Rheology and Microstructure of Cellulose Acetate - Digital Repository

224
APPAW, COLLINS. Rheology and Microstructure of Cellulose Acetate in Mixed Solvent Systems (Under the direction of Dr. Saad A. Khan and Dr. John F. Kadla) Cellulose is a natural abundant polymer used in a variety of applications. Its use however, is hampered by its poor solubility in various solvents. This is primarily due to the hydrogen bonds between the hydroxyl groups on the anhydoglucose chain. In view of that, various cellulose derivatives have been synthesized to aid dissolution and this impacts a variety of solubility characteristics. Cellulose acetates (CA) are cellulose esters that are partially substituted at the C-2, 3 and 6-positions of the anhydroglucopyranose residue. Their solubility in various solvents depends on the degree of substitution (DS) of the acetyl groups. For instance, CA is soluble in water at low DS of between 0.5-1. But it is insoluble in aqueous solutions at higher degree of substitution (DS > 1). CA is employed in various applications such as textile manufacture, tool handles, specialty papers, cigarette filters and is a polymer of choice in majority of reverse osmosis membrane preparation. These applications often exploit semi- to concentrated cellulose acetate solutions in appropriate solvents. Such systems can be induced to form aggregated structures such as gels which can be initiated by the addition of a non-solvent. Thus, depending on the solvent and non-solvent adopted, the cellulose acetate mixed solvent system can be tailored to exhibit sol-gel characteristics utilizing the inherent intra- and intermolecular interactions present in solution. However, such systems and the interactions influencing their behavior is not very well understood. In this regard, the main objectives of our study are as follows – to develop a

Transcript of Rheology and Microstructure of Cellulose Acetate - Digital Repository

Page 1: Rheology and Microstructure of Cellulose Acetate - Digital Repository

APPAW, COLLINS. Rheology and Microstructure of Cellulose Acetate in Mixed

Solvent Systems (Under the direction of Dr. Saad A. Khan and Dr. John F. Kadla)

Cellulose is a natural abundant polymer used in a variety of applications. Its use

however, is hampered by its poor solubility in various solvents. This is primarily due to

the hydrogen bonds between the hydroxyl groups on the anhydoglucose chain. In view

of that, various cellulose derivatives have been synthesized to aid dissolution and this

impacts a variety of solubility characteristics. Cellulose acetates (CA) are cellulose

esters that are partially substituted at the C-2, 3 and 6-positions of the

anhydroglucopyranose residue. Their solubility in various solvents depends on the

degree of substitution (DS) of the acetyl groups. For instance, CA is soluble in water at

low DS of between 0.5-1. But it is insoluble in aqueous solutions at higher degree of

substitution (DS > 1). CA is employed in various applications such as textile

manufacture, tool handles, specialty papers, cigarette filters and is a polymer of choice

in majority of reverse osmosis membrane preparation. These applications often exploit

semi- to concentrated cellulose acetate solutions in appropriate solvents. Such systems

can be induced to form aggregated structures such as gels which can be initiated by the

addition of a non-solvent. Thus, depending on the solvent and non-solvent adopted, the

cellulose acetate mixed solvent system can be tailored to exhibit sol-gel characteristics

utilizing the inherent intra- and intermolecular interactions present in solution.

However, such systems and the interactions influencing their behavior is not very well

understood. In this regard, the main objectives of our study are as follows – to develop a

Page 2: Rheology and Microstructure of Cellulose Acetate - Digital Repository

ternary mixed solvent system comprising of cellulose acetate, N,N dimethylacetamide

and water and manipulate the system to form aggregated structures leading to phase

separated gel network. The tools employed in this project to investigate and

characterized the macroscopic properties as well as the microstructural changes are

rheology, scanning electron microscopy (SEM) and laser scanning confocal microscopy

(LSCM).

The first part of this study involves addition of water- N,N dimethylacetamide

solutions in different ratios to bulk 20% cellulose acetate in N,N dimethylacetamide

solutions with emphasis on increasing water content in the system. Using rheology as

the main analytical tool, the steady state viscosity was found to increase with water

content increase. Above water concentrations of 19%, there is a solution to gel

transition, which also showed enhancement in dynamic viscoelastic properties with

water content increment. The SEM micrographs showed similar patterns with gels

having lower water content exhibiting larger voids in comparison to gels with higher

water content at same cellulose acetate concentration. Using LSCM, we obtained

microstructural formation with more open networks at lower water content, whiles a

more compact homogenous structure was exhibited for higher water content gel

samples.

In the second part of this study, addition of cellulose acetate to different ratios of

N,N dimethylacetamide/water solutions are investigated. At low water content, the

system showed steady state viscosity increase with water content as was observed in the

first part of this thesis. Typically beyond 19% water concentration, the systems phase

Page 3: Rheology and Microstructure of Cellulose Acetate - Digital Repository

separates into two layers consisting of a clear solution on top of a viscous bottom layer.

This is in contrast to the first part of the study where we observed a uniform rigid

material. Heating the two-phase system to 100oC and cooling back to room temperature

led to the formation of a one-phase physical gel matrix. With increasing water content,

the elastic and viscous moduli of the gels increased at constant cellulose concentration.

Finally, we investigate the gel properties with emphasis on yield stress when

mechanical stress is applied to the gels. In addition, the gel-sol transition for the gels are

investigated by subjecting them to temperature variations.

Page 4: Rheology and Microstructure of Cellulose Acetate - Digital Repository

RHEOLOGY AND MICROSTRUCTURE OF CELLULOSE ACETATE IN

MIXED SOLVENT SYSTEMS

by

COLLINS APPAW

A dissertation submitted to the Graduate Faculty of North Carolina State University

in partial fulfillment of the requirements for the Degree of Doctor of Philosophy

December, 2004

Department of Chemical and Biomolecular Engineering North Carolina State University

Raleigh, NC 27695-7905

APPROVED BY:

Dr. Saad A. Khan Dr. John F. Kadla Chair of Advisory Committee Co-Chair of Advisory Committee

Dr. Richard D. Gilbert Dr. Orlin D. Velev

Page 5: Rheology and Microstructure of Cellulose Acetate - Digital Repository

ii

Personal Biography

Collins Appaw was born in the city of Accra, which is located in Ghana, West Africa.

His desire in pursuing a career in Chemical Engineering is primarily due to his interest

in math and chemistry. His interest in research led him to pursue a Master program in

Chemical Engineering at North Carolina A & T State University after completing his

Bachelor of Science degree from the Department of Chemical Engineering, University

of Science and Technology (UST) Ghana. After completing his Master degree, he

immediately joined North Carolina State University to begin work on his doctoral

research in the Fall of 2000. He plans to work in an industrial setting beginning 2005.

Page 6: Rheology and Microstructure of Cellulose Acetate - Digital Repository

iii

Acknowledgements

I will first of all to express my deepest gratitude to all whose help in various ways

enabled me to complete this manuscript. First, I am grateful to my advisors Dr. Saad

Khan and Dr. John Kadla, for their guidance and mentorship during the entire period I

have been working on my research. I would like to thank Dr. Richard Gilbert and Dr.

Orlin Velev for their contributions to my project from their fields of expertise and also

for taking the time to serve on my committee. In addition, I would like to thank Dr.

Spontak for his assistance in confocal microscopic analysis. I would also like to

acknowledge the United States Department of Agriculture for funding this research

study.

I enjoyed working and sharing some wonderful relationships with past and present

members of the Rheology Group, Dr. Fedkiw’s post-doc members and some members

of Wood and Paper Science Department (Now at University of British Columbia,

Canada), including Matthew Burke, Adeola Ali, Jeremy Walls, Jeff Yerian, Ahmed

Abdala, Ahmed Eissa, Angelica Sanchez, Lauriane Scanu, Shamsheer Mahammed,

Sachin Tawlar, David Franskowski, Satioshi Kubo, Quizon Dai, and Cameron Morris. I

am especially indebted to my family for their love, support and encouragement through

all the years. Finally, and most importantly, I would like to say thank you to my God

and personal savior Jesus Christ.

Page 7: Rheology and Microstructure of Cellulose Acetate - Digital Repository

iv

TABLE OF CONTENTS List of Tables..............................................................................................................vii

List of Figures.............................................................................................................viii

Chapter 1. Introduction and Overview

Abstract……………………………………………………………………………….1

1.1 Introduction……………………………………………………………...........2

1.2 Goals of Project…………………………………………………..…………...5

1.3 Overview of Thesis…………………………………………………………...6

References…………………………………………………………………….............8

Chapter 2. Background and Literature Review

Abstract………………………………………………….…………………………...13

2.1 Introduction to Cellulose and Cellulosics………………………………….....14

Solution characteristics of cellulose derivatives …………………….15

Gelation of cellulose ethers and esters……………………………….16

2.2 Physical, Chemical and Solution Properties of Cellulose Acetate…………...18

2.3 Applications and New Emerging Uses of Cellulose Acetate………………...20

2.4 Gelation Behavior of Cellulose Acetate……………………………………...21

References…………………………………………………………………………....23

Chapter 3. Experimental Methods

Abstract……………………………………………………………………………....30

3.1 Rheological Characterization………………………………………………....31

Steady state rheology………………………….……………………....31

Dynamic rheology…………………………………………………….32

3.2 Cloud Point Study……………………………………………........................33

3.3 Scanning Electron Microscopy………………………………………………34

3.4 Laser Scanning Confocal Microscopy……………………………………….35

References……………………………………………………………………………37

Page 8: Rheology and Microstructure of Cellulose Acetate - Digital Repository

v

Chapter 4. Effect of Water Composition on Rheological Properties in Ternary

Cellulose Acetate Solutions

Abstract…………………………………………………………………………….46

4.1 Introduction………………………………………………………………...47

4.2 Experimental Methods…………………………………….……………….49

4.3 Results and Discussion…………………………………………………….52

Conclusions………………………………………………………………………..58

References……………………………………………………………………........60

Chapter 5. Phase Separation and Heat-Induced Gelation Characteristics of

Cellulose Acetate in a Mixed Solvent System

Abstract…………………………………………………………………………....79

5.1 Introduction…………………………………………………………..........80

5.2 Experimental Methods………………………………………………….....83

5.3 Results and Discussion……………………………………………….........85

5.4 Conclusions………………………………………………………………..93

References………………………………………………………………………....95

Chapter 6. Evolution of Microstructure and Rheological Properties in Cellulose

acetate/N,N dimethylacetamide/Water system

Abstract………………………………………………………………………….116

6.1 Introduction………………………………………………………….......117

6.2 Experimental Methods…………………………………………………..120

6.3 Results and Discussion…………………………………………………..123

6.4 Conclusions………………………………………………………….......130

References……………………………………………………………………….132

Chapter 7. Conclusions and Recommendations

7.1 Conclusions…………………………………………………………......151

7.2 Recommendations…………………………………………………........154

Page 9: Rheology and Microstructure of Cellulose Acetate - Digital Repository

vi

Appendix A. Cellulose Gels and Liquid Crystalline Polymer (LCP)

Introduction…………………………………………………………………….159

References……………………………………………………………………...185

Page 10: Rheology and Microstructure of Cellulose Acetate - Digital Repository

vii

List of Tables Appendix A

Table 1. Chiroptical properties of lyotropic AEC/AA solutions with different degree of acetylation (DA). …208 Table 2. Relaxation times as the function of chiroptical properties of lyotropic AEC/AA solutions. …209

Page 11: Rheology and Microstructure of Cellulose Acetate - Digital Repository

viii

List of Figures Chapter 1 Figure 1. Chemical structure of cellulose …10 Figure 2. Schematic representation of a fully derivatized cellulose acetate chain ...11 Figure 3. (a) Polymer in solution and (b) Aggregated network structure …12 Chapter 3 Figure 1. Viscosity as a function of shear rate depicting different material responses: Shear thickening, Newtonian, Shear thinning …40 Figure 2. Elastic (G') and viscous (G") moduli as a function of angular frequency showing a typical dynamic rheological response of polymer solutions and elastic gels …41 Figure 3. Schematic diagram of the laser setup showing a laser source, cuvette and a detector for recording the response …42 Figure 4. Schematic representation of a typical scanning electron microscopy (SEM) showing basic components of the apparatus …43 Figure 5. Schematic of a typical laser scanning confocal microscopy setup …44 Chapter 4 Figure 1. Chemical structure of a typical cellulose acetate chain with the hydroxyl groups replaced with acetyl groups …68 Figure 2. Effect of different water contents on steady shear viscosity for a CA/DMAc/water system. CA concentrations are (a) 10wt% and (b) 15wt% …69 Figure 3. Viscosity as a function of water content obtained at shear rate of 1s-1 for CA concentrations of 10 and 15wt% …70 Figure 4. Transmission intensity versus amount of water for 10 and 15wt% CA concentration in DMAc/water system …71 Figure 5. Transition of 10wt% CA/DMAc/water system from homogenous system to a cloudy system. Water content increment is showed from left to right …72

Page 12: Rheology and Microstructure of Cellulose Acetate - Digital Repository

ix

Figure 6. Effect of low and high water contents on the viscous (G″) and elastic (G′) moduli for CA/DMAc/water system. CA concentrations are (a) 10wt% and (b) 15wt% …73 Figure 7. Multifrequency plot of tan δ as function of water content for (a) 10wt% and (b) 15wt% CA/DMAc/water system …74 Figure 8. G' versus water content at a frequency of 1 rad/sec for 10wt% and 15wt% CA/DMAc/water system …75 Figure 9. SEM micrographs for CA concentration of 10 and 15wt% at two different water contents …76 Figure 10. An illustrative representation of molecular level interactions present in the system. At low water content, the solvents are dispersed forming solutions. At high water content, CA form strong and weak links with solvents in the system and leads to gel formation. Dash lines and solid lines represent weak and strong hydrogen bonds respectively …77 Chapter 5 Figure 1. Effect of different water content on the steady shear viscosity of cellulose acetate in DMAc/water cosolvent system. The total concentration of cellulose acetate is kept constant at 10wt% …104 Figure 2. Transmission intensity versus amount of water for a constant 10wt%cellulose

acetate in DMAc/water cosolvent system …105

Figure 3. Constant 10wt% cellulose acetate in DMAc/water cosolvent system showing transition from homogenous system to a two-phase system. Water content increment is showed from left to right …106 Figure 4. Viscosity pattern for the two-phase region for cellulose acetate inDMAc/water cosolvent system at two different water contents of 21.6 and 25.2wt%. Polymer concentration is a constant 10wt%. (a) Upper liquid and (b) Lower opaque layer ...107

Figure 5. Dynamic frequency spectrum of the elastic and viscous moduli for bottom viscous layer for cellulose acetate in DMAc/watercosolvent system at two different water contents of 21.6 and 25.2wt%. Total concentration of cellulose acetate is kept constant at 10wt% …108

Figure 6. Dynamic spectrum for dissolution methods 1 and 2 for 10wt% CA concentration with water contents of (a) 21.6% and (b) 25.2% ...109

Page 13: Rheology and Microstructure of Cellulose Acetate - Digital Repository

x

Figure 7. Two phase system subjected to heat treatment to 100oC after which system is cooled to room temperature …110

Figure 8. Dynamic spectra for heat-induced gelation for 10wt% cellulose acetate

DMAc/water cosolvent system at two different water contents of

21.6 and 25.2wt% …111

Figure 9. SEM images for heat-induced CA concentrations of (a) 10 and (b) 15wt% at two different water contents …112 Figure 10. Comparison of dynamic properties of heat-induced gels and dissolution method 1 gels at 10wt% cellulose acetate concentration having water content of (a) 21.6% and (b) 25.2% …113 Figure 11. SEM micrographs of heat-induced and dissolution method 1 gels for 10wt% cellulose acetate concentration having water contents of (a) 21.6% and (b) 25.2% …114 Chapter 6 Figure 1. Dynamic frequency sweep experiments for (a) CA concentration of

10wt% (b) CA concentration of 15wt% having different water contents …140

Figure 2. G′ as a function of water content for CA concentration of 10 and 15wt.%

Data were obtained at frequency of 1 rad/s …141

Figure 3. CSLM images for CA concentration of (a) 10wt% and (b) 15wt% at two

different water contents …142

Figure 4. Stress sweep experiments conducted for (a) different CA concentrations

at constant water content of 21.6wt.% (b) constant CA concentration of 15wt%

with varying water contents …143

Page 14: Rheology and Microstructure of Cellulose Acetate - Digital Repository

xi

Figure 5. Yield stress as a function of water content for varying

CA concentration …144

Figure 6. Digital image showing transition of a gel to solution at two different

temperatures for 10wt% CA concentration with water content of 21.6% …145

Figure 7. Temperature ramp experiments conducted at 1oC/min (a) for 10wt%

CA (b) 15wt% CA concentration at different water contents …146

Figure 8. Dynamic frequency sweep experiment for (a) 10wt.% CA and

(b) 15wt.% CA concentration at different temperatures …147

Figure 9. Dependence of G' and G" on temperature for (a) 10wt% CA system having

water content of 21.6wt.% (b) 15wt.% CA system having water content of 20.4wt.%.

Data obtained at a frequency of 1 rad/s …148

Figure 10. Arrhenius plot for (a) 10wt% CA concentration having water content of

21.6% (b) 15wt% CA concentration having water content of 20.4% …149

Appendix A Figure 1. Effect of different water contents on the steady shear viscosity of cellulose acetate in DMAc/water cosolvent system. The total concentration of cellulose acetate is kept constant at 10wt%. ...198 Figure 2. Viscosity pattern for the two-phase region for a constant 10wt% CA in

DMAc/water system. …199

Figure 3. Effect of different water contents on steady shear viscosity for a cellulose

acetate/DMAc/water system. CA concentration is 10wt%. ...200

Page 15: Rheology and Microstructure of Cellulose Acetate - Digital Repository

xii

Figure 4. Effect of low and high water contents on the viscous (G″) and elastic (G′) moduli for 10wt% cellulose acetate/DMAc/water system. …201 Figure 5. Helicoidal structure of chiral nematic mesophase. …202

Figure 6. Viscosity as a function of concentration for lyotropic LCPs. …203 Figure 7. Three-regime steady state shear viscosity for LCPs. …204 Figure 8. Viscosity as a function of shear rate of AEC/AA solutions …205 Figure 9. Reduced shear stress σ* vs. time for 50% AEC-4( ) and 51.5% AEC-2(∆) …206 Figure 10. Complex modulus evolution of AEC solutions vs. time. ...207

Page 16: Rheology and Microstructure of Cellulose Acetate - Digital Repository

1

CHAPTER 1

INTRODUCTION AND OVERVIEW

Abstract

In this chapter, we offer the reader a brief introduction to the characteristics of

cellulose and cellulose derivatives (cellulosics), in particular to cellulose acetate, the

cellulosic studied in this dissertation. The natural abundance and biodegradability of

cellulose together with its ability to provide unique (solution and end-use) properties

through derivatization have made cellulosic polymers attractive for usage in a wide

range of applications including textile fibers, molding powder sheets, optical

membranes among others. This thesis involves the development and characterization of

a ternary system consisting of a cellulose acetate, a solvent (N,N dimethylacetamide)

and a non-solvent (water). We examine the rheology and microstructure of the system

as a function of composition and processing parameters such as temperature. A variety

of experimental techniques are employed to gain a fundamental understanding of how

the mode of interactions impact cellulose acetate in the mixed solvent systems.

Page 17: Rheology and Microstructure of Cellulose Acetate - Digital Repository

2

1.1 Introduction

Polymers and biopolymers are used in numerous applications with the most

abundant biopolymer being cellulose.1 We see an emergence in polymeric materials

replacing traditional materials (e.g. metals) in many areas of applications in everyday

products. This range from car body parts to everyday items such as cooking utensils.

These polymers are natural or synthetic products, which science and research have

studied to investigate their applicability in various applications.

Cellulose is the most abundant natural polymer (Figure 1) that obtained its

name from Payen in 1842.2 The cellulose structure shows a chain-like extended linear

macromolecule of anhydro-D-glucopyranose units linked at the 1 and 4 positions by

glycosidic bonds.2 The use of cellulose together with its derivatives has wide spread

applications including fibers, films, plastics, coatings, suspension agents, composites.3

With the advent of synthetic polymer their use has somewhat dwindled, but some

applications still adopt cellulose derivatives as the raw materials of choice. In addition,

various studies are still been conducted to look for and expand their usage in existing

and new technologies. The inherent problem that faces users of cellulose is the general

insoluble nature in most common solvents.4 There is therefore the need to modify the

structure of cellulose to improve solubility which has led to the synthesis of various

cellulose derivatives.5

Cellulose derivatives (cellulosics) comes in all forms and structures depending

on the functional group(s) used to substitute the hydroxyl groups on the cellulose

chain.6-10 The derivatized cellulose molecule can be partially or fully substituted and

this can go a long way in influencing solution and end-use behavior. To understand the

Page 18: Rheology and Microstructure of Cellulose Acetate - Digital Repository

3

behavior of cellulosics calls obviously for intensive research efforts in various

disciplines. Knowledge of molecular level interactions will aid researchers to

manipulate their properties and impart specific properties to the cellulosic system.

Aqueous soluble cellulosics such as methylcellulose are used as food-thickening agents

etc.3 Other classes of cellulosics including cellulose acetate and cellulose nitrate are

used as cigarette filters and adhesives respectively. In spite of its potential for numerous

applications though, both cellulose and its derivatives are being relegated to the

background in preference to other polymeric materials. Thus, it is imperative to tap into

and exploit the potential for using cellulosics in existing technologies and develop new

applications for putting them to use. In view of that, we choose cellulose acetate as the

primary polymer to study.

Cellulose acetate is a derivative of cellulose, which exhibits different solubility

pattern depending on the degree of substitution of the hydroxyl units, which are

replaced with O-acetyl groups. A typical cellulose acetate structure is shown in Figure

2 which has three acetyl groups on the chain and such units are termed as cellulose

triacetate. The presence of acetyl and hydroxyl groups on the chain causes cellulose

acetate to show interesting properties in various solvents.10 Therefore, the dissolution of

cellulose acetate can be influenced depending on the affinity of a solvent for a specific

functional group. For example, the rheology of cellulose acetate in different solvents is

postulated to be affected by specific interactions such as hydrogen bonding in

solution.12 The use of cellulose acetate usually involves concentrated solutions but

unfortunately research in this area is few and far between.10 A possible explanation may

be due to the viscous nature of relatively high cellulose acetate solutions making system

Page 19: Rheology and Microstructure of Cellulose Acetate - Digital Repository

4

characterization difficult since most techniques such as traditional dynamic light

scattering work best in dilute solutions.13 Therefore, to gain a concise understanding,

which may lead to better predictions of cellulose acetate under such conditions, it is

imperative to study semi- to high concentration levels. Thus, our ternary system

comprising of cellulose acetate, N,N dimethylacetamide and water is studied at fairly

high cellulose acetate concentrations. Phase separation and sol-gel transition is an

important feature of our work since these characteristics are of considerable interest for

application purposes. An example in this regard is the pharmaceutical industry, which is

interested in phase separation characteristics in microencapsulation.14 We employ

rheological techniques as our primary tool to investigate the macroscopic properties of

our system. It must be emphasized that, rheology can also be used to predict the

microstructural properties of various systems. Rheology has also been determined as the

most direct and reliable method to determine sol-gel transition.15 Quite a number of

cellulose gel and membrane system use scanning electron microscopy (SEM) to

investigate microsctructural changes. In view of this we adopted both SEM and a

relatively new technique called laser scanning confocal microscopy (LSCM) to

complement our rheological analysis. Our hope is the use of these techniques enabled us

gain a fundamental understanding of how the mode of interactions impact cellulose

acetate in mixed solvent systems.

Page 20: Rheology and Microstructure of Cellulose Acetate - Digital Repository

5

1.2 Goals of Project

The main objective for this thesis is to examine the structure and property

relationship between cellulose acetate and mixed solvents comprising N,N

dimethylacetamide and water. The role of how polymer-cosolvent composition and

processing parameters (e.g. temperature, cosolvent mixture) have on polymer solution

properties inducing sol-gel transition in the system is examined. Furthermore, how such

changes can affect the polymer conformation and interaction leading to aggregated

structures (see Figure 3), and eventual phase-separation.

Several critical questions and issues need to be addressed in this project. They

include:

• How does polymer solvent interactions affect solution rheology and microstructure?

• Under what conditions can cellulose acetate form aggregation leading to phase-

separated gels?

• How does process parameters (e.g., temperature) and sample preparation method

influence rheology and microstructure?

The specific step-by-step goals are as followings:

1. Develop a ternary system comprising of cellulose acetate, solvent and non-

solvent

The presence of hydroxyl and acetyl groups in partially substituted cellulose acetate

molecule enable the polymer architecture to be influenced by cosolvents with different

solubility character. This prompted us to adopt two solvents, which are miscible with

each other but having different solubility behavior for cellulose acetate. N, N

dimethylacetamide is a good solvent for cellulose acetate and was chosen based on the

Page 21: Rheology and Microstructure of Cellulose Acetate - Digital Repository

6

fact that, it can dissolve cellulose acetate to appreciable high concentrations. Also, other

considerations included, high boiling point, low toxicity levels. Water the non-solvent

for this study was chosen primarily because of its strong hydrogen bonding

characteristics and obvious wide spread availability.

2. Examine the influence of cosolvent and cellulose acetate concentration

pertaining to phase separation and sol-gel transition

The influence of non-solvent and how they can cause solutions of cellulose acetate to

phase separate and eventually form elastic gel-like materials has been an important

focus of this study. Phase separation in cellulose acetate system typically occurs when

the solvent quality deteriorates. In this regard, we studied the effect of changing solvent

quality by using different ratios of the cosolvents and progressively increase water

content in the system. This process was also important to decipher the modes of

hydrogen bond associations and dissociations in the system.

1.3 Overview of Thesis

In chapter 2, we provide a review of cellulose and cellulose derivatives. The

chapter attempts to provide an investigative study of the various mechanism and mode

of interactions exhibited by cellulosic systems. Details of the experimental methods

adopted in our study are presented in chapter 3. Rheological methods are introduced and

the basic concepts pertaining to the use of scanning electron microscopy (SEM) and

laser scanning confocal microscopy (LSCM) to investigate the microstructure of the gel

system are briefly outlined.

Page 22: Rheology and Microstructure of Cellulose Acetate - Digital Repository

7

Chapter 4 deals with the rheological investigation of solutions and sol-gel

mechanism of cellulose acetate, N,N dimethylacetamide solutions by the addition of

different water concentrations (dissolution method 1). SEM studies are also performed

to characterize and explain the gel system as a result of water content changes in the

gels. In chapter 5, we examine a different route for solubilizing cellulose acetate in the

cosolvents, which involves the addition of cellulose acetate to N, N,

dimethylacetamide/water solutions (dissolution method 2). This is in contrast to the

solutions prepared in chapter 4 where we added water to cellulose acetate/ N,N

dimethylacetamide solutions. Rheology and SEM were used to investigate the solutions

and phase separated two-phase system formed by this alternative addition process. In

chapter 6, we study the behavior of gels formed from both dissolution methods with

respect to their response to stress and temperature changes In addition we depict the

microstructural properties of the gels using LSCM. Finally, we conclude the thesis with

a summary of our findings and recommendations for future work in chapter 7.

Page 23: Rheology and Microstructure of Cellulose Acetate - Digital Repository

8

References

1. Nobles, Jr. D.R.; Brown, Jr. M.B. The pivotal role of cyanobacteria in the

evolution of cellulose sysnthase and cellulose synthase-like proteins. Cellulose

2004, 11, 437-448.

2. Myasoedova, V.V. Physical Chemistry of Non-aquoeus Solutions of Cellulose

and its Derivatives. John Wiley & Sons, New York, NY 2000.

3. Gross, R.A.; Scholz, C. Biopolymers from Polysaccharides and Agroproteins.

Oxford University Press, 2001.

4. Saalwachter, K.; Burchard, W. Cellulose in new metal-complexing solvents. 2.

Semidilute behavior in Cd-tren. Macromolecules 2001, 34, 5587-5598.

5. Tsunashima, Y.; Hattori, K. Substituent distribution in cellulose acetate: Its

control and the effect on structure formation in solution. Journal of Colloid and

Interface Science 2000, 228, 279-286.

6. Kobayashi, K.; Huang, C.; Lodge, T.P. Thermoreversible gelation of aqueous

methylcellulose solutions. Macromolecules 1999, 32, 7070-7077.

7. Roschinski, C.; Kulicke, W-M. Rheological characterization of aqueous

hydroxylpropyl cellulose solutions close to phase separation. Macromolecular

Chemistry and Physics 2000, 201(15), 2031-2040.

8. Kulicke, W-M.; Kull, A.H.; Thielking, H. Characterization of aqueous

carboxymethylcellulose solutions in terms of their molecular structure and its

influence on rheological behavior. Polymer 1996, 37(13), 2723-2731.

Page 24: Rheology and Microstructure of Cellulose Acetate - Digital Repository

9

9. Boerstoel, H.; Maatman, H.; Picken, S.J.; Remmers, R.; Westerink, J.B. Liquid

crystalline solutions of cellulose acetate in phosphoric acid. Polymer 2001, 42,

7363-7369.

10. Pintaric, B.; Rogosic, M.; Mencer, H.J. Dilute solution properties of cellulose

diacetate in mixed solvents. Journal of molecular liquids 2000, 85, 331-350.

11. Gomez-Bujedo, S.; Fleury, E.; Vignon, M.R. Preparation of cellouronic acids

and partially acetylated cellournic acids by TEMPO/NaClO oxidation of water-

soluble cellulose acetate. Biomacromolecules 2004, 5, 565-571.

12. Wang, C.S., Fried, J.R., Viscoelastic properties of concentrated cellulose acetate

solutions. Journal of Rheology 1992, 36, 929-945.

13. Abdala, A.A. PhD Thesis. Solution rheology and microstructure of associative

polymers. Department of Chemical Engineering, North Carolina State

University, Raleigh, NC 2002.

14. Khalil, S.A., Phase separation of cellulose derivatives: effects of polymer

viscosity and dielectric constant of nonsolvent. Jour. of pharmaceutical Sciences

1973, 62, 1883-1884.

15. Li, L., Thermal gelation of methylcellulose in water: scaling and

thermoreversibility. Macromolecules 2002, 35, 5990-5998.

Page 25: Rheology and Microstructure of Cellulose Acetate - Digital Repository

10

O

O

O

OH

OH

HOO

OH

OH 1

2

3

4

5

6HO

Figure 1. Chemical structure of cellulose.

Page 26: Rheology and Microstructure of Cellulose Acetate - Digital Repository

11

Figure 2. Schematic representation of a fully derivatized cellulose acetate chain.

Page 27: Rheology and Microstructure of Cellulose Acetate - Digital Repository

12

(a) (b) Figure 3. (a) Polymer in solution and (b) Aggregated network structure

Page 28: Rheology and Microstructure of Cellulose Acetate - Digital Repository

13

CHAPTER 2

BACKGROUND AND LITERATURE REVIEW

Abstract

Cellulose is the principal constituent of higher plants (40-50%) and is also the main

component of various kinds of natural fibers. In as much as cellulose has many

commercial uses, cellulose is hindered by the inability to be dissolved by most

solvents. There is therefore, the need to modify the cellulose structure by

derivatization methods to partially or fully replace the hydroxyl groups on the

cellulose chain with various other functional groups. This process yields cellulose

derivatives, which exhibit better solubility characteristics and thus provide a wider

range of usage. In this regard, we present a summary of cellulose and its derivatives

and how their solution patterns are affected by different solvents. We then briefly

discuss the factors contributing to gel formation in some cellulose derivatives and

how the microstructure is affected by gelation. After which we shift the discussions

to cellulose acetate, the cellulosic being investigated in this thesis with respect to

solution and gelation characteristics in a variety of solvents.

Page 29: Rheology and Microstructure of Cellulose Acetate - Digital Repository

14

2.1 Introduction to Cellulose and Cellulosics

Cellulose is the most abundant naturally occurring biopolymer.1,2 Various

natural fibers such as cotton and higher plants have cellulose as their main

constituent.2,3,4 Cellulose consists of long chains of anhydro-D-glucopyranose units

(AGU) with each cellulose molecule having three hydroxyl groups per AGU, with the

exception of the terminal ends.2 Cellulose is insoluble in water2 and most common

solvents. The poor solubility is attributed primarily to the strong intramolecular and

intermolecular hydrogen bonding between the individual chains.1 In spite of its poor

solubility characteristics, cellulose is used in a wide range of applications including

composites, netting, upholstery, coatings, packing, paper, etc. Chemical modification of

cellulose is performed to improve processability and to produce cellulose derivatives

(cellulosics) which can be tailored for specific industrial applications as well to enable

cellulosics to be characterized in the laboratory.5 Cellulose derivitization typically

involves esterification or etherification of the hydroxyl groups on the cellulose chain.5

Cellulosics are in general strong, reproducible, recyclable and biocompatible,5 being

used in various biomedical applications such as blood purification membranes and the

like. Thus, through derivatization, cellulosics have opened a window of opportunity and

have broadened the use of cellulosics. (Cellulosics have been around for ~100years!

And actually used more as “plastics” before petroleum plastics came around)

Cellulose esters constitute the largest sector of commercially important cellulose

derivatives. Of these, both organic esters such as cellulose acetates and cellulose acetate

butyrates as well as inorganic esters such as cellulose nitrate and cellulose phosphate

are widely produced. Together with cellulose xanthates, the cellulosic intermediate in

Page 30: Rheology and Microstructure of Cellulose Acetate - Digital Repository

15

Rayon production, cellulose esters comprise more than 90% of the production capacity

in the chemical processing of cellulose. The manufacturing of cellulose esters typically

involves the reaction of cellulose with the corresponding organic / inorganic acid and a

strong acid catalyst and depending on the substituent and degree of substitution are used

as textile fibers, coatings, cigarette filters, membranes, celluloid etc.3

By contrast, cellulose ethers are manufactured through alkali reactions,5 are

typically water soluble and widely used to modify the rheological properties in

industrial applications such as paint, oil recovery, food and cosmetics.6 Some of the

most common cellulose ethers are methyl cellulose, hydroxyethyl cellulose and

carboxymethyl cellulose (CMC), with CMC dominating by far.

The utilization of cellulose and cellulose derivatives require a comprehensive

understanding of the physical and chemical properties and behavior of the various

derivatives. In view of that, more research is needed to better understand at the

molecular level the various interactions and mechanisms influencing their properties

and behavior.

Solution characteristics of cellulose derivatives

Cellulose derivatives exhibit different solution properties depending on the

solvent system and the functional group(s) used to substitute the hydroxyl group(s) on

cellulose chain. Cellulose esters such as cellulose nitrate and cellulose acetate dissolve

in a wide range of solvents.20 For instance, cellulose acetates exhibit solubility

characteristics in both aqueous and common organic solvents, such as chloroform

(CH2Cl2), acetone and DMF depending on the DS. Likewise, cellulose ethers such as

Page 31: Rheology and Microstructure of Cellulose Acetate - Digital Repository

16

methyl cellulose and other hydrophobically modified cellulosics are soluble in aqueous

solutions as well as some organic solvents depending on the degree of substitution (DS)

and preparation method.7-16 Similarly, factors such as temperature, concentration, shear

rate and degree of substitution have been found to affect the molecular structure of

hydroxypropyl cellulose (HPC) solutions.17 HPC gels have been shown to exhibit

greater elasticity with increased propylene glycol (PG) content in mixtures of water and

ethanol.18 It has been observed that for aqueous carboxymethyl cellulose (CMC)

solutions factors such as molar mass and concentration of the dissolved CMC are the

major factors affecting its rheological properties.19 In low CMC concentration solutions

Newtonian behavior was observed, while pseudoplastic, thixotropic and viscoelastic

behavior was exhibited at high concentrations.20 It has been shown that in sodium CMC

solutions crystalline regions act as binding or crosslinking sites and depending on the

solvent or salt type lead to the formation of swollen gels or aggregates in solution.21

Gelation of cellulose ethers and esters

The inception of gelation is initiated by large macromolecular associations

forming clusters, which lead to an infinite homogenous cluster extending through entire

volume of the system.22 Solution to gel transitions have been observed in various

cellulose derivatives, forming mostly physical gel systems. It is widely accepted that a

physical polymer gels is initiated by physical aggregation leading to the formation of a

three-dimensional network.22 The transitions from solutions to gels have been studied

using various techniques including rheological methods.5, 7-8 Rheology has been

Page 32: Rheology and Microstructure of Cellulose Acetate - Digital Repository

17

determined to be the most direct method to determine sol-gel transitions in polymeric

systems.12 Not only can rheology be used to investigate the gelation mechanism, it also

provides important characterization properties which invariably assist in assigning

specific applications for polymers. This technique has been applied to various cellulosic

systems to precisely determine and control gel formation. Phase separation behavior is

often exhibited by cellulosic systems.23-24 It has been shown that phase separation and

gelation are two characteristics which are closely related.25 Gelation in cellulosic

systems typically exhibits phase separation characteristics prior to gelation. Phase

separation in cellulosic solvent systems, are dependent on the concentrations of specific

derivative and solvents used.26-27 This phenomena typically occurs when solvent quality

deteriorates or changes in the polymer-solvent interactions occur due to external factors

such as temperature, non-solvent addition etc.28-29 These rearrangement of the intra- and

intermolecular interactions in solution due the above-mentioned factors may account for

gel formation in such cellulosics systems. A variety of water-soluble cellulosics exhibit

a unique thermoreversible gelation process.5-10 The unique feature of these systems is

the induction of a three-dimensional network and gel formation with increasing

temperature, then the return to an isotropic solution when cooled.6 Such

thermoreversible behavior is observed for methyl cellulose solutions and has been

attributed by some researchers to hydrophobic interactions in solution. Manipulation of

methyl cellulose gelation has been accomplished through the addition of salting-in and

salting-out salts.30-31 Sodium chloride, a typical salting-out salt has been shown to aid

gelation,30 while sodium iodide (salting-in salt) inhibited gel formation.31 Likewise, the

Page 33: Rheology and Microstructure of Cellulose Acetate - Digital Repository

18

use of ionic surfactants have been shown to greatly affect gel network strength in other

aqueous cellulose derivative systems.25

Unlike cellulose ethers, reports involving gel formation of cellulose esters are

rather limited and may be due in part to their general decline in such applications. Not

withstanding gelation has been observed in a variety of cellulose ester systems.32-33 Gels

prepared from cellulose esters usually find applications such as filters and binding

agents.34 Separation of metals in aqueous solutions have been performed using cellulose

nitrate gel matrixes,34 while cellulose phosphate gels have been synthesized to be used

in orthopedic applications especially for bone regeneration.35

2.2 Physical, Chemical and Solution Properties of Cellulose Acetate

Cellulose acetates (CA) are one of the most widely produced cellulose esters. It

is an acetylated cellulosic in which the hydroxyl groups along the cellulose chain are

fully or partially acetylated in a statistical manner.36-42 All of the industrial processes

practiced today are aimed at the manufacture of a fully substituted cellulose triacetate

(DS> 2.9, or an acetyl content of 44.8%) as the primary product.42 This is either isolated

and processed as is or partially desubstituted to “secondary acetate” with a DS of

between 1.8 and 2.5 (predominately near 2.4).42 A heterogeneous process to produce a

uniform diacetate is not currently available. The majority of CA is produced using a

solution process, although fiber acetylation is also practiced. Typically cotton linters or

softwood sulfite or prehydrolyzed sulfite pulps are used as the raw material. In solution

acetylation either glacial acetic acid alone or in combination with methylene chloride is

employed as a solvent for the CTA formed, and the reaction is catalyzed using sulfuric

Page 34: Rheology and Microstructure of Cellulose Acetate - Digital Repository

19

acid. In secondary acetate production, the CTA is not isolated rather being immediately

converted to the diacetate by the addition of water, dilute acetic acid or NaOH. This

process affectively decomposes the intermediate sulfate ester groups in the chain and

decrease the DS while still maintaining the cellulose acetate in solution. The

manufacture of CA differs in the solvent used for the acetylation and process conditions

adopted.41 Acetic acid process by far represents the route for producing the largest

amount of cellulose acetate for commercial purposes. With development of new

solvents for cellulose, new path for acetylation processes are still being investigated

since acetylation require dissolution of cellulose.42

The solution properties of cellulose acetates have been well studied and have

been shown to be influenced by the average degree of substitution and the distribution

of substituents along the chain.40 The structural properties of CA affect the rheology of

the cellulose acetate-solvent system. Cellulose acetates which have degree of

substitution (DS) of between 0.5 to 1 are soluble in water.43-44 This phenomena is

attributed to disruption of the intra- and intermolecular hydrogen bonding within the

cellulose system upon the introduction of the small number of acetyl groups.44 With

increasing DS, cellulose acetates becomes insoluble in water but show good solubility

in a variety of organic solvents, such as tetrahydrofuran (THF), acetone, N,N

dimethylacetamide (DMAc) with cellulose acetates having high DS being insoluble in

aqueous solution. Similarly, depending on the DS, concentration and solvent cellulose

acetates exhibit a variety of solution properties. Boerstoel et al. found that over a DS

range of 0.23 to 2.89, cellulose acetate was soluble in phosphoric acid and exhibited

liquid crystalline properties.45 Using dilute solutions of cellulose diacetate (CDA) in

Page 35: Rheology and Microstructure of Cellulose Acetate - Digital Repository

20

mixed solvents it has been reported that basic solvents (e.g. acetone) interact primarily

with the hydroxyl groups on the cellulose chain, while acidic solvents (e.g. formic acid)

primarily solvate the acetyl groups.46 Therefore, the use of specific solvents can induce

structural changes in solution depending on the amount of acetyl and hydroxyl groups

on partially substituted CA chains. This in some cases lead to phase separation and/or

gelation in a variety of CA solvent systems. Matsuyama et al.47 used supercritical CO2

to cause phase separation and was observed to occur instantaneously when CO2 was

added to cellulose acetate in acetone, methyl acetate, 1,3-dioxolane and 2-butanone

solutions respectively. CA samples with different molecular weight and degree of

substitution have been showed to form gels in dimethyl phthalate.48 Increase in degree

of substitution of CA and polymer concentration in the system, causes an increase in the

dynamic moduli.48

2.3 Applications and New Emerging Uses of Cellulose Acetate

The use of CA dates back to the early 1900s when the first photographic film

was produced by the Dreyfus brothers.49 Properties including good toughness, deep

gloss and a “natural” feel have made CA attractive for a number of applications. It also

has the advantage of possessing low toxicity. CA is used in a variety of industrial

applications such as textile manufacture, tool handles, speciality papers, cigarette filters

(major constituent in the U.S. for producing cigarette filter tow), among others. Factors

such as transparency, smoothness and color have allowed CA to be used in handles for

toothbrushes, umbrellas and screwdrivers. Spectacle frames have been made from

sheets of cellulose acetate, but this practice has been replaced to some extent with other

Page 36: Rheology and Microstructure of Cellulose Acetate - Digital Repository

21

synthetic thermoplastic materials. CA is increasing being utilizated in the

manufacturing of separation and filtration media in such applications as electrophoresis

and reverse osmosis membranes (ROM).50-51 As well, CA has shown good promise in

the area of drug delivery and research is underway to determine its effectiveness and

applicability.

2.4 Gelation Behavior of Cellulose Acetate

Previous work on the gelation mechanism of cellulose acetate has shown

interesting behavior with respect to the sol-gel transition. CA gels exhibit thermal

reversible properties, which depend on factors such as CA concentration, acetyl content

and the type of solvent. Unlike other cellulose esters which gel when heated, CA

solutions do not exhibit such gel characteristics. In most CA/solvent systems gelation

occurs after the CA solution is heated to a specific temperature and subsequently

cooled. Such behavior has been reported for CA/benzyl alcohol solutions.52-53 Heating

benzyl alcohol solutions of CA to 60-80oC led to gelation when the system was cooled

to 25oC. It is postulated that gel formation is induced by the existence of strong

intermolecular associations in the system.52

In polymeric systems physical gel formation can arise as a result of phase

separation. Depending on the system phase separation leads to polymer aggregation and

the formation of large macromolecular assemblies.22 In some CA/solvent systems the

addition of a non-solvent (e.g. water) can induce gelation in which phase separation

characteristics are exhibited prior to gel formation. Phase separation is typified by the

observation of a gradual to extreme cloudiness in the system and is dependent on the

Page 37: Rheology and Microstructure of Cellulose Acetate - Digital Repository

22

concentration of polymer and cosolvents used. Such phase separation induced CA

gelation is used in the production of wet phase inversion membranes.54 Here water is

the nonsolvent of choice owing to its nontoxic nature, widespread availability and

strong hydrogen-bonding characteristics. However, the application of other non-

solvents such as alcohol-water mixtures have been studied.55,56

Page 38: Rheology and Microstructure of Cellulose Acetate - Digital Repository

23

References

1. Hinterstoisser, B.; Salmen, L. Application of dynamic 2D FTIR to cellulose.

Vibrational Spectroscopy 2000, 22, 111-118.

2. Bochek, A.M. Effect of hydrogen bonding on cellulose solubility in aqueous and

nonaqueous solvents. Russian Journal of Applied Chemistry 2003, 76(11), 1761-

1770.

3. Myasoedova, V.V., Physical chemistry of non-aqueous solutions of cellulose

and its derivatives. John Wiley & Sons © 2000.

4. Gross, R. A., Scholz, C., Biopolymers from polysaccharides and agroproteins.

American Chemical Society © 2000.

5. Hon, D.N.-S., Shiraishi, N. Wood and cellulosic chemistry. Marcel Dekker, Inc.

New York 2001.

6. Hirren, M., Chevillard, C., Desbrieres, J., Axelos, M.A.V., Rinaudo, M.

Theromogelation of methylcellulose: new evidence for understanding the

gelation mechanism. Polymer 1998, 39(25), 6251-6259.

7. Bochek, A.M., Zabivalova, N.M., Lavrent’ev, V.K., Lebedeva, M.F.,

Sukhanova, T.E., Petropavlovskii, G.A., Formation of physical thermal

reversible gels in solutions of methyl cellulose in water and dimethylacetamide

and properties of films thereof. Russian Journal of Applied Chemistry 1978,

74(8), 1358-1363.

Page 39: Rheology and Microstructure of Cellulose Acetate - Digital Repository

24

8. Takahashi, M., Shimazaki, M. Formation of junction zones in thermoreversible

methylcellulose gels. Journal of Polymer Science: Part B: Polymer Physics,

2001, 39, 943-946.

9. Kobayashi, K., Huang, C., Lodge, T.P., Thermoreversible gelation of aqueous

methylcellulose solutions, Macromolecules 1999, 32, 7070-7077.

10. Li, L., Thangamathesvaran, P.M., Yue, C.Y., Tam, K.C., Hu, X., Lam, Y.C., Gel

network structure of methylcellulose in water. Langmuir 2001, 17, 8062-8068.

11. Sarkar, N. Kinetics of thermal gelation of methyl cellulose and hydroxypropyl

cellulose in aqueous solutions. Carbohydrate Polymers 1995, 26, 195-203.

12. Li, L., Thermal gelation of methylcellulose in water: scaling and

thermoreversibility. Macromolecules 2002, 35, 5990-5998.

13. Takahashi, M., Shimazaki, M., Yamamoto, J., Thermoreversible gelation and

phase separation in aqueous methyl cellulose solutions. Journal of Polymer

Science: Part B: Polymer Physics, 2001, 39, 91-100.

14. Desbrieres, J., Hirrien, M., Ross-Murphy, S.B., Thermogelation of methyl

cellulose: rheological considerations. Polymer 2000, 41, 2451-2461.

15. Hirsch, S.G.; Spontak, R.J. Temperature-dependent property development in

hydrogels derived from hydroxylpropylcellulose. Polymer 2002, 43, 123-129.

16. Sarkar, N. Thermal gelation properties of methyl and hydroxypropylmethyl

cellulose. Journal of Applied Polymer Science 1979, 24, 1073-1087.

17. Roschinski, C.; Kulicke, W-M. Rheological characterization of aqueous

hydroxylpropyl cellulose solutions close to phase separation. Macromolecular

Chemistry and Physics 2000, 201(15), 2031-2040.

Page 40: Rheology and Microstructure of Cellulose Acetate - Digital Repository

25

18. Satishkumar, S.; Chen, S.; Etzler, F. Rheological characterization of

hydroxylpropyl cellulose gels. Drug Development and Industrial Pharmacy

1999, 25(3), 153-161.

19. Kulicke, W-M.; Kull, A.H.; Thielking, H. Characterization of aqueous

carboxymethylcellulose solutions in terms of their molecular structure and its

influence on rheological behavior. Polymer 1996, 37(13), 2723-2731.

20. Ghannam, M.T.; Esmail, M.N. Rheological Properties of Carboxymethyl

Cellulose. Journal of Applied Polymer Science 1997, 64(2), 289-301.

21. Francis, P.S. Solution Properties of Water-soluble Polymers. I. Control of

Aggregation of Sodium Carboxymethylcellulose (CMC) by Choice of Solvent

and/or Electrolyte. Journal of Applied Polymer Science 1961, 5(15), 261-270.

22. Iiyina, E., Daragan, V., Self-diffusion of dimethyl sulfoxide and

dimethylformamide in solutions and gels of cellulose acetates by pulsed field

gradient NMR. Macromolecules 1994, 27, 3759-3763.

23. Guido, S., Grizzuti, N. Phase separation effects in the rheology of aqueous

solutions of hydroxylpropylcellulose. Rheol. Acta. 1995, 34, 137-146.

24. Fuji, S., Sasaki, N., Nakata, M., Rheological studies on the phase separation of

hydroxylpropylcellulose solution systems. Journal of Polymer Science: Part B:

Polymer Physics 2001, 39, 1976-1986.

25. Kjoniksen, A-L., Nystrom, B., Lindman, B. Dynamic viscoelasticity of gelling

and nongelling aqueous mixtures of ethyl(hydroxyethyl) cellulose and an ionic

surfactant. Macromolecules 1998, 31, 1852-1858.

Page 41: Rheology and Microstructure of Cellulose Acetate - Digital Repository

26

26. Vaessen, D.M.; McCormick, A.V.; Francis, L.F. Effect of phase separation on

stress development in polymeric coatings. Polymer 2002, 43, 2267-2277.

27. Reuvers, A.J., Altena, F.W., Smolders, C.A., Demixing and gelation behavior of

ternary cellulose acetate solutions. Journal of Polymer Science: Part B: Polymer

Physics 1986, 24, 793-804.

28. Matsuyama, H., Nishiguchi, M., Kitamura, Y., Phase separation mechanism

during membrane formation by dry-cast process. Journal of Applied Polymer

Science 2000, 77, 776-782.

29. Vaessen, D.M., McCormick, A.V., Francis, L.F., Effect of phase separation on

stress development in polymeric coatings. Polymer 2002, 43, 2267-2277.

30. Xu, Y., Wang, C., Tam, K.C., Li, L. Salt-assisted and salt-suppressed sol-gel

transition of nethylcellulose in water. Langmuir 2004, 20, 646-671.

31. Xu, Y., Lin, L., Zheng, P., Lam, Y.C., Hu, X. Contrallable gelation of

methylcellulose by a salt mixture. Langmuir 2004, 20, 6134-6138.

32. Schurz, J., Hedwig, T. Supermolecular gels in cellulose nitrate solutions.

Polymer 1966, 7(9), 475-477.

33. Vrancken, M.N., Ferry, J.D. Dilatometric measurements on gels of cellulose

tributyrate and cellulose nitrate. Journal of Polymer Science 1957, 24, 27-31.

34. Karatepe, A.U., Soylak, M., Elci, L. Separation/perconcentration of Cu(II),

Fe(III), Pb(II), Co(II), and Cr(III) in aqueous samples on cellulose nitrate

membrane filter and their determination by atomic absorption spectrometry.

Analytical Letters 2002, 35(9), 1561-1574.

Page 42: Rheology and Microstructure of Cellulose Acetate - Digital Repository

27

35. Granja, P.L., Pouysegu, L., Petraud, M., Jeso, B.D., Baquey, C., Barbosa, M.A.

Cellulose phosphates as biomaterials. I. Synthesis and characterization of highly

phosphorylated cellulose gels. Journal of Applied Polymer Science 2001, 82,

3341-3353.

36. Kawanishi, H., Tsunashima, Y., Horii, F., Dynamic light scattering study of

structural changes of cellulose diacetate in solution under couette flow.

Macromolecules 2000, 33, 2092-2097.

37. Tsunashima, Y., Kawanishi, H., Nomura, R., Horii, F., Influence of long-range

interactions in diffusion behavior in semidilute solution: Dynamics of cellulose

diacetate in quiescent state. Macromolecules 1999, 32, 5330-5336.

38. Kawanishi, H., Tsunashima, Y., Horii, F., A nest of structures in dynamics of

cellulose diacetate in N,N-dimetylacetamide in quiescent solution state studied

by dynamic light scattering Jour. of Chemical Physics 1998, 24, 11027-11031.

39. Tsunashima, Y., Kawanishi, H., Horii, F., Reorganization of Dynamics self-

assemblies of cellulose diacetate in solution: dynamical critical-like fluctuations

in the lower critical solution temperature system. Bioacromolecules 2002, 3,

1276-1285.

40. Zugenmaier, P. Characteristics of cellulose acetates. Macromol. Symp. 2004,

208, 81-166.

41. Steinmeier, H., H. Acetate manufacturing, process and technology. Macromol.

Symp. 2004, 208, 49-60.

Page 43: Rheology and Microstructure of Cellulose Acetate - Digital Repository

28

42. Heinze, T., Liebert, T. Chemical characteristics of cellulose acetate. Macromol.

Symp. 2004, 208, 167-237.

43. Gomez-Bujedo, S.; Fleury, E.; Vignon, M.R. Preparation of cellouronic acids

and partially acetylated cellournic acids by TEMPO/NaClO oxidation of water-

soluble cellulose acetate. Biomacromolecules 2004, 5, 565-571.

44. Bochek, A.M.; Kalyuzhnaya, L.M. Interaction of water with cellulose and

cellulose acetates as influenced by the hydrogen bond system and hydrophilic-

hydrophobic balance of the macromolecules. Macromolecular Chemistry and

Polymeric Materials 2002, 75, 989-993.

45. Boerstoel, H.; Maatman, H.; Picken, S.J.; Remmers, R.; Westerink, J.B. Liquid

crystalline solutions of cellulose acetate in phosphoric acid. Polymer 2001, 42,

7363-7369.

46. Pintaric, B.; Rogosic, M.; Mencer, H.J. Dilute solution properties of cellulose

diacetate in mixed solvents. Journal of molecular liquids 2000, 85, 331-350.

47. Matsuyama, H., Yamamoto, A., Yano, H., Maki, T., Teremoto, M., Mishima,

K., Matsuyama, K., Effect of organic solvents on membrane formation by phase

separation with supercritical CO2. Journal of Membrane Science 2002, 204, 81-

87.

48. Panina, N.I., Paneikina, T.P., Aver’yanova, V.M., Zelenev, Y.V., The

mechanical properties of gels in dimethyl phthalate of cellulose acetate of

different degress of substitution. Vysokomol. Soyed. 1978, A20, 1706-1711.

49. Carollo, P., Grospietro, B. Plastic materials. Macromol. Symp. 2004, 208, 335-

351.

Page 44: Rheology and Microstructure of Cellulose Acetate - Digital Repository

29

50. Pilon, R., Kunst, B., Sourirajan, S., Studies on the development of improved

reverse osmosis membranes from cellulose acetate-acetone-formamide casting

solution. Journal of Applied Polymer Science 1971, 15, 1317-1334.

51. Vogrin, N., Stropnik, C., Musil, V., Brumen, M., The wet phase separation: the

effect of cast solution thickness on the appearance of macrovoids in the

membrane forming ternary cellulose acetate/acetone/water system. Journal of

Membrane Science 2002, 207, 139-141.

52. Ryskina, I.I.; Aver’yanova, V.M. A study of gelation and the structure of

acetylcellulose gels in benzyl alcohol. Vysokomolekulyarnye Soedineniya 1971,

A13 (10), 2189-2194.

53. Ryskina, I.I.; Aver’yanova, V.M. Morphological study of the three dimensional

network of cellulose acetate-benzyl alcohol gelatins. Vysokomolekulyarnye

Soedineniya 1975, A17, 919-922.

54. Hao, J.N., Wang, S. Calculation of alcohol-acetone-cellulose acetate

ternaryphase diagram and their relevance to membrane formation, Journal of

Applied Polymer Science, 2001, 80, 1650-1657.

55. Tweddle, T.A., Sourirajan, S. Effect of Ethanol-Water Mixture as Gelation

Medium During Formation of Cellulose Acetate Reverse Osmosis Membranes.

Journal of Applied Polymer Science 1978, 22, 2265-2274.

56. Gildert, G.R.; Matsuura, T.; Sourirajan, S. Effect of different alcohol-water

mixtures as gelation mediums during formation o cellulose acetate reverse

osmosis membranes. Journal of Applied Polymer Science 2003, 24(1), 305-310.

Page 45: Rheology and Microstructure of Cellulose Acetate - Digital Repository

30

CHAPTER 3

EXPERIMENTAL METHODS

Abstract

This focus of this chapter is to provide a brief description of the experimental

techniques used in this thesis. A general background and the theoretical principles

behind the experimental techniques used to probe our system, which includes rheology,

scanning electron microscopy (SEM) and laser scanning confocal microscopy (LSCM)

are elaborated in this chapter. These characterization techniques provide us information

at both macroscopic and microscopic level on the solutions and gels formed in our

study.

Page 46: Rheology and Microstructure of Cellulose Acetate - Digital Repository

31

3.1 RHEOLOGICAL CHARACTERIZATION

Rheology is defined as the science involved in the study of flow and

deformation of materials.1 It is a powerful tool that can be used to characterize a wide

range of materials ranging from solutions, melts, gels, particulate systems among

others. It can also be used to provide information and characterize microstructure of

many materials.2 Rheological measurements can be performed either in steady or

dynamic modes.3,4 These two measurement techniques are discussed and for more

detailed information some useful text on rheology can be obtained from the following

sources.3, 5-7

Steady shear rheology

The steady shear measurements is performed by subjecting a sample to a steady

shear at a constant shear rate (γ•

) resulting in a generation of a shear stress (τ). The

corresponding shear stress (τ) on the sample is measured using a torque transducer. The

viscosity (η) is measured as function of the steady shear rate2(γ•

) and is defined as:

/η τ γ•

= Solutions of cellulose derivatives typically exhibit a zero shear viscosity, ηo, which is

typified by constant viscosity (Newtonian region) at low shear rates and a decrease in

viscosity (shear thinning) at high shear rates. Some polymeric systems have been shown

to exhibit a shear thickening behavior with increasing viscosity with shear rate increase.

The viscosity plots depicting different behaviors are shown in Figure 1.

Page 47: Rheology and Microstructure of Cellulose Acetate - Digital Repository

32

Dynamic rheology

This technique also called dynamic oscillatory flow involves the application of a

sinusoidally varying strain γ = γosin(ωt) in the linear viscoelastic (LVE) regime. The

symbol ω is the frequency of oscillations and γo is the strain amplitude.

The shear stress generated is also sinusoidal in nature given by the expression:

τ = τo sin(ωt+δ)

Expanding the above expression results in two components for the stress, one in phase

and one out of phase with the strain:

τ = τo cos(δ )sin(ωt) + τo sin(δ )cos(ωt)

The elastic or storage modulus, G′, is related to the stress component in phase with the

strain and is defined as:

G′ = τo cos(δ )/γo

While the viscous or loss modulus G″ is related to the stress component out of phase

with the strain and is given by:

G″ = τo sin(δ )/γo

G′ provides information on the energy stored by the sample, while G″ is related to the

energy dissipated by the sample. Thus, for a perfect elastic system, δ = 0 and G′

assumes a finite value while G″ = 0. Alternatively, for a purely viscous system, δ = 90o

with G″ having a finite value with G′ = 0. Most materials (e.g. polymer melts) exhibit

both elastic and viscous properties and are referred to as viscoelastic materials.

Therefore, for viscoelastic materials, both G′ and G″ will have non-zero values. Plots of

G′ and G″ as a function of frequency provides information regarding the structure of the

Page 48: Rheology and Microstructure of Cellulose Acetate - Digital Repository

33

material under scrutiny. A schematic representation of the frequency dependence of G′

and G″ for polymer solution (melts) and gels is shown in Figure 2. For a liquid sample

G′ and G″ have slopes close to 2 and 1 respectively at low frequencies. The point of

intersection of G′ and G″ (overlap frequency) provides information on the relaxation

time. For an elastic gel network, both moduli exhibit frequency independent behavior

with G′ significantly larger than G″. For dynamic frequency analysis to be valid, the

experiments have to be conducted in the LVE region. Dynamic strain sweep

experiments have to be performed to determine the LVE. This test involves variation of

strain amplitude while maintaining a constant frequency. The region corresponding to

the stress varying linearly with strain for the sample being investigated is termed the

linear viscoelastic (LVE) regime. One of main advantage of dynamic oscillatory

experiments is that it probes material without disrupting the microstructure. Dynamic

rheology has been used to characterize cellulose derivative systems providing important

information such as the sol-gel transition.9

3.2 CLOUD POINT STUDY

This technique employs the use of a laser set-up to investigate the on-set and

progress of phase separation in our system. Cloud point measurements are performed

using an Edmunds JDS uniphase helium-neon laser light source operating at a

maximum voltage output of 5mV. The procedure involves transferring sample into a

cuvette and placing it in the path of the laser source. The output reading is then recorded

on an Edmunds PD200 laser meter. A schematic diagram is shown in Figure 3. A

Page 49: Rheology and Microstructure of Cellulose Acetate - Digital Repository

34

similar helium-neon system was used to measure the cloud point in a study of

hydroxylpropylcellulose in mixed solvents of glycerol and water.10

3.3 SCANNING ELECTRON MICROSCOPY

Scanning electron microscopy (SEM) has the capability of allowing high depth-

of-field image, fracture surfaces and microstructural details of different samples.

Interested readers who want further review of an in-depth information on SEM can

consult the following text.11 The electron gun generates beam of electrons in a vacuum,

which is focused onto the sample by the objective lens.12 The beam is scanned over the

sample in series of lines and frames called raster by scan coils located in the

microscope. When the electron beam strikes the specimen, two types of electron

scattering termed elastic and inelastic scattering occur. This is due to interactions

between the primary beam electrons and atoms of the sample.

Elastic scattered electron is observed when the beam passes close to or impedes

with the nucleus of atom present in the material been investigated. Back scattered

electrons may result from elastic scattering and may generate secondary electrons which

causes noise in the final image.

Secondary electrons are generated during inelastic scattering when some beam

electrons collide with atoms. They typically have energies ranging 0 to many kev and

are used to generally generate three-dimensional images. An aluminum coating over a

scintillator material (Everhert-Thornley detector) is struck by the secondary electrons

generating a burst of light that strikes a photocade surface. The photoelectrons are

conducted to a photomultiplier where the number of electrons are increased. A small

Page 50: Rheology and Microstructure of Cellulose Acetate - Digital Repository

35

amount of voltage generated by the photomultiplier enters a preamplifier-amplifier unit,

which amplifies the weak signal to make detection possible. A schematic of a typical

SEM equipment set-up is shown in Figure 4. Numerous cellulose derivative gel

systems have been investigated using SEM.13-15

3.4 LASER SCANNING CONFOCAL MICROSCOPY

Laser scanning confocal microscopy (LSCM) is a technique that has been

employed for various studies including determination of cellulose fibril orientation16

and fibril angle17 in wood fibres. Characterization of some cellulose derived membranes

and gels have adopted LSCM as a tool to determine their morphologies.18-19 It has an

advantage of minimal sample preparation as compared to other techniques scanning and

transmission electron microscopy.18 The principle of LSCM is derived from

conventional light microscope but in this case a laser in place of a light source is used.

In addition, it has sensitive photomultiplier detectors and a computer set-up, which is

used control various aspects of this technique.20 It has the advantage of following

structural changes, which is an important factor in this project. LSCM has been

established as an important tool for obtaining high-resolution images and three-

dimensional images of variety of specimens. A schematic diagram of a LSCM is shown

in Figure 5. The operation involves an x-y deflection mechanism, where the laser beam

is turned into a scanning beam. It is then focused onto a small spot by an objective lens

on a fluorescent specimen. The objective lens is used to capture a mixture of reflected

and emitted fluorescent light. The diachronic mirror deviates the reflected light while

the emitted fluorescent light passes in the direction of the photomultiplier. In front of

Page 51: Rheology and Microstructure of Cellulose Acetate - Digital Repository

36

the photodetector is a confocal aperture, which serves to obstruct the so-called out-of-

focus information. This is especially important when dealing with thick specimens.

A 2-D image can be generated from a partial volume of sample centered at the

focal plane. In addition, a series of images can be obtained at successive planes by

alternating the focal plane in the z-direction. Using computer graphics, 3-D

reconstructed images can be obtained which enable users to visualize structural details

in a thin slice of a sample. Specimen to be imaged with a LSCM must be labeled with

a fluorescent probe. Major factors to be considered when selecting a fluorescence dye

are excitation/emission wavelengths of the probe, laser lines (wavelength of emission)

available and the filter sets used.4

Page 52: Rheology and Microstructure of Cellulose Acetate - Digital Repository

37

REFERENCES

1. Carreau, P.J.; De Kee, D.C.R.; Chhabra, R.P. Rheology of Polymeric Systems:

Principles and Applications. Hanser Gardner Publications, Inc. Cincinnati 1997.

2. Rohn, C.L. Analytical Polymer Rheology: Structure-Property Relationships.

Hanser Gardner Publications, Inc. Cincinnati 1995.

3. Macosko, C.W. Rheology Principles, Measurements and Applications. Wiley-

VCH Publishers, Inc. , New York 1994.

4. Pai, V.B. Tailoring Biopolymers Using Enzymes: Interactions in Xanthan-

Galactomannan Blends. PhD Thesis, Chemical Engineering Department, North

Carolina State University, Raleigh, NC 2001.

5. Barnes, H.A.; Hutton, J.F.; Walters, K. An Introduction to Rheology. Elsevier,

New York 1989.

6. Ferry, J.D. Viscoelastic properties of polymers. John Wiley 1980, New York.

7. Bird, R.B.; Armstrong, R.C.; Hassager, O. Dynamics of polymeric liquids.

Wiley 1987, New York.

8. Wang, C.S.; Fried, J.R. Viscoelastic properties of concentrated cellulose acetate

solutions. Journal of Rheology 1992, 36, 929-945.

9. Li, L. Thermal gelation of methylcellulose in water: scaling and

thermoreversibility. Macromolecules 2002, 35, 5990-5998.

10. Fuji, S.; Sasaki, N.; Nakata, M. Rheological studies on the phase separation of

hydroxylpropylcellulose solution systems. Journal of Polymer Science: Part B:

Polymer Physics 2001, 39, 1976-1986.

Page 53: Rheology and Microstructure of Cellulose Acetate - Digital Repository

38

11. Bozzola, J.J.; Russell, L.D. Electron Microscopy: Principles and Techniques for

Biologist (ebook). Sudbury, Mass & Jones & Barlett Publishers, Inc. 1999.

12. Makala, R.S. Pulsed Laser Deposition of Bi2Te3 Based Thermoelectric Thin

Films, Master Thesis, Material Science Department, North Carolina State

University, Raleigh, NC 2002.

13. Kesting, R.E.; Barsh, M.K.; Vincent, A.L. Semipermeable membranes of

cellulose acetate for desalination in the process of reverse osmosis II. Parameters

affecting membrane gel structure”, Journal of Applied Polymer Science 1965, 9,

1873-1893.

14. Pilon, R.; Kunst, B.; Sourirajan, S. Studies on the development of improved

reverse osmosis membranes from cellulose acetate-acetone-formamide casting

solution. Journal of Applied Polymer Science 1971, 15, 1317-1334.

15. Vogrin, N.; Stropnik, C.; Musil, V. Brumen, M., The wet phase separation: the

effect of cast solution thickness on the appearance of macrovoids in the

membrane forming ternary cellulose acetate/acetone/water system. Journal of

Membrane Science 2002, 207, 139-141.

16. Verbelen, J.-P.; Kerstens, S. Polarization Confocal Microscopy and Congo Red

Fluorescence: a simple and rapis method to determine the mean cellulose fibril

orientation in plants. Journal of Microscopy 2000, 198, 101-107.

17. Jang, H.F. Measurement of Fibril Angle in Wood Fibres with Polarization

Confocal Microscopy. Journal of Pulp and Paper Science 1998, 24(7), 224-230.

Page 54: Rheology and Microstructure of Cellulose Acetate - Digital Repository

39

18. Charcosset, C.; Cherfi, A.; Bernengo, J-C. Characterization of microporous

membrane morphology using confocal scanning laser microscopy. Chemical

Engineering Science 2000, 55, 5351-5358.

19. Hirsch, S.G.; Spontak, R.J. Temperature-dependent property development in

hydrogels derived from hydroxylpropylcellulose. Polymer 2002, 43, 123-129.

20. Paddock, S.W. An introduction to confocal imaging. Methods in Molecular

Biology 1999, 122, 1-9.

Page 55: Rheology and Microstructure of Cellulose Acetate - Digital Repository

40

Shear rate (1/s)10-1 100 101 102 103

Vis

cosi

ty (P

a.s)

10-2

10-1

100

Shear thickening

Newtonian

Shear thinning

Figure 1. Viscosity as a function of shear rate depicting different material

responses: Shear thickening, Newtonian, Shear thinning.

Page 56: Rheology and Microstructure of Cellulose Acetate - Digital Repository

41

Angular frequency (rad/s)10-3 10-2 10-1 100 101 102 103

G' &

G" (

Pa)

10-2

10-1

100

101

102

103

104

105

Elastic gel

Polymer solution (melt)

G'

G"

G'

G"

Figure 2. Elastic (G') and viscous (G") moduli as a function of angular frequency

showing a typical dynamic rheological response of polymer solutions and elastic gels

Page 57: Rheology and Microstructure of Cellulose Acetate - Digital Repository

42

Figure 3. Schematic diagram of the laser setup showing a laser source, cuvette and a

detector for recording the response.

Laser Source Sample PD200 laser meter

Page 58: Rheology and Microstructure of Cellulose Acetate - Digital Repository

43

Figure 4. Schematic representation of a typical scanning electron microscopy (SEM)

showing basic components of the apparatus.

Page 59: Rheology and Microstructure of Cellulose Acetate - Digital Repository

44

Figure 5. Schematic of a typical laser scanning confocal microscopy setup.

Page 60: Rheology and Microstructure of Cellulose Acetate - Digital Repository

45

CHAPTER 4

Effect of Water Composition on Rheological Properties in Ternary

Cellulose Acetate Solutions

Chapter 4 is essentially a manuscript by Collins Appaw, Richard D. Gilbert,

John F. Kadla and Saad A. Khan, prepared for submission to

Biomacromolecules.

Page 61: Rheology and Microstructure of Cellulose Acetate - Digital Repository

46

Effect of Water Composition on Rheological Properties in Ternary

Cellulose Acetate Solutions

Collins Appaw†, Richard D. Gilbert‡, and Saad A. Khan†,*

Departments of †Chemical and Biomolecular Engineering and ‡Wood & Paper Science, North Carolina State University, Raleigh, North Carolina 27695

John F. Kadla*

Advanced Biomaterials Chemistry, University of British Columbia,

Vancouver, BC V6T 1R9 Canada

Abstract

The effect of increasing water composition on the rheological and

microstructural behavior in a ternary system of cellulose acetate (CA), N,N-

dimethylacetamide (DMAc) and water is examined. In addition, cloud point studies are

conducted to determine the on-set of phase separation in this system. Increasing water

content resulted in enhanced steady shear viscosity and dynamic viscoelastic properties

at room temperature (25oC), which may be due to intensification of intermolecular

hydrogen bonds in solution. The system forms homogenous solutions at low water

content with formation of uniform turbid/cloudy system as water content is increased.

At relatively high water contents (typically >19wt %) for CA concentrations presented

in this paper, a sol-gel transition is observed with the formation of semi-solid system,

which exhibits gel-like characteristics. The sol-gel transition is primarily dependent on

water content and to some extent on CA concentration. Competitive hydrogen bond

interactions between CA and cosolvents in solution can be used to explain the sol-gel

transition.

* Corresponding authors: Saad A. Khan; Phone: 919 515-4519; Fax: 919-515-3465; e-mail

[email protected]. John F. Kadla; Phone: 604 827-5254; Fax: 604-822-9104; e-mail:

[email protected]

Page 62: Rheology and Microstructure of Cellulose Acetate - Digital Repository

47

Introduction

Cellulose and its derivatives are some of the most widely utilized natural

materials. Cellulose is a linear homopolymer consisting of β (1-4) linked

anhydroglucopyranose units (AGU).1 Extensive intra- and inter-molecular hydrogen-

bonding make cellulose insoluble in most common organic solvents.2,3 As a result,

cellulose is typically derivatized4 to facilitate dissolution and processing. Cellulose

acetates (CA) are cellulose esters5 that are partially substituted at the C-2, 3 and 6-

positions of the anhydroglucopyranose residue.4,6,7 A typical CA structure is depicted in

Figure 1 showing the replacement of the hydroxyl with acetyl groups. Other variations

of CA with have some but not all of the hydroxyl groups replaced and depending on the

acetyl groups on the chain is characterized by the term as degree of substitution (DS).

Their solubility in various solvents depends on the DS of the acetyl groups,6,8 wherein

the solubility is related to the interactions between the hydroxyl and acetyl groups

distributed throughout the CA molecules and with the solvent.6 Cellulose acetates with

a DS between 0.5 - 1 are soluble in aqueous solutions,9-10 while those with DS >1 tend

to be insoluble in aqueous medium but soluble in many organic solvent systems.

However, cellulose acetate molecules are never completely molecularly dispersed in

solution, rather existing as complex molecular associates; the extent of which depends

on the strength and amount of the intra- and inter-molecular interactions (e.g. hydrogen

bonding). For example, N,N-dimethylacetamide (DMAc), which readily dissolves CA

of DS ranging from 0.49 to 2.926 shows evidence of aggregates or associates of CA

molecules even in dilute solution; the result of long-range hydrogen bond interactions

between DMAc and CA.6,8

Page 63: Rheology and Microstructure of Cellulose Acetate - Digital Repository

48

Cellulose acetates are extensively used in filtration, membrane11-12 and encapsulation13

applications. In such systems sophisticated network structures are required. This is

typically achieved through aggregation induced phase separation.14-15 In polymeric

systems phase separation can be induced by solvent evaporation, temperature changes

or the addition of a non-solvent.16 In non-solvent induced phase separation, the

concentration of polymer, solvent and non-solvent are critical.17-18 Depending on the

system and component concentrations phase separation can lead to physical gel

formation.19 Gelation occurs as a result of non-solvent induced polymer aggregation and

the formation of large macromolecular associates and clusters.20-21 Non-solvent phase

separated gelation has been observed in several CA systems including dioxane/water22

and ethanol/water23 mixtures.

Cellulose acetates are typically used and processed in concentrated solutions.24

However, studies involving concentrated cellulose acetate solutions are relatively few,

due in part to the relatively high viscosities which make it difficult to control or

characterize properties in solution.25 Furthermore, it has been observed that basic

solvents interact primarily with the hydroxyl groups, while acidic solvents interact with

the acetyl groups.24 Therefore, depending on the solvent system CA will exhibit varying

solution properties, where the sol-gel transition and gel behavior of the resulting

material will be influenced by the interactions between the CA/solvent/non-solvent.

In this paper, we aim to characterize the sol-gel transition in a semi-concentrated

ternary system comprising CA, DMAc and water, specifically, the characteristics

leading to gelation by non-solvent (water) addition. Using steady-shear and dynamic

rheological measurements,26 the effect of non-solvent addition on viscosity and the sol-

Page 64: Rheology and Microstructure of Cellulose Acetate - Digital Repository

49

gel transition have been studied. Steady-state viscosity (η) and elastic modulus (G′)

measurements, both important factors in determining applicability of any polymer with

respect to its usage, reveal increasing water content results in phase separation and the

formation of a semi-solid system exhibiting gel-like characteristics. We present the

results observed for two CA concentrations and discuss the effect of water addition on

the intermolecular interactions present and the impact on polymer conformation and

hydrogen bond interactions that lead to aggregation and gel-induced structures.

Experimental

Materials

The cellulose acetate (CA) used in this work was purchased from Sigma-Aldrich

and it had a molecular weight of 50,000 g/mol with acetyl content of 39.7wt%. HPLC

grade N,N-dimethylacetamide (DMAc) was also purchased from Sigma-Aldrich and

together with deionized water were the cosolvents of interest. DMAc is a good solvent

for CA6 and dissolves CA forming homogenous solutions at appreciable high

concentration levels. Deionized water was utilized as the non-solvent in our system.

Sample Preparation

The initial step involved preparation of a bulk 20wt% CA concentration in

DMAc until a homogenous solution was achieved. The procedure adopted after that was

to dilute from 20wt% to specific concentrations by the addition of calculated weight

fractions of DMAc and deionized water. It must be noted that studies have been

performed on various concentrations besides those presented in this paper. Different

Page 65: Rheology and Microstructure of Cellulose Acetate - Digital Repository

50

solutions at fixed CA concentration but with varying weight fractions of water were

prepared using this procedure. The samples were mixed and kept for a day at ambient

conditions, after which samples were heated at 70oC for ten minutes to ensure complete

miscibility. Prior to heating, samples were kept under nitrogen to prevent oxidation.

Samples after heating, were kept at room temperature for at least a five days prior to use

to allow system to equilibrate. All stock solutions and dilutions were prepared on a

weight basis.

Rheological Measurements

A TA Advanced Rheometer (AR 2000) was used to measure the rheological

properties of the samples at ambient temperature (25oC). Steady state experiments

provided the viscosity of polymer samples, and is a tool that has been used in various

cellulose derivatives in solution.26-27 The configuration used, was either a cone and plate

or parallel plate geometry depending on solution viscosity. Shear rates ranging from 0.1

to 80s-1 were investigated in our system. In the dynamic experiments, a small oscillatory

shear was applied to the samples and the corresponding elastic (G′) and viscous (G″)

moduli measured as a function of frequency.26 Frequency ranges from 0.01 rad/s to 100

rad/s were conducted on samples. The shape and magnitude of the elastic and viscous

moduli as a function of frequency provides a signature of the state (e.g., solution, gel) of

a system.28 Dynamic stress sweep experiments were performed to determine the linear

viscoelastic (LVE) regime prior to the frequency sweep experiments. For the gels, we

carefully cut an appropriate circumference of samples and placed it on the lower

geometry of the rheometer. Parallel plate geometry was used to measure dynamic

viscoelastic properties of gels. The experiments were repeated with plates fitted with

Page 66: Rheology and Microstructure of Cellulose Acetate - Digital Repository

51

220 grit sandpaper to check for slippage and results showed no evidence of such an

occurrence. Experiments were repeated to ensure consistency and results were within an

error of 10%.

Cloud Point

Cloud point studies were performed using an Edmunds JDS uniphase helium-

neon laser light source operating at a maximum voltage output of 5mV. The procedure

involved passing a laser source through the sample and the subsequent recording of the

transmitted output intensity on an Edmunds PD200 laser meter. To undertake

measurements, samples were transferred to a cuvette having an optical path length of 1

cm. For samples having high viscosities including gels, samples were scooped using a

spatula and carefully transferred to the cuvette prior to measuring the intensity.

Scanning Electron Microscopy (SEM)

To investigate the microstructure of the gels formed, SEM (JEOL JSM-5900LV)

was used with accelerating voltage of 20kV and a magnification of 5000x. The protocol

was to cut small pieces of gel placed them in 3% glutaraldehyde Na acetate buffer, at

pH 6.6 (4oC). The gels were rinsed with 3 changes of 0.1M sodium acetate buffer, at pH

6.6 and then dehydrated in a graded ethanol series of 30%, 50%, 70%, 95% and 100%

ethanol, all at 4oC. Samples were then dried using critical point drying with CO2. The

dried samples were gently fractured sputter coated with Au/Pd and prior to scanning.

Page 67: Rheology and Microstructure of Cellulose Acetate - Digital Repository

52

Results and Discussion

Solution Rheology

Our initial step involved examining the solution property changes induced by

increasing water content to CA/DMAc solutions. Addition of water to bulk CA/DMAc

solutions produced a system, which shows increasing viscosity with water content. This

is illustrated in Figure 1a, we plots the viscosity profiles for a 10 wt.% CA solution for

different water content. Figure 1b shows the viscosity trend for a sample containing a

higher CA concentration from (15 wt%) and depicts a similar profile of increasing

viscosity with water content. For both samples, we observe that at low or no water

content, the viscosity exhibits a large Newtonian plateau, followed by a power-law

regime. Above a certain water content the viscosity curves change exhibiting high

viscosity at low shear rates and the gradual disappearance of the zero-shear viscosity

plateau. The disappearance of the Newtonian plateau at low shear rates suggests

development of microstructure in our sample as water content is increased. In Figure 2,

an attempt is made to highlight the progressive increase in low-shear viscosity that

occurs when water content is increased. We clearly observe that for any value of water

content the viscosity increases by almost an order of magnitude when CA concentration

increase from 10 to 15wt%. We also find that the viscosity enhancement with increasing

water content shows the same trend for both CA concentration samples. In fact, an

exponential increase in viscosity with water content is noticed.

Cloud point studies for transmission intensity as a function of water content

were conducted using a helium-neon laser light source. . A similar helium-neon system

Page 68: Rheology and Microstructure of Cellulose Acetate - Digital Repository

53

for measuring the cloud point was used to study hydroxylpropylcellulose in mixed

solvents of glycerol and water.29 Figure 3 shows that transmission remains constant

until a critical water content, whereupon a sharp drop in intensity is observed for both

10 and 15wt% CA samples. For both curves, we observe the transition corresponding to

phase separation occurring at relatively high water content. However, the cloud point

seems to shift to lower water content when the sample has a higher CA concentration of

15wt%. Thus, an increase in CA concentration cause a quicker phase separating system

since polymer-polymer and polymer-cosolvent interactions are more pronounced.

It is interesting to note that even in the water content range in which

transmission remains constant, the viscosity shows an increase with addition of water.

The increase in viscosity may be due to further hydrogen bonds being formed between

CA-DMAc-water due to enhanced intermolecular interactions in solution precipitated

by increasing water content. A detailed hypothesis for the viscosity pattern is that CA

and DMAc form hydrogen bond links when CA dissolves to form a neat clean solution.

This is may be due to primarily hydroxyl (OH) units on CA interacting with the acetyl

groups on DMAc. Addition of water causes the formation of new hydrogen bonds with

DMAc and to a much lesser extent CA, thereby intensifying the hydrogen bond

interactions in solution. This phenomenon may account for the viscosity increase

observed in our system. In addition CA concentration increment may favor additional

intermolecular interactions in solution and may account for viscosity increases

observed. Previous studies have suggested that in a CA/acetone system, strong

interactions between solvent and polymer may involve hydrogen bonding between OH

groups on cellulose acetate and carbonyl unit in acetone.30

Page 69: Rheology and Microstructure of Cellulose Acetate - Digital Repository

54

Gelation Process

Figure 5 shows a digital image of a 10wt.% CA system as a function of

increasing water content. A gradual transition from a clear polymer solution to a cloudy

system with increasing water content is observed. At low water content (up to 7.2wt%),

the system forms a clear homogenous solution but a further increase to 16.2wt% water

results in the emergence of cloudiness in the system reminiscent of a phase separating

polymer. solution. Water content increment to 21.6wt% leads to a uniform cloudy, self-

supporting gel-like material. To quantify the sol-gel behavior, we plot in Figure 6a the

elastic (G′) and viscous (G″) moduli as a function of frequency for two of the water

compositions shown in Figure 4: 7.2 and 21.6wt.%. At the low water composition of

7.2wt%, we find G″ to be larger than G′ at low frequencies with both showing strong

frequency dependence reminiscent of a polymer solution. At the highest water content

of 21.6wt.%, there is transition with both moduli increasing substantially by several

orders of magnitude. In addition, G′ is larger than G″ over the entire frequency range

studied and both are relatively independent of frequencies, features characteristic of an

three dimensional elastic gel.31-36 Figure 6b shows the dynamic frequency spectrum of a

15wt% CA sample at two different water content. At the lower water content of

6.8wt.%, the sample exhibits solution characteristics whereas at a water content of 20.4

wt.%, we observe behavior characteristic of a gel network. We also observe that the

magnitude of the moduli, are higher for this higher CA concentration sample.

Therefore, as CA concentrations increases, stronger gels are formed due to greater

intermolecular interactions between CA and cosolvents.

Page 70: Rheology and Microstructure of Cellulose Acetate - Digital Repository

55

To pinpoint the sol-gel transition as a function of water content, we plot a

multifrequency plot of tan δ as a function of water content for two samples containing

10 and 15wt% CA (Figure 7). Determination of gel-point using this method was

developed by Winter and Chambon and is appropriately termed as the Winter-Chambon

criterion. This procedure has been applied in a variety of polymeric systems37-41 and an

application of this method is extended in our work to measure the water content where

the sol-gel transition occurs. The point of intersection of the various curves at different

frequencies is indicative of the gel-point.40-41 Data for Figure 7 were obtained from plots

similar to that in Figure 6 (dynamic frequency plots) performed at different water

contents. For both samples we find the tan δ curves at different frequencies to converge

for a specific water content. However, unlike conventional gelation in which these

curves diverge after intersecting, our system shows the curves to remain together and

independent of frequency following the convergence. Assuming the initial point of

convergence to be the gel point, we find gelation to occur at a lower content (20 versus

23 wt.%) for the more concentrated CA system. Thus, even though water content

increment is the key to induce sol-gel properties in our system, polymer concentration

also plays a role, albeit minor, in influencing gelation.

We examine in Figure 8 the systematic variation of G′ as a function of water

content. We find that G′ increases by over 4 orders of magnitude as water content

increases 0 to 25 wt% for both the 10 and 15wt% CA containing samples. Interestingly,

curves for both CA concentrations show the same sigmoidal shape with the one

containing 15wt% CA exhibiting a higher value, inline with earlier observation of

viscosity (Figure 3). As such, we observe two distinct regimes in terms of G′

Page 71: Rheology and Microstructure of Cellulose Acetate - Digital Repository

56

enhancement. As the water content increases, G′ initially increases slowly followed by a

regime of sharp increase with formation of gel. Following gelation, G’ increases slowly

Figure 9 depicts SEM micrographs of CA gels obtained under different water

contents and CA concentrations. Increasing water content from 21.6 to 25.2wt% for CA

concentration of 10wt% results in particulate structure with smaller voids in comparison

to lower water content gels. The more uniform and homogenous microstructure having

smaller voids may account for the stronger gel properties observed when water content

is increased for CA concentration of 10wt%. When CA concentration is increased to

15wt%, we observe a similar structure, which are closely packed together with the voids

shrinking in size with water content. The more tightly packed voids for water content of

23.8wt% provide us with a more rigid and compact gel network and may be due

additional bond links formed due to water content increase. We observe that for both

concentrations when water content is increase, the system portray a more continuous

which have less holes in them. This may also account for the stronger gel properties

when we increase water content. Thus, for both 10 and 15wt%, we observe dynamic

visoelastic properties enhanced when we increase water composition at constant CA

concentration.

Schematic representations of the molecular level interactions, which may be

responsible for the sol-gel process are portrayed in Figure 10. The hypothesis for this

process may be due to the fact that, CA is dispersed in DMAc (black dots) initially

forming hydrogen bond linkages with each other. Initial addition of small amounts of

water (white dots) to CA and DMAc solution causes some DMAc to interact with

water. But a greater amount of DMAc may still be interacting to a greater extent with

Page 72: Rheology and Microstructure of Cellulose Acetate - Digital Repository

57

CA thereby rendering solution homogenous. With water content increment, a

competition for DMAc between CA and water occurs. Thus, as we increase water

content, CA chains resist the displacement from DMAc, with water also holding on

firmly to as much DMAc available for bonding. In addition some water may be weakly

bonded to “free” CA in the system. This segregation effect caused by the interactions in

the system lead to CA experiencing both intra- and intermolecular interactions causing

CA to disperse through entire volume of sample. These interactions may be may

balanced and at equilibrium this leads to the formation of a material exhibiting gel-like

properties at critical water contents. Therefore, our assumption is that, this complex

interaction process causes a complete dispersion of CA and lead to aggregates of CA

uniformly spread throughout the entire system.

Experiments conducted on some cellulose derivatives give indications that

intermolecular interactions may be one of the important mechanisms for gel

formation.42 An earlier study of CA behavior in solvents that selectively interact with

specific functional groups on the CA chain may provide more insight to the sol-gel

transition process present in our system.43 Its our belief that similar selective

interactions between CA and cosolvents may account for the gelation process. Our

theory for the transition is deduced base on the fact that, the acetyl groups on DMAc

preferentially interact with OH groups on the CA molecule. The addition of a non-

solvent in this case water to CA/DMAc solutions, weakens this interactive bond and

shifts some DMAc molecules to interact rather with the OH groups on water. The bond

breaking process gets progressively worse with water content increase and lead to

intramolecular interactions between unbonded or “free” CA chains some of which have

Page 73: Rheology and Microstructure of Cellulose Acetate - Digital Repository

58

been displaced from DMAc. This process is also accompanied by intermolecular

interactions between CA units, DMAc and water. It is postulated that at equilibrium the

interactions mentioned lead to uniform aggregation of CA in the entire volume and

invariably gelation occur. Thus, in this system CA may be experiencing intermolecular

interaction as well as intramolecular interactions as we increase the amount of water in

solution. We have thus been able to formulate a protocol to convert CA in a mixed

solvent system from a homogenous solution to a material with “gel-like” properties

using water as a non-solvent.

Conclusions

In this study, we examined the effect of adding water to CA/DMAc solutions

using steady and dynamic rheology. Addition of small quantities of water led to

formation of weakly entangled system resulting in enhanced steady shear viscosity at

room temperature. Beyond critical water contents typically >19wt% dependent on CA

concentration, the system transforms and forms gel network structure. Prior to gelling,

the sample phase separated depicting a cloudy/turbid solution. The sol-gel transition

depends to high extent on the water content and the gelation process is found to occur at

lower water content when CA concentration is increased. The dynamic viscoelastic

properties were also more pronounced when CA concentration is increased. The

hypothesis for the gelation process is believed to occur due to when water and CA

competing for DMAc in the system. Due to initial formation of CA-DMAc bonds

formed, CA resists this bond breaking process. The strong attraction of water for

Page 74: Rheology and Microstructure of Cellulose Acetate - Digital Repository

59

DMAc, leads to “free” unbonded CA units in the system leading to gelation. Thus, we

have been able to tailor a process for converting solutions of ternary systems

comprising of CA/DMAc/water to gels. We hope our study will lead to better

understanding the interactions present in CA mixed solvent systems paving the way for

such systems to be adopted and utilized pertaining to its applications in industry.

Acknowledgement

The authors gratefully acknowledge funding support for this work from United

States Department of Agriculture (USDA).

Page 75: Rheology and Microstructure of Cellulose Acetate - Digital Repository

60

References

(1) Kadla, J.F., Gilbert, R.D., Cellulose structure: a review. Cellulose Chemistry

and Technology 2000, 34, 197-216.

(2) Saalwachter, K., Burchard, W., Cellulose in new metal-complexing solvents. 2.

Semidilute behavior in Cd-tren. Macromolecules 2001, 34, 5587-5598.

(3) Hattori, K., Cuculo, J.A., Hudson, S.M., New solvents for cellulose:

hydrazine/thiocyanate salt system. Journal of Polymer Science: Part A: Polymer

Chemistry 2002, 40, 601-611.

(4) Tsunashima, Y., Hattori, K., Substituent distribution in cellulose acetate: Its

control and the effect on structure formation in solution. Jour. of Colloid and

Interface Sci., 2000, 228, 279-286.

(5) Gross, R.A., Scholz, C. Biopolymers from Polysaccharides and Agroproteins.

Oxford University Press, 2001.

(6) Kawanishi, H., Tsunashima, Y., Okada, S., Horii, F. Change in chain stiffness in

viscometric and ultracentrifugal fields: cellulose diacetate in N,N-

dimethylacetamide dilute solution. Journal of Chemical Physics 1998, 108 (14),

6014-6025.

(7) Tsunashima, Y., Kawanishi, H., Horii, F. Reorganization of dynamic self-

assemblies of cellulose diacetate in solution: Dynamical critical-like fluctuations

in mower critical solution temperature system. Biomacromolecules 2002, 3,

1276-1285.

Page 76: Rheology and Microstructure of Cellulose Acetate - Digital Repository

61

(8) Kawanishi, H., Tsunashima, Y., Horii, F., Dynamic light scattering study of

structural changes of cellulose diacetate in solution under couette flow.

Macromolecules 2000, 33, 2092-2097.

(9) Gomez-Bujedo, S., Fleury, E., Vignon, M.R., Preparation of cellouronic acids

and partially acetylated cellournic acids by TEMPO/NaClO oxidation of water-

soluble cellulose acetate, Biomacromolecules 2004, 5, 565-571.

(10) Bochek, A.M., Kalyuzhnaya, L.M., Interaction of water with cellulose and

cellulose acetates as influenced by the hydrogen bond system and hydrophilic-

hydrophobic balance of the macromolecules. Macromolecular Chemistry and

Polymeric Materials 2002, 75, 989-993.

(11) Kesting, R.E., Barsh, M.K., Vincent, A.L., Semipermeable membranes of

cellulose acetate for desalination in the process of reverse osmosis II. Parameters

affecting membrane gel structure”, Journal of Applied Polymer Science Jour. of

Applied Polymer Science 1965, 9, 1873-1893.

(12) Pilon, R., Kunst, B., Sourirajan, S., Studies on the development of improved

reverse osmosis membranes from cellulose acetate-acetone-formamide casting

solution. Journal of Applied Polymer Science 1971, 15, 1317-1334.

(13) Khalil, S.A., Phase separation of cellulose derivatives: effects of polymer

viscosity and dielectric constant of nonsolvent. Jour. of pharmaceutical Sciences

1973, 62, 1883-1884.

(14) Vogrin, N., Stropnik, C., Musil, V., Brumen, M., The wet phase separation: the

effect of cast solution thickness on the appearance of macrovoids in the

Page 77: Rheology and Microstructure of Cellulose Acetate - Digital Repository

62

membrane forming ternary cellulose acetate/acetone/water system. Journal of

Membrane Science 2002, 207, 139-141.

(15) Kunst, B., Sourirajan, S., Evaporation rate and equilibrium phase separation data

in relation to casting conditions and performance of porous cellulose acetate

reverse osmosis membranes. Journal of Applied Polymer Science 1970, 14,

1983-1996

(16) Matsuyama, H., Nishiguchi, M., Kitamura, Y., Phase separation mechanism

during membrane formation by dry-cast process. Journal of Applied Polymer

Science 2000, 77, 776-782.

(17) Vaessen, D.M., McCormick, A.V., Francis, L.F., Effect of phase separation on

stress development in polymeric coatings. Polymer 2002, 43, 2267-2277.

(18) Reuvers, A.J., Altena, F.W., Smolders, C.A., Demixing and gelation behavior of

ternary cellulose acetate solutions Jour. of Polymer Sci.: Part B: Polymer

Physics 1986, 24, 793-804.

(19) Hao, J.H., Wang, S., Calculation of alcohol-acetone-cellulose acetate ternary

phase diagram and their relevance to membrane formation. Journal of Applied

Polymer Science 2001, 80, 1650-1657.

(20) Iiyina, E., Daragan, V., Self-diffusion of dimethyl sulfoxide and

dimethylformamide in solutions and gels of cellulose acetates by pulsed field

gradient NMR. Macromolecules 1994, 27, 3759-3763.

(21) Alie, C., Pirard, R., Pirard, J-P., Preparation of low-density xerogels from

mixtures of TEOS with substituted alkoxysilanes. II. Viscosity study of the sol-

gel transition. Journal of non-crystalline solids 2003, 320, 31-39.

Page 78: Rheology and Microstructure of Cellulose Acetate - Digital Repository

63

(22) Altena, F.W., Smolders, C.A., Phase separation phenomena in solutions of

cellulose acetate. I. Differential scanning calorimetry of cellulose acetate in

mixtures of dioxane and water Jour. of Polymer Science: Polymer Symposium

1981, 69, 1-10.

(23) Tweddle, T.A., Sourirajan, S., Effect of Ethanol-Water Mixture as Gelation

Medium During Formation of Cellulose Acetate Reverse Osmosis Membranes.

Journal of Applied Polymer Science 1978, 22, 2265-2274.

(24) Wee, W.-K., Mackley, M.R., The rheology and processing of a concentrated

cellulose acetate solution. Chemical Engineering Science 1998, 53, 1131-1144

(25) Pintraric, B., Rogosic, M., Mencer, H.J., Dilute solution properties of cellulose

diacetate in mixed solvents. Journal of Molecular Liquids 2000, 85, 331-350.

(26) Lizaso, E., Munoz, M.E., Santamaria, A., “Formation of Gels in Ethylcellulose

Solutions.An Interpretation from Dynamic Viscoelastic Results”,

Macromolecules 1999, 32, 1883-1889.

(27) Maestro, A., Gonzalez, C., Gutierrez, J.M., Shear thinning and thixotropy of

HMHEC and HEC water solutions. Journal of rheology 2002, 46,1445-1457.

(28) Eissa, A.S., Bisram, S., Khan, S.A., Polymerization and gelation of whey protein

isolates at low pH using transglutaminase enzyme. Journal of Agricultural and

Food Chemistry 2004, 52, 4456-4464.

(29) Fuji, S., Sasaki, N., Nakata, M., Rheological studies on the phase separation of

hydroxylpropylcellulose solution systems. Journal of Polymer Science: Part B:

Polymer Physics 2001, 39, 1976-1986.

Page 79: Rheology and Microstructure of Cellulose Acetate - Digital Repository

64

(30) Griswold, P.D., Cuculo, J.A., An experimental study of the relationship between

rheological properties and spinnability in the dry spinning of cellulose acetate-

acetone solutions. Journal of Applied Polymer Science 1974, 18, 2887-2902.

(31) Pai, V., Srinivasarao, M., Khan, S. A. Evolution of Microstructure and Rheology in Mixed

Polysaccharide Systems. Macromolecules 2002, 35, 1699-1707.

(32) Wilder, E. A., Hall, C. K., Khan, S. A., Spontak, R. J. Effects of Composition and Matrix

Polarity on Network Development in Organogels of Poly(ethylene glycol) and

Dibenzylidene Sorbitol. Langmuir 2003, 19(15), 6004-6013.

(33) Raghavan, S. R., Walls, H. J., Khan, S. A. Rheology of Silica Dispersions in Organic

Liquids: New Evidence for Solvation Forces Dictated by Hydrogen Bonding. Langmuir

2000, 16(21), 7920-7930.

(34) Chauvelon, G., Doublier, J-L. Buleon, A., Thibault, J-F., Saulnier, L.,

Rheological properties of sulfoacetate derivatives of cellulose. Carbohydrate

Research 2003, 338, 751-759.

(35) Sakar, N., Kinetics of thermal gelation of methylcellulose and hydroxypropyl

methylcellulose in aqueous solutions, Carbohydrate Polymers 1995, 26, 195-

203.

(36) Lazaridou, A., Biliaderis, C.G., Izydorczyk, M.S., Molecular size effects on

rheological properties of oat β-glucans in solution and gels. Food Hydrocolloids

2003, 17, 693-712.

(37) Dumitras, M., Friedrich, C. Network formation and elasticicty evolution on

dibenzylidene sorbitol/poly(propylene oxide) physical gels. Journal of Rheology

2004, 48(5), 1135-1146.

Page 80: Rheology and Microstructure of Cellulose Acetate - Digital Repository

65

(38) Cossar, S., Nichetti, D., Grizzuti, N. A rheological study of the phase transition

in thermoplastic polyurethanes. Critical gel behavior and microstructure

development. Journal of Rheology 2004, 48(3), 691-703.

(39) Power, D.J., Rodd, A.B., Paterson, L., Boger, D.V. Gel transition studies on

nonideal polymer networks using small amplitude oscillatory rheometry.

Journal of Rheology 2004, 42(5), 1021-1036.

(40) Fahrlander, M., Fuchs, K., Mulhaupt, R., Friedrich, C. Linear and nonlinear

rheological properties of self-assembling tectons in polypropylene matrices.

Macromolecules 2003, 36, 3749-3457.

(41) Chiou, B-S., Raghavan, S.R., Khan, S.A. Effect of colloidal fillers on the cross-

linking of a uv-curable polymer: gel point rheology and the Winter-Chambon

criterion. Macromolecules 2001, 34, 4526-4533.

(42) Nystrom, B., Walderhaug, H., Hansen, F.K., Rheological behavior during

thermoreversible gelation of aqueous mixtures of ethyl(hydroxyethyl) cellulose

and surfactants. Langmuir 1995, 11, 750-757.

(43) Timofeyeva, G.N., Aver”yanova, V.M., Influence of the nature of the solvent on

the processes of structuring in solutions of cellulose acetates. Vysokomol. Soyed.

1982, A24 (11), 2189-2194.

Page 81: Rheology and Microstructure of Cellulose Acetate - Digital Repository

66

Figure Captions

Figure 1. Chemical structure of a typical cellulose acetate chain with the hydroxyl

groups replaced with acetyl groups

Figure 2. Effect of different water contents on steady shear viscosity for a

CA/DMAc/water system. CA concentrations are (a) 10wt% and (b) 15wt%.

Figure 3. Viscosity as a function of water content obtained at shear rate of 1s-1 for CA

concentrations of 10 and 15wt%.

Figure 4. Transmission intensity versus amount of water for 10 and 15wt% CA

concentration in DMAc/water system.

Figure 5. Transition of 10wt% CA/DMAc/water system from homogenous system to a

cloudy system. Water content increment is showed from left to right.

Figure 6. Effect of low and high water contents on the viscous (G″) and elastic (G′)

moduli for CA/DMAc/water system. CA concentrations are (a) 10wt% and (b) 15wt%.

Figure 7. Multifrequency plot of tan δ as function of water content for (a) 10wt% and

(b) 15wt% CA/DMAc/water system.

Figure 8. G' versus water content at a frequency of 1 rad/sec for 10wt% and 15wt%

CA/DMAc/water system.

Figure 9. SEM micrographs for CA concentration of 10 and 15wt% at two different

water contents.

Figure 10. An illustrative representation of molecular level interactions present in the

Page 82: Rheology and Microstructure of Cellulose Acetate - Digital Repository

67

system. At low water content, the solvents are dispersed forming solutions. At high

water content, CA form strong and weak links with solvents in the system and leads to

gel formation. Dash lines and solid lines represent weak and strong hydrogen bonds

respectively.

Page 83: Rheology and Microstructure of Cellulose Acetate - Digital Repository

68

Figure 1. Chemical structure of a typical cellulose acetate chain with the hydroxyl

groups replaced with acetyl groups.

Page 84: Rheology and Microstructure of Cellulose Acetate - Digital Repository

69

10-2 10-1 100 101 102

Vis

cosi

ty (P

a.s)

100

101

0.0 wt%

3.6 wt%

7.2 wt%

10.8 wt%

14.4 wt%

Shear rate (s-1)

10-2 10-1 100 101 102

Vis

cosi

ty (P

a.s)

101

102

0.0 wt%

3.4 wt%

6.8 wt%

10.2 wt%

13.6 wt%

Figure 2. Effect of different water contents on steady shear viscosity for a

cellulose acetate/DMAc/water system. CA concentrations are (a) 10wt%

and (b) 15wt%.

Page 85: Rheology and Microstructure of Cellulose Acetate - Digital Repository

70

Water content (wt%)0 2 4 6 8 10 12 14 16

Vis

cosi

ty (P

a.s)

100

101

102

15%

12.5%

10%

Figure 3. Viscosity as a function of water content obtained at shear rate of 1s-1 for

CA concentrations of 10, 12.5 and 15wt.%.

Page 86: Rheology and Microstructure of Cellulose Acetate - Digital Repository

71

Water content (wt.%)

0 5 10 15 20 25 30

Tran

s. In

tens

ity (a

b. u

nits

)

10-5

10-4

10-3

10-2

10-1

100

101

10%15%

Figure 4. Transmission intensity versus amount of water for 10 and 15wt.% CA

concentration in DMAc/water system.

Page 87: Rheology and Microstructure of Cellulose Acetate - Digital Repository

72

Figure 5. Transition of 10wt% CA/DMAc/water system from homogenous

system to a cloudy system. Water content increment is showed from left to right.

0.0wt% 7.2wt% 16.2wt% 21.6wt%

Page 88: Rheology and Microstructure of Cellulose Acetate - Digital Repository

73

10-3 10-2 10-1 100 101 102 103

G' (

Pa)

10-3

10-2

10-1

100

101

102

103

104

105

G' (21.6%)

G" (21.6%)

G" (7.2%)

G' (7.2%)

Frequency (rad/s)

10-3 10-2 10-1 100 101 102

G' &

G" (

Pa)

10-2

10-1

100

101

102

103

104

105

106

G' (20.4%)

G" (20.4%)

G" (6.8%)

G' (6.8%)

Figure 6. Effect of different water contents on the viscous (G″) and elastic

(G′) moduli for CA/DMAc/water system. CA concentrations are (a)

10wt% and (b) 15wt%.

Page 89: Rheology and Microstructure of Cellulose Acetate - Digital Repository

74

0 5 10 15 20 25

tan

d

10-1

100

101

102

103

0.631 rad/s1 rad/s1.58 rad/s2.5 rad/s3.9 rad/s6.3 rad/s10 rad/s

Water content (wt%)0 5 10 15 20 25 30

tan

d

10-1

100

101

102

0.631 rad/s1 rad/s1.58 rad/s2.5 rad/s3.9 rad/s6.3 rad/s10 rad/s

Figure 7. Multifrequency plot of tan δ as function of water content for (a) 10wt%

and (b) 15wt% CA/DMAc/water system

Page 90: Rheology and Microstructure of Cellulose Acetate - Digital Repository

75

Water content (wt%)

0 5 10 15 20 25 30

G' (

Pa)

10-2

10-1

100

101

102

103

104

105

106

15wt%

10wt%

Figure 8. G' versus water content at a frequency of 1 rad/sec for 10wt%

and 15wt% CA/DMAc/water system.

Page 91: Rheology and Microstructure of Cellulose Acetate - Digital Repository

76

(a) 21.6% 25.2%

(b) 20.4% 23.8% Figure 9. SEM micrographs for CA concentration of (a) 10 and (b) 15wt% at two

different water contents.

Page 92: Rheology and Microstructure of Cellulose Acetate - Digital Repository

77

Low water content

High water content Figure 10. An illustrative representation of molecular level interactions present in the

System. At low water content, the solvents are dispersed forming solutions. At high

water content, CA form strong and weak links with solvents in the system and leads to

gel formation. White and black circles depict water and N,N dimethylacetamide

respectively. Dash lines and solid lines represent weak and strong hydrogen bonds

respectively.

Page 93: Rheology and Microstructure of Cellulose Acetate - Digital Repository

78

CHAPTER 5

Phase Separation and Heat-Induced Gelation Characteristics of

Cellulose Acetate in a Mixed Solvent System

Chapter 5 is essentially a manuscript by Collins Appaw, Richard D. Gilbert,

John F. Kadla and Saad A. Khan, prepared for submission to Macromolecules.

Page 94: Rheology and Microstructure of Cellulose Acetate - Digital Repository

79

Phase Separation and Heat-Induced Gelation Characteristics of

Cellulose Acetate in a Mixed Solvent System

Collins Appaw†, Richard D. Gilbert‡, and Saad A. Khan†,*

Departments of †Chemical and Biomolecular Engineering and ‡Wood & Paper Science, North Carolina State University, Raleigh, North Carolina 27695

John F. Kadla,*,§

Advanced Biomaterials Chemistry, University of British Columbia,

Vancouver, BC V6T 1R9 Canada

Abstract

The rheological behavior of cellulose acetate (CA) in mixed solvents comprising

of N,N dimethylacetamide (DMAc) and water is examined at different CA

concentrations and water content. Addition of a constant weight of CA to different

ratios but constant weight of DMAc/water solutions results in enhancement of

viscoelastic properties as the ratio of water is increased at room temperature (25oC). The

system displays a unique behavior exhibiting a two-phase polymer-rich viscous mixture

at the bottom of an almost clear liquid layer at relatively high water content (>19wt%)

based upon the CA concentrations studied. Whiles the viscous bottom exhibit a weak

“gel-like” property, the top liquid layer exhibits an almost Newtonian behavior. Heat

treatment results in gelation for the two-phase system. Cloud point studies also showed

the onset of phase separation characteristics as water content is increased. The phase

separation and gelation characteristics can be explained based on hydrogen bond

interactions being intensified between CA and cosolvents in solution.

*Corresponding authors: Saad A. Khan; Phone: 919 515-4519; Fax: 919-515-3465; e-mail [email protected]. John F. Kadla; Phone: 604 827-5254; Fax: 604-822-9104; e-mail: [email protected]

Page 95: Rheology and Microstructure of Cellulose Acetate - Digital Repository

80

Introduction

Cellulose a renewable biodegradable natural polymer1-3 is highly insoluble in a

wide range of solvents.4 The lack of solubility is attributed to the intra- and

intermolecular hydrogen bond interactions between the hydroxyl units on the chains.4-6

Due to this inherent problem, various derivatives of cellulose have been prepared by

substituting the hydroxyl groups with different functional groups.7 This invariably

impact different solubility characteristics to cellulose derivatives. Cellulose derivatives

such as cellulose ethers an example being methyl cellulose (MC) are soluble in aqueous

medium.8-15 It is presumed that, at low temperatures MC is water-soluble due to water

molecules presumably forming “cage-like” structures around the hydrophobic methoxyl

groups.8 It is important to note that depending on degree of substitution (DS) and

preparation methods, a variety of other solvents besides water can dissolve MC.15

Cellulose acetate (CA) is a semiflexble polymer16 that has been used in the

manufacture of textiles, moldings, filters for cigarettes, fibres, selective membrane,

bioactive and biocompatible materials, etc. One method for preparing CA is through

acetylation of cellulose by substituting the hydroxyl groups on the cellulose chain.17-19

This process yields CA with different solubility characteristics dependent on the degree

of substitution of the acetyl groups. The process of direct acetylation to manufacture

cellulose acetates (CAs), yield partially water-soluble CAs at same low DS.17 This is

due to the fact that, nonuniform distribution of the substituents are formed as a result

which is a factor inhibiting solubility of low DS water-soluble CAs.17 For CA to be

soluble in water at low DS (0.5 to 1), it should be prepared only through homogenous

saponification.17 Thus, CA depending on the DS can be soluble in a variety of solvents.

Page 96: Rheology and Microstructure of Cellulose Acetate - Digital Repository

81

However, most commercial cellulose acetates are not soluble in water19 due to

preparation methods and DS. The interactions between hydroxyl and acetyl groups on

partially substituted CA molecule and the solvent polarity, influences its solubility.20-21

The presence of such functional groups render CA to exhibit intra- and intermolecular

long-range hydrogen bond interactions dependent on solvent.20 These interactions can

lead to solution to gel phenomena which is termed as gelation.

Gelation in polymer-solvents solutions can be induced by various techniques. In

some instances, the process of gelation occurs when solutions are heated and

subsequently cooled. Heat-induced gelation has been observed in a variety of systems

ranging from food to protein samples.22-24 Heat-induced gel formation have also been

observed in CA benzyl alcohol solutions25-27 typified by heating and cooling the

solutions to form gels.

The gelation process can be induced by addition of a gelation medium (non-

solvent) to CA/solvent system. An example is the manufacture of reverse osmosis

membranes, which often employs cellulose acetate as an important raw material.28-29

The majority of porous membranes are formed by casting the homogenous polymer

solution in a non-solvent. A non-solvent is a solvent that exhibits poor solubility

characteristics or does not possess the capability to dissolve the polymer in question.

Phase separation characteristics usually occur prior to gel/membrane formation in such

ternary systems.30-31 Phase separation is mainly due to the polymer been segregated due

to various interactions in solution. The polymer and cosolvent concentrations pay a vital

role in phase separation formation in ternary systems.32 Phase separation accompanied

by cloudiness has been observed prior to gel and/or membrane formation in some

Page 97: Rheology and Microstructure of Cellulose Acetate - Digital Repository

82

cellulose derivative solvent systems. An earlier work we conducted on the same ternary

system involved addition of water to bulk CA/N,N dimethylacetamide solutions showed

enhanced viscoelastic properties when water content was increased.33 The system

exhibited phase separation prior to gelation when water content was increased

depending on CA concentration. This study therefore, serves as an extension of our

investigations involving ternary CA/solvent/non-solvent systems.

In this study, we take advantage of the good miscibility of N,N

dimethylacetamide with water and investigate the influence different mixtures

comprising of these two solvents will have on CA. We study particularly the influence

of increasing water in the system since it is a non-solvent for CA, while N,N

dimethylacetamide show good solvent properties for CA.20 Dynamic and steady-state

rheology are primarily the tools used to study the solution and gels formed in our study.

The important concept of rheology is the ability to predict liquid and solidlike

properties since polymeric solutions typically exhibit both behaviors.34 This

characterization technique has been used in a variety of cellulose based systems.8,10,33,35

It was observed that addition of CA to different compositions of DMAc/water results in

a two-phase system which was formed at relatively high water contents. Heating the

two-phase system resulted in a single phase gel network. In that regard, we present our

results showing the effect of water content increase and its effect on intra- and

intermolecular hydrogen bond interactions present in the system.

Page 98: Rheology and Microstructure of Cellulose Acetate - Digital Repository

83

Experimental

Materials

Cellulose acetate (CA) with a molecular weight of 50,000 g/mol and acetyl

content of 39.7wt% was used in this study. HPLC grade N,N dimethylacetamide

(DMAc) and deionized water were the cosolvents adopted for this study. DMAc was

chosen as the solvent used to dissolve CA due its low toxicity and can effectively

dissolve CA to high concentration levels. Both CA and DMAc were purchased from

Sigma-Aldrich and used as received.

Sample Preparation

The mixed solvent system was prepared by adding CA to mixed fractions of

DMAc and water. The initial dissolution step involved preparation of different ratios

but constant weight of DMAc/water solutions. A constant weight of CA was then added

to these solutions containing different ratios of DMAc/water. In this study, we

primarily examine two representative CA concentrations, 10 and 15wt%, although

other concentrations have been investigated. After addition of CA, the mixture was

stirred and kept at room temperature for one day. Subsequently, to ensure an inert

atmosphere, the system was purged with nitrogen and heated for 10 minutes at 70oC to

solubilize samples. All solutions were prepared on a weight basis. We term this process

of sample preparation dissolution method 2 to differentiate it from our earlier sample

preparation technique33 (dissolution method 1) which yielded gels at relatively high

water contents (dissolution method 1). In dissolution method 1, water was added to

already prepared bulk solutions of CA/DMAc. The procedure involved, a progressive

increase in water content with the eventual formation of a gel system at water content

Page 99: Rheology and Microstructure of Cellulose Acetate - Digital Repository

84

19% and above. A comparison between the gelation properties of the two dissolution

methods will be discussed in later sections of this discussion.

Rheological Measurements

A TA Advanced Rheometer (AR 2000) was used to measure the rheological

properties of the samples at ambient temperature (25oC). The viscosity of the samples

was obtained from the steady state experiments using a cone and plate geometry.

Dynamic stress sweep experiments were performed to determine the linear viscoelastic

(LVE) regime. After determining the LVE, a small oscillatory shear was applied to the

samples and the corresponding elastic (G′) and viscous (G″) moduli measured as a

function of frequency. Frequency ranges from 0.01 rad/s to 100 rad/s were conducted

on gel samples. The shape and magnitude of the elastic and viscous moduli as a

function of frequency provides a signature of the state (e.g., solution, gel) of a system.36-

41 Parallel plate geometry was used to measure dynamic viscoelastic properties of

extremely viscous mixtures and gels. Experiments were repeated to ensure

reproducibility and consistency.

Cloud Point

Cloud point measurements were performed using an Edmunds JDS uniphase

helium-neon laser light source operating at a maximum voltage output of 5mV. Similar

measurements using a helium-neon system set up has been adopted to investigate

hydroxylpropylcellulose in mixed solvents of glycerol and water.42 The CA solutions in

DMAc/water were transferred to a cuvette having an optical path length of 1 cm. The

laser light was passed through the solutions and the transmitted output intensity

measured on an Edmunds PD200 laser meter.

Page 100: Rheology and Microstructure of Cellulose Acetate - Digital Repository

85

Scanning Electron Microscopy (SEM)

The microstructure of some gels was investigated using SEM (JEOL JSM-

5900LV) with accelerating voltage of 20kV and a magnification of 5000x. The protocol

was to cut a small piece of gel and place them in 3% glutaraldehyde Na acetate buffer at

pH 6.6 (4oC). The gels were rinsed with 3 changes of 0.1M sodium acetate buffer at pH

6.6 and then dehydrated in a graded ethanol series of 30%, 50%, 70%, 95% and 100%

ethanol, all at 4oC. Critical point drying with CO2 was used to dry samples. The dried

samples were gently fractured and sputter coated with Au/Pd before scanning, to obtain

the microstructure.

Results and Discussion

Solution Behavior

The rheological and cloud point behavior are highlighted in this section

regarding the addition of a constant weight of CA to N,N dimethylacetamide

(DMAc)/water cosolvents of different composition. Figure 1 shows the steady state

viscosity of samples as a function of shear rate for varying composition of water at a

constant CA concentration of 10wt%. The trend depicted is a Newtonian behavior with

shear thinning at high shear rates. With increasing water content, there is an appearance

of shear thinning at low shear rates and suggest the onset of microstructural

development. The viscosity curves also shifts to higher values when water content is

increased. Similar trend of increasing viscosity with water content has been observed

Page 101: Rheology and Microstructure of Cellulose Acetate - Digital Repository

86

for other CA concentrations studied but for simplicity it is not been presented in this

report.

The transmission intensity as a function of water content of an analogous sample

is portrayed in Figure 2. The intensity remains constant until a critical water content is

achieved, after which a drop in intensity is observed accompanied with cloudiness in

solution. The region showing constant intensity corresponds to a visually clear

homogenous solution. Above the water contents presented in Figure 2, the system phase

separates into a clear liquid solution on top and an opaque viscous polymer-rich mixture

at the bottom. Thus, reading of the transmitted intensity of the system was rendered

impossible at this juncture.

The transition from a clear homogeneous solution to a phase system with

increasing water content is shown more clearly in the digital images of Figure 3. In this

figure which corresponds to a 10wt.% CA in DMAc/water cosolvent with increasing

water proportion, we initially observe a clear solution at a water content of 0%. At a

higher water content of 16.2%, we observe onset of cloudiness. Upon further addition of

water (21.6%), the system separates into a top clear and bottom opaque phase. We an

intensification of the two-phase system with viscous opaque phase becoming more

pronounced at a water content of 25.2%This process was slow and equilibrium between

the two phases was established after about four days. It is of great interest to examine

the characteristics of these two phases. We start with a discussion of the rheological

behavior of the two-phase system in subsequent section. The increase in viscosity prior

to phase separation as observed in Figure 1 may be due to increased hydrogen bonds in

solution induced by adding water, thereby leading to greater intermolecular interactions

Page 102: Rheology and Microstructure of Cellulose Acetate - Digital Repository

87

between CA and cosolvents. The hypothesis for the viscosity pattern is that, water and

DMAc are completely miscible and addition of CA to the solvent system, cause both

water and to a much extent DMAc to form complex hydrogen bond interaction with

CA. We postulate the steps that hydroxyl (OH) groups on the water molecules interact

initially with the acetyl groups on DMAc. The introduction of CA lead to further partial

hydrogen bond links formed through the entire system primarily due to the OH and

acetyl groups present on the anhydroglycose chain of CA. This phenomenon may

account for the viscosity increase observed in our system. Interestingly, an investigation

conducted in dilute systems of cellulose diacetate/acetone/formamide solutions

suggested that, even though cellulose diacetate is not soluble in formamide, it swells in

formamide which is indicative of some form of interaction present.43 This highlights our

assertion that, water and CA interact partially in solution even though water is a non-

solvent.

Rheology of Phase-Separated System

Figure 4 shows viscosity pattern of the upper clear and lower opaque layer of a

two-phase system. Results for two representative samples of different water content are

shown. We find both samples to exhibit a Newtonian viscosity suggesting this layer to

be predominantly composed of water and DMAc. The viscosity pattern for the viscous

layer depicted in Figure 4b shows a shear thinning profile for all ranges of shear rate

studied. We also find the magnitude of the viscosity for this lower layer to be higher

than that of the top layer by several orders. Figure 4b also reveals that increasing water

content from 21.6 to 25.2% results in a slight increase in the viscosity as well as a

Page 103: Rheology and Microstructure of Cellulose Acetate - Digital Repository

88

steeper slope at low shear rates. The absence of a low-shear plateau suggests the

presence of microstructure whereas the increase in slope at 25.2% suggest an

enhancement of this at the higher water content. It is interesting to note that while the

viscosity increases with water content for the bottom layer, the reverse is observed for

the upper liquid. These trends are consistent and suggest that at the higher water content

of 25.2%, the bottom layer contains relatively more water than the sample at 21.6 water

resulting in a higher viscosity. Consequently, the top layer contains relatively more

DMAc than water for the higher (25.2%) water content sample resulting in a lower

viscosity as DMAc has a lower viscosity than water.

We probe further the bottom opaque layer through dynamic rheology with

Figure 5 exhibiting the elastic (G′) and viscous (G’’) moduli as a function of frequency.

We find G’ to be higher than G″ over all the frequencies studied. In addition, the

moduli are weakly dependent on frequency. These features are characteristics of a gel

network, albeit it may be termed a weak one by some authors. 44-46 An increase in water

content from 21.6 to 25.2% increases the moduli by almost an order of magnitude

results in enhanced dynamic viscoelastic properties. Both G′ and G″ increased slightly

over the frequency range studied which is reminiscent of a weak gel. In addition weak

gels often have the magnitude of G″ < ten times that G′ which is exhibited by the

viscous system.45

The phenomenon of the two-phase behavior at high water contents leads to the

idea of a competitive interaction for DMAc since both CA and water have affinity for

DMAc. Because water is completely miscible with DMAc, it is envisaged water forms

stronger bonds with DMAc as compared to CA. Therefore, at low water content, both

Page 104: Rheology and Microstructure of Cellulose Acetate - Digital Repository

89

CA and water may both interact favorably with DMAc. But as we increase the water

content, some CA chains are displaced from DMAc leading to some degree of

aggregation in the system. Further explanation could be due to the fact that, as water

composition increases, water exhibits stronger intermolecular interaction with DMAc.

With increasing water content, we tend to have a poor solvent system and CA may form

rigid conformation due to intramolecular interactions prevailing over intermolecular

attractions. It is anticipated that, intramolecular associations with various CA chains are

therefore intensified and lead to formation of a viscous mixture at the bottom of system.

Thus the two-phase system formed may be due to water exhibiting a higher hydrogen

bond preference for DMAc in comparison to CA. The dense viscous system is

envisaged may be a combination of predominantly CA, with some amount of DMAc

and to a less extent water weakly bonded with the CA chains in solution. It must be

noted that the polymer concentration for the viscous mixture when isolated from liquid

layer, was computed to be approximately 16wt%.

Similar phase separated two-phase system was observed in polymer solutions up

to 20% CA in dioxane/water mixture, formed upon aging of turbid solutions at room

temperature.47 Experiments conducted on cellulose derivatives give indications that

intermolecular interactions may be one of the important mechanisms for gel formation.6

Thus in this system CA may be experiencing a greater intramolecular interaction in

comparison to intermolecular interactions as we increased the amount of water in

solution.

Page 105: Rheology and Microstructure of Cellulose Acetate - Digital Repository

90

Comparison of Gels from Dissolution Methods 1 and 2

An important issue to consider is a comparison of the gel rheology of the bottom

layer with the gels produced in our earlier work using a different sample preparation

protocol (referred to as Method 1).33 This earlier approach, described also in the

experimental section, involved addition of water to prepared CA/DMAc solutions in

contrast to our present approach involving addition of CA to a cosolvent of DMAc and

water (coined Method 2). Method 1 led to uniform gel formation at high water content

and a detailed explanation for this phenomena is provided elsewhere.33 Nevertheless,

the dynamic moduli of the two gels are compared in Figure 6 for identical system

conditions. We find that that gels formed through dissolution method 1 exhibit slightly

higher elastic and viscous moduli for the two representative water contents of 21.6%

and 25.2%. This is counterintuitive as one would expect G’ of gels from dissolution

method 2 to be higher as the effective CA concentration in the bottom layer for

dissolution method 2 is higher than the corresponding gel from dissolution method 1.

Since our results cannot be explained on the basis of concentration effects, we can

suggest other possibilities. First, the gel from the bottom layer of the two phase system

may have relatively lower water content. Second, the gel microstructures in the two

cases may be different. The observed phenomenon may thus be due to the complete

dispersion of CA after gelation in dissolution method 1, which provides better gel

properties. In contrast, in dissolution method 2, the aggregation of CA clusters with

small amounts of cosolvent weakly attached produces a system with weaker gel

characteristics. Further work needs to be done to elucidate the exact nature of the

Page 106: Rheology and Microstructure of Cellulose Acetate - Digital Repository

91

interactions. We refrain from delving into this further as it beyond the scope of this

work..

After effect of Heat Treatment

The two-phase system formed at high water contents were subjected to

temperature increment spanning from 60 to 100oC for 30 minutes and cooled back to

room temperature. This procedure was performed to investigate if temperature could

induce the system to exhibit a heat-induced gelation process and form a single-phase gel

system. Figure 7 shows digital images of the a) initial heated two phase system, b)

system heated 100oC sample cooled back to ambient temperature after heating. We find

that the system heated to 100oC showed a complete gelled texture with little or no

solvent around the gel matrix upon cooling. In that context, our discussions will center

primarily on the initial gel layer of the two-phase system, and the transformed gel

matrix upon cooling to room temperature after heating to 100oC for 30minutes. Figure 8

shows the dynamic viscoelastic profile at 10wt% for two different water contents of

21.6 and 25.2 wt%. We see a slight pronouncement (about double) in the viscoelastic

properties with water content increment. Thus, the aggregation process leading to

gelation is enhanced to some extent when water content is increased.

Figure 9a shows SEM micrographs for heat-induced gels for 10wt.% CA

concentrations at two different water contents 21.6% and 25.2%. We observe a

particulate microstructure with pores with both the size of the particulates and the pores

showing a small decrease with increasing water content. For a CA concentration of

15wt.% (Figure 9b), the microstructure for the lower water containing system (20.4%)

Page 107: Rheology and Microstructure of Cellulose Acetate - Digital Repository

92

reveal similar overall features as that observed in Figure 9a. However, increase in water

content to 23.8% leads to a pronounced change in microstructure and depicts a

predominantly fibrular morphology.

It is envisaged that heating the two-phase system to 100oC breaks the majority

of the interactive bonds formed between water and DMAc initially, thus making

unbonded DMAc assessable to CA in solution facilitating their contact. Therefore the

hydroxyl groups on CA and the acetyl groups on DMAc form strong intermolecular

interactions. After cooling back to room temperature, water competes with CA for the

DMAc. This leads bond cleavage of some CA molecules with some of these molecules

exhibiting a spike in intramolecular interactions. But at the same time CA molecules

also form intermolecular interactions with DMAc and to a lesser extent water. A

balance of these interactions causes dispersion of CA uniformly through the entire

system leading to gelation. An earlier study of a ternary system comprising of same

constituents we conducted showed a similar gelation process with increasing water

content at room temperature.33

Comparison of Heat-induced and Dissolution method 1 Gels

Figure 10 compares the dynamic rheological properties of gels formed using

dissolution method 1 and heating of two-phase samples. Figure 10a shows almost

identical dynamic viscoelastic properties for 10wt% CA concentration containing the

same amount of water (21.6%). A different pattern is however observed at the higher

water content of 25.2% (Figure 10b) with the G’ and G’’ of the heat-induced gels

showing a significantly lower value in comparison to the gels prepared using dissolution

Page 108: Rheology and Microstructure of Cellulose Acetate - Digital Repository

93

method 1. This was also evident qualitatively from the texture and visual observation of

the gels. At high water contents, the observed phenomenon may be due to the complete

dispersion of CA after gelation in dissolution method 1, which provides better gel

properties. On the other hand, the heat-induced gelation process, the aggregation of CA

clusters with small amounts of cosolvent weakly attached produces a system with

weaker gel characteristics.

In Figure 11, we observe the SEM images which show similar microstructural

characteristics for the heat-induced and dissolution method 1 gels. The texture for the

gels in dissolution method 1 assumes a more rough texture with close-packed structure

compared to the heat-induced gels at 10wt% CA concentration and identical water

contents. In addition, with water content increment, we observe the voids become

smaller in size. This microstructural formation may allude to the stronger and rigid gel

characteristics exhibited by dissolution 1 containing 2.8g (25.2%) water composition

and the disparity that is reflected in the dynamic response in Figure 12b

Conclusions

In this study, we examine the effect of sample preparation protocol on CA

systems in mixed solvents of DMAc and water. At relatively low water content, the

solutions are homogenous and exhibit an increase in viscosity with enhanced water

content in the cosolvent. It is evident that sequence of addition of CA and solvents

greatly influence the phase separation mechanism in our system. By preparing solutions

of CA in different ratios of DMAc and water, we produce a system, which phase

Page 109: Rheology and Microstructure of Cellulose Acetate - Digital Repository

94

separates at high water contents (>19wt%) for a specific polymer concentrations. A

two-phase system consisting of a dense polymer-rich gel-like bottom phase and a clear

liquid upper layer are formed. Heat treatment of the two-phase system at a high

temperature (~100oC) causes the formation of a uniform gel structure whereas heating it

to a lower temperature causes the system to revert back to a two-phase system. By

increasing water content in the system, the intramolecular interaction between CA units

is increased. In addition CA forms strong and weak links with DMAc and water

respectively causing gelation to occur. This pattern of hydrogen bond interactions may

account for the increase in viscoelastic properties in our system.

Acknowledgement

The authors gratefully acknowledge funding support for this work from United

States Department of Agriculture (USDA).

Page 110: Rheology and Microstructure of Cellulose Acetate - Digital Repository

95

References

(1) Kamide, K., Saito, M. Effect of total degree of substitution on molecular

parameters of cellulose acetate. European Polymer Journal 1984, 20(9), 903-

914.

(2) Kulicke, W-M.; Kull, A.H.; Thielking, H. Characterization of aqueous

carboxymethylcellulose solutions in terms of their molecular structure and its

influence on rheological behavior. Polymer 1996, 37(13), 2723-2731.

(3) Bochek, A.M. Effect of hydrogen bonding on cellulose solubility in aqueous and

nonaqueous solvents. Russian Journal of Applied Chemistry 2003, 76(11), 1761-

1770.

(4) Saalwachter, K.; Burchard, W. Cellulose in new metal-complexing solvents. 2.

Semidilute behavior in Cd-tren. Macromolecules 2001, 34, 5587-5598.

(5) Hattori, K., Cuculo, J.A., Hudson, S.M., New solvents for cellulose:

hydrazine/thiocyanate salt system. Journal of Polymer Science: Part A: Polymer

Chemistry 2002, 40, 601-611.

(6) Itagaki, H., Tokai, M., Physical gelation process for cellulose whose hydroxyl

groups are regioselectively substituted by fluorescent groups”, Polymer 1997,

38, 4201-4205.

(7) Tsunashima, Y., Hattori, K., Substituent distribution in cellulose acetate: Its

control and the effect on structure formation in solution. Jour. of Colloid and

Interface Sci., 2000, 228, 279-286.

Page 111: Rheology and Microstructure of Cellulose Acetate - Digital Repository

96

(8) Kobayashi, K., Huang, C., Lodge, T.P., Thermoreversible gelation of aqueous

methylcellulose solutions, Macromolecules 1999, 32, 7070-7077.

(9) Sarkar, N., Thermal gelation properties of methyl and hydroxypropylmethyl

cellulose. Journal of Applied Polymer Science 1979, 24, 1073-1087.

(10) Sarkar, N., Kinetics of thermal gelation of methyl cellulose and hydroxypropyl

cellulose in aqueous solutions. Carbohydrate Polymers 1995, 26, 195-203.

(11) Takahashi, M., Shimazaki, M., Yamamoto, J., Thermoreversible gelation and

phase separation in aqueous methyl cellulose solutions. Journal of Polymer

Science: Part B: Polymer Physics, 2001, 39, 91-100.

(12) Li, L., Thangamathesvaran, P.M., Yue, C.Y., Tam, K.C., Hu, X., Lam, Y.C., Gel

network structure of methylcellulose in water. Langmuir 2001, 17, 8062-8068.

(13) Desbrieres, J., Hirrien, M., Ross-Murphy, S.B., Thermogelation of methyl

cellulose: rheological considerations. Polymer 2000, 41, 2451-2461.

(14) Takahashi, M., Shimazaki, M., Formation of junction zones in thermoreversible

methyl cellulose gels. Journal of Polymer Science: Part B: Polymer Physics,

2001, 39, 943-946.

(15) Bochek, A.M., Zabivalova, N.M., Lavrent’ev, V.K., Lebedeva, M.F.,

Sukhanova, T.E., Petropavlovskii, G.A., Formation of physical thermal

reversible gels in solutions of methyl cellulose in water and dimethylacetamide

and properties of films thereof. Russian Journal of Applied Chemistry 1978,

74(8), 1358-1363.

Page 112: Rheology and Microstructure of Cellulose Acetate - Digital Repository

97

(16) Meibner, D., Einfeldt, J., Schareina, T. New results concerning the behavior of

cellulose acetate in solutions and films by means of CD measurements.

Biopolymers 1999, 50, 163-166.

(17) Bochek, A.M., Kalyuzhnaya, L.M., Interaction of water with cellulose and

cellulose acetates as influenced by the hydrogen bond system and hydrophilic-

hydrophobic balance of the macromolecules. Macromolecular Chemistry and

Polymeric Materials 2002, 75, 989-993.

(18) Heinze, T., Liebert, T. Chemical characterics of cellulose acetate.

Macromolecular Symposia 2004, 208, 167-237.

(19) Steinmeier, H. Acetate manufacturing, processing and technology.

Macromolecular Symposia 2004, 208, 49-60.

(20) Kawanishi, H., Tsunashima, Y., Okada, S., Horii, F. Change in chain stiffness in

viscometric and ultracentrifugal fields: cellulose diacetate in N,N-

dimethylacetamide dilute solution. Journal of Chemical Physics 1998, 108 (14),

6014-6025.

(21) Kawanishi, H., Tsunashima, Y., Horii, F., Dynamic light scattering study of

structural changes of cellulose diacetate in solution under couette flow.

Macromolecules 2000, 33, 2092-2097.

(22) Ikeda, S. Heat-induced gelation of whey proteins observed by rheology, atomic

force microscopy, and raman scattering spectroscopy. Food Hydrocolliods 2003,

17, 399-406.

Page 113: Rheology and Microstructure of Cellulose Acetate - Digital Repository

98

(23) Kavanagh, G.M., Clark, A.H., Ross-Murphy, S.B. Heat-induced gelation of

globular proteins. Part 5. Creep behavior of β-lactoglobulin gels. Rheologica

Acta 2002, 41, 276-284.

(24) Adachi, M., Chunying, H., Utsumi, S., Effects of designed sulfhydryl groups

and disulfide bonds into soybean proglycinin on its structural and heat-induced

gelation. Journal of Agricultural and Food Chemistry 2004, 52, 5717-5723.

(25) Ryskina, I.I., Aver’yanova, V.M. A study of gelation and the structure of

acetylcellulose gels in benzyl alcohol. Vysokomol. Soyed. 1971, A13, 2189-

2194.

(26) Ryskina, I.I., Aver’yanova, V.M. Morphological study of the three dimensional

network of cellulose acetate-benzyl alcohol gelatins. Vysokomol. Soyed. 1975,

A17, 919-922.

(27) Ryskina, I.I., Aver’yanova, V.M. Gelation and structure formation processes of

cellulose acetate gels in benzyl alcohol.

(28) Kesting, R.E., Barsh, M.K., Vincent, A.L., Semipermeable membranes of

cellulose acetate for desalination in the process of reverse osmosis II. Parameters

affecting membrane gel structure”, Journal of Applied Polymer Science Jour. of

Applied Polymer Science 1965, 9, 1873-1893.

(29) Pilon, R., Kunst, B., Sourirajan, S., Studies on the development of improved

reverse osmosis membranes from cellulose acetate-acetone-formamide casting

solution. Journal of Applied Polymer Science 1971, 15, 1317-1334.

(30) Vogrin, N., Stropnik, C., Musil, V., Brumen, M., The wet phase separation: the

effect of cast solution thickness on the appearance of macrovoids in the

Page 114: Rheology and Microstructure of Cellulose Acetate - Digital Repository

99

membrane forming ternary cellulose acetate/acetone/water system. Journal of

Membrane Science 2002, 207, 139-141.

(31) Kunst, B., Sourirajan, S., Evaporation rate and equilibrium phase separation data

in relation to casting conditions and performance of porous cellulose acetate

reverse osmosis membranes. Journal of Applied Polymer Science 1970, 14,

1983-1996.

(32) Vaessen, D.M., McCormick, A.V., Francis, L.F., Effect of phase separation on

stress development in polymeric coatings. Polymer 2002, 43, 2267-2277.

(33) Appaw, C., Gilbert, R.D., Kadla, J.F., Khan, S.A. Effect of water composition

on rheological properties in ternary cellulose acetate solutions. (not submitted)

2004.

(34) Ramachandran, S., Chen, S., Etzler, F. Rheological characterization of

hydroxylpropylcellulose gels. Drug Development and Industrial Pharmacy

1999, 25(2), 153-161.

(35) Ryo, Y., Nakai, Y., Kawaguchi, M., Viscoelastic measurements of silica

suspensions in aqueous cellulose derivative solutions. Langmuir 1992, 8, 2413-

2416.

(36) Macosko, C.W. Rheology Principles, Measurements and Applications. Wiley-

VCH Publishers, Inc. , New York 1994.

(37) Pai, V., Srinivasarao, M., Khan, S. A. Evolution of Microstructure and Rheology

in Mixed Polysaccharide Systems. Macromolecules 2002, 35, 1699-1707.

Page 115: Rheology and Microstructure of Cellulose Acetate - Digital Repository

100

(38) Wilder, E. A., Hall, C. K., Khan, S. A., Spontak, R. J. Effects of Composition

and Matrix Polarity on Network Development in Organogels of Poly(ethylene

glycol) and Dibenzylidene Sorbitol. Langmuir 2003, 19(15), 6004-6013.

(39) Raghavan, S. R., Walls, H. J., Khan, S. A. Rheology of Silica Dispersions in

Organic Liquids: New Evidence for Solvation Forces Dictated by Hydrogen

Bonding. Langmuir 2000, 16(21), 7920-7930.

(40) Chauvelon, G., Doublier, J-L., Buleon, A., Thibault, J-F., Saulnier, L.,

Rheological properties of sulfoacetate derivatives of cellulose, Carbohydrate

Research 2003, 338, 751-759.

(41) Lazaridou, A., Biliaderis, C.G., Izydorczyk, M.S., Molecular size effects on

rheological properties of oat glucans in solution and gels. Food Hydrocolloids

2003,17, 693-712.

(42) Fuji, S., Sasaki, N., Nakata, M., Rheological studies on the phase separation of

hydroxylpropylcellulose solution systems. Journal of Polymer Science: Part B:

Polymer Physics 2001, 39, 1976-1986.

(43) Pintaric, B., Rogosic, M., Mencer, H.J., Dilute solution properties of cellulose

diacetate in mixed solvents. Journal of Molecular Liquids 2000, 85, 331-350.

(44) Griswold, P.D., Cuculo, J.A., An experimental study of the relationship between

rheological properties and spinnability in the dry spinning of cellulose acetate-

acetone solutions. Journal of Applied Polymer Science 1974, 18, 2887-2902

(45) Ikeda, S., Nishinari, K., “Weak-gel”-type rheological properties of aqueous

dispersions of nonaggregated k-Carrageenan helices. Journal of Agric. Food

Chem. 2001, 49, 4436-4441.

Page 116: Rheology and Microstructure of Cellulose Acetate - Digital Repository

101

(46) Tsitsilianis, C. Viscoelastic properties of physical gels formed by associative

telechelic polyelectrolytes in aqueous media. Macromolecules 2002, 18, 3662-

3667.

(47) Altena, F.W., Smolders, C.A., Phase separation phenomena in solutions of

cellulose acetate. I. Differential scanning calorimetry of cellulose acetate in

mixtures of dioxane and water Jour. of Polymer Science: Polymer Symposium

1981, 69, 1-10.

Page 117: Rheology and Microstructure of Cellulose Acetate - Digital Repository

102

Figure Captions

Figure 1. Effect of different water content on the steady shear viscosity of cellulose

acetate in DMAc/water cosolvent system. The total concentration of cellulose acetate is

kept constant at 10wt%.

Figure 2. Transmission intensity versus amount of water for a constant

10wt%cellulose acetate in DMAc/water cosolvent system.

Figure 3. Constant 10wt% cellulose acetate in DMAc/water cosolvent system showing

transition from homogenous system to a two-phase system. Water content increment

is showed from left to right.

Figure 4. Viscosity pattern for the two-phase region for cellulose acetate

inDMAc/water cosolvent system at two different water contents of 21.6 and 25.2wt%.

Polymer concentration is a constant 10wt%. (a) Upper liquid and (b) Lower opaque

layer.

Figure 5. Dynamic frequency spectrum of the elastic and viscous moduli for bottom

viscous layer for cellulose acetate in DMAc/watercosolvent system at two different

water contents of 21.6 and 25.2wt%. Total concentration of cellulose acetate is kept

constant at 10wt%.

Page 118: Rheology and Microstructure of Cellulose Acetate - Digital Repository

103

Figure 6. Dynamic spectrum for dissolution methods 1 and 2 for 10wt% CA

concentration with water contents of (a) 21.6% and (b) 25.2%

Figure 7. Two phase system subjected to heat treatment to 100oC after which system is

cooled to room temperature.

Figure 8. Dynamic spectra for heat-induced gelation for 10wt% cellulose acetate

DMAc/water cosolvent system at two different water contents of 21.6 and 25.2wt%.

Figure 9. SEM images for heat-induced CA concentrations of (a) 10 and (b) 15wt% at

two different water contents.

Figure 10. Comparison of dynamic properties of heat-induced gels and dissolution

method 1 gels at 10wt% cellulose acetate concentration having water content of (a)

21.6% and (b) 25.2%.

Figure 11. SEM micrographs of heat-induced and dissolution method 1 gels for 10wt%

cellulose acetate concentration having water contents of (a) 21.6% and (b) 25.2%.

Page 119: Rheology and Microstructure of Cellulose Acetate - Digital Repository

104

Shear rate (s-1)

10-3 10-2 10-1 100 101 102 103

Vis

cosi

ty (P

a.s)

100

101

0%

3.6%

7.2%

10.8%

14.4%

Figure 1. Effect of different water content on the steady shear viscosity of cellulose

acetate in DMAc/water cosolvent system. The total concentration of cellulose acetate is

kept constant at 10wt%.

Page 120: Rheology and Microstructure of Cellulose Acetate - Digital Repository

105

Water content (wt.%)

0 2 4 6 8 10 12 14 16 18 20

Tran

s. In

tem

sity

(ad.

uni

ts)

10-1

100

101

Figure 2. Transmission intensity versus amount of water for a constant 10wt%

cellulose acetate in DMAc/water cosolvent system.

Page 121: Rheology and Microstructure of Cellulose Acetate - Digital Repository

106

Figure 3. Constant 10wt% cellulose acetate in DMAc/water cosolvent system

showing transition from homogenous system to a two-phase system. Water

content increment is showed from left to right.

0% 21.6%16.2% 25.2%

Page 122: Rheology and Microstructure of Cellulose Acetate - Digital Repository

107

Shear rate (1/s)101 102 103

Vis

cost

y (P

a.s)

10-3

10-2

10-1

25.2%

21.6%

Shear rate (1/s)10-4 10-3 10-2 10-1 100 101

Vis

cost

y (P

a.s)

102

103

104

105

106

25.2%

21.6 %

Figure 4. Viscosity pattern for the two-phase region for cellulose acetate

inDMAc/water cosolvent system at two different water contents of 21.6 and 25.2wt%.

Polymer concentration is a constant 10wt%. (a) Upper liquid and (b) Lower opaque

layer.

Page 123: Rheology and Microstructure of Cellulose Acetate - Digital Repository

108

Frequency (rad/s)10-3 10-2 10-1 100 101 102 103

G' &

G" (

Pa)

102

103

104

105

106

G' (25.2%)

G" (25.2%)

G' (21.6%)

G" (21.6%)

Figure 5. Dynamic frequency spectrum of the elastic and viscous moduli for bottom

viscous layer for cellulose acetate in DMAc/watercosolvent system at two different

water contents of 21.6 and 25.2wt%. Total concentration of cellulose acetate is kept

constant at 10wt%.

Page 124: Rheology and Microstructure of Cellulose Acetate - Digital Repository

109

Frequency (rad/s)10-3 10-2 10-1 100 101 102 103

G' &

G" (

Pa)

102

103

104

105

G' (21.6%, method 1)

G' (21.6%, method 2)

G" (21.6%method 1)

G" (21.6%, method 2)

Frequency (rad/s)10-3 10-2 10-1 100 101 102 103

G' &

G" (

Pa)

103

104

105

106

G' (25.2%, method 1)

G' (25.2%, method 2)G" (25.2%, method 1)

G" (25.2%, method 2)

Figure 6. Dynamic spectrum for dissolution methods 1 and 2 for 10wt% CA

concentration with water contents of (a) 21.6% and (b) 25.2%

Page 125: Rheology and Microstructure of Cellulose Acetate - Digital Repository

110

Heated to100oC Cooled to room temperature Figure 7. Two phase system subjected to heat treatment to 100oC. After which system

is cooled to room temperature.

Page 126: Rheology and Microstructure of Cellulose Acetate - Digital Repository

111

Frequency (rad/s)10-3 10-2 10-1 100 101 102 103

G' &

G" (

Pa)

103

104

105

G' (25.2%)

G' (21.6%)

G" (25.2%)

G" (21.6%)

Figure 8. Dynamic spectra for heat-induced gelation for 10wt% cellulose acetate

DMAc/water cosolvent system at two different water contents of 21.6 and

25.2wt%.

Page 127: Rheology and Microstructure of Cellulose Acetate - Digital Repository

112

(a) 21.6% 25.2%

(b) 20.4% 23.8% Figure 9. SEM images for heat-induced CA concentrations of (a) 10 and (b) 15wt% at

two different water contents.

Page 128: Rheology and Microstructure of Cellulose Acetate - Digital Repository

113

Frequency (rad/s)10-3 10-2 10-1 100 101 102 103

G' &

G" (

Pa)

103

104

105

G' (Heat-Induced)

G' (method 1)

G" (Heat Induced)

G" (method 1)

Frequency (rad/s)10-3 10-2 10-1 100 101 102 103

G' &

G" (

Pa)

103

104

105

106

G' (dissolution method 1)

G' (heat induced)

G" (dissolution method 1)

G" (heat induced)

Figure 10. Comparison of dynamic properties of heat-induced gels and dissolution

method 1 gels at 10wt% cellulose acetate concentration having water content of (a)

21.6% and (b) 25.2%.

Page 129: Rheology and Microstructure of Cellulose Acetate - Digital Repository

114

(a) 21.6% 25.2%

(b) 21.6% 25.2%

Figure 11. SEM micrographs of (a) heat-induced and (b) dissolution method 1 gels for

10wt% cellulose acetate concentration having water contents of 21.6% and 25.2%.

Page 130: Rheology and Microstructure of Cellulose Acetate - Digital Repository

115

CHAPTER 6

Evolution of Microstructure and Rheological Properties in Cellulose

acetate/N,N dimethylacetamide/Water system

Chapter 6 is essentially a manuscript by Collins Appaw, John F. Kadla and Saad A.

Khan, prepared for submission to Biomacromolecules.

Page 131: Rheology and Microstructure of Cellulose Acetate - Digital Repository

116

Evolution of Microstructure and Rheological Properties in Cellulose acetate/N,N dimethylacetamide/Water system

Collins Appaw and Saad A. Khan*

Department of Chemical and Biomolecular Engineering, North Carolina State University, Raleigh, North Carolina 27695

John F. Kadla*

Department of Wood & Paper Science, Advanced Biomaterials Chemistry, University of British Columbia, Vancouver, BC V6T 1R9 Canada

Abstract

Dynamic rheology is used to examine gelation properties induced by increasing water

content in cellulose acetate (CA)/N,N dimethylacetamide (DMAc)/water system. We

examine yield stress of the gels by performing stress sweep experiments by alternatively

increasing water content and CA concentration both resulting in higher yield stresses.

Temperature ramp and dynamic frequency experiments are also performed to determine

the thermal stability of the physical gels and provide us with the gel-sol transition point

(Tgs). The resulting systems exhibit “solution-like” properties beyond specific

temperatures. The changes in dynamic viscoelastic properties may be caused by

enhanced hydrogen interactions in the system when CA concentration and primarily

water content or are increased.

*Corresponding authors: Saad A. Khan; Phone: 919 515-4519; Fax: 919-515-3465; e-mail

[email protected]; John F. Kadla; Phone: 604 827-5254; Fax: 604-822-9104; e-mail:

[email protected]

Page 132: Rheology and Microstructure of Cellulose Acetate - Digital Repository

117

Introduction

Cellulose derivatives exhibit variety of solution properties dependent on nature

and type of functional groups used to substitute the hydroxyl groups on the

anhydroglucose units in cellulose. The use of solvents which interact with specific

functional groups may cause precise intermolecular linkages and can allow particular

process conditions be established.1 By utilizing particular substituents some derivatives

can be water-soluble or insoluble in water but soluble in some solvent systems, which

can be single or combinations of different solvent mixtures. Typical derivatives, which

are soluble in aqueous solutions include methyl cellulose (MC)2-10, hydroxypropyl

methyl cellulose (HPMC)3-4 including other hydrophobically modified cellulose

derivatives.

Cellulose acetate (CA), a derivative of cellulose is a polymer that has a

semiflexible chain11 characterized by partially or fully substituting the hydroxyl (OH)

groups on the C-2,3 and 6 positions with O-acetyl groups.12-14 It exhibits different

solubility behavior in various solvents dependent on the degree of substitution (DS) of

the acetyl units.13-14 For instance, cellulose acetates having a DS of between, 0.5 to 1 are

soluble in water.15-16 Uniform distribution of the substituents have been found to be

essential for CA to be water-soluble.16 It is suggested that, solubility in water also

depends on the substitution of the OH groups at the C3 position.16 A study on dilute

cellulose diacetate (CDA) in mixed solvents, indicate that solvents capable of dissolving

CDA can be classified into two main groups.17 There are basic solvents (e.g. acetone)

which interact primarily with hydroxyl groups on the chain and acidic solvents (e.g.

formic acid) that primarily solvate the acetyl groups.17 Thus, the use of specific solvents

Page 133: Rheology and Microstructure of Cellulose Acetate - Digital Repository

118

could impact partially substituted CA chains to exhibit different solution characteristics

which can lead to interesting system properties changes such as viscosity modifications,

gelation etc.

Gel and membrane formation have been observed in different cellulose

derivative solvent systems.2-9, 18-21 Most cellulose derivative gel systems if not all, form

physical gel network. The process of physical gelation is not well understood in

comparison to chemical gelation due to the transient nature of gel network junctions.22

These network junctions holding physical gels can be created and destroyed by thermal

agitation of the polymeric system.5,8 For example, aqueous systems of MC and HPMC

exhibits thermoreversible gelation by forming a three-dimensional network structure.3

This unique gelation process involves formation of gels when the aqueous systems of

these derivatives are heated and reversibly form solutions when cooled.2-9 The

phenomena may be attributed primarily to hydrophobic interaction between molecules

having methoxyl subsititutes.3 Unlike MC aqueous systems, low DS (0.5 to 1) CA

aqueous solutions do not gel when heated.16 Gelation in some CA/organic solvent

systems have been observed19, 23,25 and has been attributed to be mainly caused by

strong intermolecular contacts.19 CA containing different amounts of acetic acid in

benzyl alcohol has been shown to exhibit gel formation.18-20 The gelation temperature

was discovered to depend on molecular weight and quantity of combined acetic acid.18

These gels are found to exhibit structural changes dependent on solvent quality.18 5%

solutions of CA in benzyl alcohol prepared at 60-80oC formed gels when solutions are

cooled to 25oC.20 CA samples with different molecular weight and degree of

substitution containing combined acetic acid concentrations of 54% and 60% also lead

Page 134: Rheology and Microstructure of Cellulose Acetate - Digital Repository

119

to gel formation in dimethyl phthalate.21 Increase in degree of substitution and polymer

concentration in the system causes an increase in the dynamic moduli.21

The process of gelation can also be induced by the addition of a non-solvent to a

CA/solvent system or by adding CA to a mixture of solvent/non-solvent. This gelation

process may involve phase separation prior to gelling. It is dependent on not only on the

polymer characteristic, but also on specific polymer/solvent and polymer/polymer

interactions.23 A previous study we conducted which involves addition of water to

CA/DMAc solutions shows gelation occurring at relatively high water contents

(>18wt%) for CA concentrations investigated.24 The system phase separated which was

determined by cloud point studies prior to gelling.24 Ternary systems of

CA/solvent/water have be found to exhibit gelation using solvents comprising of

acetone, dioxane and tetrahydrofuran (THF).25 The gelation process in a study

conducted on CA/dioxane/water system was found to be very slow26, a phenomenon

observed in our previous study.24 It must be stated that, investigations in gel formation

in CA systems are few and far between. Possibly due to the complexity involved and/or

the relatively high concentrations of polymer which may be needed for gelation to

occur.

We seek to take advantage of the good solvent properties of N,N

dimethylacetamide (DMAc) for CA and the subsequent addition of a non-solvent

(water) to induce gelation in our system. In our previous work on same system, we

established protocols needed to covert solutions to gels.24 By tuning the polymer

concentration and using appropriate combination of both solvents, we determine

conditions pertaining to physical gel formation.24 This paper examines the dynamic

Page 135: Rheology and Microstructure of Cellulose Acetate - Digital Repository

120

rheological properties of the gels formed by adding relatively high water content to

CA/DMAc solutions. Rheological methods have been used to characterize various

cellulose and cellulose derivatives in aqueous and organic solvent systems.2,4,6-7,24,27 It

provides important information on gel properties such changes in dynamic viscoelastic

properties and is the most direct method for determining sol-gel transition.28 Therefore

rheology can also be used to study the evolution and changes during the transition from

solution to gel network and vice versa. The main focus of this paper is to examine from

the standpoints the effect of CA concentration and solvent quality on the gel network

structure. We initially investigate the changes in dynamic viscoelastic properties by

varying CA concentration and water content. The nonlinear viscoelastic behavior is also

examined to determine mechanical strength and fracture properties of these gels. This is

a technique that has been used in studying a considerably wide range of materials.29-31

Investigations in nonlinear viscoelastic characteristics of cellulosic fiber gels showed

the gel suspensions to exhibit nonlinear behavior at strains above 1%.29 Finally, we

examine the gels with respect to their thermal stability, which provide us with the gel to

sol transition temperature (Tgs). Heating the gels and its effect on the dynamic

viscoelastic properties changes enabled us to determine the gel to sol transition point.

Experimental

Materials

Cellulose acetate (CA) with a molecular weight of 50,000 g/mol and acetyl

content of 39.7wt% was used. HPLC grade N,N dimethylacetamide (DMAc) and

Page 136: Rheology and Microstructure of Cellulose Acetate - Digital Repository

121

deionized water were the cosolvents used for this study. DMAc is a good solvent for

CA and dissolves CA to appreciable high concentrations. Both CA and DMAc were

purchased from Sigma-Aldrich and used as received.

Sample Preparation

Bulk CA concentrations in DMAc were prepared at a constant CA concentration

of 20wt%. The procedure was to add calculated weight fractions of DMAc to the

aliquots of the 20wt% stock solutions. Sample were mixed and kept for a day at ambient

conditions. Specific amounts of deionized water was then added to the resulting

solutions, after which samples were heated at 70oC for 10 minutes to ensure complete

miscibility The relatively high water contents (>19wt%) used in this study ensured that

all samples analyzed were gels. Prior to heating, samples were maintained under

nitrogen atmosphere to prevent oxidation of CA solutions. Samples after heating were

kept at room temperature for at least five days prior to use to allow system to

equilibrate. All stock solutions and dilutions were prepared on a weight basis.

Rheological Measurements

A TA Advanced Rheometer (AR 2000) was used to measure the dynamic

rheological behavior of the samples at various temperatures. In these experiments, an

oscillatory shear was applied to the samples and the corresponding elastic (G′) and

viscous (G″) moduli measured as a function of either frequency or stress amplitude.

Initial experiments at a constant frequency but increasing stress amplitude were

performed to determine the linear viscoelastic (LVE) for the frequency sweep

experiments. The stress sweep experiments also provided information on the yield

stress/strain of the gels30-31, as discussed later in the paper. Subsequently experiments

Page 137: Rheology and Microstructure of Cellulose Acetate - Digital Repository

122

were conducted as a function of frequency at a low strain/stress level within the LVE

regime. The magnitude and shape of corresponding elastic (G′) and viscous (G″) moduli

measured as a function of frequency provides a signature of the state of a system (e.g.,

gel or liquid).31-33 To determine the gel to sol transition as a function of increasing

temperature, two types of experiments were conducted, temperature ramp and

frequency sweep at different temperatures. Experiments were performed at different

temperatures between 25oC and 100oC. Temperature ramp experiments were conducted

at a heating rate of 1oC/min and frequency of 1 rad/s to obtain the temperature range

where gel-sol transitions occur. Dynamic frequency sweep experiments at different

temperatures were also performed and changes in viscoelastic properties used to

estimate gel-sol transition . For all the experiments, the disk shaped gels were carefully

cut and carefully placed on the lower plate of parallel plate geometry (20cm diameter or

radius) [did you check for slip using sandpaper or different gaps?]. For temperatures

greater that 25oC, the samples were coated around the circumference with

polydimethylsiloxane (PDMS) oil to prevent solvent evaporation.

Confocal Scanning Laser Microscopy (CSLM)

In the CSLM experiments, a confocal laser scanning system (Leica TCS SP)

connected to an inverted microscope (Leica DM IBRE) was used to scan the gels.

Confocal microscopic analysis have previously been adopted to characterize

membranes of mixed cellulose esters (cellulose nitrate/cellulose acetate).35-36 Prior to

preparing our gels, our solutions were tagged with calcofluor white which is a cellulose-

selective fluorescent dye.34 Measurements were performed with 100x numerical

aperture 1.4 oil immersion objective. Optical sectioning was performed as a function of

Page 138: Rheology and Microstructure of Cellulose Acetate - Digital Repository

123

the depth of the gels along the z-axis. The excitation source we used was a UV laser

(350 nm, 361 nm; it might excite at both wavelengths simultaneously). The spatial

resolution of the images was 512x512 pixel. The image analysis was done using Adobe

Photoshop 7.0 (Adobe Systems Inc.).

Results and Discussion

Rheology and Microstructure

We first investigate changes in the dynamic viscoelastic properties at 25oC for

two conditions: constant CA concentration with varying water content and constant

water content with changing CA concentration. Figure 1a shows frequency spectrum of

the elastic (G′) and viscous (G″) moduli of 10wt% CA samples in DMAc/water

containing different water content. Figure 1b reveals similar data for samples with

15wt.% CA. For the sake of clarity, G″ results are shown only for one representative

sample. We observe G′ to be relatively flat and greater than G″ (this is true for all

samples although data for only one is shown), features indicative of a gel network.37-41

In addition, we observe G′ to increase with increasing CA content or water.

The effects of water content and CA concentration are depicted more clearly in

Figure 2. Figure 2 shows G′ (at a fixed frequency of 1 rad/s) as a function of water

content for a CA concentration of 10 and 15wt%. With successive increase in water

content, G′ shifts to higher values, with an overall increase greater than an order of

magnitude. We observe over a five-fold increase in G′ as CA concentration is varied

from 7.5wt% to 15wt%. The increasing trend in G’ observed in Figure 2 suggests

Page 139: Rheology and Microstructure of Cellulose Acetate - Digital Repository

124

formation of a progressively stronger network with either an increase in water content

or CA concentration.

Direct evidence of a sample spanning gel network is observed from confocal

micrographs shown in Figure 3 obtained under different water content for two different

CA concentrations. In Figure 3a, at low CA concentration (10wt.%) and water content

(21.6%), we observe a very open network. The microstructure also seems somewhat

aggregated. At a higher water content (23.2%) but at the same CA concentration, we

find the network to be less open and the microstructure more uniform. In Figure 3b at a

higher CA concentration (15%), we see a significantly different microstructure. A

denser network with a fine and uniform structure is observed. Incorporation of

additional water to this system accentuates these features. The increased network

connectivity with enhanced CA or water content is consistent with the higher G’

observed in rheological experiments.

To explain the phenomenon of increase in G′, we must revisit our earlier work

where we established conditions needed for sol-gel formation.24 In the previous study,

we found that at low water contents, the system forms a clear solution. We envisage that

at low water contents, water does not possess the capacity to break a lot of the

interactive bonds formed between CA and DMAc. Gelation occurs when critical water

content is attained dependent on CA concentration, beyond which competitive

interaction for DMAc between water and CA becomes more or less balanced. Thus

some of the bonds formed between CA and DMAc are broken off and a lot more DMAc

become accessible to water. CA therefore exhibits an increase in intramolecular

interaction between different chains whiles “holding” onto as much DMAc as possible.

Page 140: Rheology and Microstructure of Cellulose Acetate - Digital Repository

125

It is speculated that the system involves CA and water forming strong bonds with

DMAc, with partial weak links also formed between CA and water. This leads to

aggregates of CA spanning uniformly through the entire system. In view of the

explanation provided, the trend in Figures 1a & b may be due to water content increase

promoting the intramoleculer interaction between CA units forming stronger aggregates

leading to the formation of a more rigid gel network. Similarly, increase in CA

concentration may lead to significant intermolecular interactions being formed between

CA, DMAc and water promoting aggregation of CA, which causes stronger gel

formation. Tactile and visual observations respectively showed gels exhibiting harder

texture and more whitish appearance with increase in CA concentration and water

content.

Yield Behavior

Several approaches have been described in the literature on measuring the “yield

stress” of complex materials, and references and brief descriptions of some these

approaches can be found in a recent study from our group.42 In this work we undertake

one such approach for measuring yield stress to examine relative magnitude and trends

of different systems. Gels having different CA concentrations and water contents were

subjected to increasing oscillatory shear stress values to determine the yield behavior at

room temperature. An example is displayed in Figure 4a where stress sweep are

conducted for different CA concentrations at a constant water content of 21.6%. Figure

4b shows similar experiments performed at constant CA concentration of 15wt% with

varying water contents of 20.4 and 23.8%. Both G′ and G″ at low stress are relatively

Page 141: Rheology and Microstructure of Cellulose Acetate - Digital Repository

126

independent of applied stress, which indicates materials are in the linear viscoelastic

regime. However, with increasing stress, there is a breakdown of the interactive bonds

in the materials and we see a gradual decrease in G′. Beyond a critical stress value (τc),

the structure ruptures and/or yields leading to a sharp drop in G′. To obtain the yield

stress, two asymptotic lines are drawn through the initial and post-breakdown G′ values

(as shown in Figure 4a and 4b). Interestingly, G″ shows a gradual increase and a

maximum at the onset where G′ begins to decline. This process may be caused by

viscous properties being enhanced as the bonds get broken with gradual stress elevation.

Since essentially the gel formation is attributed to intramolecular interaction between

CA units, with additional links formed between DMAc and water. At the initial stages

of gel rupture, some of the bonds between the CA units may be broken leading to “free”

CA units being formed at some sections in the gel. Thus, there is an introduction of

some viscous components, which may account for the initial increase in G″. Eventually

after G′ and G″ crossover and the critical stress is exceeded, we observe a sharp decline

in both dynamic properties after which the material completely breaks down.

For all the gels studied, we see the nonlinear and yield behavior shift to higher

values with increasing CA concentration and/or water content (Figure 4a and 4b).

Figure 5 shows the effect of these two variables on yield stress more clearly. We find

that the yield stress increases substantially, and in some cases by an order of magnitude,

when water content or CA concentration is increased. The shift in the onset of non-

linearity and yield stress to higher values can be attributed to the formation of stronger

network as CA concentration or water content is increased, thereby requiring a greater

stress to disrupt the microstructure.

Page 142: Rheology and Microstructure of Cellulose Acetate - Digital Repository

127

Effects of Temperature

Since our system forms physical gels, the networks junctions should be

disrupted by the application of thermal energy.5,8 In this regard, we examined the

transition of our gels to liquids by subjecting them to temperature increments from

room temperature upwards. Figure 6 is a digital image showing the transition of a gel to

solution at temperatures of 25 and 80oC for 10wt% CA concentration containing 21.6%

water fraction. We see a transition and a change in the gel characteristics from a

completely opaque at 25oC to a system that shows an almost clear solution at 80oC. This

pictorial transition correlates well with rheological experiments, which showed gel-sol

properties at approximate conditions. Similar images were observed for varying water

content and CA concentrations. The dependence of G′ and G″ on temperature43 provides

good indications of the transition from solution to gel and vice versa. As such, we

conducted two types of experiments to examine the gel-sol transition: temperature ramp

and isothermal dynamic frequency experiments at different temperatures. Figure 7a

shows temperature ramp experiments conducted at 1oC/min for CA concentrations of

10wt% at two different water contents of 19.8 and 21.6%. The first region was

characterized by a relatively low dependence on temperature for both G′ and G″,

showing viscoelastic behavior of a gel network. However, beyond 40oC, there is a rapid

decrease in G′ and G″ until a crossover point at about 73oC and 81oC for 19.8 and

21.6% water compositions respectively is observed. Following the crossover, G″

exceeds G′. We term this crossover point the gel-sol transition temperature. The trend

for a 15wt.% system follow a similar pattern with the only difference being the

Page 143: Rheology and Microstructure of Cellulose Acetate - Digital Repository

128

crossover of G′ and G″ which occur at higher temperatures of about 79oC and 90oC for

19.8 and 21.6% water compositions respectively. Obviously, one could argue that a

better way to obtain this transition point would have been to use the Winter-Chambon

approach 22,37,44-45 In fact, we attempted to obtain the gel-sol temperature by plotting the

loss tangent (tan δ) measured at various frequencies as a function of temperature.22,37,45

But the procedure was aborted because there were no precise convergence at a set

temperature. The inability of the Winter Chambon criteria to hold has been observed for

other systems as well and can be attributed to absence of self-similarity, heterogeneity.46

To ascertain the validity of the temperature ramp experiments, dynamic

frequency experiments over a wide range of temperatures were performed. Figure 8a

exhibits the dynamic frequency spectrum of G′ and G″ for a 10wt% CA sample

concentration at different temperatures. For clarity, we plot only three temperatures to

show the transition from gel to solution. At room temperature the system shows

relatively frequency independent dynamic properties. As the temperature is increased to

60oC, both G′ and G″ drop substantially. However, G′ is still larger than G″ and both

moduli are frequency independent at low frequencies, suggesting that a gel network still

persists. As the temperature is elevated to 80oC, both moduli are frequency dependent

and there is a crossover between G″ and G’. These features indicate the system is

behaving solution-like. A similar plot is depicted in Figure 8b for CA concentration of

15wt%. The sequence of gel-sol transition is similar to that Figures 7a & b with the

difference being the transition to a solution-type behavior, which is observed above

85oC for the higher CA concentration system. To check how our gel-sol transition

temperature obtained from temperature ramp experiments compare with results

Page 144: Rheology and Microstructure of Cellulose Acetate - Digital Repository

129

obtained from the frequency spectrum at different temperature, we plot in Figure 9

G’and G’’ (at ω=1 of Figure 8) as a function of temperature. We observe a decrease in

both dynamic moduli until approximately 77oC when G" crosses over and exceeds G’.

This transition temperature, which is slightly lower than 82oC obtained from the

temperature ramp experiment (Figure 8a), is expected as the sample is equilibrated at

each temperature in this case. In the temperature ramp experiment, the sample does not

have sufficient time to equilibrate at each temperature. This pushes the transition to

manifest itself at a higher temperature. This trend is observed in Figure 9b for 15wt.%

sample where the crossover of G’ and G” occur 85oC which is lower in comparison to

temperature ramp experiment which had the G” crossing G’ at 90oC. In any event, the

results are quite close and depict the same trend.

A semi-logarithmic plot of G′ versus reciprocal temperature for a 10 and 15wt.%

system containing 21.6% and 20.4% water respectively is shown in Figure 10. From the

Arrhenius plot, the apparent activation energy calculated from the slope is found to be

approximately 113 kJmol-1 for 10wt.% sample, while the apparent activation energy is

approximately 92 kJmol-1 for the 15wt.% system. Thus, using the apparent activation

energy and the pre-exponential factor, G′ at different temperatures can be estimated.

Similar Arrhenius plots can be used to estimate G′ at different CA concentration and

water contents.

Page 145: Rheology and Microstructure of Cellulose Acetate - Digital Repository

130

Conclusions

This study presents a detailed examination of the effects of CA concentration

and water on the rheological behavior of CA gels in mixed solvents of water and

DMAc. Linear, and non-linear rheological behavior together with direct observation of

microstructure through confocal microscopy are illustrated. In this study, gel formation

induced by addition of relatively high water content 19% and above in CA/DMAc

solutions have been described using rheological methods. The elastic (G’) and viscous

moduli (G”) have been found to increase with increasing water content and CA

concentration. This increase in elastic character has been attributed to CA units

undergoing a progressively increase in intramolecular interactions whiles forming

relatively strong and weak intermolecular interactions with DMAc and water

respectively. The yield stress followed a similar pattern with gels exhibiting greater

yield stresses with water content and CA concentration increase, which can be

attributed to similar reasoning as stated earlier. Confocal images showed a system

exhibiting a more closely knit structure with water content and CA concentration

increase. The increased homogenous closely packed network with enhanced CA or

water content is consistent with higher dynamic viscoelastic properties observed in

rheological experiments.

The gel-sol transition temperature (Tgs) for all gels were between 70oC to 90oC,

with Tgs showing a progressive increase with CA concentration at comparable water

content. The Tgs has been found to also shift to high values with increasing water

content at constant CA concentration. This trend emphasizes the earlier explanation of

stronger gel network formed with increase in water content and CA concentration, thus

Page 146: Rheology and Microstructure of Cellulose Acetate - Digital Repository

131

increasing the Tgs. The results show that, the gel strength and for that matter the

rheology of the gels are highly dependent on water content and CA concentration and

both properties play important roles in determining how gels respond to stress and

temperature variation.

Acknowledgement

The authors gratefully acknowledge funding support for this work from United

States Department of Agriculture (USDA).

Page 147: Rheology and Microstructure of Cellulose Acetate - Digital Repository

132

References

(1) Timofeyeva, G.N., Aver’yanova, V.M., Influence of the nature of the solvents

on the processes of structuring in solutions of cellulose acetates. Vysokomol.

Soyed. 1982, A24, 2268-2274.

(2) Kobayashi, K., Huang, C., Lodge, T.P., Thermoreversible gelation of aqueous

methylcellulose solutions, Macromolecules 1999, 32, 7070-7077.

(3) Sarkar, N., Thermal gelation properties of methyl and hydroxypropylmethyl

cellulose. Journal of Applied Polymer Science 1979, 24, 1073-1087.

(4) Sarkar, N., Kinetics of thermal gelation of methyl cellulose and hydroxypropyl

cellulose in aqueous solutions. Carbohydrate Polymers 1995, 26, 195-203.

(5) Takahashi, M., Shimazaki, M., Yamamoto, J., Thermoreversible gelation and

phase separation in aqueous methyl cellulose solutions. Journal of Polymer

Science: Part B: Polymer Physics, 2001, 39, 91-100.

(6) Li, L., Thangamathesvaran, P.M., Yue, C.Y., Tam, K.C., Hu, X., Lam, Y.C., Gel

network structure of methylcellulose in water. Langmuir 2001, 17, 8062-8068.

(7) Desbrieres, J., Hirrien, M., Ross-Murphy, S.B., Thermogelation of methyl

cellulose: rheological considerations. Polymer 2000, 41, 2451-2461.

(8) Takahashi, M., Shimazaki, M., Formation of junction zones in thermoreversible

methyl cellulose gels. Journal of Polymer Science: Part B: Polymer Physics,

2001, 39, 943-946.

(9) Bochek, A.M., Zabivalova, N.M., Lavrent’ev, V.K., Lebedeva, M.F.,

Sukhanova, T.E., Petropavlovskii, G.A., Formation of physical thermal

reversible gels in solutions of methyl cellulose in water and dimethylacetamide

Page 148: Rheology and Microstructure of Cellulose Acetate - Digital Repository

133

and properties of films thereof. Russian Journal of Applied Chemistry 1978, 74,

1358-1363.

(10) Kalinina, N.A., Bochek, A.M., Silinskaya, I.G., Nud’ga, L.A., Petrova, V.A.,

Structure of aqueous methyl and propyl methyl cellulose solutions in the initial

stages of gelation. Russian Journal of Applied Chemistry 2001, 9, 1568-1572.

(11) MeiBner, D., Einfrldt, J., Schareina, T., New results concerning the behavior of

cellulose acetate in solutions and films by means of CD measurements.

Biopolymers 1999, 50, 163-166.

(12) Kawanishi, H., Tsunashima, Y., Okada, S., Horii, F., Change in chain stiffness

in viscometric and ultracentrifugal fields: cellulose diacetate in N,N-

dimethylacetamide dilute solution. Journal of Chemical Physics 1998, 108 (14),

6014-6025.

(13) Kawanishi, H., Tsunashima, Y., Horii, F., Dynamic light scattering study of

structural changes of cellulose diacetate in solution under couette flow.

Macromolecules 2000, 33, 2092-2097.

(14) Tsunashima, Y., Kawanishi, H., Nomura, R., Horii, F., Influence of long-range

interactions in diffusion behavior in semidilute solution: Dynamics of cellulose

diacetate in quiescent state. Macromolecules 1999, 32, 5330-5336.

(15) Gomez-Bujedo, S., Fleury, E., Vignon, M.R., Preparation of cellouronic acids

and partially acetylated cellournic acids by TEMPO/NaClO oxidation of water-

soluble cellulose acetate, Biomacromolecules 2004, 5, 565-571.

(16) Bochek, A.M., Kalyuzhnaya, L.M., Interaction of water with cellulose and

cellulose acetates as influenced by the hydrogen bond system and hydrophilic-

Page 149: Rheology and Microstructure of Cellulose Acetate - Digital Repository

134

hydrophobic balance of the macromolecules. Macromolecular Chemistry and

Polymeric Materials 2002, 75, 989-993.

(17) Pintaric, B., Rogosic, M., Mencer, H.J., Dilute solution properties of cellulose

diacetate in mixed solvents. Journal of molecular liquids 2000, 85, 331-350.

(18) Ryskina, I.I., Aver’yanova, V.M., A study of gelation and the structure of

acetylcellulose gels in benzyl alcohol. Vysokomol. Soyed. 1971, A13, 2189-

2194.

(19) Ryskina, I.I., Aver’yanova, V.M., Morphological study of the three dimensional

network of cellulose acetate-benzyl alcohol gelatins. Vysokomol. Soyed. 1975,

A17, 919-922.

(20) Glikman, S.A., Aver’yanova, V.M., Khomutov, L.I., Structure of

acethylcellulose solutions. Vysokomol. Soyed. 1963, 5, 598-604.

(21) Panina, N.I., Paneikina, T.P., Aver’yanova, V.M., Zelenev, Y.V., The

mechanical properties of gels in dimethyl phthalate of cellulose acetate of

different degress of substitution. Vysokomol. Soyed. 1978, A20, 1706-1711.

(22) Nystrom, B., Walderhaug, H., Hansen, F.K., Rheological behavior during

thermoreversible gelation of aqueous mixtures of ethyl(hydroxyethyl) cellulose

and surfactants. Langmuir 1995, 11, 750-757.

(23) Nunes, S.P., Wolf, B.A., On the cooccurence of demixing and thermoreversible

gelation of polymer solutions. 3. overall view. Macromolecules 1987, 20, 1952-

1957.

(24) Appaw, C., Gilbert, R.D., Kadla, J.F., Khan, S.A., Sol-gel transition of cellulose

acetate in a mixed solvent system. (unpublished results) 2004.

Page 150: Rheology and Microstructure of Cellulose Acetate - Digital Repository

135

(25) Reuvers, A.J., Altena, F.W., Smolders, C.A., Demixing and gelation behavior of

ternary cellulose acetate solutions Jour. of Polymer Sci.: Part B: Polymer

Physics 1986, 24, 793-804.

(26) Altena, F.W., Smolders, C.A., Phase separation phenomena in solutions of

cellulose acetate. I. Differential scanning calorimetry of cellulose acetate in

mixtures of dioxane and water. Journal of Polymer Science: Polymer

Symposium 1981, 69, 1-10.

(27) Ryo, Y., Nakai, Y., Kawaguchi, M., Viscoelastic measurements of silica

suspensions in aqueous cellulose derivative solutions. Langmuir 1992, 8, 2413-

2416.

(28) Li, L., Thermal gelation of methylcellulose in water: scaling and

thermoreversibility. Macromolecules 2002, 35, 5990-5998.

(29) Carriere, C.J., Inglett, G.E., Constitutive analysis of the nonlinear viscoelastic

properties of cellulosic fiber gels produced from corn or oat hulls. Food

Hydrocolloids 2003, 17, 605-614.

(30) Pai, V.B., Khan, S.A., Gelation and rheology of xanthan/enzyme-modified guar

blends. Carbohydrate Polymers 2002, 49, 207-216.

(31) Eissa, A.S., Bisram, S., Khan, S.A., Polymerization and gelation of whey protein

isolates at low pH using transglutaminase enzyme. Journal of Agricultural and

Food Chemistry 2004, 52, 4456-4464.

(32) Chauvelon, G., Doublier, J-L., Buleon, A., Thibault, J-F., Saulnier, L.,

Rheological properties of sulfoacetate derivatives of cellulose, Carbohydrate

Research 2003, 338, 751-759.

Page 151: Rheology and Microstructure of Cellulose Acetate - Digital Repository

136

(33) Lazaridou, A., Biliaderis, C.G., Izydorczyk, M.S., Molecular size effects on

rheological properties of oat glucans in solution and gels. Food Hydrocolloids

2003,17, 693-712.

(34) Hirsch, S.G.; Spontak, R.J. Temperature-dependent property development in

hydrogels derived from hydroxylpropylcellulose. Polymer 2002, 43, 123-129.

(35) Charcosset, C.; Cherfi, A.; Bernengo, J-C. Characterization of microporous

membrane morphology using confocal scanning laser microscopy. Chemical

Engineering Science 2000, 55, 5351-5358.

(36) Charcosset, C.; Bernengo, J-C. Characterization of microporous membranes

morphology using confocal scanning laser microscopy in fluorescence mode.

European physical Journal Applied Physics 2000, 12, 195-199.

(37) Macosko, C.W. Rheology Principles, Measurements and Applications. Wiley-

VCH Publishers, Inc. , New York 1994.

(38) Pai, V., Srinivasarao, M., Khan, S. A. Evolution of Microstructure and Rheology

in Mixed Polysaccharide Systems. Macromolecules 2002, 35, 1699-1707.

(39) Wilder, E. A., Hall, C. K., Khan, S. A., Spontak, R. J. Effects of Composition

and Matrix Polarity on Network Development in Organogels of Poly(ethylene

glycol) and Dibenzylidene Sorbitol. Langmuir 2003, 19(15), 6004-6013.

(40) Raghavan, S. R., Walls, H. J., Khan, S. A. Rheology of Silica Dispersions in

Organic Liquids: New Evidence for Solvation Forces Dictated by Hydrogen

Bonding. Langmuir 2000, 16(21), 7920-7930.

Page 152: Rheology and Microstructure of Cellulose Acetate - Digital Repository

137

(41) Egermayer, M., Norman, J., Piculell, L., Gels of hydrophobically modified

hydroxyl cellulose cross-linked by amylose: competition by added surfactants.

Langmuir 2003, 19, 10036-10043.

(42) Walls, H.J., Caines, S.B., Sanchez, A.M., Khan, S.A. Yield stress and wall slip

phenomena in colloidal silica gels. Journal of Rheology 47, 2003, 847-868.

(43) Weng, L., Zhang, L., Ruan, D., Shi, L., Xu, J., Thermal gelation of cellulose in a

NaOH/Thiourea aqueous solution. Langmuir 2004, 20, 2086-2093.

(44) Shay, S., Raghavan, S.R., Khan, S.A. Thermoreversible Gelation in Aqueous

Dispersions of Colloidal Particles Bearing Grafted Poly(ethylene oxide) Chains.

Journal of Rheology 45, 2001, 913-928.

(45) Egermayer, M., Karlberg, M., Piculell, L., Gels of hydrophobically modified

ethyl(hydroxyethyl) cellulose cross-linked by amylose: effects of hydrophobe

architecture. Langmuir 2004, 20, 2203-2214.

(46) Chiou, B-S., Raghavan, S.R., Khan, S.A. Effect of colloidal fillers on the cross-

linking of a uv-curable polymer. Gel point rheology and the winter-chambon

criterion. Macromolecules 2000, 34, 4526-4533.

Page 153: Rheology and Microstructure of Cellulose Acetate - Digital Repository

138

Figure Caption

Figure 1. Dynamic frequency sweep experiments for (a) CA concentration of

10wt% (b) CA concentration of 15wt% having different water contents.

Figure 2. G′ as a function of water content for CA concentration of 10 and 15wt.%

Data were obtained at frequency of 1 rad/s.

Figure 3. CSLM images for CA concentration of (a) 10wt% and (b) 15wt% at two

different water contents.

Figure 4. Stress sweep experiments conducted for (a) different CA concentrations

at constant water content of 21.6wt.% (b) constant CA concentration of 15wt%

with varying water contents.

Figure 5. Yield stress as a function of water content for varying CA concentration.

Figure 6. Digital image showing transition of a gel to solution at two different

temperatures for 10wt% CA concentration with water content of 21.6%.

Figure 7. Temperature ramp experiments conducted at 1oC/min (a) for 10wt%

CA (b) 15wt% CA concentration at different water contents.

Page 154: Rheology and Microstructure of Cellulose Acetate - Digital Repository

139

Figure 8. Dynamic frequency sweep experiment for (a) 10wt.% CA and

(b) 15wt.% CA concentration at different temperatures.

Figure 9. Dependence of G' and G" on temperature for (a) 10wt% CA system having

water content of 21.6wt.% (b) 15wt.% CA system having water content of 20.4wt.%.

Data obtained at a frequency of 1 rad/s.

Figure 10. Arrhenius plot for (a) 10wt% CA concentration having water content of

21.6% (b) 15wt% CA concentration having water content of 20.4%.

Page 155: Rheology and Microstructure of Cellulose Acetate - Digital Repository

140

G' &

G" (

Pa)

102

103

104

105

106

G" (19.8%)

G' (19.8%)

G' (21.6%)

G' (23.4%)

G' (25.2%)

Frequency (rad/s)10-2 10-1 100 101 102

G' &

G" (

Pa)

103

104

105

106

G" (18.7%)

G' (18.7%)

G' (20.4%)

G' (22.1%)

G' (23.8%)

Figure 1. Dynamic frequency sweep experiments for (a) CA concentration of

10wt% (b) CA concentration of 15wt% having different water contents.

Page 156: Rheology and Microstructure of Cellulose Acetate - Digital Repository

141

Water content (%)18 20 22 24 26

G' (

Pa)

103

104

105

106

15wt.%

10wt.%

Figure 2. G′ as a function of water content for CA concentration of 10

and15wt.%. Data were obtained at frequency of 1 rad/s.

Page 157: Rheology and Microstructure of Cellulose Acetate - Digital Repository

142

(a) 21.6%g 25.2%

(b) 20.4% 23.8%

Figure 3. CSLM images for CA concentration of (a) 10wt% and (b) 15wt% at two

different water contents.

Page 158: Rheology and Microstructure of Cellulose Acetate - Digital Repository

143

Stress (Pa)100 101 102 103 104 105

G' &

G" (

Pa)

103

104

105

G' (15%)

G' (12.5%)

G' (10%)

G" (15%)

G" (12.5%)

G" (10%)

Stress (Pa)100 101 102 103 104 105

G' &

G" (

Pa)

103

104

105

106

G' (23.8%)

G' (20.4%)

G" (23.8%)

G" (20.4%)

Figure 4. Stress sweep experiments conducted for (a) different CA concentrations

at constant water content of 21.6wt.% (b) constant CA concentration of 15wt%

with varying water contents.

Page 159: Rheology and Microstructure of Cellulose Acetate - Digital Repository

144

Water content (%)16 18 20 22 24 26 28

Yie

ld S

tress

(Pa)

101

102

103

104

7.5%

10%12.5%

15%

Figure 5. Yield stress as a function of water content for varying

CA concentration.

Page 160: Rheology and Microstructure of Cellulose Acetate - Digital Repository

145

25oC 80oC

Figure 6. Digital image showing transition of a gel to solution at two different

temperatures for 10wt% CA concentration with water content of 21.6%.

Page 161: Rheology and Microstructure of Cellulose Acetate - Digital Repository

146

G' &

G" (

Pa)

10-1

100

101

102

103

104

105

G' (21.6%)

G" (21.6%)

G' (19.8%)

G" (19.8%)

Temperature (oC)

0 20 40 60 80 100 120 140

G' &

G" (

Pa)

10-1

100

101

102

103

104

105

G' (20.4%)

G" (204%)

G' (18.7%)

G" (18.7%)

Figure 7. Temperature ramp experiments conducted at 1oC/min (a) for 10wt%

CA (b) 15wt% CA concentration at different water contents.

Page 162: Rheology and Microstructure of Cellulose Acetate - Digital Repository

147

G' &

G" (

Pa)

100

101

102

103

104

105

G' (25c)

G" (25C)

G' (60C)

G" (60C)G" (80C)

G' (80C)

Frequency (rad/s)10-2 10-1 100 101 102

G' &

G" (

Pa)

100

101

102

103

104

105

106

G' (25C)

G" (25C)

G' (60C)

G" (60C)G" (90C)

G' (90C)

Figure 8. Dynamic frequency sweep experiment for (a) 10wt.% CA and

(b) 15wt.% CA concentration at different temperatures.

Page 163: Rheology and Microstructure of Cellulose Acetate - Digital Repository

148

20 30 40 50 60 70 80 90

G' &

G" (

Pa)

100

101

102

103

104

105

G'

G"

Temperature (oC)

20 30 40 50 60 70 80 90 100

G' &

G" (

Pa)

101

102

103

104

105

G'

G"

Figure 9. Dependence of G' and G" on temperature for (a) 10wt% CA system having

water content of 21.6wt.% (b) 15wt.% CA system having water content of 20.4wt.%.

Data obtained at a frequency of 1 rad/s.

Page 164: Rheology and Microstructure of Cellulose Acetate - Digital Repository

149

1/T (K-1)

2.7x10-32.8x10-32.9x10-33.0x10-33.1x10-33.2x10-33.3x10-33.4x10-3

ln G

'

0

2

4

6

8

10

12

y = 13546x - 35.612R2 = 0.9878

1/T (1/K)2.7x10-32.8x10-32.9x10-33.0x10-33.1x10-33.2x10-33.3x10-33.4x10-

ln G

'

4

5

6

7

8

9

10

11

12

y = 11080x - 25.925R2 = 0.9915

Figure 10. Arrhenius plot for (a) 10wt% CA concentration having water content of

21.6% (b) 15wt% CA concentration having water content of 20.4%.

Page 165: Rheology and Microstructure of Cellulose Acetate - Digital Repository

150

CHAPTER 7

CONCLUSIONS AND RECOMMENDATIONS

Page 166: Rheology and Microstructure of Cellulose Acetate - Digital Repository

151

7.1 Conclusions

In this thesis study, we have investigated the various protocols pertaining to the

effect of varying water content (non-solvent) on solutions of cellulose acetate (CA) and

N,N dimethylacetamide (solvent). The initial challenge in this project was the selection

of the appropriate solvent to dissolve CA. We chose N,N dimethylacetamide because of

its high boiling point, low toxicity and the ability to dissolve CA to appreciably high

concentration levels, since we were interested in studying semi- to concentrated CA

solutions. The second phase involved the selection of a non-solvent which can interact

in solution to generate phase-separated gel systems. In this regard, water was then

chosen as the non-solvent due its strong hydrogen bonding characteristics and the

obvious wide spread availability. We have shown in our investigations that solvent

quality, most importantly water composition, plays a major role in influencing polymer

rheology and microstructure. Our work has explored the changes in solutions and

eventual transition to gel formation by increasing water content in our system. In view

of these findings, some of the important conclusions that can be inferred from this

dissertation are summarized in the following paragraphs.

• Both CA concentration and water content have been shown to have profound

effects on solution viscosity. Increasing water content at constant CA

concentration and vice versa resulted in enhanced steady state viscosity. In the

same vein, increment in CA concentration depicted a similar trend of viscosity

increase. This phenomenon is attributed to the enhanced intermolecular

interactions between CA and cosolvents. Further explanation can be due to the

Page 167: Rheology and Microstructure of Cellulose Acetate - Digital Repository

152

fact that additional hydrogen bond linkages in solution are formed causing some

entanglements in solution, thus, causing the solution to become more viscous

with increase in water content or CA concentration.

• The process of dissolving CA in the cosolvents had a strong influence on the

mode of phase separation leading to gelation. Addition of water to CA/N,N

dimethylacetamide (dissolution method 1) solutions led to the formation of

cloudy samples at intermediate water content depending on CA concentration.

Formation of gels with uniform structures was observed at high water contents

above the intermediate water composition. The gels became more elastic with

increase in water composition and CA concentration depicted by enhanced

dynamic viscoelastic properties. In contrast, addition of CA to different ratios

but constant weight of water/N,N dimethylacetamide (dissolution method 2)

showed a two-phase system depicting a low viscosity liquid layer on top of a gel

layer at identical high water contents in comparison to dissolution method 1.

The hypothesis speculates the two different phase separated systems to occur

mainly due to competitive hydrogen bond interactions in the system. It is

postulated that, water competes with CA for N,N dimethylacetamide in solution

and at high water contents, different mode of segregation of CA lead to different

phase separated systems.

• Heating the two-phase system in dissolution method 2 above a certain

temperature resulted in the formation of a uniform gel network. It is

Page 168: Rheology and Microstructure of Cellulose Acetate - Digital Repository

153

hypothesized that the elevation in temperature breaks some of the N,N

dimethylacetamide-water hydrogen bonds enabling unbonded CA units to obtain

access to “free” N,N dimethylacetamide. Thus, when the system is cooled to

room temperature, water has to compete for N,N dimethylacetamide with CA

causing a gel system to be formed due to dispersion of CA through the entire

volume of the matrix.

• Gel-sol transition which was influenced by water and CA composition was

observed in the gel. Gels with high water content and/or CA concentration had

the transition occurring at much higher temperatures. The yield behavior also

portrayed a similar pattern where greater stress/strain was needed to disrupt the

microstructure of the network. This is mainly due to the stronger gel network

formed at such conditions attributed to greater level of hydrogen bond

interactions. These stronger gels invariably will need an increase in external

factors to break (stress application) or revert them solutions (temperature

elevation) as was observed. This hypothesis was supported by examining

different water contents and CA concentrations.

• Gel microstructure showed a more homogenous network with increasing water

content and CA concentration. Both laser scanning confocal microscopy

(LSCM) and scanning electron microscopy (SEM) depicted this trend. This was

consistent with rheological studies and confirmed that gel strength was enhanced

when water or CA concentration was increased.

Page 169: Rheology and Microstructure of Cellulose Acetate - Digital Repository

154

7.2 Recommendations

This research has examined the various aspects of changing CA and cosolvent

compositions in ternary systems of CA, N,N dimethylacetamide and water. We have

thus employed a variety of techniques to understand the rheological and microstructural

properties. However, there are various areas of research which need to be explored to

fully understand the molecular level interactions accounting for solution and gel

characteristics observed.

Differential scanning calorimetric study of gels

Our main characterization technique was rheology. This was supplemented with

scanning electron microscopy (SEM) and confocal laser scanning microscopy (CLSM)

to correlate microstructure with rheology. Differential scanning calorimetry (DSC) is

recommended as an additional tool to determine the thermal properties of the gel

formed using a heating and cooling process.1,2 The shift and peaks of the relative

thermal heat capacity will provide information on gelation process3,4 in relation to water

content and concentration of CA.

Addition of other non-solvents

Since a variety of solvents interact differently with acetyl and hydroxyl groups

on the CA chains5, addition of other non-solvents to CA-N,N dimethylacetamide

solutions in place of water is recommended. We observed in our work enhanced steady

state viscosity and a sol-gel system transition which primarily depended on water

content in the system. Therefore, a class of solvents that can be explored as an initial

Page 170: Rheology and Microstructure of Cellulose Acetate - Digital Repository

155

step in any future work, are organic alcohols. Essentially, alcohols and water share a

common characteristic in that both have a hydroxyl group attached to side group. Even

though water exhibits much stronger hydrogen bond association in solution, the addition

of alcohols may provide additional information by showing a fundamental difference in

solution properties. Because of the variety of alcohols available, it is envisioned that the

influence of the side chain attached to the hydroxyl group (e.g. methyl, ethyl etc) will

also allow investigators determine how they impact the system. In addition, solvents

such the diols which are compounds that contain two hydroxyl groups can also be

utilized. All the solvents recommended exhibit to some extent hydrogen bonding

characteristics and based on the strength of the hydrogen bonds, we can evaluate how

CA-N,N dimethylacetamide bonds are affected in solution. The effect of alcohols as

non-solvents have shown different systems in relation to microstructure being formed in

a variety of CA solvent system.6

Another class of solvents that will also provide some useful information is

solvents that practically exhibit no hydrogen bonding character in solution. These would

include hexane and octane among others.

Effect of salts

The influence of salt on the suppressing or enhancing polymer solution, phase

separation7 and gel properties is an important future direction, and preferably will shed

light and expand our knowledge of the hydrogen bond interactions between CA and the

cosolvents, especially water in the system. Progressive increase in the valence number

of the cation in a particular class of salt will affect the ordering of salt around the water

Page 171: Rheology and Microstructure of Cellulose Acetate - Digital Repository

156

molecules. This will invariably affect the intermolecular hydrogen bonds already

formed, and based on salt-type and concentration, phase separation and gelation can be

induced to occur at low or high non-solvent content. Previous studies on CA acetonitrile

solutions containing different salts have shown viscosity increase to depend on the

valence number of the cation.8

To aid or suppress gelation, another approach would be to classify the salt-type

into “salting-in” and “salting-out” since the mechanism of interaction in solution is

different. The use of NaCl and NaI, typical salting-out and salting in salts respectively3-

4, will enable us to determine the extent to which gelation is affected. For instance,

NaCl exhibit strong interactions with water molecules3 and may suppress or aid gel

formation.

Effect of pH

The modification of pH in solution is an interesting feature which can affect the

strength of bonds in the system influencing solution and gel properties. The initial step

can be the change in pH of water prior to its addition to the polymer-solvent system.

Therefore, increasing or decreasing the pH can allow us to develop a pattern of altering

system characteristics.

Page 172: Rheology and Microstructure of Cellulose Acetate - Digital Repository

157

References

1. Altena, F.W., Smolders, C.A., Phase separation phenomena in solutions of

cellulose acetate. I. Differential scanning calorimetry of cellulose acetate in

mixtures of dioxane and water Jour. of Polymer Science: Polymer Symposium

1981, 69, 1-10.

2. McCrystal, C.B., Ford, J.L., Rajabi-Siahboomi, A.R. A study on the interaction

of water and cellulose ethers using differential scanning calorimetry.

Thermochimica Acta 1997, 294, 91-98.

3. Xu, Y., Wang, C., Tam, K.C., Li, L. Salt-assisted and salt-suppressed sol-gel

transition of nethylcellulose in water. Langmuir 2004, 20, 646-671.

4. Xu, Y., Lin, L., Zheng, P., Lam, Y.C., Hu, X. Contrallable gelation of

methylcellulose by a salt mixture. Langmuir 2004, 20, 6134-6138.

5. Pintaric, B.; Rogosic, M.; Mencer, H.J. Dilute solution properties of cellulose

diacetate in mixed solvents. Journal of molecular liquids 2000, 85, 331-350.

6. Tweddle, T.A., Sourirajan, S. Effect of Ethanol-Water Mixture as Gelation

Medium During Formation of Cellulose Acetate Reverse Osmosis Membranes.

Journal of Applied Polymer Science 1978, 22, 2265-2274.

7. Robb, I.D. Polymer-small molecule interactions. Textile and Fiber Engineering

1998, 994, 8-13.

8. Tager, A.A., Lirova, B.I., Vasyanina, N.S. Ionic-molecular reactions and the

viscosity of cellulose acetate solutions in acetonitrile. Polymer Science U.S.S.R

1977, 10(11), 2885-2893.

Page 173: Rheology and Microstructure of Cellulose Acetate - Digital Repository

158

APPENDIX A

CELLULOSE GELS AND LIQUID CRYSTALLINE POLYMER (LCP)

Appendix is essentially a manuscript by John F. Kadla, Qizhou Dai, Collins Appaw,

and Saad A. Khan, prepared for submission to Encyclopedia of Chemical Processing

(EHCP).

Page 174: Rheology and Microstructure of Cellulose Acetate - Digital Repository

159

CELLULOSE GELS AND LIQUID CRYSTALLINE POLYMER (LCP)

John F. Kadla1, Qizhou Dai1, Collins Appaw2 and Saad Khan2

1Advanced Biomaterials Chemistry, University of British Columbia,

Vancouver, BC V6T 1R9 CANADA

2Department of Chemical Engineering, North Carolina State University,

Raleigh, NC 27695 USA

KEYWORDS: Cellulose, cellulose derivatives, liquid crystal, chiroptical property,

rheology, relaxation, phase separation, gelation, hydrogen bond, sol-gel

ABSTRACT

INTRODUCTION

CELLULOSE

Cellulose constitutes the most abundant, renewable polymer resource. It has been

estimated [1] that the yearly photosynthesis of biomass is 170 billion tons, 40% of

which consists of polysaccharides; mainly cellulose and starch. Today, with the

availability of an enormous variety of synthetic polymers, cellulose and its derivatives

are somewhat overshadowed. Nevertheless, cellulose occupies a unique place in the

annals of high polymers. It was one of the first polymers studied starting with Anselm

Page 175: Rheology and Microstructure of Cellulose Acetate - Digital Repository

160

Payen’s investigations [2]. “Many of the basic principles were worked out in the course

of cellulose investigations” [3]. Payen first recognized cellulose as a definitive

substance and coined the name cellulose. Today it is still widely investigated. Synthesis

of cellulose derivatives and regeneration of cellulose, along with the physical chemistry

of cellulosic solutions, including those, which are mesomorphic, constitute areas of

active research. The use of cellulose and its derivatives in a diverse array of

applications, such as fibers, films, plastics, coatings, suspension agents, composites,

wood and paper products, continue to grow on a worldwide basis. Relatively newer uses

of cellulose and cellulose derivatives include cellulosic membranes for hemodialysis

and hemafiltration, [4] chiral isomer separations, [5] and calorie-free fat substitutes [6].

Cellulose is the main constituent of higher plants, including wood, cotton, flax,

kemp, jute, bagasse, ramie, cereal straws, etc. Industrially, the principle sources of

cellulose are wood, cotton fiber and cotton linters. Cellulose is also produced by a type

of acetic acid-producing bacterium, Acetobacter xylinum. [7] Algae such as Valonia,

Chladophora, Rhizoclonium and Microdictyon also produce highly crystalline cellulose

I. There are also several celluloses of animal origin, of which tunicin, a cell wall

component of ascidians, has been extensively studied [8].

Composition and Conformation

In 1838, Payen first proposed the elemental composition of cellulose to be

C6H10O5, [2] classifying it as a carbohydrate. A number of studies [9-12] have shown

cellulose is a chain-like extended linear macromolecule of β-D-glucopyranose units

linked by (1, 4) glycosidic bonds in which the anhydroglucopyranose units exist in the

Page 176: Rheology and Microstructure of Cellulose Acetate - Digital Repository

161

lowest energy 4C1-chair conformation. [13] It is generally accepted that in the cellobiose

repeat unit (being 10.3Å) each anhydroglucopyranose residue is displaced 180o with

respect to its neighbor. Thus cellulose has a two-fold screw axis. [14]

O

O

O

OH

OH

HOO

OH

OH 1

2

3

4

5

6HO

The rigidity of cellulose chain comes from the steric repulsion between atoms on

the adjacent units and intra-chain hydrogen bonding. For cellulose, the conformation of

the chain depends on the rotation of the two bonds linked to each glucosidic oxygen

atom. Results from conformational energy indicate that cellulose chain can have only

limited number of energetically favorable conformations and all stable and metastable

conformations are extended helices [15,16]. Three hydroxyl groups on each unit form a

large number of inter- and intra- chain hydrogen bonds. The intra-chain hydrogen

bonds, C(3)OH—O(5’) and C(6)OH—O(2’)OH, limit the rotations of adjacent rings

and contribute to the rigidity of the cellulose backbone. As a result of these strong inter-

and intra-chain hydrogen bonding, cellulose has poor solubility in common solvents.

CELLULOSE GELS

Solution to gel transitions has been observed in various cellulose derivatives.

Most cellulose derivative gel systems if not all, form physical gel networks. The process

Page 177: Rheology and Microstructure of Cellulose Acetate - Digital Repository

162

of physical gelation is not well understood in comparison to chemical gelation due to

the transient nature of gel network junctions [17]. Gelation in cellulosic systems in

some circumstances exhibits phase separation characteristics prior to gel network

formation. The phenomenon of phase separation depends to a large extent on the

concentrations of polymer and solvents used [18, 19].

Gelation properties in aqueous cellulosics

Some cellulosic systems, which are soluble in aqueous solutions exhibit unique

thermoreversible gelation by forming three-dimensional gel networks [20]. The system

when heated is transformed into gels and reversibly forms solutions when cooled [20-

27]. Various studies have been conducted particularly on methyl cellulose (MC) to

investigate the gelation mechanism. Various reasons have been proposed to explain this

phenomenon and particularly for MC and there exist some level of controversy detailing

the gelation mechanism [20]. It is reported that heating of MC aqueous solutions form

aggregates due to the hydrophobic regions being exposed when the “cage-like”

structures formed by the water molecules are broken [21]. Lizaso et al. [28] indicated

that thermoreverisble gels were formed in solutions of ethylcellulose in diester solvents.

Gelation in a novel azobenzene functionalized hydoxypropyl methylcellulose (AZO-

HPMC) have been suggested to be mainly caused by hydrophobic associations between

HPMC chains [29].

Cellulose acetate gel systems

Page 178: Rheology and Microstructure of Cellulose Acetate - Digital Repository

163

Cellulose acetate (CA) exhibits dual solubility characteristics depending on the

DS. CA a derivative of cellulose is characterized by partially or fully substituting the

hydroxyl (OH) groups on the C-2,3 and 6 positions with O-acetyl groups [30-31]. It has

been observed that cellulose acetates having DS of 0.5 to 1 are soluble in aqueous

solutions [33, 34]. Studies have shown that solubility in aqueous medium may depend

on the uniform distribution of the substituents and substitution of the OH groups at the

C3 position [34]. On the other hand, when the DS > 1, CA tend to be insoluble in

aqueous solutions but soluble in various other solvents. Boerstoel et al.[35] found that

over a DS range of 0.23 to 2.89, cellulose acetate was also soluble in phosphoric acid

exhibiting liquid crystalline properties. Pintaric et al. [36] mentioned in a study on

dilute solutions of cellulose diacetate (CDA) in mixed solvents, that basic solvents (e.g.

acetone) interact primarily with hydroxyl groups on the chain, whiles acidic solvents

(e.g. formic acid) primarily solvate the acetyl groups. Therefore, the use of specific

solvents can induce structural changes in solution due to the presence of both acetyl and

hydroxyl groups on partially substituted CA chains.

Unlike MC aqueous systems, low DS (0.5 to 1) CA aqueous solutions do not gel

when heated [34]. Gelation in some CA/organic solvent systems have been observed to

be mainly caused by strong intermolecular contacts [37]. CA containing different

amounts of acetic acid in benzyl alcohol has been shown to exhibit gel formation [37-

39]. The gelation temperature was discovered to depend on molecular weight and

quantity of combined acetic acid [37]. These gels are found to exhibit structural changes

dependent on solvent quality [37]. 5% solutions of CA in benzyl alcohol prepared at 60-

80oC formed gels when solutions are cooled to 25oC [39]. The gelation process can be

Page 179: Rheology and Microstructure of Cellulose Acetate - Digital Repository

164

induced by the addition of a non-solvent. Ternary systems consisting of

CA/solvent/water exhibits gelation, adopting solvents comprising of acetone, dioxane

and THF [19]. A study on CA/dioxane/water system was found to gel even though the

process was very slow [40]. Chauvelon et al. [41] reported weak gel-like systems being

formed in water-soluble cellulose acetate sulfate (CAS) at high concentrations (above 7-

8 g/l).

Rheology of cellulosic solutions and gels

Rheology has been used in quite a number of cellulosic solutions and gel

systems [21, 22, 24, 41, 42, 43]. Wang and Fried [42] postulated that specific

interactions such as hydrogen bonding between CA and solvents may affect rheological

properties in solution. Dynamic rheology provide information on viscoelastic property

changes and is the most direct method to study sol-gel transformation [44]. This

technique has used in most MC gel system to investigate and probe gel properties [21,

22, 24].

Figure 1. Effect of different water contents on the steady shear viscosity of cellulose

acetate in DMAc/water cosolvent system. The total concentration of cellulose acetate is

kept constant at 10wt%.

Steady state viscosity measurements for the study conducted by adding CA to

DMAc/water (dissolution method 1) fractions showed viscosity with increasing water

content at CA concentration of 10wt% (Figure 1). At low water content the system

Page 180: Rheology and Microstructure of Cellulose Acetate - Digital Repository

165

portray a Newtonian pattern with low shear thinning at high shear rates. But at

relatively high water contents, the viscosity profiles show shear thinning being

developed at low shear rates. It is postulate that the hydroxyl (OH) groups on the water

molecules interacting strongly with the acetyl groups on DMAc. Therefore, the addition

of CA to the solvent system, cause both water and to a much extent DMAc to form

complex hydrogen bond interaction with CA in the system leading to intermolecular

interaction intensification. Thus increasing the viscosity with water content increment.

The system phase separates forming a two-phase region at high water contents

(>19wt%). The upper layer depicts a clear solution, which exhibits Newtonian behavior

whiles the polymer-rich viscous bottom show a shear-thinning behavior over the whole

range of shear rate studied (Figure 2).

Figure 2. Viscosity pattern for the two-phase region for a constant 10wt% CA in

DMAc/water system

The phenomenon of the two-phase behavior at high water contents leads to the

idea of a competitive interaction for DMAc since both CA and water have affinity for

DMAc. It is perceived that water and DMAc form strong hydrogen bond links due to

the high composition of water. Therefore CA is hindered from gaining access to DMAc

and may be weakly bonded to DMAc and possibly some water molecules. This causes

CA to exhibit a greater level of intramolecular interactions, which leads to formation of

a viscous mixture at the bottom of system.

Page 181: Rheology and Microstructure of Cellulose Acetate - Digital Repository

166

Figure 3. Effect of different water contents on steady shear viscosity for a cellulose

acetate/DMAc/water system. CA concentration is 10wt%.

By utilizing a different route of sample preparation, which involved addition of

water to homogenous CA/DMAc solutions (dissolution method 2), the system showed

similar viscosity increase with water content (Figure 3) at CA concentration of 10wt%.

It is envisaged the intensification of bonds interactions was the main factor influencing

the increment in viscosity. Beyond critical water content (>19wt%), the system

transforms from a homogenous to a uniform cloudy rigid gel-like material. This sol-gel

transition is shown in Figure 4 where G′ and G″ as a function of frequency for two

different water compositions at CA concentration of 10wt%. At the low water

composition, we find G″ to be larger than G′ over the frequency range with both

showing dependence on frequency a behavior typically depicted by a polymer solution.

But at the high water composition shown, G′ is larger than G″ indicating that elastic

properties are dominant over viscous properties. Furthermore, both G′ and G″ are

relatively independent of frequency indicating the formation of a gel network. Thus the

system can be tuned from a solution to a gel by increase in the non-solvent composition.

Figure 4. Effect of low and high water contents on the viscous (G″) and elastic (G′)

moduli for 10wt% cellulose acetate/DMAc/water system.

Page 182: Rheology and Microstructure of Cellulose Acetate - Digital Repository

167

It is envisaged that when bulk CA/DMAc solutions are prepared, the CA form

strong hydrogen bonds with DMAc in solution. As we increase the amount of water in

the system, CA resist water from displacing the hydrogen bond formed initially but

water due to its strong hydrogen bond characteristics can break off some of the CA-

DMAc bonds. Thus we have this complex hydrogen bond formation involving CA-

DMAc-water with some weak bonds formed between CA and water. This causes

aggregates of CA to span entire sample leading to formation of rigid, uniform gel-like

structures. Gelation process may also be described as the aggregation of particles during

the growth process of these clusters, which collide to from gels [45].

CELLULOSIC LIQUID CRYSTALLINE POLYMERS (LCPs)

Formation of Liquid Crystals

Liquid crystalline phase, or mesophase, is a special state of the matter that

possesses long-range order on a macroscopic scale but can flow like fluids. Generally,

molecules are required to have asymmetric shapes to form liquid crystalline phases. For

polymers, the mesogenic groups (moieties to form liquid crystalline phases) are linked

either in the main chain with flexible segments or on the side chain attached to the

flexible main chain. It is these mesogenic groups that provide the liquid crystalline

properties for the polymers. Some polymers without mesogenic groups are also capable

of forming liquid crystalline phases. These polymers have a rigid or semirigid backbone

arising from the steric effects or intra-molecular hydrogen bonding which restrict the

flexibility of the chain.

Page 183: Rheology and Microstructure of Cellulose Acetate - Digital Repository

168

Flory first predicted that linear rod-like polymer would form ordered phases in

concentrated solutions [46]. By using a lattice model, Flory’s theory predicted that the

rigid polymer solution would separate into two phases - isotropic and anisotropic, at a

concentration given by:

)21(8*

xxVp −=

where Vp* is the critical volume fraction of the polymer, x is the axial ratio, given by x

= l/d (rod length to diameter). The phase separation can be attributed to a favorable

entropy of forming an ordered phase at high polymer concentration.

Both entropy and enthalpy are found to contribute to the stability of nematic

mesophases, although for LCPs the dominant factor is the entropy term, which is

controlled by the size and shape of the molecule [47, 48]. The original theory, which

assumed that the polymer chains were completely rigid and rod-like, does not reflect

real polymers. Even the most rigid polymers have some degree of flexibility. Besides,

polydispersity, distribution of substituted side groups as well as polymer-solvent

interactions will affect the flexibility of the polymer.

Formation of Lyotropic Liquid Crystalline Phase of Cellulose

As discussed above, due to the strong inter- and intra-chain hydrogen bonding,

cellulose has poor solubility in common solvents. Solubility problems limit the

formation of lyotropic liquid crystalline solutions of cellulose. Only a few solvents are

known to dissolve cellulose and form the anisotropic phase.

Page 184: Rheology and Microstructure of Cellulose Acetate - Digital Repository

169

For cellulose solution in N-methyl-morpholine-N-oxide (NMNO)/H2O, the

formation of anisotropic phase was reported [49] without determining the nature of the

mesophase. Patel et al. reported the mesophase formation of cellulose in trifluoroacetic

acid (TFA)/chlorinated alkane solvent system and labeled as chiral nematic [50].

However, due to the high acidity and reactivity of TFA, cellulose is degraded or

derivatized. Cuculo reported the formation of mesophase of cellulose in liquid

NH3/NH4SCN [51]. Both nematic and chiral nematic phases were observed at different

cellulose concentrations. Later, Hattori and Cuculo observed the anisotropic phase

formation of cellulose in hydrazine/thiocyanate salt solvent system [52] and

ethylenediamine/thiocyanate salt solvent system [53]. The nature of these mesophases

was not determined.

Dimethylacetamide (DMAc)/LiCl system is a stable cellulose solvent system.

The dissolution of cellulose comes from cellulose complexes and hydrogen bonding

[54]. However, due to the solubility limit of cellulose, a high concentration (> 15%) is

unlikely to achieve [55]. The critical concentration to form mesophase (10 – 15%)

depends on the salt concentration of the solvent.

Hydrolysis of natural cellulosic materials by sulfuric acid or phosphoric acid produced

highly crystalline well-defined rods – cellulose nanocrystallites (or cellulose

whiskers)[56]. The cellulose crystallites formed colloidal suspension in water and it was

stabilized by the negative surface charge. The suspension formed ordered chiral nematic

phase at a critical concentration as low as 1-2% volume fraction [57].

Formation of Lyotropic Liquid Crystalline Phase of Cellulose Derivatives

Page 185: Rheology and Microstructure of Cellulose Acetate - Digital Repository

170

The hydroxyl groups on cellulose can be easily substituted to form ethers and

esters, resulting in variety of cellulose derivatives. The introduction of substituents on

cellulose disrupts the intra-chain hydrogen bonds and brings new steric interactions,

which results in the change of rigidity of the cellulose backbone and solubility of the

cellulosic polymer. In generally, cellulose derivatives have better solubility than

cellulose in a larger range of common solvents.

Following the discovery of mesophases formation in aqueous solutions of

hydroxypropyl cellulose (HPC) [58], a variety of other cellulose derivatives have been

reported to form liquid crystals [59-61]. In general, lyotropic liquid crystalline formed

by cellulose derivatives are achieved in highly concentrated solutions, ranging from 20

– 70% by weight. Due to the semi-flexible nature of cellulosics, these polymers have a

critical concentration to form mesophase greater than that of rigid rods predicted by

Flory’s theory. The critical concentration for phase separation has been demonstrated to

depend on the nature of substituents, the degree of substitution (DS, defined as the

average number of hydroxyl groups substituted on each anhydroglucopyranose unit),

the distribution of substituents for partially substituted cellulose derivatives,

temperature and solvents. This indicates that the stiffness of cellulose derivatives in

solutions is a function of steric interactions between adjacent units, intra-chain

hydrogen bonding, and intermolecular interactions of polymer-polymer and polymer-

solvent.

Intermolecular Interactions in Lyotropic Cellulosic Liquid Crystalline Solutions

Page 186: Rheology and Microstructure of Cellulose Acetate - Digital Repository

171

The formation of liquid crystalline phase is highly dependent on specific

polymer-solvent and polymer-polymer interactions. The transition from isotropic to

ordered phase of lyotropic systems of semi-rigid polymers often represents a delicate

balance of polymer-polymer and polymer-solvent interactions [61]. Not all cellulose

derivatives form liquid crystalline phase at high concentrations. Actually, only those

with appropriate solubility in a particular solvent can form liquid crystalline solutions.

In some cellulosic/solvent systems, increasing the polymer concentration from the semi-

dilute state, where already microgel appear, leads directly to the gel state, with no

evidence of liquid crystallinity, even though the backbone of these derivatives has

similar stiffness compared to those derivatives that produce mesophases. Other

cellulose derivatives, due to strong polymer-polymer interactions, do not have high

enough solubility in a particular solvent to achieve the concentrations necessary for

mesophase formation.

The solvent effect on the stiffness of cellulose backbone is clearly demonstrated

by difference in the clearing temperature of anisotropic ethylcellulose (EC) solutions in

acetic acid (AA), dichloroacetic acid (DCA) and the mixture of these two [62]. The

higher clearing temperature of EC/DCA mesophase indicates that EC is more rigid in

DCA than in AA. The effect of polymer-solvent interactions on polymer rigidity was

further shown using shear-relaxation studies. After shearing a rigid polymeric liquid

crystal, a band texture (regular bright and dark lines under cross polarizers), formed in

which the direction of the bands was perpendicular to the shearing direction. The band

structure formed due to a periodic packing of rigid/semi-rigid polymer chains in a

zigzag fashion. This orientation process is a transient phenomenon. After a period of

Page 187: Rheology and Microstructure of Cellulose Acetate - Digital Repository

172

time the polymer chains relax and the band texture disappears. The bandwidth and the

relaxation time of the bands depend on the rigidity of the polymer chain. In fact, at the

same concentration, EC chains are more rigid in DCA than in acrylic acid or glacial

acetic acid, resulting in wider bands and longer relaxation time [63].

Chiroptical Properties of Cellulosic Liquid Crystals

Most lyotropic liquid crystalline solutions of cellulose derivatives form chiral

nematic (cholesteric) phase. Some can form nematic phase under certain critical

conditions [61, 62, 64]. These chiral nematic mesophases possess a unique structure in

which the alignment of the cellulosic molecules is at a slight angle to one another to

result in a helicoidal supramolecular structure (Figure 5). The helicoidal structure is

described by a pitch p (or its inverse, the twist p-1) and the corresponding handedness of

the twist. Pitch is the distance between two layers that have the same molecular

orientation and is determined by p = λ0/ñ where λ0 is the reflection wavelength and ñ

is the mean refractive index. The right-handed helicoidal structure is conventionally

assigned to a positive pitch (p > 0) and left-handed helicoidal structure to a negative

pitch (p < 0)[59, 65]. The nematic phases can be considered as a special state of chiral

nematic phases with no handedness and infinite pitch. Chiral nematic phases are formed

when optically active molecules are incorporated into the nematic state. Chiral nematic

phases exhibit unique optical properties, such as selective reflection of incident light

and optical rotation, because of their helicoidal supramolecular structure [59].

Figure 5. Helicoidal structure of chiral nematic mesophase

Page 188: Rheology and Microstructure of Cellulose Acetate - Digital Repository

173

The anisotropic phase of cellulose in DMAc/LiCl [61] and anisotropic cellulose

nanocrystallite suspension [66] are left-handed. Cellulose in TFA/chlorinated alkane is

right-handed, which may be due to the esterification reaction of TFA with cellulose

[50]. Most cellulose derivatives form right-handed mesophases. Guo [67] hypothesized

that the handedness of cellulosic mesophase was somewhat determined by the size of

substituents, with small groups to form left-handed phase and bulky groups to form the

right one. This statement was supported by the fact that methylcellulose and fully

acetylated methylcellulose formed only left-handed mesophase [67] but acetylated

ethylcellulose showed the reversal of handedness, from left-handed to right-handed,

with increasing acetyl content [64]. However, the handedness of cellulosic mesophase is

not only determined by the size of side groups. Several bulky derivatives, such as

phenylacetoxy cellulose in CH2Cl2 [68], cellulose tris(4-chlorocarbanilate) in diethylene

glycol monomethyl ether (DMME) [69] and 6-O-trityl-2,3-O-hexylcellulose in THF

[70], were reported to form left-handed phase. The handedness and pitch of cellulosic

mesophases result from complicated inter- and intra- molecular interactions in the

solutions.

The observed handedness and pitch of the helicoidal structure is sensitive to

temperature, concentration, substitutions, solvent, ions as well as external field [71, 72].

The effect of temperature, concentration, molecular weight, solvent, substitution on the

chiroptical properties of the helicoidal structure have been comprehensively reviewed

by Zugenmaier [60, 61], Guo [59] and Gray [65]. Due to the complicated inter-

Page 189: Rheology and Microstructure of Cellulose Acetate - Digital Repository

174

molecular interactions caused by these factors, except for molecular weight, all other

factors do not seem to follow any particular patterns without exceptions.

An effective way to study the specific interactions in lyotropic liquid crystalline

solutions is to utilize regioselective substitution of cellulose. Kondo [73] reported that

free hydroxyls at the C3 position played an important role in determining the

handedness of the EC mesophase in CH2Cl2. Breaking intramolecular hydrogen

bonding by increasing the substitution at C3-OH position cause the cellulose backbone

to become more flexible, which may account for the handedness change. Similarly, the

twisting power of the mesophase of cellulose regioselective tricarbanilate derivatives

was less influenced by the substitution in C6-OH, more by the C2-OH and probably

most by C3-OH [74].

For chiral molecules, equal amount of L- and R- isomers will compensate the

opposite twisting power and form a racemic mixture without optical activities – nematic

phase, as formation of nematic phase by poly-(γ-benzyl-L-glutamate) (PBLG) and poly-

(γ-benzyl-D-glutamate) (PBDG) [75]. For cellulosics, due to the existence of multiple

chiral centers in each unit, the interactions among these chiral centers are complicated.

With the same average degree of acetylation, pure AEC and mixture of EC and fully

acetylated AEC showed different chiroptical properties in chloroform [76]. However,

the inversion in handedness provides a useful routine to untwist the helicoidal structure

and to form nematic phases. This “untwisting” of the helicoidal structure, characterized

by an infinite pitch and the absence of macroscopic chirality, has been observed

experimentally. Zugenmaier et al. found that at an appropriate ratio of two solvents in

which the handedness of the individual mesophase is opposite produced a nematic-like

Page 190: Rheology and Microstructure of Cellulose Acetate - Digital Repository

175

ordered phase [61, 62]. By introducing a random distribution of phenyl and 3-Cl-phenyl

side groups into cellulose tricarbanilate, which cause opposite handedness of

mesophases in the same solvent individually, a phase with a zero twist formed [62].

Guo reported that for lyotropic AEC mesophases, at a critical temperature or a degree of

substitution (DS), no chiral order was observed due to the property that the reversal in

handedness with increasing acetylation of AEC [65]. By mixing a certain among of EC

and fully acetylated AEC polymers in the same solvent, which form left- and right-

handed mesophase individually, a nematic phase may be formed [76].

Rheology of Cellulosic Liquid Crystalline Solutions

The microstructure and polymer-solvent interactions of lyotropic cellulosic

mesophases can be derived from the rheological studies. The lyotropic PLC solution is a

complicated system and a wide range of unusual rheological phenomena has been

observed. Following the first observation of mesophase formation of HPC/H2O, this

system has been most widely investigated in rheology. Rheological studies in other

cellulosic mesophases, e.g., cellulose [77], ethylcellulose [62, 78, 79] and cellulose

tricarbanilate [61], have also been reported.

Concentration dependence of viscosity

Isotropic solutions show a monotonic increase in shear viscosity with the

increasing concentration. The viscosity increases to a maximum when the isotropic to

anisotropic transition is approached. Upon formation of anisotropic phase the viscosity

begins to decrease. After which viscosity increases exponentially as the concentration

Page 191: Rheology and Microstructure of Cellulose Acetate - Digital Repository

176

continues to increase (Figure 6). In the isotropic state, the hydrodynamic volume is

large due to the random polymer orientation, which restricts the polymer diffusivity and

causes an increase in viscosity. Upon anisotropic phase formation, the aligned polymer

leads to a small hydrodynamic volume. The viscosity decreases since rotational

diffusion is much easier with a net orientation.

Figure 6.Viscosity as a function of concentration for lyotropic LCPs

However, Zugenmaier found that the correlation of the viscosity maximum with

the formation of the anisotropic phase was valid only when the shear rate was low

(shear rate 0)[61]. At a higher shear rate, the concentration at which the maximum of

viscosity occurs shifts to a smaller value. When the shear rate is high enough to cause

shear-induced orientation (pseudo-nematic phase), entanglements of the random

distributed polymers are loosened. The viscosity maximum disappears and a monotonic

increase of viscosity vs. concentration is observed. This indicates that the viscosity of

the nematic or pseudo-nematic mesophase is less sensitive to concentration than that of

the chiral nematic phase [61].

Anisotropic suspensions of cellulose nanocrystallites demonstrate similar

viscosity-concentration relationship. The viscosity vs. concentration curve has a

maximum, which disappears at high shear rates [80].

Steady flow behavior

Page 192: Rheology and Microstructure of Cellulose Acetate - Digital Repository

177

The steady flow behavior is the most thoroughly studied rheological property.

Onogi and Asada hypothesized the universal existence of three shear flow regimes [81]

to describe the viscosity of PLCs: a shear thinning regime at low shear rates (Region I),

a Newtonian plateau at intermediate shear rates (Region II), and another shear thinning

regime at low shear rates (Region III) (Figure 7). The three-regime curve was observed

in the anisotropic aqueous HPC solution [82] and HPC in acetic acid [83]. Asada

proposed that some PLC systems do not show all regimes because not every regime lies

in the accessible shear rate range. In fact, it appears to be unusual to be able to access all

regimes for PLCs including cellulosics. Whether the three-regime flow curve is indeed

universal or not is still questionable.

Figure 7. Three-regime steady state shear viscosity for LCPs

So far, for cellulosic mesophases with all three regimes, microstructures in all

three regimes have been well documented by experiments. Region I is generally

believed to reflect a defect structure, or piled polydomain texture. The texture is a

supramolecular disorder of the nematic structure and is composed of nematic domains

possessing little or no macroscopic orientation [82, 84]. This region is characterized by

distortional elasticity associated with spatial variation in the director (average local

molecular orientation) field. Region II reflects a “dispersed polydomain” structure.

Region III is believed to be characterized by non-linear effects of flow on the molecular

orientation. In region I and II, the flow is not strong enough to affect the molecular

orientation. In Region III, the flow field is very strong that shear induced molecular

Page 193: Rheology and Microstructure of Cellulose Acetate - Digital Repository

178

orientation becomes important. By birefringence measurement, the molecular

orientation is a monotonic increasing function of steady state shear rate for anisotropic

HPC/H2O solutions [85] and HPC/m-cresol solutions [86].

The most striking phenomenon in the steady flow of PLCs is the occurrence of

negative first normal stress difference (N1) – two sign changes in N1 as a function of

shear rate. In contrast, isotropic solutions only show positive N1 at all shear rates.

Negative N1 has been reported in HPC/H2O system [87-89] and HPC/m-cresol system

[90].

The negative N1 results from a coupling between molecular tumbling under flow

and the local molecular-orientation distribution [87]. At low shear rates, the director

tumbles with the flow and N1 will be positive. At intermediate shear rates, nonlinear

viscoelastic effects are important. The director tumbling competes with the steady

director alignment along the flow and director oscillates about a steady value. N1

becomes negative. At very high shear rates, the director aligns along the flow and N1 is

positive again. It is generally believed that the shear rate at which N1 is minimum halts

director tumbling and aligns the molecules [90]. However, negative N1 behavior

disappears at very high polymer concentration due to the polymer-polymer friction

interactions [87].

Relaxation behavior

The relaxation behavior, or the transient behavior of cellulosic liquid crystalline

solutions upon the cessation of steady flow, is unique for PLCs. There are two kinds of

relaxations. The bulk stresses relax quickly while the structures relax over a much

Page 194: Rheology and Microstructure of Cellulose Acetate - Digital Repository

179

longer time. Moldenaers and Mewis suggested that two levels of structures exist [91].

Stress relaxation reflects the fast relaxation at molecule level arrangements and is

independent of the previous shear rate. Structure relaxation reflects the gradual changes

of the textures. This slow process is unique in that its time scale is inversely

proportional to the previous shear rate. When the strain recovery (or recoil) after

cessation of steady state flow is plotted against time multiplied by the previous shear

rate, the curves superimpose one another [92].

It is generally believed that under high rate flow, the chiral nematic mesophase

aligns along the flow direction and uncoils to form nematic structure. Upon cessation of

flow, the chiral nematic phase will reform and the molecular orientation will decrease

[93]. By using evolution of the dynamic moduli [94] and birefringence [95] as a

function of time, the structural change can be investigated. For lyotropic HPC/H2O

solutions, upon flow cessation, molecular orientation decreases to a globally isotropic

condition at all rates [95] and the dynamic moduli increase to a maximum [94].

However, the final relaxed state depends on the shear history. After high shear rates, the

solutions evolve towards an 'equilibrium' state with a high modulus, while after low

shear rates, the solutions relax to the 'equilibrium' state with a low modulus [85]. The

low-modulus state is ordered and evolves out of a state that has no macroscopic upon

cessation of the flow. The high-modulus state is much less ordered, although it evolves

from a rather well flow-aligned state upon cessation of the flow [96].

The formation of band texture is another type of relaxation phenomenon, as

discussed in previous sections. For HPC/H2O solutions, the band texture is only

observed when the molecules have been well orientated in the shear direction [87]. A

Page 195: Rheology and Microstructure of Cellulose Acetate - Digital Repository

180

critical lower shear rate limit exists due to the stability of chiral nematic textures. An

upper shear rate limit also exists above which no bands are formed [97]. The upper

critical shear rate is associated with the flow-aligning region of molecular dynamics.

Band texture formation is driven by the release of energy, which has been stored in the

mesophase during shear [87]. The evolution of the band texture depends on the previous

shear rate applied to mesophase. The rate of evolution of the band texture increases,

remains constant or decreases according to whether the mesophase is shear at low,

intermediate or high rates [98]. When the band texture appears during recoil, the

appearance of band texture stops the strain recovery process until the band texture

disappears, and then recovery continues. The presence of the band texture during recoil

enhances the strain recovery [99].

Relaxation Behavior of Lyotropic (Acetyl)(ethyl)cellulose / Acrylic Acid Solutions

with Different Chiroptical Properties

At a high shear rates the chiral nematic structure changes to a flow-induced

nematic phase. However, the shear-oriented phase is easy to disrupt after removing

shear force, which is due to the driving force for the liquid crystalline solution to

form the more thermodynamically stable chiral nematic structure [93]. For a

pseudo-nematic lyotropic solution, after oriented along the high shear rate direction,

the driving force to reform helicoidal structure may be limited and the relaxation

persists for very long time. So far, only few reports are available on the relations

Page 196: Rheology and Microstructure of Cellulose Acetate - Digital Repository

181

between the relaxation behavior and chiroptical properties of the mesophases. In the

case of PBG solutions containing a single optical isomer (PBLG or PBDG) or

racemic mixture of these two, there is no phenomena that cholestercity has any

significant effects on the relaxation [95]. This is due to the fact that the mesophases

of both the optical isomers and the racemic mixture have large pitch and are

nematic-like. In our recent study, relaxation behaviors upon shear force removal

were investigated for lyotropic solutions of (acetyl)(ethyl)cellulose (AEC) in acrylic

acid (AA) with different chiroptical properties.

Ethylcellulose (EC) and AEC can form cholesteric liquid crystalline solutions in

many solvents and were found to change handedness as well as pitch with the change of

degree of acetylation. In acrylic acid, EC forms left-handed mesophase. Fully acetylated

AEC forms right-handed mesophase [100]. At a certain critical degree of acetylation

(DA*) the pitch becomes infinite and nematic-like and untwisted structure forms. When

the degree of acetylation of AEC increases across DA*, the handedness of cholesteric

liquid crystalline solutions changes from left-handedness to right-handedness. The pitch

of lyotropic solution of EC or fully acetylated AEC in AA is about one tenth of that of

PBG solutions. Therefore, the effect of chiroptical properties on relaxation behavior is

expected to be more pronounced in AEC system than in PBG system.

The values of pitch for lyotropic AEC solutions in AA (50% w/w) are in Table

1. AEC-2 is a pure AEC with medium DA. AEC-3 is the mixture of EC and fully

acetated AEC, which has the same average DA as AEC-2. However, in the liquid

crystalline solutions, even at the same concentration, AEC-2 and AEC-3 have different

pitch and handedness. This phenomenon was also observed in lyotropic

Page 197: Rheology and Microstructure of Cellulose Acetate - Digital Repository

182

AEC/chloroform system [76]. The difference in chiroptical properties may come from

the complex interactions of multiple chiral centers existing in each repeating unit on

cellulose chain, not to form simple racemic mixtures as in PBG system.

Table 1. Chiroptical properties of lyotropic AEC/AA solutions with different degree of

acetylation (DA)

Steady state flow

At the same concentration, the viscosities of AEC solutions are different. The

larger the pitch, the smaller the viscosity [64]. To eliminate the influence of viscosity of

LCP solutions on the relaxation behaviors, the concentrations of AEC solutions were

adjusted so that all solutions had similar viscosities under shear. The steady state shear

curves of AEC/AA are shown in Figure 8. It is a typical three-region curve. With the

increase of shear rate, a shear-thinning is followed by a Newtonian plateau, then

followed by another shear-thinning region at high shear rate. The Newtonian plateau is

in the shear rate range 1 – 5 S-1. With the increase of pitch, the curve becomes flat and

the three regions are not as distinct as the one with smaller pitch. When the shear rate is

between 0.5s-1 and 6s-1, the four AEC solutions had the same viscosity.

Figure 8. Viscosity as a function of shear rate of AEC/AA solutions

Page 198: Rheology and Microstructure of Cellulose Acetate - Digital Repository

183

Stress relaxation.

Mercer and Weymann [101] proposed an equation for the time-

dependent viscosity η(t) of thixotropic suspensions:

(η(t)-ηf)/(ηi-ηf) = η* = A1exp(-t/τ1) + A2exp(-t/τ2) (1)

where ηi and ηf are viscosities at shear rate γi and γf, A1 and A2 are constants, t is time,

and τ1 and τ2 are relaxation times.

Based on Mercer and Weymann’s Equation, Suto et al. [102] found that the

stress relaxation was a two-stage process for lyotropic EC/m-cresol solutions and

proposed another equation for lyotropic liquid crystals:

(σ(t) - σf) / (σi - σf) = σ* = A1exp(-t/τ1) when (t < tc)

= A2exp(-t/τ2) when (t > tc) (2)

where tc is a characteristic time. ln(σ*) is the reduced shear force.

The stress relaxation behavior of some AEC/AA solutions is in Figure 9. Only

one relaxation process was observed, not a two-relaxation process as Suto found. This

may be due to the fact that the first stage relaxation process is too fast to be observed.

Therefore, the stress relaxation behavior can be represented as:

(σ(t) - σf) / (σi - σf) = σ* = Aexp(-t / τ) (3)

Figure 9. Reduced shear stress σ* vs. time for 50% AEC-4( ) and 51.5% AEC-2(∆)

The relaxation times of different AEC solutions can be evaluated according to

equation 3 by the least-square method. The relaxation time and the chiroptical

Page 199: Rheology and Microstructure of Cellulose Acetate - Digital Repository

184

properties of AEC solutions are listed in Table 2. The larger the pitch, the longer the

time for shear force to relax. The relaxation times for AEC solutions with similar pitch

but different handedness are approximately the same.

Table 2. Relaxation times as the function of chiroptical properties of lyotropic AEC/AA

solutions

Dynamic modulus evolution upon cessation of flow

The development of storage and loss moduli after flow cessation is a useful tool

to analyze structural relaxations on LCPs. Upon flow cessation, the flow-induced

orientation is lost. The evolution of the moduli with time is due to the reformation of a

chiral nematic phase that had become nematic under flow. Figure 10 shows the

relaxation in complex modulus with time.

Figure 10. Complex modulus evolution of AEC solutions vs. time

Upon flow cessation, the larger the pitch, the slower the modulus evolved. The

modulus of right-handed mesophase (AEC-4) developed faster than that of the left-

handed one (AEC-1) with the similar pitch. The difference in relaxation behavior of

AEC solutions may be due to the smaller driven force for the mesophases with larger

pitch (nematic-like) to reform helicoidal structures of chiral nematic phases from

nematic under flow.

Page 200: Rheology and Microstructure of Cellulose Acetate - Digital Repository

185

REFERENCE

1. Krassig, H. A., Polymeric monographs, Vol. 11, Fordon and Breach,

Philadelphia, 1992.

2. Payen, A., Compt. Rend., 1, 1052, 1125 (1838).

3. Ott, E., Spurlin, H. M., Graffin, M. W., Cellulose Derivatives, Interscience, N.

Y., 1954, pg. 1.

4. Kamide, K., Iijima, H., in Cellulosic Polymers, Blends and Composites, R. D.

Gilbert, Ed., Hansen, N. Y., 1994, pg. 189.

5. Noether, H. D., in Celluloisic Polymers, Blends and Composites, R. D. Gilbert,

Ed., Hansen, N. Y., 1994, pg. 217.

6. Rawls, R., Chemical and Engineering News, xx, 6 (1996).

7. Yamauaka, S., Watanabe, K., in Cellulosic Polymers, Blends and Composite, R.

D. Gilbert, Ed., Hanser, N. Y., 1994, pg. 207.

8. Chanzy, H., in Cellulose Sources and Exploitation, J. F. Kennedy, G. O.

Phillips, P. A. Williams, Eds., Ellis Horwood, England, 1990, pg. 3.

9. Irvine, J. C.; Hirst, E. L., J. Chem. Soc., 1923, 123, 518; (b) Freudenberg, K.;

Plankenhorn, E.; Boppel, H., Ber., 1938, 71, 2435. (C.A. 32, 5656)

10. Spencer, C. C., Celluloechemie, 1921, 10, 61. (C.A. 35, 1042)

11. Haworth, W. N., Helv. Chim. Acta. 1928, 11, 534. (C.A. 28, 6313)

12. Staudinger, H., Die hochmolekularen organischen Verbindungen - Kautschuk

and Cellulose, 1932, Springer Verlag, Berlin (2nd Edn 1960). (C.A. 26, 3513)

13. Rao, V. S. N.; Sundararajan, P. R.; Ramakrishnan, C.; Ramachandran, G. N.,

Conformation of Biopolymers, 1957, Ramachandran, G. N. Ed. Academic Press.

Page 201: Rheology and Microstructure of Cellulose Acetate - Digital Repository

186

London, p 721. (b) Goebel, K. D.; Harvie, C. E.; Brant, D. A., Appl. Polym.

Symp., 1976, 28, 671. (c) Mitchell, A. J.; Higgins, Tetrahedron, 1965, 21, 1109.

14. Mann, J.; Marrinan, H. J., J. Polym. Sci., 1958, 32, 357; (b) Marrinan, H. J.,

Mann, J., J. Polym. Sci., 1956, 21, 301.

15. Simon, I.; Scheraga, H.A.; Manley, R.S.J. Structure of Cellulose. 1. Low-Energy

Conformations of Single Chains. Macromolecules, 1988 21(4):983-990,

16. Pizzi, A.; Eaton, N. The Structure of Cellulose by Conformational-Analysis.3.

Crystalline and Amorphous Structure of Cellulose-I. J. Macromol. Sci.-Chem.

1985, A22: 105-137.

17. Nystrom, B.; Walderhaug, H.; Hansen, F.K. Rheological behavior during

thermoreversible gelation of aqueous mixtures of ethyl(hydroxyethyl) cellulose

and surfactants. Langmuir 1995, 11, 750-757.

18. Vaessen, D.M.; McCormick, A.V.; Francis, L.F. Effect of phase separation on

stress development in polymeric coatings. Polymer 2002, 43, 2267-2277.

19. Reuvers, A.J., Altena, F.W., Smolders, C.A., Demixing and gelation behavior of

ternary cellulose acetate solutions. Journal of Polymer Science: Part B: Polymer

Physics 1986, 24, 793-804.

20. Sarkar, N. Thermal gelation properties of methyl and hydroxypropylmethyl

cellulose. Journal of Applied Polymer Science 1979, 24, 1073-1087.

21. Kobayashi, K.; Huang, C.; Lodge, T.P. Thermoreversible gelation of aqueous

methylcellulose solutions. Macromolecules 1999, 32, 7070-7077.

22. Sarkar, N. Kinetics of thermal gelation of methyl cellulose and hydroxypropyl

cellulose in aqueous solutions. Carbohydrate Polymers 1995, 26, 195-203.

Page 202: Rheology and Microstructure of Cellulose Acetate - Digital Repository

187

23. Takahashi, M.; Shimazaki, M.; Yamamoto, J. Thermoreversible gelation and

phase separation in aqueous methyl cellulose solutions. Journal of Polymer

Science: Part B: Polymer Physics, 2001, 39, 91-100.

24. Li, L.; Thangamathesvaran, P.M.; Yue, C.Y.; Tam, K.C.; Hu, X.; Lam, Y.C. Gel

network structure of methylcellulose in water. Langmuir 2001, 17, 8062-8068.

25. Desbrieres, J.; Hirrien, M.; Ross-Murphy, S.B. Thermogelation of methyl

cellulose: rheological considerations. Polymer 2000, 41(4), 2451-2461.

26. Haque, A.; Morris, E.R. Thermogelation of methylcellulose. Part I: molecular

structures and processes. Carbohydrate Polymers 1993, 22, 161-173.

27. Hirren M.; Chevillard, C.; Desbrieres, J.; Axelos, M.A.V.; Rinaudo, M.

Thermogelation of methylcellulose: new evidence for understanding the gelation

mechanism. Polymer 1998, 39(25), 6251-6259.

28. Lizaso, E.; Munoz, M.E.; Santamaria, A. Formation of Gels in Ethylcellulose

Solutions. An Interpretation from Dynamic Viscoelastic Results.

Macromolecules 1999, 32, 1883-1889.

29. Hu, X.; Zheng, P.J.; Zhao, X.Y.; Li, L.; Tan, K.C.; Gna, L.H. Preparation,

characterization and novel photoregulated rheological properties of azobenzene

functionalized cellulose derivatives and their α-CD complexes. Polymer 2004,

45(18), 6219-6225.

30. Kawanishi, H.; Tsunashima, Y.; Okada, S.; Horii, F. Change in chain stiffness in

viscometric and ultracentrifugal fields: cellulose diacetate in N,N-

dimethylacetamide dilute solution. Journal of Chemical Physics 1998, 108 (14),

6014-6025.

Page 203: Rheology and Microstructure of Cellulose Acetate - Digital Repository

188

31. Kawanishi, H.; Tsunashima, Y.; Horii, F. Dynamic light scattering study of

structural changes of cellulose diacetate in solution under couette flow.

Macromolecules 2000, 33, 2092-2097.

32. Tsunashima, Y.; Kawanishi, H.; Nomura, R.; Horii, F. Influence of long-range

interactions in diffusion behavior in semidilute solution: Dynamics of cellulose

diacetate in quiescent state. Macromolecules 1999, 32, 5330-5336.

33. Gomez-Bujedo, S.; Fleury, E.; Vignon, M.R. Preparation of cellouronic acids

and partially acetylated cellournic acids by TEMPO/NaClO oxidation of water-

soluble cellulose acetate. Biomacromolecules 2004, 5, 565-571.

34. Bochek, A.M.; Kalyuzhnaya, L.M. Interaction of water with cellulose and

cellulose acetates as influenced by the hydrogen bond system and hydrophilic-

hydrophobic balance of the macromolecules. Macromolecular Chemistry and

Polymeric Materials 2002, 75, 989-993.

35. Boerstoel, H.; Maatman, H.; Picken, S.J.; Remmers, R.; Westerink, J.B. Liquid

crystalline solutions of cellulose acetate in phosphoric acid. Polymer 2001, 42,

7363-7369.

36. Pintaric, B.; Rogosic, M.; Mencer, H.J. Dilute solution properties of cellulose

diacetate in mixed solvents. Journal of molecular liquids 2000, 85, 331-350.

37. Ryskina, I.I.; Aver’yanova, V.M. A study of gelation and the structure of

acetylcellulose gels in benzyl alcohol. Vysokomolekulyarnye Soedineniya 1971,

A13 (10), 2189-2194.

Page 204: Rheology and Microstructure of Cellulose Acetate - Digital Repository

189

38. Ryskina, I.I.; Aver’yanova, V.M. Morphological study of the three dimensional

network of cellulose acetate-benzyl alcohol gelatins. Vysokomolekulyarnye

Soedineniya 1975, A17, 919-922.

39. Glikman, S.A.; Aver’yanova, V.M.; Khomutov, L.I. Structure of

acethylcellulose solutions. Vysokomolekulyarnye Soedineniya 1963, 5(4), 598-

604.

40. Altena, F.W.; Smolders, C.A. Phase separation phenomena in solutions of

cellulose acetate. I. Differential scanning calorimetry of cellulose acetate in

mixtures of dioxane and water. Journal of Polymer Science: Polymer

Symposium 1981, 69, 1-10.

41. Chauvelon, G.; Doublier, J-L.; Buleon, A.; Thibault, J-F.; Saulnier, L.

Rheological properties of sulfoacetate derivatives of cellulose. Carbohydrate

Research 2003, 338, 751-759.

42. Wang, C.S.; Fried, J.R. Viscoelastic properties of concentrated cellulose acetate

solutions. Journal of Rheology 1992, 36, 929-945.

43. Ryo, Y.; Nakai, Y.; Kawaguchi, M. Viscoelastic measurements of silica

suspensions in aqueous cellulose derivative solutions. Langmuir 1992, 8, 2413-

2416.

44. Li, L. Thermal gelation of methyl cellulose in water: scaling and

thermoreversibility. Macromolecules 2002, 35, 5990-5998.

45. Alie, C.; Pirard, R.; Pirard, J-P. Preparation of low-density xerogels from

mixtures of TEOS with substituted alkoxysilanes. II. Viscosity study of the sol-

gel transition. Journal of Non-Crystalline Solids 2003, 320, 31-39.

Page 205: Rheology and Microstructure of Cellulose Acetate - Digital Repository

190

46. Flory, P.J. Molecular Theory of Liquid Crystals. Adv. Polym. Sci. 1984, 59:1-

35.

47. Flory, P.J.; Ronca, G. Theory of Systems of Rodlike Particles. 1. Athermal

Systems. Mol. Cryst. Liq. Cryst. 1979, 54(3-4): 289-309.

48. Warner, M.; Flory, P.J. The Phase-Equilibrium in Thermotropic Liquid-

Crystalline Systems. J. Chem. Phys. 1980, 73(12): 6327-6332.

49. Chanzy, H.; Peguy, A.; Chaunis, Y.S.; Monize, P. Oriented Cellulose Films and

Fibers from a Mesophase System. J. Polym. Sci., Polym. Phys. Ed. 1980, 18(5):

1137-1144.

50. Patel, D.L.; Gilbert, R.D. Lyotropic Mesomorphic Formation of Cellulose in

Trifluoroacetic-Acid Chlorinated-Alkane Solvent Mixtures at Room-

Temperature. J. Polym. Sci. Polym. Phys. Ed. 1981, 19 (8): 1231-1236.

51. Chen, Y.S.; Cuculo, J.A. Lyotropic Mesophase of cellulose in Ammonia

Thiocyanate Solution. J. Polym. Sci. Polym. Chem. Ed. 1986, 24(9): 2075-2084.

52. Hattori, K.; Cuculo, J.A.; Hudson, S.M. New Solvents for Cellulose:

Hydrazine/Thiocyanate Salt System. J. Polym. Sci. Polym. Chem. Ed. 2002,

40(4): 601-611.

53. Hattori, K.; Abe, E.; Yoshida, T.; Cuculo, J.A. New Solvents for Cellulose. II.

Ethylenediamine/Thiocyanate Salt System. Polym. J. 2004, 36 (2): 123-130.

54. McCormick, C.L., Callais, P.A., Hutchinson, B.H. Solution Studies of Cellulose

in LiCl and N,N-Dimethylacetamide. Macromolecules 1985, 18(12): 2394-2401.

Page 206: Rheology and Microstructure of Cellulose Acetate - Digital Repository

191

55. Bianchi, E.; Ciferri, A.; Conio, G.; Coseni, A.; Terbojevick, M. Mesophase

Formation and Chain Rigidity in Cellulose and Derivatives. 4. Cellulose in N,N-

Dimethylacetamide LiCl. Macromolecules 1985, 18(4): 646-650.

56. Dong, X.M.; Revol, J.-F.; Gray, D.G.; Effect of Microcrystalline Preparation

Conditions on the Formation of Colloid Crystals of cellulose. Cellulose 1998,

5(1): 19-32.

57. Fleming, K.; Gray D.G.; Matthews, S. Cellulose Crystallites. Chem. Eur. J.

2001, 7(9): 1831-1835.

58. Werbowyj, R.S.; Gray, D.G. Liquid-Crystalline Structure in Aqueous

Hydroxypropyl Cellulose Solution. Mol. Cryst. Liq. Cryst. 1976, 34(4): 97-103.

59. Guo, J.-X. and Gray, D.G. “Lyotropic cellulosics Liquid Crystals”, in Cellulosic

Polymers, Blends and Composites, Gilbert, R.D. Ed., Hanser, N.Y., 1994, p.25-

46.

60. Zugenmaier, P., “Cellulosic Liquid Crystals”, in Handbook of Liquid Crystals.

Vol. 3, High Molecular Weight Liquid Crystals, Chapter IX, Demus, D.,

Goodby, J., Gray, G.W., Spiess, H.-W., Vill, V. Ed., Wiley-VCH, N.Y., 1998, p.

453-482.

61. Zugenmaier, P. “Polymer Solvent Interactions in Lyotropic Liquid Crystalline

Cellulose Derivatives Systems, in Cellulosic Polymers, Blends and Composites,

Gilbert, R.D. Ed., Hanser, N.Y., 1994, p.71-94.

62. Zugenmaier, P.; Haurand, P. Structural and Rheological Investigations on the

Lyotropic, Liquid Crystalline System: O-Ethylcellulose-Acetic acid-

dichloroacetic acid. Carbohy. Res. 1987, 160: 369-380.

Page 207: Rheology and Microstructure of Cellulose Acetate - Digital Repository

192

63. Zhao, C.T.; Zhang, G.L.; Cai, B.L.; Xu, M. Solvent Composition Dependence of

Band Morphology in Sheared Lyotropic Ethylcellulose Liquid Crystals.

Macromol. Chem. Phys. 1998, 199(8): 1485-1488.

64. Guo, J.X.; Gray, D.G. Effect of Degree of Acetylation and Solvent on the

Chiroptical Properties of Lyotropic (Acetyl)(ethyl)cellulose Solutions. J. Polym.

Sci., Polym. Phys. 1994, 32(15):2529-2537.

65. Gray, D.G. and Harkness, B.R., “Chiral nematic mesophase of lyotropic and

thermotropic cellulose derivatives”, in Liquid Crystalline and Mesomorphic

Polymers, V.P. Shibev and L. Lam Ed., 1994, Springer-Verlag, N.Y., p298-323.

66. Dong, X.M.; Gray, D.G. Induced Circular Dichroism of Isotropic and

Magnetically-Oriented Chiral Nematic Suspensions of Cellulose Crystallites.

Langmuir 1997, 13(11): 3029-3034.

67. Guo, J.X.; Gray, D.G. Effect of Degree of Acetylation and Solvent on the

Chiroptical Properties of Lyotropic (Acetyl)(Ethyl)Cellulose Solutions. J.

Polym. Sci., Polym. Chem. 1994, 32: 889-896.

68. Pawlowski, W.P.; and Gilbert, R.D. The Liquid-Crystalline Properties of

Selected Cellulose Derivatives. J. Polym. Sci., Polym. Phys. 1988, 26: 1101-

1110.

69. Klohr, E.; Zugenmaier, P. Polymer-Solvent Interaction of Molecular Dispersed

and Supramolecular structures of Cellulose Urethanes. Macromol. Symp. 1997,

120: 219-230.

Page 208: Rheology and Microstructure of Cellulose Acetate - Digital Repository

193

70. Harkness, B.R.; Gray, D.G. Left-Handed and Right-Handed Chiral Nematic

Mesophase of (Trityl)(alkyl)cellulose Derivatives. Can. J. Chem. 1990, 68(7):

1135-1139.

71. Nishio, Y.; Kai, T. Controlling the Selective Light Reflection of a Cholesteric

Liquid Crystal of (Hydroxypropyl)cellulose by Electrtical Stimulation.

Macromolecules 1998, 31(7): 2384 -2386.

72. Wang, L.; Huang, Y. Structure Characteristics and defects in Ethyl-Cyanoethyl

Cellulose / Acrylic Acid Cholesteric Liquid Crystalline System.

Macromolecules 2000, 33(19): 7062-7065.

73. Kondo, T.; Miyamoto, T. The Influence of Intramolecular Hydrogen Bonds on

Handedness in Ethylcellulose/CH2Cl2 Liquid Crystalline Mesophases. Polymer

1998, 39(5): 1123-1127.

74. Siekmeyer, M.; Zugenmaier, P. Solvent dependence of Lyotropic Liquid-

Crystalline Phases of Cellulose Tricarbanilate. Macromol. Chem., 1990, 191(5):

1177-1196.

75. Taratuta, V.G.; Srajer, G.M.; Meyer, R.B. Parallel Alignment of Poly(y-Benzyl

Glutamate) Nematic Liquid-Crystal at a Solid-Surface. Mol. Cryst. Liq. Cryst.

1985, 116(3-4): 245-252.

76. Cowie, J.M.G.; Rodden, G.I. Blending as a Method of Tuning reflection

Wavelength and Helical Twisting Sense in Films and Composites of Liquid

Crystalline Cellulose Derivatives. Polymer 2002, 43(12): 3415-3419.

Page 209: Rheology and Microstructure of Cellulose Acetate - Digital Repository

194

77. Navard, P. and Haudin, J.M, “Rheological behavior of Isotropic and Lyotropic

Solutions of Cellulose”, in Cellulose: structure, modification, and hydrolysis,

Young, R.A. and Rowell, R.M. Ed., Wiley, N.Y., 1985, pp.247-261.

78. Santamaria, A.; Lizaso, M.I.; Munoz, M.E. Rheology of Ethylcellulose

Solutions. Macromol. Symp.1997, 114:109-119.

79. Lizaso, M.I.; Munoz, M.E.; Santamaria, A. Transient Rheological behavior of

Lyotropic Solutions of Ethyl Cellulose in m-Cresol. Rheol. Acta 1999, 38: 108-

116.

80. Bercea, M.; Navard, P. Shear Dynamics of Aqueous Suspension of Cellulose

Whiskers. Macromolecules 2000, 33(16): 6011-6016.

81. Onogi, S. and Asada, T., Rheology and Rheo-Optics of Polymer Liquid

Crystals”, in Rheology, Ed. Astarita, G. and Nicolais, Plenum, N.Y., 1980, pp.

127-147.

82. Walker, L.; Wagner, N. Rheology of region I flow in a Lyotropic Liquid-Crystal

Polymer: The Effects of Defect Texture. J. Rheol. 1994, 48(5): 1525-1547.

83. Metzner, A.B.; Priilutski, G.M. Rheological Properties of Polymeric Liquid-

Crystals. J. Rheol. 1986, 30(2): 661-691.

84. Ugaz, V.M.; Cinader, D.K.; Burghardt, W.R. Origins of Region I Shear Shear

Thinning in Model Lyotropic Liquid Crystalline Polymers. Macromolecules

1997, 30(5): 1527-1530.

85. Hongladarom, K.; Secakusuma, V.; Burghardt, W.R. Relation between

Molecular Orientation and Rheology in Lyotropic Hydroxypropylcellulose

Solutions. J. Rheol. 1994, 38(5): 1505-1523.

Page 210: Rheology and Microstructure of Cellulose Acetate - Digital Repository

195

86. Caputo, F.E.; Burghardt, W.R. Real-Time 1-2 Plane SAXS measurements of

Molecular Orientation in Sheared Liquid Crystalline Polymers. Macromolecules

2001, 34(19): 6684-6694.

87. Ernst, B. and Navard, P. Band Textures in Mesomorphic

(Hydroxypropyl)cellulose Solutions. Macromolecules 1989, 22:1419-1422.

88. Grizzuti, N.; Cavella, S.; Cicarelli, P. Transient and Steady-State Rheology of

Liquid Crystalline Hydroxypropylcellulose Solution. J. Rheol. 1990, 34(8):

1293-1310.

89. Baek, S.-G.; Magda, J.J.; Larson, R.G.; Hudson, S.D. Rheological Differences

among Liquid-Crystalline Polymers. II. Disappearance of Negative N1 in

Densely Packed Lyotropes and Thermotropes. J. Rheol. 1994, 38(5): 1473-1503.

90. Baek, S.-G.; Magda, J.J.; Cementwala, S. Normal Stress Difference in Liquid-

Crystalline Hydroxypropylcellulose Solutions. J. Rheol. 1993, 37(5): 935-945.

91. Mewis, J.; Moldenaers, P. Transient Rheological behavior of a Lyotropic

Polymeric Liquid-Crystal. Mol. Cryst. Liq. Cryst. 1987, 153:291-300.

92. Vermant, J.; Mortier, M.; Moldenaers, P.; Mewis, J. An evaluation of the

Larson-Doi Model for Liquid Crystalline Polymers Using Recoil. Rheol. Acta

1999, 38: 537-547.

93. Asada, T.; Toda, K.; Onogi, S. Deformation and Structural Re-formation of

Lyotropic Cholesteric Liquid Crystal of Hydroxylpropeycellulose + Water

System. Mol. Cryst. Liq. Cryst. 1981, 68:231-246.

Page 211: Rheology and Microstructure of Cellulose Acetate - Digital Repository

196

94. Grizzuti, N.; Moldnaers, P.; Mortier, M.; Mewis, J. On the time-Dependency of

the Flow-Induced Dynamic Moduli of a Liquid Crystalline

Hydroxypropylcellulose Solution. Rheol. Acta 1993, 32(3): 218-226.

95. Burghardt, W.R. Molecular Orientation and Rheology in Sheared Lyotropic

Liquid Crystalline Polymers. Macromol. Chem. Phys. 1998, 199:471-488.

96. Godinho, M.H.; van der Klink, J.J.; Martins, A.F. Shear-History Dependent

‘Equilibrium’States of Liquid-Crystalline Hydroxypropylcellulose Solutions

Detected by Rheo-Nuclear Magnetic Resonance. J. Phys.: Condens. Matter

2003, 15(32): 5461-5468.

97. Vermant, J.; Moldenaers, P.; Mewis, J.; Picken, S.J. Band Formation upon

Cessation of Flow in Liquid-Crystalline Polymers. J. Rheol. 1994, 38(5): 1571-

1589.

98. Harrison, P.; Navard, P. Investigation of the Band texture Occurring in

Hydroxypropylcellulose Solutions Using Rheol-Optical, Rheological and Small

Angle Light Scatterign Techniques. Rheol. Acta 1999, 38(6): 569-593.

99. Riti, J.-B.; Navard, P. Texture during Recoil of Anisotropic

Hydrosypropylcellulose Solutions. J. Rheol. 1998, 42(2): 225-237.

100. Shimamoto, S.; Uraki, Y.; Sano, Y. Optical Properties and Photopolymerization

of Liquid Crystalline (Acetyl)(ethyl)cellulose/Acrylic Acid System. Cellulose.

2000, 7(4): 347-358.

101. Mercer, H.A.; Weymann, H.D. Structure of Rhixotropic Suspensions in Shear-

Flow. 3. Time-Dependent Behavior. Trans. Soc. Rheol. 1974, 18(2): 199-218.

Page 212: Rheology and Microstructure of Cellulose Acetate - Digital Repository

197

102. Suto, A.; Sasaki, K.; Tateyama S. Stress-relaxation behavior of Liquid-

Crystalline Solutions of Ethyl Cellulose with Different Molecular-Weight in –

m-Cresol. Angew. Makromol.Chem. 1990, 179: 203-216.

Page 213: Rheology and Microstructure of Cellulose Acetate - Digital Repository

198

Shear rate (s-1)

10-3 10-2 10-1 100 101 102 103

Vis

cosi

ty (P

a.s)

100

101

0%

3.6%

7.2%

10.8%

14.4%

Figure 1. Effect of different water contents on the steady shear viscosity of cellulose

acetate in DMAc/water cosolvent system. The total concentration of cellulose acetate is

kept constant at 10wt%.

Page 214: Rheology and Microstructure of Cellulose Acetate - Digital Repository

199

Shear rate (1/s)10-3 10-2 10-1 100 101 102 103 104

Vis

cosi

ty (P

a.s)

10-3

10-2

10-1101

102

103

104

105

106

Figure 2. Viscosity pattern for the two-phase region for a constant 10wt% CA in

DMAc/water system.

Page 215: Rheology and Microstructure of Cellulose Acetate - Digital Repository

200

Shear rate (s-1)

10-3 10-2 10-1 100 101 102 103

Vis

cosi

ty (P

a.s)

100

101

14.4%

10.8%

7.2%

3.6%0.0%

Figure 3. Effect of different water contents on steady shear viscosity for a cellulose

acetate/DMAc/water system. CA concentration is 10wt%.

Page 216: Rheology and Microstructure of Cellulose Acetate - Digital Repository

201

Frequency (rad/s)

10-3 10-2 10-1 100 101 102

G' &

G"(

Pa)

10-4

10-3

10-2

10-1

100

101

102

103

104

105

G' (21.6%)

G" (21.6%)

G" (7.2%)

G' (7.2%)

Figure 4. Effect of low and high water contents on the viscous (G″) and elastic (G′)

moduli for 10wt% cellulose acetate/DMAc/water system.

Page 217: Rheology and Microstructure of Cellulose Acetate - Digital Repository

202

Figure 5. Helicoidal structure of chiral nematic mesophase

Page 218: Rheology and Microstructure of Cellulose Acetate - Digital Repository

203

Figure 6. Viscosity as a function of concentration for lyotropic LCPs

Page 219: Rheology and Microstructure of Cellulose Acetate - Digital Repository

204

Figure 7. Three-regime steady state shear viscosity for LCPs

Page 220: Rheology and Microstructure of Cellulose Acetate - Digital Repository

205

Figure 8. Viscosity as a function of shear rate of AEC/AA solutions

10

100

1000

0.1 1 10 100

Shear Rate (s-1)

Vis

cosi

ty (P

a*s)

46%AEC-148%AEC-350%AEC-451.5%AEC-2

Page 221: Rheology and Microstructure of Cellulose Acetate - Digital Repository

206

Figure 9. Reduced shear stress σ* vs. time for 50% AEC-4( ) and 51.5% AEC-

2(∆)

AEC-4R2 = 0.9908

R2 = 0.9711AEC-2

-8.0

-7.0

-6.0

-5.0

-4.0

-3.0

-2.0

0 10 20 30 40 50 60

Time (s)

Red

uced

She

ar S

tres

s

Page 222: Rheology and Microstructure of Cellulose Acetate - Digital Repository

207

Figure 10. Complex modulus evolution of AEC solutions vs. time

1.E+02

1.E+03

1.E+04

1.E+05

1.E+06

1.E+01 1.E+02 1.E+03 1.E+04

Time (s)

|G*|

(Pa)

AEC-3

AEC-4

AEC-2

AEC-1

Page 223: Rheology and Microstructure of Cellulose Acetate - Digital Repository

208

Table 1. Chiroptical properties of lyotropic AEC/AA solutions with different degree

of acetylation (DA)

AEC Sample Sample Description Degree of Acetylation Pitch*

AEC-1 Low DA 0.13 - 360nm

AEC-2 Medium DA 0.34 - 7.4µ

AEC-3 AEC/EC Mixture 0.34 + 1.2µ

AEC-4 High DA 0.50 + 370nm

+: right-handed; -: left-handed

Page 224: Rheology and Microstructure of Cellulose Acetate - Digital Repository

209

Table 2. Relaxation times as the function of chiroptical properties of lyotropic

AEC/AA solutions

Sample 46% AEC-1 51.5% AEC-2 48% AEC-3 50% AEC-4

Pitch - 500nm - 5.9µ + 1.3µ + 370nm

Relax. Time (sec) 14.5 28.4 16.7 14.9