coat the AKD droplets. To remain

stable, the AKD droplets remained larger to decrease their specific surface area. In

the 1 % starch solution, the starch was completely un-clustered, making each starch

molecule available for coating the AKD, thus producing smaller drop lets. A standard

radius of gyration for starch in literature is given as 108 nm [13]. With the di lute

solution, we come close to achieving this value (see Appendix B for a review of

starch clustering as a function of solution concentration and sonication). When our

26

results are represented graphically as in Figure 2.3, we can demonstrate that the AKD

drop lets are more dependent on the concentration of starch than the actual proportions

of AKD and starch.

In the end, we chose to use a 2% starch solution, and consistently obtained a

particle diameter of 0.8 microns. These emulsions were then further used for all

Photometrie Dispersion Analyzer experiments, adsorption kinetics, and experiments

on the laboratory twin-wire sheet former.

2.4.4 Camparisan ta Cammercial Emulsian

A sample of a commercial emulsion was also tested for particle size with the

Malvem Mastersizer. Under the same water and pH conditions, its average particle

size was 0.45 microns in diameter. However, its electrophoretic mobility was higher,

at +1.8 [Ü.un/s)/(V/cm)]. In comparison, our AKD emulsion had a slightly larger

diameter and a weaker charge, but the values were similar, and we concluded that the

emulsions were comparable.

2.5 Photometrie Dispersion Analyzer Experiments

To study AKD and PCC interactions in a wet-end environment, we began

observations at the laboratory scale, with experiments using the Photometrie

Dispersion Analyzer (PDA) (Rank Brothers Ltd., UK). In typical papermaking

operations, the proportions of PCC and AKD added are about 10-15% wt PCC/wt

fiber, and 0.15% wt AKD/wt fiber; thus the ratio of these two additives would be

approximately 10-15 mg AKD/ g PCC.

27

The Photometrie Dispersion Analyzer provides a sensitive (but qualitative)

indication of the changes in the state of aggregation of a suspension. A more in-depth

discussion of the princip les behind the PDA operation is given in Appendix B.

2.5.1 Methods and Materials

Precipitated calcium carbonate (PCC) Albacar HO (Specialty MineraIs Inc.)

was used in aIl experiments. This PCC has an average partic1e diameter of 1.3 ~m and

a high specifie surface area of ~12 m2/g, due to its scalenohedral structure. This

compound and its behavior with fibers and other chemicals has been fully

characterized elsewhere [15, 6, 16, 7, 8].

The AKD emulsion was created in our laboratory; its properties were outlined

at the beginning of this chapter. AIl experiments were performed in tap water with a

conductivity of280 ~S/cm, which resembles industrial waters [17, 18].

A 5g/100mL stock solution of precipitated calcium carbonate was prepared in

tap water. A portion was measured and added to a one-liter solution of tap water to

produce a 500ppm suspension of PCc. The concentration of PCC was held constant

between experiments, while the quantity of AKD varied. AKD was added 30 seconds

after the start of each experiment. Tubing of 3mm I.D. was used for aIl trials to

minimize the effect of shear and to allow a larger flowrate and floc size. Operating at

a higher shear rate may break up flocs and narrower tubing would give an incorrect

indication of flocculation rates. The suspension was circulated at 150 mL/min, and

stirred at 100 rpm in the beaker. Figure 2.4 shows a schematic ofthis setup.

28

Stirrer

Suspension

Figure 2.4: PDA arrangement with circulation and pump

Subtle differences exist for various flowrates and stirring speeds; the effects of

changing either of these settings are discussed further in this section. OveraU, aU

flocculation patterns were similar, so intermediate values were chosen for both the

stirring speed and the flowrate. Settings on the PDA apparatus were held constant at

RMS Gain: 0.0 and DC Gain: 6.80.

As the solution passes through the PDA light detector, the apparatus measures

several values, inc1uding the turbidity and distribution of partic1es in a 1 mm2 cross-

sectional area. Data was obtained as the RATIO of the RMS/DC signal over time. A

signal was recorded for 10 minutes.

RATIO = RMS De (2.4)

The RATIO signal is an indirect measure of the level of flocculation. The

root-mean-square of fluctuations is very low when there is no aggregation, and

increases with the size of the flocs (and decreases with the number of particles)

29

(Figure 2.5). Since the De signal also changes as aggregation occurs, the RATIO

value takes this into account.

Low RATIO

••• •• • • • • • • • • • ••• • • ••• Figure 2.5: RATIO for dispersed versus aggregated particles.

2.5.2 PDA Parameters

Prior to our flocculation experiments, we investigated the influence of stirring

speed and flowrate (in tubing) on PDA results. We explored the effect these

parameters had on the flocculation rate and final floc size. The stirring speed was

varied between 100, 150, 200 rpm, and the flowrate through the tubing was set to

150, 200, or 250 mL/min. Figure 2.6a shows the differences when varying the

flowrate, while Figure 2.6b presents flocculation with various stirring speeds. Since

all curves are for a value of 10 mg AKD/g PCC (at which we obtained optimal

flocculation), we are able to investigate the effect of shear alone. At an intermediate

flowrate of 150 mL/min, flocs form and remain stable. Above this value, flocs are

broken up by shear forces, as indicated by the decease in flocculation. Below this (not

shown), the floc rate is limited by the insufficient stirring and fewer collisions.

Likewise, at low stirring speeds, the rate of collisions and flocculation decreases, and

at high values, breakup occurs. In general, changing the stirring and flowrate

parameters shifts the dynamic equilibrium between flocculation and floc break-up.

30

We chose intennediate values for both; 100rpm and 150 mL/min flowrate.

This is equivalent to a relatively low shear rate of approximately 20 S-1 in the beaker.

Shear is greater in the tubing (650 S-I) due to the small tube diameter.

1.6 Flowrate: [mL/min]

1.4 150

1.2

....... ::::> 1.0

« -- 0.8 0

+=' co 0.6 ~

0.4

0.2

0.0 0 100 200 300 400 500 600 700

time (8)

Figure 2.6a: Effect ofvarying flowrate on the floc fonnation in PDA for a stirring rate of 150 rpm.

31

1.6

1.4

1.2

--::> 1.0

<C - 0.8 0 ;; CO 0.6 0::

100 rpm

0.4

0.2

0.0 0 100 200 300 400 500 600 700

time (8)

Figure 2.6b: Effect ofvarying stirring speed in beaker on the floc formation in PDA for a flowrate of 150 mL/min

2.5.3 Results & Discussion

2.5.3.1 PCC Homoflocculation

When the signal from PCC in the water was measured without AKD present, a

small increase in signal was detected. This accounts for the minimal flocculation that

occurs since positively-charged PCC acquires a negative charge in tap water.

Nevertheless, no significant flocculation occurs under these conditions. Figure 2.7

shows this trend as the curve labeled "0 mg".

2.5.3.2 PCC and AKD Flocculation

Tests performed with PCC and various additions of AKD emulsion showed

that there was a maximum in flocculation that occurred at 10 mg AKD/g PCC (Figure

2.7). (As stated earlier, this is in the range of typical addition levels in industrial

32

settings.) In all PDA figures, the left axis represents the RATIO of the root-mean­

square/ DC signal, in arbitrary units (based on the gain settings of the PDA). This

value indicates the relative floc size. For larger flocs, there are less flocs for the same

number of particles initially, resulting in a larger RATIO value. Note that the

maximum ratio in these tests reaches 1.2. All additions of AKD are per gram of PCC.

Below 10 mg AKD/g PCC, flocculation was slower and reached a lower final

value. At higher AKD dosages, the initial slope was maximum, (the same as that for

10 mg/g PCC), but flocculation reaches a lower plateau. Tests were repeated to verify

that this behavior is reproducible.

Since PCC is white with a high refractive index, the PDA signal primarily

reflects the aggregation of PCC. That is, we have no real evidence that AKD and PCC

have heteroflocculated, but rather that PCC has been flocculated. This is particularly

interesting since we discovered later that the flocculation inferred by the PDA

consisted of only clusters of PCC and starch molecules. This was confirmed with

experiments testing the flocculation of PCC by cationic starch, and likewise, by AKD

free of starch.

The results shown below suggest that there exists an optimal AKD addition

dosage, at which half of the PCC is covered (by starch), allowing for maximum

flocculation. Beyond this, the PCC becomes over-coated, leading to a quick initial

flocculation, but the flocs cannot continue to grow since they are fully coated. Figure

2.7 shows this phenomenon with the 50 mg line. Although restabilization may be

expected in these conditions, van der Waals forces are able to form weak bonds

between fully-coated PCC particles, allowing them to homoflocculate.

33

-::> ~ 0

:;::: ca

c:::

1.4

1.2

1.0

0.8

0.6

0.4

0.2

....... ~ .... ~""'_~IW~ ........ ~ ......... ~ _ ....... __ .. ~ 0 mg

l..A"''-~

0.0 0 100 200 300 400 500 600 700

time (s)

Figure 2.7: PCC flocculation with AKD emulsion. Quantities indicated are AKD dosages [mg AKDI g PCC].

2.5.2.3 PCC and Starch Flocculation

Figures 2.8 and 2.9 demonstrate the PCC flocculation with the addition of

cationic starch. The starch was sonicated to un-c1uster starch molecules. At low

additions (less than lmg starchlg PCC), the flocculation is slow, and reaches a plateau

height relative to the amount of starch added. Above this optimum dosage, the initial

flocculation rate remains the same, but the flocs are not able to grow. This implies

that for PCC fully coated by starch, no energy barrier is present in the interaction

energy between two partic1es. Thus, cationic starch do es not cause steric or

electrostatic stability, and its adsorption does not eliminate aggregation of PCC, since

van der Waals forces are still present. However, the bond strength between fully-

coated PCC partic1es is weaker than that for partially coated partic1es, as indicated

schematically in Figure 2.10.

34

=> ~ o

:;:::; ct!

0:::

2.0 -r---------------------------,

1.5

1.0 0.25 mg

• ""Iane "'.e."" 0.5 r:cr 0.1 mg starch

0.0 U.lII~-_"T"'----,---.,__--_._--__,_---,__---;

o 100 200 300 400 500 600 700

time (s)

Figure 2.8: PCC flocculation with sonicated starch (low starch additions) Labels indicate [mg starch/g PCC].

2.0 -r---------------------------,

1.5

-::j <è -0

1.0 :;:::; ct!

0:::

0.5

30 mg, 50 mg

0.0 ..,'---_"T"'----,---.,__--_._--__,_---,__---; o 100 200 300 400 500 600 700

time (s)

Figure 2.9: PCC flocculation with sonicated starch (high starch additions) Labels indicate [mg starch/g PCC].

35

starch

a) Bridging by starch b) Aggregation by van der Waals forces

steric repulsion

Vint

distance

b) van der Waals attractions

a) van der Waals and electrostatic attractions

1

Fig. 2.10: Schematics of energy of interaction between two PCC partic1es: a) partially coated by starch- van der Waals and electrostatic attractions b) fully coated- van der Waals attractions and steric repulsions

2.5.2.4 Effects of Starch Filtration

Next, the AKD emulsion was filtered through a 0.45Jlm filter. AKD droplets,

approximately 0.8Jlm in diameter, were retained by the filter, then re-dispersed in the

same displaced volume of DI water, so the same concentrations remained. A mass

balance on the AKD and starch revealed that approximately half the starch was

adsorbed to the AKD drop lets whereas the remaining half was free in solution or

washed from the AKD partic1es during filtration. This filtration was repeated with the

36

filtered AKD solution "AKD1" to obtain "STARCH2" and "AKD2". A schematic of

the filtering is presented in Figure 2.11.

AKD

il l > 1

ni >~ o STARCH 2 ).!J LJ o STARCH 3 )

Figure 2.11: Successive filtration schematic of AKD droplets and starch stabilizer

PDA experiments were performed on both the filtrate and the AKD droplets. Results

with the filtered starch were identical to the ones obtained with sonicated starch in the

previous section. Experiments with the filtered AKD are described below.

Figure 2.12 shows that the AKD partic1es, now filtered and depleted of

cationic starch, hardly flocculate with PCC, even at high quantities such as 50 mg

AKD/g PCC. In other words, upon a second filtration of the AKD emulsion, only

minimal flocculation was observed, similar to the behavior of PCC alone.

The electrophoretic mobility of the filtered AKD was measured, and the value

was found to be lower than that of the original emulsion (Table 2.2). This is expected

since AKD itself carries no charge.

37

-::J ci

0.6.,...----------------------,

0.5 <D 0 ~ ?~~œt"OO

o ~~ 0 50 mg o~

oqr}) • ....

~W • • ~nl"IiIlaF."" 004

"-'" 0.3 o ~... . 20 mg cf) _~-8' 0.2 <D 0.,/

0.0 -f-l-''---,....---,----r-----,;------,----,.----I

o 100 200 300 400 500 600 700

time (s)

Figure 2.12: PCC flocculation with "AKD2" (AKD filtered twice to remove starch). Quantities indicate [mg AKD/g PCC].

It was conc1uded that the flocculation of PCC was due primarily to the free

starch in solution, and in part by starch transfer from AKD partic1es to PCC. A final

interesting point is that an optimum of 1 mg starch corresponds very c10sely with the

10 mg AKD emulsion optimum dosage we found earlier. Since the ratio of

AKD:starch is 1:0.15, these two values are similar. This suggests that it is truly the 1

mg of starch within a sample of 10 mg AKD that induces PCC flocculation.

Table 2.2 Electrophoretic Mobilities and Partic1e Diameters of AKD emulsions

AKDo AKD1 AKD2 AKD3

EM [(r.tm/s)/(V/cm)] +1.33 +1.21 +1.01 NIA

dAKD (/-lm) 0.826 0.834 0.838 0.849

38

..-..

It is concluded that the PCC flocculation occurs due to starch only, and that

AKD and PCC do not interact. Instead, the starch-coated AKD particles adsorb to the

PCC, and polymer transfer occurs, leading to flocculation. Starch transfer also

explains why flocculation by starch is on the same time-scale as that with AKD.

2.6 Scanning Electron Microscopy Analysis

To visualize the interactions between AKD and PCC, and to verify results

obtained by PDA, images were taken with the scanning electron microscope. AKD

and PCC were added in different amounts to determine their flocculation properties.

As before, samples were prepared by diluting the emulsion and depositing a drop onto

an SEM pin. When the droplet had dried, it was sputter coated with a gold film.

Photographs were taken at 5,OOOX and 10,OOOX magnification.

Figure 2.13 shows PCC particles as seen with the SEM. The photographs

clearly show the scalenohedral structure of the particles, as well as their relative size.

Figure 2.13: SEM image of PCC particles (scale = 1 micron)

39

Figure 2.14 below presents an AKD emulsion at 1O,000X magnification. We

can see the spherical AKD drop lets with diameters just under 1 micron. As mentioned

before, we must assume that the indentations and irregularities in their spherical

shape are due to the process of coating at an extremely low pressure.

Figure 2.14: SEM image of AKD emulsion (scale = 1 micron)

Next, we added AKD and PCC in various proportions on a pm. Using

standard additive quantities (lg PCC and 10mg AKD), AKD appears 100 times 1ess

frequently than PCC, and is not easily seen in the photographs. N everthe1ess, no

evidence of AKD-PCC heteroflocculation was observed at these addition levels

(optimal additions).

Figure 2.15 a,b and c show PCC and AKD at a 1:1 ratio. Although this is

higher than standard quantities for AKD, we do not see evidence of AKD-PCC

interactions. We observe homofloccu1ation of both AKD and PCC, though there were

no interactions between the two species.

40

a

b

c

Figure 2.15(a,b,c): PCC and AKD at a 1:1 ratio

41

Figure 2.16: SEM image of AKD emulsion after filtering out free starch; AKD drop lets aggregate.

In Figure 2.16, we show an AKD emulsion where the starch has been filtered away.

This photograph illustrates how the removal of starch results in the instability of the

individual AKD partic1es, causing them to aggregate.

2.7 Conclusions

In this chapter, we have outlined a procedure for creating an AKD emulsion

free from extraneous substances. Its properties were similar to those of a commercial

emulsion, and we proceeded to use it for further testing. U sing the Photometrie

Dispersion Analyzer, we found that the AKD emulsions flocculate with PCC at an

optimal dosage of 10mg AKD/g PCC. However, later we found that flocculation is

actually caused by the cationic starch acting as the AKD stabilizer.

No visual proof was found that AKD and PCC form flocs. The SEM

photographs show that PCC and AKD each homoflocculate; little to no

heteroflocculation occurs. This, in combination with the results obtained from the

42

PDA tests, suggests that PCC and AKD do not directly interact. Instead, starch

removal from the AKD particles aids the homoflocculation of PCC (via polymer

bridging). Starch transfer causes the AKD particles to aggregate since their stabilizing

polymer (starch) has been depleted. It can therefore be concluded that cationic starch

has a higher affinity for PCC than for AKD.

It is interesting to consider what these findings imply. Previously, it was

thought that PCC directly hydrolyzes AKD. Now, we see evidence that AKD and

PCC do not heteroflocculate, but that polymer transfer occurs from AKD to PCC,

followed by PCC homoflocculation. This indicates that the alkaline environment is

the most significant factor contributing to AKD hydrolysis. More importantly, we

find that AKD and PCC are not likely to interact in the whitewater cycle. This

dismisses the theory that PCC will lower AKD retention. PCC does decrease the

sizing of a sheet by increasing the specifie surface area and forming more pores.

However, this would not influence AKD single-pass retention.

2.8 Acknowledgements

Many thanks to Helen Campbell for assistance with the Scanning Electron

Microscope at McGill University and to Lou Cusmich for help with the electronic

data acquisition setup for the PDA. 1 would also like to thank Patrick Lim Soo for his

assistance with the Dynamic Light Scattering.

43

2.9 References

[1] Bottorff, K.J., "AKD sizing mechanism: A more definitive description", Tappi Journal, 77(4) 105-16 (1994).

[2] Bottorff, K.J. and Sullivan, M.J., "New insights into the alkylketene dimer (AKD) sizing mechanism", Nordic Pulp & Paper Research Journal, 8(1) 86-95 (1993).

[3] Scott, W.E., "Wet End Chemistry", Tappi Press, Atlanta, (1996). [4] Brungardt, C.L., "Studies indicate wet-end additive effects on AKD retention,

reaction efficiency", Pulp & Paper, 75(6) 47-49 (2001). [5] Isogai, A., "Effect of cationic polymer addition on retention of alkylketene

dimer", Journal ofPulp and Paper Science, 23(6) 276-281 (1997). [6] Kamiti, M. and Van de Ven, T.G.M., "Kinetics of deposition of calcium

carbonate particles onto pulp fibers", Journal ofPulp and Paper Science, 20(7) 199-205 (1994).

[7] Vanerek, A., Alince, B. and van De Ven, T.G.M., "Interaction of calcium carbonate fillers with pulp fibers: effect of surface charge and cationic polyelectrolytes", Journal of Pulp and Paper Science, 26(9) 317-322 (2000).

[8] Vanerek, A., Alince, B. and Van De Ven, T.G.M., "Colloidal behavior of ground and precipitated calcium carbonate fillers: effects of cationic polyelectrolytes and water quality", Journal of Pulp and Paper Science, 26(4) 135-139 (2000).

[9] Chew, Y.S., Peng, G., Roberts, J.C., Xiao, H., Nurmi, K. and Sundberg, K., "Characterization of the stability of cationic starch stabilised AKD emulsions", Preprints - 5th International Paper and Coating Chemistry Symposium, Montreal, QC, Canada, 331-337 (2003).

[10] Mohlin, K., Leijon, H. and Holmberg, K., "Spontaneous emulsification of alkyl ketene dimer", Journal of Dispersion Science and Technology, 22(6) 569-581 (2001).

[11] Roberts, J.C."Neutral and Alkaline Sizing" in Paper Chemistry Ed.Roberts, J.C. New York: Blackie Chapman & Hall, (1991).

[12] Esser, A. and Ettl, R."On the mechanism ofsizing with alkyl ketene dimer (AKD): physico-chemical aspects of AKD retenti on and sizing efficiency~" in Fundamentals of Papermaking Materials, Transactions of the Il th Fundamental Research Symposium Cambridge, UK, (1997). 997-1020.

[13] Shirazi, M., Van de Ven, T.G.M. and Garnier, G., "Adsorption of Modified Starches on Pulp Fibers", Langmuir, 19(26) 10835-10842 (2003).

[14] Shirazi, M., Van de Ven, T.G.M. and Garnier, G., "Adsorption of Modified Starches on Porous Glass", Langmuir, 19(26) 10829-10834 (2003).

[15] Cechova, M., Alince, B. and van de Ven, T .G.M., "Stability of ground and precipitated CaC03 suspensions in the presence of polyethylene oxide and kraft lignin", Colloids and Surfaces, A: Physicochemical and Engineering Aspects, 141(1) 153-160 (1998).

[16] Suty, S., Alince, B. and van de Ven, T.G.M., "Stability of Ground and Precipitated CaC03 Suspensions in the Presence of Polyethylenimine and Salt", Journal of Pulp and Paper Science, 22(9) 321-326 (1996).

44

[17] Donat, V., van de Ven, T.G.M. and Paris, J., "Distribution of dissolved and colloidal substances in the fonning and press sections of a paper machine", Journal of Pulp and Paper Science, 29(9) 294-298 (2003).

[18] Huber, P., Carre, B., Mauret, E. and Roux, J.-C., "The influence of fine elements build-up in the short-circulation on fibre flocculation." Preprints -International Paperand Coating Chemistry Symposium, 5th, Montreal, QC, Canada, (2003).

45

CHAPTER3

AKD INTERACTIONS WITH cP AM AND FIBERS

3.1 Abstract

Interactions of AKD with cP AM were examined using both anionic and

cationic AKD emulsions. PDA experiments showed that cationic polyacrylamide did

not induce flocculation of cationic AKD (stabilized by cationic starch). Conversely,

anionic AKD was found to flocculate by cP AM. Kinetic adsorption tests were

performed to determine the rate of AKD deposition onto fibers under shear

conditions.

We found that without addition of retention aid, 75% of cationic AKD

particles deposited onto fibers, with detachment beginning after a peak in deposition.

With the addition of cP AM, the retention increased to 100%. Although cP AM

molecules do not flocculate AKD on their own, cP AM adsorbed to fibers is capable

of adsorbing onto AKD by a mechanism known as asymmetric polymer bridging. We

propose that the AKD-cP AM -fiber bond is stronger than the one between starch­

coated AKD and fibers, explaining the complete retenti on and absence of particle

detachment.

Anionic AKD emulsions have become common in industry. In testing the

adsorption kinetics with this emulsion, we found that the absence of a retenti on aid

resulted in very low AKD retention. Addition of cP AM rapidly increases the rate of

AKD deposition onto the fibers even at high dosages, where we would expect steric

repulsions to limit the deposition. This is likely a kinetic effect, since the AKD

deposition onto fibers is faster than the coating of AKD particles by cP AM.

47

3.2 Introduction

In further studying the retention mechanism of AKD, we investigated AKD

interactions with a cationic retenti on aid. It is not enough to consider only AKD and

fiber interactions, since in practice, retention aids also interact with these compounds.

In a paper mill, retention aids are added to improve fines retenti on, increase

drainage, control deposits, improve filler retention as well as to improve operations

[1]. To function properly, the retention polymer must adsorb onto the surfaces of the

fibers and the particles to be retained. The adsorption of cationic polymers onto

anionic surfaces, such as cellulose fibers and clay, occurs rapidly because there is no

barrier to adsorption. Polymers are usually added at a point close to the headbox to

minimize disruption of the flocs and to maximize retention. It is believed that upon

adsorption, the conformation of the adsorbed polymer facilitates the formation of

polymer bridges between surfaces, promoting the retention of fines [2]. Changes in

conformation of the polymer occur after adsorption, usually making bridging less

effective, although fuis mechanism is not well understood.

It is concluded by one researcher that the improvement of AKD retenti on is

not caused by direct adsorption of retenti on aids on AKD particles, but by AKD

adsorption on fines and fibers, leading to co-flocculation of fines, fibers and AKD

particles [3]. Another study states that retention is thought to be caused by

heterocoagulation of cationic AKD particles to negatively charged fiber surfaces,

where cationic polymers act as fixatives for AKD [3, 4]. Using cPAM, one study

showed that AKD deposition reached a maximum after 20 seconds, indicating that

AKD retention is a reversible process. This observation suggests that the subsequent

48

decrease in AKD deposition is due to particle detachment/charge reversaI, indicating

starch transfer. The results proved that AKD retention is optimized if the contact time

between AKD and fibres is short, and that of AKD-fibres and cationic polymers is

short as well [4]. The absence of re-attachment was explained by the deactivation of

the adsorbed polymer layer via re-conformation and/or degradation [5].

Although the surface charge of both AKD and cationic retention aids lS

positive, the retention aids have a significant influence on retenti on behavior [3].

Previous work conducted on the interactions between AKD and retention aids has

shown that the AKD surface charge has a profound effect on AKD retention on its

own, but as soon as a retenti on aid (cPAM) is added, it is a less significant factor [6].

Others agree that adding cationic polymer to AKD and fibers leads to a higher

retention, which increases sizing [7]. One suggestion to explain this behavior is that

the increase in AKD retenti on is due to an increase in fines retention, though this

point has not been fully explored. Another theory assumes that at a neutral pH, AKD

particles become amphoteric from the adsorption of anionic materiaI, and in this

manner a cationic retenti on aid is able to bridge the negatively charged areas [7].

Typically, AKD is stabilized by cationic starch, since the fibers are negatively

charged. This enables the AKD to deposit on the fibers by electrostatic attraction. On

the other hand, the use of cationic retenti on aids allows anionic AKD emulsions to be

used as well. In sorne paper mills, AKD is stabilized with anionic polymer, and

although the AKD and fibers naturally repel each other, the cationic retention aid

permits bridging between the two.

49

Anionic AKD emulsions are advantageous because they can be stored at a

50% solids content, whereas the cationic emulsions are not stable above 15 or 20%

solids content. This is most likely due to the selection of stabilizing chemicals in

either case. This drastically reduces the cost associated with transporting and storing

AKD at a mill. N evertheless, AKD and fibers are both negatively charged, so a

precise balance of polyelectrolytes/retention aids must be added to optimize

performance. Conversely, cationic AKD has good retention over a broader range of

retention-aid dosages. In this chapter, we discuss experiments performed with both

cationic and anionic emulsions.

Many researchers predict that cationic AKD emulsions will interact with

cP AM, owing to the amphoteric nature of stabilizing agents in the emulsion. In these

cases, cationic starch and lignosulfonic acid are commonly used as stabilizing

polymers [8]. However, the explanation of this mechanism is vague. Furthermore, it

would not apply in our case, since only cationic starch is used as a stabilizer.

3.3 Materials and Methods

3.3.1 Materials

F or trials with cationic AKD emulsions, we used the emulsions described in

Chapter 2. These AKD particles have a particle diameter of 0.8 f.lm, and an

electrophoretic mobility of + 1.33 [(f.lmls)/(V/cm)]. Anionic AKD emulsions were

produced using carboxymethylcellulose (CMC) as the anionic stabilizer. This

chemical (received from Aldrich Chemicals) has a molecular weight of 700,000 Da.

The same procedures were used for emulsification, using 0.15g CMC: 1 g AKD. The

50

anionic dispersions were monodisperse with an average diameter of 0.7 J.lm. AlI

emulsions were diluted to a 3g/L solution, to extract precise amounts for testing smalI

quantities. Tap water, with a pH of 6.4 and a conductivity of 280 J.lS/cm, was used in

all experiments.

We used Percol292 as the cationic polyacrylamide (CIBA, Inc.), which has a

degree of substitution of 20% and a molecular weight of 3xl06 Da. Several other

polyacrylamide polymers were tested, but since they all gave similar results, they are

not discussed here.

The fibers we used were bleached softwood kraft fibers (BSK) obtained from

Domtar. Fibers were dispersed and diluted to produce a 0.6% consistency pulp (the

same consistency used for the twin-wire sheet former).

3.3.2 Methods: Photometrie Dispersion Analyzer Tests (PDA)

PDA tests were carried out in a IL beaker with tap water, using the same

configuration as presented in Figure 2.4. One gram of AKD (in emulsion) was found

to produce enough turbidity such that accurate changes in flocculation could be

detected. The amount of cPAM added was between 0-10 g/g AKD. A typical amount

used in industry is 1 mg cPAM/g fiber (lg cPAM/g AKD). Stirring was constant at

100 rpm while the dispersion was circulated at a flow rate of 150 mL/min. Cationic

polyacrylamide was added 30 seconds after the start of each mn. Since the refractive

index of AKD is lower than that of PCC, the apparatus needed to be adjusted to make

it more sensitive. The RMS Gain was increased to 5.3 to c1early detect changes in the

state of flocculation. The RATIO output was monitored for ten minutes.

51

The PDA apparatus was used to investigate the behavior of AKD with cPAM.

Since both compounds are positively charged, we expected no interactions between

them, though it was necessary to verify. In addition, anionic AKD emulsions were

tested to compare how a negatively charged emulsion might behave.

3.3.3 Methods: Adsorption Kinetics

Additional experiments were performed to determine the kinetics of AKD

deposition on fibers in the presence of cPAM and under shear. A 0.6%wt fiber

suspension was added to 500 mL of tap water. This consistency was chosen because

it was the concentration used in the twin-wire sheet former experiments, and was held

constant since fiber concentration can have an effect on flocculation.

AKD was added to the fiber suspension being stirred. Fifteen seconds later, a

specified amount of cP AM was injected; this was designated as time zero. A 200-

mesh screen (74 !-lm openings) was used to extract a 2-mL sample free from fibers

and fines. Samples were obtained at 15s, 30s, Imin, 3 min, 5min, and every 5 minutes

thereafter. A UV-VIS Cary lE spectrophotometer (Varian, Inc.) was calibrated at

350nm; this wavelength was chosen for its good transmittance of AKD. A filtrate

from the fiber suspension prior to the addition of AKD was used in the reference cell.

After AKD addition to fiber suspension, we measured the absorbance of the filtrate,

from which we obtained the amount of AKD that was not retained, and then

calculated the amount of AKD that had deposited onto the fibers, as a function of

time and cP AM addition. A small experimental error occurred due to the slightly

negative readings obtained by the spectrophotometer at zero concentrations.

52

3.4 Results

3.4.1 AKD Flocculation

Since we expected an anionic AKD emulsion (a-AKD) to be flocculated by

cationic polyacrylamide, we tested this first. The flocculation that was detected with

the anionic emulsion was not detected with our original cationic emulsion and cP AM.

This confirmed our initial premise that cationic AKD and cP AM do not interact.

Figure 3.1 shows vials containing Ig anionic AKD and different quantities of

cP AM. With the addition of cPAM, it is not until the 50-mg/g-addition point that

destabilization occurs. At lower additions, such as those used in our experiments,

anionic-AKD and cP AM flocculate to a much lesser degree. Large flocs, as seen in

vial #7, only form at high quantities of cP AM.

Figure 3.1: Photograph of vials with Ig a-AKD and various quantities ofcPAM; Amounts (0-50) indicate [mg cPAMIg anionic AKD].

53

3.4.2 Cationic-AKD Deposition on Fibers

Kinetic adsorption tests were performed with the cationic AKD emulsion,

fibers, and retention aid. An chemicals were added in quantities that approximate

industrial papermaking conditions. AKD addition was held constant at 10 mg/g

fibers, while cPAM addition varied from 0-lOmg/g fibers. As shown in Figure 3.2

below, in the absence of cPAM, AKD retention reaches approximate1y 75%, and

begins to slowly decrease to about 60%. When cP AM is added, AKD retenti on

reaches 100% (lOmg AKD/g fibers).

10 .~;;"'-.'~-~~-:'::':::-~':J ~"'~.:.-. :-===~t== .... --.. .. ,/ ..................... /'/' .............•............•........

.... / /..-v /' 8 : /.' ./'

. -:l1li-../_. f -. J '-'III-.L. ~ /. -.-... -'-'l1li-.-E J/ : / '-.-tIII.-.- -111-.-.-

6 JJ. .1 / § If /' .r ~ Ï.~/ / ~ 4 J /. Cl Cl ~ «

2

[mg cPAM/g fiber5]

_. -l1li. _. control

-+- 0.05mglg ..... 0.1 mglg

- .... - 0.25 mglg -"-"1'-" 2 mglg --+- 4 mglg

- __ - 5mglg

O __ ----.-----.---~/~--~==~==~==~ o 200 400 600 900 1200 1500 1800

Time (5)

Figure 3.2: AKD deposition onto fibers, at 10 mg AKD/g fiber addition, and varying cP AM addition. Inset: [mg cP AMI g fibers]

At an concentrations of cP AM, the initial rate of flocculation is the same, as AKD

deposition reaches ~20% retention in the same time. At low cPAM concentrations

(below 0.25 mg/g) , the time to achieve maximum adsorption is short (about three

minutes), whereas at high cP AM concentrations, the time required for the AKD to

54

reach 100% is prolonged, by up to 10 minutes for dosages above 4 mg/go The four

main features of the figure above are described below.

1. Without cP AM, AKD deposition goes through a maximum

In the absence of a retention aid, cationic AKD partic1es are deposited on

negative fibers by heterocoagulation [8, 7]. The AKD partic1es, which are coated with

cationic starch, adsorb to fibers through a weak bond causing sorne AKD partic1es to

detach, leaving starch on the fibers (Figure 3.3).

( + )AKD coated with cationic starch adsarbs ta fibre surface

Balance between attachment and detachment

Figure 3.3: AKD deposition to fiber surface

After a maximum peak in AKD deposition, the AKD partic1es begin to detach.

The transfer of starch from AKD to fibers explains why the deposition decreases after

a certain time.

2. cP AM increases AKD deposition, despite the fact that cPAM does not adsorb on

AKD coated by cationic starch

In the presence of cP AM, starch-coated AKD partic1es can adsorb either to the

bare fiber or to the cP AM previously adsorbed to the fiber (Figure 3.4 a,b).

55

Ca) low [cPAM]

AKD( +) adsorbs to fibre surface

AKD detachment / AKD held more strongly by cPAM

Cb) high [cPAM]

cPAM adsorbs to fibre surface

AKD adsorbs to cPAM

Figure 3.4: AKD deposition on fibers a) at low cPAM dosages; b) at high cP AM dosages

It may seem contradictory that AKD deposits on a cP AM molecule adsorbed

to the fiber surface, especially since we found in the previous section that cationic

AKD particles and cP AM do not interact. However, this is possible by asymmetric

polymer bridging. The theory behind this phenomenon is presented below.

Heteroflocculation is not expected when a polymer is added to a mixture of

two types of particles and polymer adsorption occurs on only one type. A previous

study by van de Ven et al. proposed a mechanism of "asymmetric polymer bridging"

[9]. This theory states that heteroflocculation can occur even for systems in which the

polymer does not adsorb on one of the two flocculating species, provided that the

polymer upon adsorption on Particle 1 acquires the ability to adsorb on Particle 2

(Figure 3.5). We find that cPAM in free solution does not adsorb to AKD, but upon

adsorption to fibers, cPAM can adsorb onto AKD.

56

A + ~

c

no adsorption

A + ~

::-:-: ......... '.:-»;:::.:.: <. ~ :::::::.:. ---"'~~ __ sm

+

A: cPAM B:AKD C: fiber

Figure 3.5: Asymmetric Polymer Bridging Mechanism

Whether a freely dissolved polymer adsorbs onto a particle depends on the

competition between a gain in enthalpy and a loss of entropy upon adsorption.

Adsorbed macromolecules have fewer configurations, so the loss in entropy is less

upon adsorption onto another surface. This permits a mole cule to bridge two particles

even when it is adsorbing onto only one of them. If the polymer adsorbs onto only

one surface, one would expect that steric stability is not possible and, as a result,

heteroflocculation should occur at aH polymer concentrations. This phenomenon has

been observed with polyethylene oxide (PEO) and clay with fibers [9], and here with

AKD and cP AM with fibers.

Also puzzling is why the AKD-cPAM-fiber bonds would be stronger than the

AKD-fiber bonds. InitiaHy, the starch-coated AKD particle adsorbs to the cPAM

already adsorbed on the fiber. One explanation for the increased strength involves the

reconfiguration of starch and its displacement from the gap. If the starch were to

spread, as illustrated in Figure 3.6, AKD would be in direct contact with the cPAM,

most likely producing a stronger bond. This spreading not only increases the bond

57

strength, but secures the AKD partic1e so it is less likely to detach. In addition, AKD

is still able to quickly adsorb to fibers independently of cP AM. In solutions with low

cP AM concentrations, a combination of these weak and stronger bonds exists, as

shown in Figure 3.4a.

(a) (b)

+- starch

/ cPAM

fiber

Weak bond: cPAM has little Strong bond: cPAM is likely to have affinity for cationic starch stronger affinity for AKD than for starch

(Sorne starch rnay transfer to fiber)

Figure 3.6: Increased Bond Strength by Starch Re-conformation ( a) Initial adsorption of starch to cP AM (b) Spreading of starch allowing for direct AKD adsorption to cP AM

3. The initial kinetics (up to 20%) is nearly independent of cP AM dosage

AKD deposits on bare fibers and cP AM-coated fibers at the same rate,

suggesting that in both cases there is no (or a negligible) energy barrier in the AKD-

fiber interaction that might slow deposition. At first glance, this was surprising since

AKD does not adsorb on cP AM, but was explained by asymmetric polymer bridging

in the previous section.

Equation 3.1 describes the kinetics for regular homoflocculation. An optimum

in flocculation efficiency (a) is reached when the fractional surface area covered (8)

is K

58

a = 2B(I-B)

For heteroflocculation, the flocculation efficiency equals:

a = BI (1- B2 ) + B2 (1- BI )

(3.1)

(3.2)

In asymmetric polymer bridging, we find the following relationship [9]:

(3.3)

slllce 92 = O. When complete fiber coverage (B=I) is achieved, the maximum

adsorption efficiency (a) is reached, indicating the maximum rate of flocculation.

Any dosage corresponding to full coverage or more will yield the same rate a. We

would expect high concentrations of cP AM to produce identical responses for initial

flocculation, as seen in Figure 3.2. Since there is no energy barrier, the initial rate at

aIl cP AM dosages, as weIl as for AKD depositing on bare fiber patches is the same.

As more time elapses, the curves digress due to other processes, such as polymer

rearrangement, which is discussed below.

4. With ex cess cPAM, the reaching of maximum retenti on is delayed

When a high concentration of cP AM is added to the AKD-fiber suspension,

the fiber surface is coated exclusively by cPAM molecules (Figure 3.4b). The

sluggish process of attaining 100% AKD retenti on occurs due to re-conformation and

a slower, but stronger process of AKD adsorption to cP AM. One possible scenario is

that the expulsion of starch from the gap (Figure 3.6) is fast at low cPAM dosages,

but slow at high dosages. For low dosages, the starch can spread directly onto the

fiber, without competition from neighboring cP AM particles. This is consistent with

the explanation of starch transfer from AKD to the fiber.

59

Another scenario is that in aIl cases, starch is excluded from the gap, but the

bond strength increases when the polymer conformation becomes less expanded with

time. The re-conformation is fast at low polymer dosages, but slow at high dosages. It

was shown previously that rearrangement time is a strong function of fractional

coverage [10]. Polymers rearrange quickly on low surface concentrations, but slowly

on crowded surfaces, since neighboring molecules interfere with the rearrangement.

For extremely crowded surfaces, non-equilibrium results for a long time. High cPAM

dosages are not suitable for industrial conditions where the residence times are very

short.

3.4.3 Anionic-AKD Deposition on Fibers

Since AKD deposition on fibers without cP AM is not complete, we explored

what would happen with AKD stabilized with anionic polymer. The behavior we

observed for anionic AKD dispersions differed significantly from that of the cationic

dispersions. Results from a-AKD deposition experiments are presented in Figure 3.7.

If we divide this graph into its three types of curves based on retention aid

dosage, we see three distinct results. When no retenti on aid is added, AKD adsorption

is nearly zero since the negatively charged AKD particles do not interact with the

negatively charged fibers. Adsorption begins with the addition of cP AM. As the

cP AM dosage is increased, the a-AKD deposition on fibers increases. When

intermediate amounts of cPAM such as 0.2 and 0.3 mg/g are added, fibers are

partially coated, and AKD deposition increases with the addition of cP AM. The

60

maximum flocculation rate is reached above 0.3 mg cPAM/g when fibers are fully

coated. For full fiber coverage, 100% AKD deposition is reached within 5 minutes.

Furthermore, the period to reach 100% retenti on with high dosages is not

protracted as with cationic AKD. AKD particles adsorb to the fibres by cPAM

bridging and remain adsorbed due to a strong bond strength between AKD-cPAM-

fibers. In this situation standard heteroflocculation behavior is exhibited .

.-.. ~ ~ 10 1+= Cl

o ~ 8 « Cl E o

c o ~ o g. 4 o o ~ .2 2 c o ~

---=--:-:: :::::~, -' _. _. -II _.

t .·:.- "'V''' -::-:::'-"'-=." -17 :::;.:.~'::::='=-~:;.:......-=I=~:.:::-' .... '-V'_"

/: -~---# --~/ _-------T ---""----...,-J/ /,""'- e ............ e· ...................... .

~. t- ......................................... . 1.' /.

.1 ./

, [mg cPAM/g fiber]

• Control ........ e ........ 0.2 mg/g ---T--- 0.3 mg/g - .. -v._ .. - 1.7 mg/g ---- 5.0 mg/g -'-111-'- 6.6 mg/g

oe------.-----.----~~~--.-----.----.----~

o 200 400 600 900 1200 1500 1800

Time (5)

Figure 3.7: Adsorption kinetics ofanionic AKD emulsion to fibers, in the presence ofePAM. AKD addition: 10mg/g fiber; Inset: [mg cPAM/g fiber]

We have shown previously that anionic AKD and cP AM do heteroflocculate.

If there is an excess of cP AM and all fibers and AKD are fully coated, sterie

repulsions should occur. In other words, the excess cP AM in solution would coat the

AKD particles making them positively charged and unable to adsorb to the positive

fibers. Instead, this is likely a kinetic effect. Of the three processes taking place: i.e.

(i) polymer adsorption on fibers; (ii) AKD deposition on fibers; and (iii) polymer

61

adsorption on AKD, (i) is the fastest, followed by (ii) and (iii). Thus at high cP AM

dosages, AKD deposits on fibers before cP AM coats the AKD particles. This explains

how AKD can deposit on the fibers, even at high cP AM concentrations.

Altematively, if the AKD particles were coated completely by cP AM, van der

Waals forces exist which would bind the AKD to the fibers. However, the first

scenario is more likely.

3.5 Conclusions

We have drawn several interesting conclusions by performing experiments to

monitor the AKD deposition kinetics onto fibers. First, cationic AKD is not

flocculated by cP AM. This was expected since both molecules have a positive charge

and repel one another. Adsorption kinetic experiments revealed that cP AM increases

AKD retenti on by assisting AKD deposition on fibers. A mechanism was proposed

for this behavior. Asymmetrical polymer bridging explains both the increase in AKD

retenti on and the kinetics of AKD deposition by cP AM. Although the initial kinetics

(up to 20%) is nearly independent of cP AM dosage, excess cP AM de1ays the reaching

of the maximum possible retenti on due to starch and cP AM reconformation. Lastly,

we found that the addition of cP AM results in an increase in the bond strength

between AKD and fibers, thereby eliminating AKD detachment.

For adsorption kinetic tests performed with anionic emulsions, two types of

behavior were observed. When cP AM is added in quantities insufficient to coat the

full fiber surface, cP AM increases deposition of AKD onto fibers in relation to the

amount added. With fibers fully coated with cP AM, AKD deposition is fast and

62

remains at 100%, due to the increase in bond strength. Steric repulsions do not occur,

since the kinetics of AKD adsorption onto fibers is faster than cP AM adsorption onto

AKD molecules.

3.6 Acknowledgements

The author would like to thank Dr. T. van de Ven for valuable discussion, and A.

Vanerek for assistance with experiments.

63

3.7 References

[1] Allen, L.H., Polverari, M., Levesque, B. and Francis, W., "Effects of system closure on retention- and drainage-aid performance in TMP newsprint manufacture", Tappi Journal, 188-195 (1999).

[2] Tanaka, H., Swerin, A. and Odberg, L., "Transfer of cationic retention aid from fibers to fine particles and cleavage of polymer chains under wet-end papermaking conditions ", Tappi Journal, 76(5) 157-163 (1993).

[3] Esser, A. and Ettl, R. "On the mechanism of sizing with alkyl ketene dimer (AKD): physico-chemical aspects of AKD retenti on and sizing efficiency." in Fundamentals of Papermaking Materials, Transactions of the Il th Fundamental Research Symposium Cambridge, UK, (1997). 997-1020.

[4] Champ, S. and Ettl, R., "The dynamics of Alkyl Ketene Dimer (AKD) retention." Preprints -5th International Paper and Coating Chemistry Symposium, Montreal, QC, Canada, 285-291 (2003).

[5] Asselman, T. and Garnier, G., "Polymer Transfer During Fines Detachment Under Turbulent Flow: Mechanism and Implication", Journal ofPulp and Paper Science, 27(2) 60-65 (2001).

[6] Johansson,1. and Lindstrom, T., "A Study on AKD-size retention, reaction and sizing efficiency. Part 1: The effects ofpulp bleaching on AKD sizing", Nordic Pulp & Paper Research Journal, 19(3) 330-335 (2004).

[7] Isogai, A., "Effect of cationic polymer addition on retenti on of alkylketene dimer", Journal of Pulp and Paper Science, 23(6) 276-281 (1997).

[8] Johansson, J. and Lindstrom, T., "A Study on AKD-size retenti on, reaction and sizing efficiency. Part 2: The effects of electrolytes, retention aids, shear forces and mode of addition of AKD sizing", Nordic Pulp & Paper Research Journal, 19(3) 336-344 (2004).

[9] van de Ven, T.G.M. and Alince, B., "Heteroflocculation by Asymmetric Polymer Bridging", Journal of Colloid and Interface Science, 181 73-78 (1996).

[10] Polverari, M. and van de Ven, T.G.M., "Dynamic light scattering of suspensions ofPEO-coated latex particles", Colloids and Surfaces, A: Physicochemical and Engineering Aspects, 86 209-228 (1994).

64

CHAPTER4

INTERACTIONS OF AKD AND OTHER PAPERMAKING ADDITIVES

ON A LABORATORY TWIN-WIRE SHEET FORMER MACHINE

4.1 Abstract

Using the twin-wire sheet former, we tested AKD retenti on on a time scale

resembling industrial paper machines. The twin wire former enabled us to perform a

detailed mass balance on fibers and precipitated calcium carbonate. Furthermore, the

sheet produced was analyzed for AKD content, giving us the percentage of bound and

unbound AKD retained under various experimental conditions. AKD had a low

retention on the TWF, likely due to the choice of AKD stabilizer and its properties.

No distinguishable differences in AKD retenti on were found between the first-pass

and simulated second-pass conditions. These finding suggests that PCC does not

diminish the retention efficiency of AKD. We also found that cPAM increases the

retention of AKD. Since cationic polyacrylamide do es not adsorb onto starch-coated

AKD, it further supports the mechanism of asymmetric polymer bridging described in

Chapter 3.

The results from these initial tests provide valuable information about the

interactions of wet-end chemicals on larger scale, and validate the results obtained on

a smalliaboratory scale as discussed in Chapters 2 and 3.

66

4.2 Introduction

The laboratory twin-wire sheet former (TWF) is an invaluable tool for

studying particle retention under realistic papermaking conditions. The TWF is a

replica of a vertiformer and was completed at McGill University in 1998 [1]. Since

then, this instrument has been used for several projects. Mitsui has used the TWF to

determine the optimal retenti on of PCC, and for the study of fines retention in

combination with PEO and cPAM [2]. In addition, Pigeon determined fines retention

as a function of TWF parameters [3]. Several students from École Polytechnique de

Montréal have used the machine for experiments with PEO and fines retenti on [4, 5].

In these previous trials, the focus was determining the influence of additives

on the retenti on of fines, as well as characterizing machine drainage. Since the TWF

is equipped with three addition ports, it can accurately portray the real-time

interactions of fibers and fines with retenti on aids, fillers, and other additives. In this

research, we conduct the first such experiment with sizing agents.

4.3 Machine Description

The TWF is composed of three separate sections (headbox approach, drainage

and web collection sections) that work in series to during a complete experiment (see

Figure 4.1). The fabrics can attain a speed of5.3 mis, which is close to the same time­

scale of an industrial paper machine.

67

Headbox Section

Drainage Section

Figure 4.1: Laboratory Twin-Wire Sheet Former (source: C. Hammock)

Web Collection

Section

The headbox approach loop is composed of a 400L holding tank and a pump

(Figure 4.2). The pulp suspension is stirred and circulated to ensure even mixing. A

mixer and baffles inside the holding tank ensure that the fibers do not settle to the

bottom of the tank. The centrifugaI pump controls the flowrate within the loop, as

weIl as the speed at which the pulp suspension is delivered onto the wire. At time

zero, the three-way valve is switched to divert the flow of pulp to the headbox,

denoting the start of a test mn.

The drainage section consists of the fast-rolling fabrics, foils and vacuum

boxes (see Figure 4.3). The water drains at various drainage points and is collected in

bins. The resulting paper sheet is blown off the wire into a collection bin and a

portion of this sample is bloWJ:l onto a roll to form a sheet that can then be dried and

68

Air Purge Valve

Stock

Air Purge Valve

CentrifugaI Pump

Three-Way Valve

Headbox

Figure 4.2: Headbox Approach Section ofTwin-wire Sheet Former

Guide Roll

Long Fabric

Tension Roll

Short

Guide Roll

Tension Roll

Figure 4.3: Drainage Section of Twin-wire Sheet Former

69

tested. The water drains by pressure blades (foils) and vacuum suction. More on the

exact design can be found in the two theses by Hammock [1, 6]. At the end of each

run, an 8% consistency paper sheet is collected (Figure 4.5). A complete manual for

the operation of the TWF is included as Appendix C.

Many agree that the interactions that occur in the primary pass are more

important than combined consecutive passes in determining the paper characteristics,

so it is vital to find what occurs during this first pass [7]. Because the TWF has no

recirculation (whitewater) loop, it is an excellent tool for studying true first-pass

retention. Whitewater from each drainage point is collected in individual collection

bins, making it simple to calculate the first-pass retention of components. Since

papermakers are encouraged to close their whitewater loops, this true first pass

retenti on value is nearly impossible to obtain in an industrial setting. This is even

more important considering that the water entering a typical paper machine has

already been recirculated and contains the non-retained particles from the first-pass.

70

(

Figure 4.4: Photograph of the twin-wire sheet fonner machine

Figure 4.5: Photograph of a paper sheet produced: the sheet is very thin (38 g/m2) and

broken in many places, but adequate for testing AKD retention

71

4.4 Materials and Methods

A dried bleached softwood kraft pulp obtained as dried sheets from Domtar

was used in aIl the experiments. The pulp sheets were wetted and tom into small

pieces (15cm x 15 cm) and soaked ovemight in tap water. The pulp was then

disintegrated to produce a low consistency (1 %) suspension from which a portion was

transferred to the holding tank. For each experiment, the concentration in the holding

tank was adjusted to 0.6% ±0.03 consistency, without any further processing. In sorne

experiments, PCC was also added to the holding tank at a concentration of 15% wt/wt

offibers.

Cationic polyacrylamide (Percol 292) supplied by CIBA, Inc was used as the

retention aid. This compound has a degree of substitution of 20% and a molecular

weight of 3x 106 Da. Solutions of 3.2 glL were prepared by wetting the cP AM with a

drop of ethanol and dissolving in de-ionized water. The solution was stirred for a

minimum of 2 hours.

Albacar HO, supplied by Specialty MineraIs, was used as the precipitated

calcium carbonate (PCC). This filler particle has an average particle diameter of 1.3

f..lm and a specifie surface area of 12 m2/g, as reported by the manufacturer. The PCC

stock was prepared by suspending the PCC in tap water and added to the holding tank

containing the pulp suspension. This mixture was stirred to ensure that the PCC and

pulp suspension were weIl dispersed.

AKD from our laboratory emulsions was used in aIl TWF experiments. These

AKD particles have a particle diameter of 0.8 f..lm and an electrophoretic mobility of

+1.33 [(f..lmls)/(V/cm)]. Typical additions of AKD in industry are 1mg/g fibres, which

72

is usually the concentration required to achieve appreciable sizing (600ppm AKD) on

the final paper sheet. These additions assume a 60% retenti on value, producing

sufficient AKD for adequate sheet sizing.

AKD is added to the first addition port when combined with cP AM, or in the

second addition port when tests were conducted without cPAM. Figure 4.6 shows the

arrangements used.

TI~ 60 ms

tx!D~ 210 ms

AKD cPAM 1 HEAD BOX

Figure 4.6: Additive locations and residence times on the TWF

Various tube lengths connecting the three-way valve to the headbox are

available for use with the TWF. Other researchers have used longer tubes to extend

the residence time of the additives [2]. In this study, we use only the shortest tube that

minimizes the distance between the three-way valve and the headbox. This length of

1.2m provides a residence time of 0.21 seconds from the second addition port to the

time flow reaches the headbox. Chemicals added to the first addition port (before the

three-way valve) have a residence time of 0.27 seconds. For colloidal particles, this

time scale is more than sufficient for chemical interactions to occur [8]. Mitsui has

found that residence times and tube lengths can be adjusted to optimize machine

operation [2]. These adjustments were not required in this study since we were

interested in obtaining results on a comparative basis.

73

A test fUll consisted of switching the three-way valve to redire ct the flow to

the drainage section. Simultaneously, the pumps for the additives were activated to

deliver the chemicals. The pulp suspension was pumped onto the drainage wire at a

flowrate of 4.8 kg/s for 22.5 seconds, the duration of each test run.

4.5 Analysis of Results

For each test run, seven drainage points and the wet web sheet were sampled

for filler content, fiber content, and AKD concentration. These values gave a

complete mass balance of each fUll.

PCC and fiber concentrations from the seven collection points and the wet

web were determined in a two-step process. By evaporating water from the sample,

we obtained the combined concentration of PCC and fibers. Next, the sample was

ashed 1 at 525°C to isolate the PCC alone. The concentration of fibers was calculated

from the difference of these two values.

Since AKD is not a pure material (the chain length varies), it is difficult to

analyze because it decomposes into various compounds [9]. Several methods to

analyze the AKD content have been employed, each with its advantages and

drawbacks. These methods incIude isotopic labeling [10], spectophotometric methods

lacing AKD with a fluorescent dye [11], capillary gas chromatography [12], as well

as extraction and ketone analysis, pyro1ysis gas chromatography, and visible/UV

spectroscopy [13].

1 According to Tappi Stand Method #413 om-93

74

In our analysis, a FID-GC2 method was used to determine the AKD retained

in the sheet [9]. A thin wet sheet was produced during one rotation of the metal drum

of the TWF. Sheets were taken off the drum, pressed and dried for 20 minutes at

105°C on a flat-bed drier and dried ovemight at standard conditions. Samples were

sent to Paprican, Point Claire, for AKD analysis using a solvent extraction Soxtec

apparatus. The sample was then analyzed using FID-GC to determine the quantity of

AKD in the sheet. Results were expressed as the bound, unbound, and total AKD

concentrations in ppm (mg AKD/ kg offibers).

Sizing efficiency was determined by performing contact angle measurements

on samples. The wet web was dried to remove water and mechanically pressed into

1.5cm-wide round discs with a Imm thickness. The discs were cured ovemight in a

sealed vessel to reduce AKD los ses by vaporization. The Contact Angle System OCA

instrument (Dataphysics, Germany) was used to record the rate of water penetration

into a dise. The contact angle was calculated as a function of time with software using

the Laplace-Young model.

4.6 Outline of Experiments

4.6.1 Reproducibility experiments

A set of experiments using PCC and fibers was performed initially to verify

the reproducibility of the twin-wire sheet former. Three runs were performed, which

were then analyzed for mass balance and filler retention. Figures 4.7 and 4.8 below

show the results from these trials. The three lines represent three experimental runs

2 Flame Ionization Detector- Gas Chromatography

75

under identical conditions. They produce the same machine drainage and consistency

for each run. Likewise, the fiUer retention of these three tests is similar for each.

These trials were run without the addition of a retention aid, explaining the low PCC

retention. Performing a statistical analysis, we find that our results were accurate to ±

0.5%, which is suitable for our purposes. In aU subsequent experiments, two runs

were performed under each condition; aU values presented are the average of these

two.

C 12.-----~----~----~----~------~----~----. o 'iii c 2i. 10 CI) ~ CI)

.:/ : .. / ..... ; ............... . /

/ ! 8 ......., ·············.,,;A', ........... . ~ 6 ................ L ............. _,~~<: .......... . ~ 4 ....... . ..... ; .. ,~/~~ ..... , ............. . (jj iii c o

Ü -Q) Q)

.s:::. CI)

./ :

2 ................ : .. .......... ;Jt7C.~ ...... : ................. ~ .............. ·T .. ······· ...... : ............... . : ,/: : : . : • :;.-0 : :: : . . . .

: : o~----,r-----r·-----+·----~------r-----T---~

Position in Drainage Section of TWF

Figure 4.7: Water Drainage along the TWF for three duplicate runs

76

10~--------------------------------.

8

2

Run#1 Run#2 Run#3

Figure 4.8: FiUer retenti on in the wet-web for three duplicate trials on the TWF

4.6.2 PCC Retention with Cationic Retention Aid

Previous trials by Mitsui showed that 0.8 mg cPAM/g fiber produces the

highest PCC retenti on, based on a 15% wt PCC/wt fiber addition [2]. In our

experiments, we use the same PCC amount and found that 0.8 mg/g results in a 55%

retention of PCc. Without the addition of retention aid, the PCC retenti on is minimal,

averaging 5%. Figure 4.9 demonstrates how the addition of cationic polyacrylamide

increases the PCC deposition onto fibers.

77

70.-------------------------------~

60 ---. ~ -- 50 C o ~ 40 C Q)

+-' Q) 30

0::: Ü 20 Ü a..

10

o mg/g 0.8 mg/g

cPAM addition [mg cPAM/g fiber]

Figure 4.9: Influence of cP AM on PCC retention on the TWF

4.6.3 First-Pass Retention vs. Second Pass Retention with AKD

Tests were performed adding PCC, AKD, and cPAM, varying the addition to

resemble either a first-pass or second-pass ron through a paper machine. The second-

pass conditions were simulated by combining the AKD emulsion together with

additional PCC at the first addition port. The AKD-PCC suspension was stirred for at

least 1 hour, to simulate how AKD and PCC would interact in whitewater conditions.

Results show that the AKD retention was extremely low, such that the second-pass

retention of AKD could not be distinguished from the first pass retention, possibly

due to the low AKD levels retained. Altematively, these results could indicate that

AKD and PCC do not interact, suggesting that PCC does not affect AKD retention.

78

4.6.4 AKD Retention at Various Additions

A set of tests was performed to determine the amount of AKD necessary to

achieve an adequate degree of sizing in the sheet. Tests were performed with only

fibers and AKD. It was found that at typical additions of 1 mgAKD/g fiber, the

retention was only 3.2% AKD. The same low retention percentage was observed at 5

mg/g and 10 mg/g additions. Typically, 600ppm AKD in the sheet is necessary to

impart sufficient sizing. Using 1 mg AKD/g fibers, we ob tain retenti on of 35 ppm;

using 5 mg/g, we obtain 150 ppm. This low retention necessitated the addition often

times the normal amount of AKD (10 mg/g) to achieve good sizing in the sheet. Even

then, we could not retain the necessary quantity of AKD without the use of retention

aids.

4.6.5 AKD Retention with cPAM

Tests were also conducted with fibers, AKD and cPAM. In contrast to results

in the tests described in section 4.6.4, we found that cPAM increased the AKD

retention in the sheet, from 3.2% without cPAM, to 9% with addition of 0.8 mg

cPAM/g fiber, as seen in Figure 4.1. This is consistent with results found in Chapter 3

regarding AKD and cP AM adsorption kinetics with fibers. The rate of water

penetration into a sample of the wet web was determined using the contact angle

instrument. For 5 mg AKD/g fiber water penetration was instantaneous. 10 mg

AKD/g fiber were required to produce a sufficient contact angle (>80° for more than

one minute). These findings with the twin wire former further support our theory of

79

asymmetric polymer bridging as a mechanism for AKD retention lU the paper

machine.

10.-------------------------------~

C 8 o

+=i C 2 6 Cl)

0::: Cl 4 :::s::: <C '$.2

o mg/g 0.8 mg/g

cPAM addition [mg cPAM/g fiber]

Figure 4.10: The effect of cP AM on AKD retenti on on the TWF

4.7 Conclusions

Using the twin-wire sheet former, we were able to conduct retention tests on a

time scale resembling industrial paper machines. The twin wire former enabled us to

perform a detailed mass balance on fibers and calcium carbonate. Furthermore, the

sheet produced was analyzed for AKD content, giving us the percentage of AKD

retained under various experimental conditions. The three main conclusions obtained

from these studies are as follows:

1. AKD has a very low retention on the TWF. We hypothesize that this is due to

the low electrophoretic mobility of the AKD particles.

80

2. No distinguishable differences in AKD retenti on were found between first­

pass and simulated second-pass retention of AKD. These finding suggests that

PCC does not diminish the retention efficiency of AKD.

3. Cationic polyacrylamide increases the retention of AKD. Since cPAM does

not adsorb onto starch-coated AKD; this finding further supports the

mechanism of asymmetric polymer bridging described in Chapter 3.

The results suggest that the TWF has not yet been optimized for the study of AKD

retention. However, the results from these initial tests do provide extremely valuable

information about the interactions of wet-end chemicals on a larger scale, and

validate the results obtained on a small laboratory scale as discussed in Chapters 2

and 3.

4.8 Acknowledgements

We would like to thank B. Sitholé, J. Pimentel, and A. Gagné at Paprican, Pointe

Claire, for the AKD analysis. Thank you to Marcius De Oliveira, Loius Godbout, and

Alois Vanerek for help in running the twin-wire sheet former machine, and most of

aIl to Chris Hammock for training in machine operation.

81

4.9 References

[1] Hammock, C., "Development ofa laboratory-scale twin-wire sheet former", M.Eng. thesis, Chemical Engineering, McGill University, (1998).

[2] Mitsui, K., "Mechanism of Fines Retention and Drainage with a Polyacrilamide/Bentonite Retention Aid", thesis, Chemical Engineering, McGill University, (2000).

[3] Pigeon, C., "The effects of a PEO-cofactor-bentonite system in papermaking", M.Sc. thesis, Chemistry, Mc Gill University, (2001).

[4] Hermann, A., "Étude de l'enchevêtrement et de l'efficacité de l'oxyde de polyethylene utilizé comme agent de rétention sur des machines à papier", M.Sc. thesis, Chemical Engineering, École Polytechnique, (2004).

[5] Périn-Levasseur, Z., "Étude expérimentale d'un système de rétention des particules fines dans un formeur de laboratoire ", M.Eng. thesis, Chemical Engineering, Ecole Polytechnique, (2004).

[6] Hammock, C., "Mechanisms of drainage, retenti on and formation during web forming in papermaking", PhD thesis, Chemical Engineering, McGill University, (2005).

[7] Smook, G.A., "Handbook for Pulp & Paper Technologists", Angus Wilde Publications, Vancouver, 2nd Ed., (1992).

[8] Alince, B., "Time factor in pigment retention", Tappi Journal, 79(3) 291-294 (1996).

[9] Sithole, B.B., Nyarku, S. and Allen, L.H., "Chromatographie methods for the determination of alkylketene dimer in pulp and paper matrixes", Analyst (Cambridge, United Kingdom) , 120(4) 1163-9 (1995).

[10] Johansson, J. and Lindstrom, T., "A Study on AKD-size retention, reaction and sizing efficiency. Part 2: The effects of electro1ytes, retenti on aids, shear forces and mode of addition of AKD sizing", Nordic Pulp & Paper Research Journal, 19(3) 336-344 (2004).

[11] Esser, A. and Ettl, R."On the mechanism ofsizing with alkyl ketene dimer (AKD): physico-chemical aspects of AKD retention and sizing efficiency." in Fundamentals of Papermaking Materials, Transactions of the Il th Fundamental Research Symposium Cambridge, UK, (1997). 997-1020.

[12] Dart, P.J. and McCalley, D.V., "Determination of alkylketene dimer sizing agent products in paper by capillary gas chromatography", Analyst (Cambridge, United Kingdom), 115(1) 13-16 (1990).

[13] Jaycock, M.J. and Roberts, J.C., "A new procedure for the analysis of alkyl ketene dimers in paper", Paper Technology, 35(4) 38-42 (1994).

82

CHAPTER5

CONCLUSIONS

5.10verview

Since AKD has a low first-pass retention (~40%), a large portion enters the

whitewater cycle and has time to interact with many types of particles including PCC.

Although the mechanisms behind AKD sizing and reactions with cellulose have been

studied in-depth, mechanisms of AKD retention are poody understood. It is also

unknown how interactions in the whitewater affect subsequent passes of AKD.

The objectives ofthis research were three-fold. First, we attempted to characterize

the interactions between AKD and PCC by simulating wet-end conditions. Secondly, we

investigated the interactions between AKD and cPAM. Thirdly, we performed

experiments on a large-scale twin-wire sheet former to determine the chemical

interactions of AKD in real-time.

We prepared an AKD emulsion stabilized by cationic starch, free from extraneous

substances. After verifying that its properties were similar to those of a commercial

emulsion, we used our laboratory emulsions for further testing. Using the Photometric

Dispersion Analyzer, we found that the AKD emulsions flocculate PCC at an optimal

dosage of 10 mg AKD/g PCC. However, subsequently we found that flocculation is

caused by the cationic starch acting as the AKD stabilizer.

No visual proof was found that AKD and PCC form flocs. The SEM photographs

show that PCC and AKD each homoflocculate; however, little to no heteroflocculation

occurs. This, in combination with the results obtained from the PDA tests, suggests that

PCC and AKD do not directly interact. Instead, removal of starch from AKD aids PCC

homoflocculation (via polymer bridging). Starch transfer causes the AKD particles to

84

/" aggregate, since their stabilizing polymer (starch) has been depleted. It can therefore be

concluded that cationic starch has a higher affinity for PCC than for AKD.

More importantly, we can conclude that AKD and PCC are unlike1y to interact in

the whitewater circulation loop. This dismisses the idea that PCC will lower AKD

retention. Although PCC does decrease the sizing of a sheet by increasing the specific

surface area, this would not influence AKD single-pass retention.

The second part of our research dealt with the adsorption kinetics of AKD onto

fibers. First, we found that cPAM does not flocculate cationic AKD. This result was

expected since both additives have a positive charge and repel one another. Adsorption

kinetic experiments revealed that cP AM increases the retention of AKD by assisting its

adsorption to fibers. Asymmetrical polymer bridging is able to explain the increase in

AKD retention, as well as predict trends in the flocculation kinetics with cP AM. The

main conclusions drawn from this study are:

1. Without cP AM, AKD deposition goes through a maXImum, after which

detachment occurs.

2. Cationic polyacrylamide increases the AKD deposited on fibers by asymmetric

polymer bridging, and increases the bond strength between AKD and the fiber,

eliminating detachment of AKD mole cules under laboratory conditions.

3. The initial kinetics of AKD deposition (up to 20%) is nearly independent of

cP AM dosage.

4. With excess cPAM, attaining the maximum retention is delayed due to starch

and/or cPAM re-conformation.

85

For anionic emulsions, cP AM increases AKD deposition on fibers in relation to

the amount added, for less than full fiber coverage. With fibers fully coated with cP AM,

AKD deposition is fast and remains at 100% due to the increase in bond strength. Steric

repulsion does not occur, since the kinetics of AKD deposition onto fibers is faster than

cP AM adsorption onto AKD particles.

Our third focus was to measure the AKD interactions on a time scale relevant to

industrial paper machines, using the twin-wire sheet former (TWF). The TWF enabled us

to perform a detailed mass balance on fibers and calcium carbonate. The sheet produced

was analyzed for AKD content, giving the percentage of AKD retained under various

operating conditions. Unfortunately, AKD retention on the TWF is very low; this may be

because the AKD particles had a low electrophoretic mobility. Altematively, varying the

tube length in the headbox approach section may have improved retenti on by changing

the residence times of the additives. No distinguishable differences in AKD retenti on

were found between fresh AKD and that which had been mixed with PCC prior to the

test. These findings suggest that PCC does not diminish the retention efficiency of AKD.

Lastly, it was found that cPAM increases the retention of AKD and does so on a very fast

time scale. Because cPAM does not adsorb onto starch-coated AKD, this further supports

the asymmetrical polymer bridging mechanism.

It was found that PCC and AKD do not directly interact. Rather, AKD retention

~-', is increased by the addition of cP AM. The mechanism behind these interactions has been

86

shown to be due to asymmetric polymer bridging. Cationic polyacrylamide aids the AKD

retention by first adsorbing to fibers, after which it is able to adsorb onto AKD particles.

5.2 Recommendations for Future Work

Steps have been taken to make the twin-wire sheet former an effective tool in

measuring AKD retention on a larger scale, though much work remains to be done in this

respect. It is clear that AKD retention values were not optimized. Further experiments

should be conducted on the TWF with an AKD emulsion containing a different

stabilizing polymer. The use of a cationic starch with a higher degree of substitution

would increase AKD retention by increasing the electrophoretic mobility of the emulsion.

Altematively, amphoteric emulsions, containing both anionic and cationic stabilizing

polymers, could be used to stabilize the emulsion. Furthermore, trials may also be

performed using anionic AKD emulsions.

87

APPENDICES

Appendix A: Princip les ofFlocculation

Particles attract each other for a variety of reasons. In the presence of

polymers, particles may aggregate via polymer bridging or patch aggregation. With

other types of additives, charge neutralization or coagulation caused by a higher ionic

strength may occur.

The method of aggregation determines much of the properties of the flocs, so

it is important to know the mechanism of particle flocculation. It is known that in

papermaking, precipitated calcium carbonate (PCC) will deposit on to other filler

particles or fibres via polymer bridging (with cationic polyacrylamide as the

polymer). Thus, studying this flocculation behavior can improve our understanding of

AKD and filler retenti on in the paper web.

Precipitated calcium carbonate (PCC) is positively charged in de-ionized

water, but reverses its charge due to impurities in the whitewater. Thus, in a

papermaking environment, PCC assumes a negative charge. Since fibres are also

negatively charged, cationic retenti on aids are added to bridge the two. Several types

of flocs may be formed, depending on charge density and the properties of the

polymer, as weIl as the shear rate that the particles experience.

The main variable that dictates whether collisions will be dominated by ortho-

or peri-kinetic interactions is the Peclet (Pe) number. This dimensionless quantity is a

ratio of shear and diffusion forces:

[k: Boltzmann constant, T: temperature (K), G: shear rate, l1:viscosity, a: particle

radius] Thus, when Pe» 1, orthokinetic rates are dominant in determining the

89

collision frequency. The value ofPe dictates the type of collisions that will occur and

the kinetic model that can be used to predict particle behavior. In PDA experiments,

the Peclet number is well above a thousand, indicating that shear forces.

Photometrie Dispersion Analyzer (PDA) Operation

The photometrie dispersion analyzer was used to study PCC and AKD

flocculation. The PDA consists of a light emitting diode (LED) at 850nm and a

photo-detector (PIM) photodiode that continuously monitors the transmitted light.

The frequency of measurements ranges from 6-2000 Hz. A plastic block allows for

the precise fitting of tubing that lies in the light path. A suspension flowing through

the photosensitive cell is illuminated by a narrow beam of light perpendicular to the

flow. The volume illuminated by the light beam is of the order of Imm3. The PDA

works on the principle that in any emulsion or suspension, there exist small local

fluctuations in composition which follow the Poisson distribution, such that the

standard deviation about the mean is the square root of the average number of

particles in the sample volume (i.e. in the light beam).

The current from the photodiode is converted to a voltage signal, which has a

large DC component, with a smaIl fluctuations. The DC component is a measure of

the average transmitted light intensity and is dependent on the turbidity of the

suspension. The AC component arises from random variations in the number of

particles in the sample. The AC (fluctuating) component is separated from the DC

component and amplified by a factor of up to 500. Because the suspension flows

through the ceIl, the actual sample in the light beam is continuaIly renewed, and local

90

variations in particle number concentration give fluctuations in the transmitted light

intensity. These fluctuations cease when the flow is stopped.

The root-mean-square (RMS) value of the fluctuating signal is related to the

average number concentration and size of the suspended particles. For fairly uniform

suspensions, estimates of particle size and number concentration can be made, but the

main use of the PDA2000 is in monitoring flocculation and dispersion processes. The

RMS value of the fluctuating signal increases when particle aggregation occurs.

Measurable changes in the RMS value occur long before visible signs of aggregation.

Conversely, when aggregates are disrupted, the RMS value decreases, reaching a

minimum when disaggregation (or dispersion) is complete. The DC value (related to

the average value of the turbidity) is much less sensitive to changes in the state of

aggregation. Another output is the ratio ofRMS to DC reading:

RATIO = RMS De

This RATIO output is useful for routine monitoring of flocculation or dispersion

processes. The RATIO varies with solids concentration in a straightforward manner,

and is mostly unaffected by contamination of the tube walls in the flow cell or by drift

in the electronic components. These effects cause changes in the DC value (and can

be a serious problem in flow-through turbidity measurements). However, the RMS

value of the voltage fluctuations is changed in the same proportion, so that the ratio of

the two values do es not change. This means that the ratio output can be monitored for

long periods without the need for re-calibration.

Concentrations from a few ppm to several percent solids can be monitored.

For more concentrated suspensions, sorne form of in-line dilution would be

91

necessary. The limits of partic1e size are roughly 0.5 to 100J.1m. In non-uniform

suspensions, larger partic1es give the dominant contribution to the RMS and ratio

values. The concentration and partic1e size limits are inter-related to sorne extent.

Thus, for larger partic1es, lower concentrations can be detected.

92

Appendix B. Starch Clustering and E([ects ofShear and Sonication

The influence of sonication and stirring on starch particle was tested. Particle

size was measured using a Brookhaven Goniometer BI-200SM with particle size

software. This dynamic light scattering apparatus works by passing polarizable

partic1es through a beam of light. The electric field induces partic1es to radiate light in

all directions by a dynamic light scattering method.

Our studies found that starch size is a function of starch concentration and

sonication. Initially, starch c1usters were large (greater than 1 micron). Stirring

decreases the partic1e size, however, sonication is much more effective breaking up

c1usters. In fact, we find that sonication reduces the starch to single particles. The

results from the dynamic light scattering and software analysis can be seen in the

graphs below. The graphs show the partic1e size distribution on the right, while the

left column presents the calculated value for the mean partic1e diameter. Figure (A)

shows the starch size without any sonication. In contrast, Figure (B) demonstrates the

partic1e size after 3 minutes of sonication. Figure (C) shows a sample with the same

starch concentration, sonicated for 5 minutes.

Furthermore, dynamic light scattering was performed for vanous

concentrations of starch solutions. The results demonstrated that as the concentration

is decreased, the partic1e size decreases as weIl. These results confirm our premise

that a more di lute starch solution will result in smaller AKD partic1es, since more

starch is available to coat its surface.

93

A

Sample ID Cato237 100 Operator ID Agatha Elapsed Time 00:08:39 Mean Diam. 1786.1 (nm) ... Rel. Var. 2.638 Skew 2.008 o .j.U~=*::It+L..u..4-U~.LL~ RmsError 3.8445e-03 5.00e-01 50000.0

Diameter (nm)

d G(d) C(d) d G(d) C(d) d G(d) C(d) 1.00 10 1 562.34 88 59 1.78 0 1000.00100 72 3.16 10 3 1778.28 83 83 5.62 0 3 3162.28 0 83

'10.00 0 3 5623.41 63 92 17.78 16 5 10000.00 63 100 31.62 17 7 56.23 18 10

100.00 93 22 177.83 97 35 316.23 93 47

B

Sample ID Cato237 Operator ID Agatha Elapsed Time 00:10:25 Mean Diam. 310.6 (nm) ... Rel. Var. 0.279 Skew -0.233 RmsError 1.7224e-03 5.00e-01 5000.0

Diameter (nm)

d G(d) C(d) d G(d) C(d) d G(d) C(d)

1.00 0 0 259.76 67 51 1.66 0 0 430.61 100 97 2.75 0 0 713.82 7 100 4.56 0 0 1183.30 0 100 7.55 6 3 1961.57 0 100

12.52 0 3 3251.70 0 100 20.75 0 3 34.40 18 11 57.02 18 19 94.53 0 19

156.70 a 19

,~

94

~', C

Sample ID Cato237 100 Operator ID Agatha

>-Elapsed lime 00:12:36 "" (1.1

Mean Diam. 292.7 (nm) ~ c: 50 ID

Rel. Var. 0.203 1;;

Skew -0.021 0 RmsError 3.1776e-03 5.0 5000.0

Diameter (nm)

d G(d) C(d) d G(d) C(d) d G(d) C(d)

1.00 0 0 214.94 55 41

1.63 0 0 350.22100 92

2.65 0 0 570.64 17 100

4.33 0 0 929.81 0 100

7.05 0 0 1515.03 0 100

11.49 0 0 2468.60 0 100 18.71 0 0 30.49 0 0 49.68 25 13

80.96 0 13 131.91 0 13

95

Al!pendix C. Twin-Wire Sheet Former Machine (TWF) Instructions

A detailed description for operating the twin-wire sheet former is given below:

DAYl • Tear one sheet ofpulp into small pieces for each circular drum (2 drums). • Fill each drum to ~ 100 L (preferably with hot water) and let soak overnight.

DAY 2 • Dilute pulp slurry to 160 L in each barrel. • Attach mixer to barrel. Ensure that it is angled into drum and off-centered to

create a circular flow pattern and let mix during the day. • At the end, fill up the barrels to the 180 L mark.

Note: Safety Precautions • Always wear ear protection and safety goggles when operating the twin-wire

former • Never wear loose clothing that can become stuck in the machine • Never work alone in the room when running the TWF

DAY 3 • Turn the mixers back on to re-agitate the pulp suspensions. • Switch up the main power source • Switch on PUMP and PRESS power source boxes

• Close yellow valve (headbox drainage pipe) under the holding tank • Open the green air valve (this controls the 3-way valve). • Check that 3-way (S) valve is ON (recirculation mode).

• Fill the holding tank with tap water from the hose until water begins pouring out the vent in the bottom portion of the tube. Plug this with the brass plug or an addition port. Wrap plug with teflon tape before inserting.

• Fill headbox to the 15 cm mark with water. • Turn on mixer (dial is on top of mixer) to gently agitate the solution. • Transfer pulp slurry into headbox bucket by buckets. Fill the tank with 4 expts

worth ofpulp (60L x 3 = 180 L); di lute with tap water up to ~55cm mark. • As pulp is being added, slowly turn on the pump. Then raise pump motor to

945 rpm. While recircu1ating the solution, the upper portion of the tube will begin to fill and start to suck air into the air vent. Once this is happening, close the air vent by turning it to a horizontal position.

• Connect the gray tube from 3-way valve to impingement box (headbox). • Use green hose to fill water level to the 59.5 cm mark. • Let the circulation loop fUll for about 5-10 minutes while the rest of the

equipment is being setup.

96

/-',

Equipment setup

1. Plug in the large scale and move the other scale to the center of the room. 2. Tare the large tank on the scale. Make sure none of the pipes are resting on the

large collection tank when taring. Also position the web collection container. This should be tilted against the machine, and inclined with a block of wood.

3. Attach the three pipes leading to the three smaller collection tanks 4. Set the three vacuums in place; (first attach the pipes, next the heads). 5. The next part of the equipment setup is the air supply. 6. In the back, tum on both air tanks. This should give a pressure of 220kPa for

the inner and 425kPa on the outer fabric, producing 2-3 kN/m on the fabric. 7. Open the valve for the web blowing. However, ensure that the ON valves are

only at a 45degs, and not completely ON. 8. Control Box: Switch on Press slowly. 9. While this is happening, with the spray nozzle, wash the forming fabric with

water, in an area so as to prevent any water from collecting in the tanks. This is best done at the beginning of the forming sections, just after the headbox impingement, and at the wire separation point prior to web collection.

Running a single experiment

1. Place the safety polycarbonate-glass shield into position. 2. Tum ON the PUMP control and ramp up to 945. (Note: the pump should be

running at this value for a few minutes prior to starting the experiment to create a well-distributed pulp suspension in the headbox.

3. Tum ON the PRESS control and ramp up to 1750 (a fabric speed of approx 5.3 mis.) Go quickly through 1000-1300, as the machine vibrates in this range.

4. Flip the "fibre collection" switch up. 5. Next flip the "vacuum" switch up. This will activate the first two vacuums

immediately. After a few seconds, the third, final vacuum will be activated. At this point, the experiment can be started.

6. Simultaneously flip the s-valve switch to OFF and press stopwatch. 7. After 22.5 seconds, tum ON the s-valve and stop the time. Record the time

duration of experiment.

Shutdown procedure ..... After each run

1. 2.

3.

4. 5. 6. 7.

Stop rolls by ramping down the PRESS dial, and then switching to OFF. Keep the pump motor running, but ramp the dial so that it slows to a minimum. Disconnect aIl the pipes leading to the collection tanks and vacuums. Shake aIl remaining water out of the pipes. Next, weigh aIl collection tanks, and take fluid samples from each. Also sample the solution from the storage tank at this point. Empty the contents of the collection tanks into the sink For the web collection: place in a plastic container and seal air-tight.

97

Immediate Values (wei2;ht} Samll.Ies fJ!r Anall!.,sis 1 ° fonning roll 1 ° fonning roll waters

2° fonning roll 2° fonning roll waters 3° fonning roll 3° fonning roll waters Initial and final height of headbox 1 st 2nd 3rd vacuum weight , , 1 st 2nd 3rd vacuum waters , , Duration of run (s) Wet web collection mass Wet web bin weight Headbox Consistency

Preparation for the next experiment

1. Slow down the pump motor to 173 (minimum). Then turn off so that the surface level of the water can become still.

2. Measure the height of the water in the storage tank. 3. To this amount add 60 L of solution from the large mixing drum. 4. Fill the level up to 59.5 cm again. 5. Resume circulation with the pump. 6. Replace aIl tanks and pipes as they were before.

Shutdown procedure .... At the end of the day

1. Perfonn same shutdown as described above .. .in addition ... 2. Empty the remains of the headbox. This is done by opening the valve leading

the red hose to empty into the floor drain. AIso, a bucket may be placed to divert flow from the flow exiting the 3-way valve.

3. Clean the impingement box with the water spray nozzle. 4. Also make sure to clean the fabrics and rolls, etc. (while running the fabric at

low speed). 5. Fill the storage tank a bit with tap water and recirculate water through pump

and the pipes to clean the system. 6. Turn off the air supply in the back. Air release: tum off the knob on the

cylinders, and next the black handles corresponding to the top and bottom belt (these are located on the back of the TWF).

7. Air release: turn off the yellow air valve next to the control panel. However, sorne air will still be trapped between the valve and the mini-motor. To release this air, pull out the tiny metal pin on the top side of the mini-motor. A loud gush of air will be released.

o Tum OFF the power supply boxes for the PRESS and PUMP. o Turn OFF the MAIN power supply box.

98

Storage Procedures

o Store collection tanks and vacuums upside down.

Variations with the pipe length

o A short pipe connecting the 3-way valve to the impingement box is the basic setup for the machine. Depending on various addition sequences, longer pipes may be used. This will extend the residence time of the solution, and also provide several ports along the pipe at which it is possible to add different additives to the pulp slurry.

DAY 4

o Pulp consistency measurements o weight filter paper o Pour contents of plastic bottle into a funnel and drain through. o Place the filter paper in an ovemight. o Weigh the filter paper the next day twice, to ensure accurate readings.

o Web consistency measurements o Weigh 4 aluminum dishes oPIace 1 sample in each dish o Leave in the oyen ovemight o Measure oyen-dry weight in the moming

o Fines measurements o Take solution from headbox and pour thru the 76J.l1l1 filter screen. o Weigh a piece of filter paper o Take the residual fibers that remain on the mesh and pour them onto a

filter paper in a funnel. o Place the filter paper in an ovemight. o Weigh the filter paper the next day twice, to ensure accurate readings.

o Ash content measurements o Weigh 4 crucibles oPiace 1 sample (either process water or wet web) in each dish o Leave in the muffle fumace ovemight at 525°C o Measure oyen-dry weight in the moming.

99