experimental and modeling investigation of cellulose

194
Purdue University Purdue e-Pubs Open Access Dissertations eses and Dissertations Spring 2015 Experimental and modeling investigation of cellulose nanocrystals polymer composite fibers Si Chen Purdue University Follow this and additional works at: hps://docs.lib.purdue.edu/open_access_dissertations Part of the Chemical Engineering Commons , and the Materials Science and Engineering Commons is document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] for additional information. Recommended Citation Chen, Si, "Experimental and modeling investigation of cellulose nanocrystals polymer composite fibers" (2015). Open Access Dissertations. 435. hps://docs.lib.purdue.edu/open_access_dissertations/435

Upload: others

Post on 20-Apr-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Experimental and modeling investigation of cellulose

Purdue UniversityPurdue e-Pubs

Open Access Dissertations Theses and Dissertations

Spring 2015

Experimental and modeling investigation ofcellulose nanocrystals polymer composite fibersSi ChenPurdue University

Follow this and additional works at: https://docs.lib.purdue.edu/open_access_dissertations

Part of the Chemical Engineering Commons, and the Materials Science and EngineeringCommons

This document has been made available through Purdue e-Pubs, a service of the Purdue University Libraries. Please contact [email protected] foradditional information.

Recommended CitationChen, Si, "Experimental and modeling investigation of cellulose nanocrystals polymer composite fibers" (2015). Open AccessDissertations. 435.https://docs.lib.purdue.edu/open_access_dissertations/435

Page 2: Experimental and modeling investigation of cellulose

Graduate School Form 30Updated 1/15/2015

Page 3: Experimental and modeling investigation of cellulose
Page 4: Experimental and modeling investigation of cellulose

i

EXPERIMENTAL AND MODELING INVESTIGATION OF CELLULOSE

NANOCRYSTAL POLYMER COMPOSITE FIBERS

A Dissertation

Submitted to the Faculty

of

Purdue University

by

Si Chen

In Partial Fulfillment of the

Requirements for the Degree

of

Doctor of Philosophy

May 2015

Purdue University

West Lafayette, Indiana

Page 5: Experimental and modeling investigation of cellulose

ii

Page 6: Experimental and modeling investigation of cellulose

iii

ACKNOWLEDGEMENTS

I would like to express my sincere gratitude to my advisors, Prof. Byron Pipes,

Prof. Jeffrey Youngblood, and Prof. Robert Moon for their invaluable guidance and

support: Prof. Pipes for his continued encouragement, broadening the scope of my work,

and personal and professional mentoring. Prof. Youngblood for our caffeine-driven

insightful discussions. Prof. Moon for his insight and helpful suggestions, as well as

teaching me to be prepared for Murphy’s Law.

I would like to thank my committee members, Prof. James Caruthers, and Prof.

Michael Harris, for their valuable expertise, discussions, and recommendations.

I would also like to thank Dr. Greg Schueneman, Steve Lacher and Dave Eustice

at the USDA Forest Products Laboratory, for their expertise and experimental support;

and Materials Engineering department at Purdue University for support and the usage of

characterization equipment.

Finally, I wish to thank my friends and family, especially my parents, and my

fiancé, Nathan, for their love and support.

This work has been supported by The U.S. Department of Defense, Air Force

Office of Scientific Research (Grant No. FA9550-11-1-0162) and USDA Forest Service

(Grant number 11-JV-11111129-118).

Page 7: Experimental and modeling investigation of cellulose

iv

TABLE OF CONTENTS

Page

LIST OF TABLES ........................................................................................................... viii

LIST OF FIGURES ........................................................................................................... ix

NOMENCLATURE ........................................................................................................ xiii

ABSTRACT ................................................................................................................... xviii

CHAPTER 1. INTRODUCTION ................................................................................. 1

1.1 Motivation for Work ............................................................................................ 1

1.2 Research Objectives ............................................................................................. 4

1.3 Organization of This Work .................................................................................. 5

Bibliography ................................................................................................................... 6

CHAPTER 2. EFFECTS OF CELLULOSE NANOCRYSTALS ON RANDOMLY ORIENTED AND ALIGNED CELLULOSE ACETATE ELECTROSPUN FIBERS ..... 7

2.1 Introduction .......................................................................................................... 7

2.2 Methods and Materials ....................................................................................... 12

2.2.1 Materials ..................................................................................................... 12

2.2.2 Experimental Setup ..................................................................................... 13

2.2.3 Characterization of Nanofibers ................................................................... 15

2.3 Discussion and Results ....................................................................................... 16

2.3.1 CNC Dispersion .......................................................................................... 16

2.3.2 Surface Morphology of Electrospun CA-CNC Fibers ................................ 19

2.3.3 Thermal Properties of Electrospun CA-CNC Fibers .................................. 23

2.3.4 Mechanical Properties of Electrospun CA-CNC Fibers ............................. 23

2.4 Conclusion .......................................................................................................... 26

Bibliography ................................................................................................................. 27

Page 8: Experimental and modeling investigation of cellulose

v

Page

CHAPTER 3. EFFECTS OF CRYSTAL ORIENTATION ON CELLULOSE NANOCRYSTALS-CELLULOSE ACETATE NANOCOMPOSITE FIBERS PREPARED BY DRY SPINNING .................................................................................. 29

3.1 Introduction ........................................................................................................ 30

3.2 Experimental Section ......................................................................................... 36

3.2.1 Materials ..................................................................................................... 36

3.2.2 Preparation of CNC-CA Solutions.............................................................. 37

3.2.3 Preparation of CNC-CA Fibers ................................................................... 37

3.2.4 Characterization of CA-CNC Composite Fiber .......................................... 38

3.2.4.1 Rheology .............................................................................................. 38

3.2.4.2 Microscopy .......................................................................................... 41

3.2.4.3 2D X-ray Diffraction ........................................................................... 41

3.2.4.4 Mechanical Testing .............................................................................. 43

3.2.4.5 Micromechanical Modeling ................................................................. 43

3.3 Results and Discussion ....................................................................................... 45

3.3.1 Suspension Properties and Fiber Spinning ................................................. 45

3.3.2 Fiber Morphologies ..................................................................................... 47

3.3.3 Birefringence of Fibers ............................................................................... 49

3.3.4 CNC Orientation ......................................................................................... 51

3.3.5 Fiber Mechanical Properties ....................................................................... 56

3.3.6 Micromechanical Modeling of CA-CNC Nanocomposite Fibers .............. 60

3.4 Conclusion .......................................................................................................... 63

Bibliography ................................................................................................................. 64

CHAPTER 4. DRY SPUN POLYVINYL ALCOHOL FIBERS REINFORCED WITH CELLULOSE NANOCRYSTALS ....................................................................... 69

4.1 Introduction ........................................................................................................ 69

4.2 Materials and Methods ....................................................................................... 73

4.2.1 Materials ..................................................................................................... 73

4.2.2 Methods....................................................................................................... 73

4.2.2.1 Preparation of PVA-CNC Fibers ......................................................... 73

4.2.2.2 Microscopy of PVA-CNC Fibers ........................................................ 74

4.2.2.3 Mechanical Testing of PVA-CNC Fibers ............................................ 75

4.2.2.4 Micromechanical Modeling of PVA-CNC Fibers ............................... 75

Page 9: Experimental and modeling investigation of cellulose

vi

Page

4.3 Results and Discussion ....................................................................................... 76

4.3.1 Morphologies and Surface Structure of PVA-CNC Fiber .......................... 76

4.3.2 Mechanical Properties of PVA-CNC Fibers ............................................... 79

4.3.3 Micromechanical Modeling of PVA-CNC Fibers ...................................... 86

4.4 Conclusion .......................................................................................................... 87

Bibliography ................................................................................................................. 88

CHAPTER 5. MODELING OF DRY SPUN POLYMER - CELLULOSE NANOCRYSTAL COMPOSITE FIBERS....................................................................... 91

5.1 Introduction ........................................................................................................ 91

5.2 Model Development ........................................................................................... 95

5.2.1 Model Assumptions .................................................................................... 95

5.2.2 Transport Equations .................................................................................... 97

5.2.2.1 Equation of Continuity ........................................................................ 97

5.2.2.2 Equation of Momentum ....................................................................... 98

5.2.2.3 Equation of Energy .............................................................................. 99

5.2.3 Constitutive Model.................................................................................... 100

5.2.4 Material Properties and Correlations ........................................................ 104

5.2.4.1 Density ............................................................................................... 104

5.2.4.2 Heat and Mass Transfer Coefficients ................................................ 104

5.2.4.3 Activity Coefficient ........................................................................... 105

5.2.4.4 Vapor Pressure ................................................................................... 106

5.2.4.5 Air Drag Coefficient .......................................................................... 106

5.2.4.6 Heat Capacity and Latent Heat of Evaporation ................................. 106

5.2.4.7 Glass Transition Temperature ........................................................... 107

5.2.4.8 Solution Viscosity.............................................................................. 108

5.2.5 Experimental Materials and Methods ....................................................... 109

5.2.5.1 Materials ............................................................................................ 109

5.2.5.2 Preparation of CA-CNC Fibers ......................................................... 109

5.2.6 Spinning Parameters ................................................................................. 110

5.2.7 Numerical Method .................................................................................... 110

5.3 Results and Discussion ..................................................................................... 113

5.3.1 Newtonian Model Prediction .................................................................... 113

Page 10: Experimental and modeling investigation of cellulose

vii

Page

5.3.2 Effect of Spinning Parameters .................................................................. 120

5.3.2.1 Mass Flow Rate ................................................................................. 123

5.3.2.2 Temperature of Spinning Dope and Air ............................................ 123

5.3.2.3 Winding Speed and Air Velocity ....................................................... 131

5.3.3 Effect of Fillers ......................................................................................... 135

5.3.3.1 Effect of Collimated Fibers ............................................................... 135

5.3.3.2 Effect of fiber orientation distribution ............................................... 142

5.3.3.3 Comparison to experimental data ...................................................... 149

5.4 Conclusion ........................................................................................................ 149

Bibliography ............................................................................................................... 151

CHAPTER 6. CONCLUSIONS AND FUTURE WORK ........................................ 154

6.1 Electrospun CA-CNC Fibers ............................................................................ 154

6.2 Effects of CNC on Dry Spun CA-CNC Fibers ................................................ 155

6.3 Dry Spun Polyvinyl Alcohol Fibers ................................................................. 156

6.4 Modeling of Dry Spun Fibers .......................................................................... 157

APPENDICES

Appendix A: Equation of Continuity .......................................................................... 160

Appendix B: Equation of Momentum ........................................................................ 163

Appendix C: Equation of Energy ................................................................................ 167

Bibliography ............................................................................................................... 169

VITA ............................................................................................................................... 170

Page 11: Experimental and modeling investigation of cellulose

viii

LIST OF TABLES

Table .............................................................................................................................. Page

Table 2-1 Summary of various published electrospun polymer-CNC nanocomposite fiber studies ....................................................................................................................... 10

Table 3-1 Processing parameters for CNC-CA composite fibers ..................................... 40

Table 3-2 Consistency coefficient (K), flow behavior index (n) as determined by the Power Law model for different experiment samples ........................................................ 48

Table 3-3 Diameter, mechanical properties and order parameters of CNC-CA composite fibers ................................................................................................................ 59

Table 5-1 Processing conditions for CNC-CA composite fibers .................................... 111

Table 5-2 Material Properties of CA-CNC Composite Fibers........................................ 112

Page 12: Experimental and modeling investigation of cellulose

ix

LIST OF FIGURES

Figure ............................................................................................................................. Page

Figure 1-1 (a) Chemical Structure of the cellulose unit, (b) representative cellulose micro

fibril section (c) cellulose nanocrystals after removal of amorphous regions. ................... 3

Figure 2-1 TEM Image of CNC suspension deposited on a TEM grid ............................ 14

Figure 2-2 Flow Birefringence of dried CNC solutions under crosspolarizers ................ 17

Figure 2-3 (a) Crossed polarizer image of solvent transferred CNC ................................ 18

Figure 2-4 SEM Images of Pure Cellulose Acetate Fibers: (a) random fiber mat and (b) aligned fibers ............................................................................................................... 20

Figure 2-5 CA-CNC electrospun fiber showing defects such as large beads and inconsistent diameters ....................................................................................................... 21

Figure 2-6 SEM Image of a 1 wt% CNC CA Fiber. Surface cracking due to the electron beam is observed on the surface ....................................................................................... 22

Figure 2-7 TGA thermograms of the CA-CNC electrospun fibers ................................... 24

Figure 2-8 Mechanical Properties of CA-CNC electrospun fibers as a function of CNC concentration ..................................................................................................................... 25

Figure 3-1 Overview of elastic modulus vs. tensile strength of various CN-polymer filaments reported in literature and other common filament materials. Colored regions indicate CN studies ........................................................................................................... 35

Figure 3-2 Schematic of the dry fiber spinning experimental setup ................................. 39

Figure 3-3 Viscosity vs. shear rate data of various solutions.. ......................................... 46

Figure 3-4 Optical microscopy image of (a) 17 wt% CNC-CA fiber, SEM images of (b) CA fiber (c) 10 wt% CNC-CA fiber (d) 44 wt% CNC-CA fiber ................................ 50

Figure 3-5 Image of CNC-CA fibers between cross polarizers ........................................ 52

Page 13: Experimental and modeling investigation of cellulose

x

Figure ............................................................................................................................. Page

Figure 3-6 2DXRD diffractograms of (a) a solution cast neat CNC film with randomly oriented CNCs , (b) shear cast neat CNC films with highly aligned CNC, and (c-f) are CNC-CA fibers with various CNC contents ..................................................................... 54

Figure 3-7 Herman’s order parameter as a function of CNC content in fiber .................. 55

Figure 3-8 Mechanical properties of as spun CNC-CA fibers: a) Young’s modulus, b) tensile strength and c) elongation, at break as a function of CNC%. ........................... 58

Figure 3-9 Predictions of the Young’s modulus using isostrain, isostress, Halpin-Tsai equation and the modified orientation Cox model with aspect ratios of 5 and 15 plotted along with experimental data. ............................................................................... 61

Figure 4-1 (a)-(f) Optical microscopy image of PVA-CNC fibers with CNC concentrations of 0, 1, 2, 7, 15, 25 wt% respectively. ...................................................... 77

Figure 4-2 (a)-(f) Crossed polarizer optical microscopy image of PVA-CNC fibers with various CNC concentrations. (g) shows a 25 wt% PVA-CNC fiber exhibiting the “skinning” effect ......................................................................................................... 78

Figure 4-3 SEM images of (a) PVA fiber, (b) 2 wt% PVA-CNC fiber (c) 20 wt% PVA-CNC fiber (d) 25 wt% PVA-CNC fiber ............................................................................ 80

Figure 4-4 Mechanical properties of as spun PVA-CNC fibers: a) Young’s modulus, b) elongation at break and c) tensile strength, as a function of CNC %. .......................... 81

Figure 4-5 Examples of static force vs. strain curves of PVA-CNC fibers ...................... 83

Figure 4-6 Predictions of the Young’s modulus using isostrain, isostress, Halpin-Tsai equation and the Cox model with aspect ratios of 15 plotted along with experimental data. ................................................................................................................................... 85

Figure 5-1 Schematic representation of a dry spinning system ........................................ 96

Figure 5-2 Diameter profile as a function of distance from the spinneret for two cases: fibers with and without drawing ..................................................................................... 114

Figure 5-3Velocity profile as a function of distance from the spinneret for two cases: fibers with and without drawing ..................................................................................... 115

Figure 5-4 Concentration fraction profile as a function of distance from the spinneret for two cases: fibers with and without drawing. ND and D stands for non-drawing and drawing respectively. ...................................................................................................... 116

Page 14: Experimental and modeling investigation of cellulose

xi

Figure ............................................................................................................................. Page

Figure 5-5 Temperature profile as a function of distance from the spinneret for two cases: fibers with and without drawing ..................................................................................... 117

Figure 5-6 Viscosity profile as a function of distance from the spinneret for two cases: fibers with and without drawing ..................................................................................... 118

Figure 5-7 Glass transition temperature profile as a function of distance from the spinneret for two cases: fibers with and without drawing .............................................. 119

Figure 5-8 Diameter profile as a function of distance from the spinneret for cases with varying mass flow rate .................................................................................................... 121

Figure 5-9 Concentration profile as a function of distance from the spinneret for cases with varying mass flow rate ............................................................................................ 122

Figure 5-10 Diameter profile as a function of distance from the spinneret for cases with varying spinning dope temperature ................................................................................. 124

Figure 5-11 Concentration profile as a function of distance from the spinneret for cases with varying spinning dope temperature......................................................................... 125

Figure 5-12 Diameter profile as a function of distance from the spinneret for cases with varying quenched air temperature ................................................................................... 126

Figure 5-13 Velocity profile as a function of distance from the spinneret for cases with varying quenched air temperature ................................................................................... 127

Figure 5-14 Temperature profile as a function of distance from the spinneret for cases with varying quenched air temperature........................................................................... 128

Figure 5-15 Diameter profile as a function of distance from the spinneret for cases with varying take-up velocity. ................................................................................................ 129

Figure 5-16 Velocity profile as a function of distance from the spinneret for cases with varying take-up velocity. ................................................................................................ 130

Figure 5-17 Diameter profile as a function of distance from the spinneret for cases with varying quenched air velocity. ........................................................................................ 132

Figure 5-18 Velocity profile as a function of distance from the spinneret for cases with varying quenched air velocity. ........................................................................................ 133

Figure 5-19 Temperature profile as a function of distance from the spinneret for cases with varying quenched air velocity. ................................................................................ 134

Page 15: Experimental and modeling investigation of cellulose

xii

Figure ............................................................................................................................. Page

Figure 5-20 Diameter profile as a function of distance from the spinneret for cases with varying final CNC volume concentration. ...................................................................... 136

Figure 5-21 Velocity profile as a function of distance from the spinneret for cases with varying final CNC volume concentration. ...................................................................... 137

Figure 5-22 Concentration profile as a function of distance from the spinneret for cases with varying final CNC volume concentration. (a)-(f) refer to a volume fraction of 0%, 8.4%, 17.1%, 25.2%, 34.5%, and 44.2% respectively. ................................................... 138

Figure 5-23 Temperature profile as a function of distance from the spinneret for cases with varying final CNC volume concentration. .............................................................. 139

Figure 5-24 Glass transition temperature profile as a function of distance from the spinneret for cases with varying final CNC volume concentration. ............................... 140

Figure 5-25 Normalized probability distribution for filler orientation for 4 different experimental composite fibers: 44.2%, 25.2%, 17.1% and 8.4% CNC respectively. .... 143

Figure 5-26 Effect of fiber orientation on elongational viscosity ................................... 144

Figure 5-27 Effect of fiber orientation on diameter profiles ........................................... 146

Figure 5-28 Effect of fiber orientation on velocity profiles ............................................ 147

Figure 5-29 Diameter Comparison between theoretical and experimental data ............. 148

Page 16: Experimental and modeling investigation of cellulose

xiii

NOMENCLATURE

A Cross Sectional Area of the Fiber

as Activity Coefficent of the Solvent

ci Volume Fraction of the Component i

Cf Coefficient of Friction

Cp Heat Capacity

Cpi Heat Capacity of the Component i

oPsC Ideal Gas State Heat Capacity

CA Cellulose Acetate

CNC Cellulose Nanocrystal

D Diameter

DMAc N,N-dimethylacetamide

DMA Dynamic Mechanical Analysis

DMF Dimethylformamide

DSC Differential Scanning Calorimetry

E Activation Energy

E Elastic Modulus

Ei Elastic Modulus of the Component i

F Geometric Constant used in Equation 5.12

Page 17: Experimental and modeling investigation of cellulose

xiv

g Gravity

gps Cellulose Acetate-Acetone Interaction Parameter

Gm Shear Stiffness of the Matrix

v Molar Latent Heat of Vaporization

h Heat Transfer Coefficient

K Consistency Index

K Ratio of Difference in Thermal Expansivity

ka Thermal Conductivity of Air

ky Mass Transfer Coefficient

L Length

Mi Molecular Weight of the Component i

Nu Nusselt Number

n Flow Behavior Index

satsP Saturated Partial Pressure of the Solvent

P Pressure

Pn Degree of Polymerization

PEO Polyethylene Glycol

PMMA Poly(Methyl Methacrylate)

PVA Polyvinyl Alcohol

R Radius

R Universal Gas Constant

Re Reynolds Number

Page 18: Experimental and modeling investigation of cellulose

xv

S Herman’s Order Parameter

SEM Scanning Electron Microscopy

T Temperature

Ta Temperature of Air

Tc Critical Temperature

Tg Glass Transition Temperature

Tgi Glass Transition Temperature of the Component i

Tr Reduced Temperature

Ts Standard Temperature

TEM Transmission Electron Microscopy

THF Tetrahydrofuran

TGA Thermogravimetric Analysis

Vi Volume Fraction of the Component i

Vi Molar Volume of the Component i

vz Axial Velocity

va Air Velocity

W Mass Flow Rate

Wi Mass Flow Rate of the Component i

wi Weight Fraction of the Component i

,sy

Greek

Shear Rate

Page 19: Experimental and modeling investigation of cellulose

xvi

F Elongation at Failure

Viscosity

o Zero Shear Viscosity

0 Fiber Orientation Efficiency Factor in Equation 3.8

1 Fiber Length Efficiency Factor in Equation 3.8

11 Axial Elongational Viscosity

12 In-Plane Shearing Viscosity

22 Transverse Elongational Viscosity

a Viscosity of Air

Constant Controlled by Geometric Packing in Equation 3.9

Parameter Used in Pipes Model defined by Equation 5.12

i Partial Specific Volume of the Component i

3.14159265359…

Density

a Density of Air

i Density of the Component i

F Ultimate Tensile Strength

Stress Tensor

ij Stress Tensor in the ij Element

Acentric Factor

Page 20: Experimental and modeling investigation of cellulose

xvii

Subscripts

a Air

f Filler

m Matrix

p Polymer

r r Direction in Cylindrical Coordinates

s Solvent

z z Direction in Cylindrical Coordinates

Direction in Cylindrical Coordinates

Page 21: Experimental and modeling investigation of cellulose

xviii

ABSTRACT

Chen, Si. Ph.D., Purdue University, May 2015. Experimental and Modeling Investigation of Cellulose Nanocrystal Polymer Composite Fibers. Major Professor: R. Byron Pipes.

Cellulose nanocrystals (CNCs) are a class of newly developed and sustainable

nanomaterial derived from cellulose-based materials such as wood. There have been

substantial research efforts to utilize these materials as reinforcing agents. However, in

order to develop CNC nanocomposites with industrial applications, it is necessary to

understand how addition of CNCs affect the properties of the polymer nanocomposite. In

the present work, several approaches, experimental and theoretical, are presented in an

effort to characterize and understand the effect of CNCs on the properties of polymer

CNC fibers.

Two experimental methods were used to develop cellulose acetate (CA) and CNC

composite fibers: electrospinning and dry spinning. Polyvinyl alcohol (PVA) and CNC

dry spun composite fibers were also produced in order to study the effect of CNCs in

additional polymeric systems. Surface morphology of these fibers was analyzed by

scanning electron microscopy and optical microscopy. Elastic modulus, elongation, and

ultimate tensile strength were measured by dynamic mechanical analysis. CNC

dispersion was assessed by optical microscopy under cross polarizers. 2D X-ray

Page 22: Experimental and modeling investigation of cellulose

xix

diffraction was used along with the Herman’s order parameter to quantify the level of

orientation of the CNCs within the dry spun fibers.

It was found that the orientation of the CNCs within the fiber directly correlates to

the mechanical properties of the composite. Continual improvement in the tensile

strength and elastic modulus of the fibers were observed until a critical concentration.

Further increases in CNC concentration did not yield corresponding improvement and led

to large variations in fiber performance, likely due to increased defect density and/or

aggregation of CNCs. Empirical micromechanical models Halpin-Tsai equation and an

orientation modified Cox model were used to predict the fiber performance and compared

with experimental results.

Finally, a one dimensional model is presented using fundamental equations of

momentum, mass and energy, along with a modified Pipes viscous constitutive equation

to predict the diameter, velocity, component concentration, and temperature profile of the

dry spinning system. Sensitivity analysis was performed on various spinning parameters

such as mass flow rate, temperature of the spinning dope and air, the velocity of take-up

speed, and the velocity of the quenched air. In addition, the effect of CNCs’ orientation

on the system was also investigated. It is found that perfectly collimated CNCs increased

the elongational viscosity of the system, thus resulting in a thicker fiber.

The methods and results presented in this work help describe the behavior of

polymer-CNC composites and may serve as a useful tool for process optimization of such

materials. Based on these results, this study shows that for near collimated states of

orientation, the CNCs provide an excellent form of polymer reinforcement and may

contribute to the field as a renewable, lightweight and strong nanomaterial.

Page 23: Experimental and modeling investigation of cellulose

1

CHAPTER 1. INTRODUCTION

1.1 Motivation for Work

Cellulose nanocrystals (CNCs) are a class of newly developed and sustainable

nanomaterial derived from cellulose-based materials such as wood. Reported to have a

greater axial elastic modulus than Kevlar, and a greater tensile strength than steel,1 they

can be used in a wide range of applications including material packing, flexible circuits,

solar panels, and armor for defense application.2 The abundance of their source material

and low cost of production are also attractive traits for industrial scale applications.

CNCs are produced by controlled acid hydrolysis of cellulose-based materials such

as plants. The removal of amorphous regions of the cellulose leaves behind CNCs in the

form of highly crystalline, rod-like particles. (Figure 1.1) The shape, size and dimensions

of the CNCs are primarily determined by the source of cellulose, as well as the hydrolysis

conditions, with diameter ranging from 2-10 nm and length from 50-500 nm.3 Literature

has reported CNCs with an axial elastic modulus between 110-220GPa and strength of

7.5-7.7GPa.1 By aligning these crystals while they are acting as reinforcing agents, the

mechanical properties of the composite in the alignment direction can be greatly

increased. Due to the comparable dimensions between the nanofiller and the polymer

matrix molecules, the nanocomposite material response may differ with the principles of

macro-scale composites.4

Page 24: Experimental and modeling investigation of cellulose

2

However, despite moderate success, the effective polymer reinforcement due to

CNC’s large variability has not been well understood.1

Cellulose acetate (CA) is the acetate ester of cellulose. Since its’ invention in

1905, it has remained a commonly used polymer in clothing, filters and diapers and is

typically produced by dry-spinning. The worldwide market is projected to 1.05 million

metric tons by 2017. Major industrial companies include Celanese Corporation, Daicel

Corporation, Primester, and Eastman Chemical Company among others.5 Polyvinyl

alcohol (PVA) is another commonly used polymer derived from the controlled hydrolysis

of poly(vinyl acetate). PVA is water soluble, biodegradable, nontoxic, and biocompatible,

thus has a wide range of applications from filtration, membranes, to 3D printing. The

fiber form of PVA, sometimes known as vinylon, is widely used in textiles, ropes, and

shoes. Short fibers of PVAs are often used as reinforcement in concrete.6 Incorporating

CNCs into CA and PVA fiber can greatly improve the mechanical properties of the fiber

while still maintaining the eco-compatibility of the polymer to create a light weight and

strong nanocomposite.

Despite the intensive research efforts in CNC nanocomposites in the past decade,5

the number of industrial applications of CNC nanocomposites is relatively small. In order

to develop CA-CNC or PVA-CNC nanocomposites that can find industrial applications,

it is necessary to identify, control and understand how addition of CNCs leads to

improvement of polymer mechanical properties.

Page 25: Experimental and modeling investigation of cellulose
Page 26: Experimental and modeling investigation of cellulose

4

1.2 Research Objectives

The aim of this work is to identify, control and understand how the addition of CNCs

affect polymer-CNC nanocomposite fiber formation and properties, as well as,

identifying the parameters and mechanisms that are fundamental to mechanical property

improvement. To achieve this goal, several research objectives were established:

- Develop CA-CA composite nanofibers using electrospinning. Characterize the

resulting fiber morphology and properties. Identify and optimize variables that

affect mechanical properties of the nanofiber.

- Use CNC as a reinforcing agent in a CA dry-spun fiber using a high shear rate

spinneret. Characterize the resulting fiber morphology and properties.

Characterize the dispersion of CNCs within the polymer matrix, both

qualitatively, and quantitatively. Propose and validate hypothesis and models that

predicts mechanical property improvements by the addition of CNCs to CA.

- Use CNC as a reinforcing agent in a polyvinyl alcohol dry-spun fiber using a high

shear rate spinneret. Characterize the resulting fibers and compare to CA-CNC

composite fibers.

- Develop a model of the ternary dry spinning system that describes the effect of

CNCs on the dynamics of the dry spinning process such as diameter and velocity

profiles.

Page 27: Experimental and modeling investigation of cellulose

5

1.3 Organization of This Work

In this work, each of the chapters will describe one of the research objectives

above. The literature background and state of the art in each topic will be discussed

separately in each topic. Chapter 2 presents the methods used to develop CA-CNC

electrospun nanofibers, and the effects of CNCs on these fibers. Chapter 3 presents the

development of dry spun cellulose acetate (CA) fibers using cellulose nanocrystals

(CNCs) as reinforcements. A systematic characterization of dispersion of CNCs in the

polymer fiber and their effect on the nanocomposites’ mechanical properties is described.

Chapter 4 describes a similar system to which polyvinyl alcohol is used as the matrix. A

one-dimensional model describing the ternary dry spinning system consisting of cellulose

acetate, acetone and cellulose nanocrystals is discussed in Chapter 5. The last chapter

includes a summary of this work and its principal conclusions.

Page 28: Experimental and modeling investigation of cellulose

6

Bibliography

1. Moon, R. J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J., Cellulose nanomaterials review: structure, properties and nanocomposites. Chemical Society Reviews 2011, 40 (7), 3941-3994.

2. Habibi, Y.; Lucia, L. A.; Rojas, O. J., Cellulose Nanocrystals: Chemistry, Self-Assembly, and Applications. Chem Rev 2010, 110 (6), 3479-3500.

3. Beck-Candanedo, S.; Roman, M.; Gray, D. G., Effect of reaction conditions on the properties and behavior of wood cellulose nanocrystal suspensions. Biomacromolecules 2005, 6 (2), 1048-1054.

4. Bhattacharya, S. N.; gupta, R. K.; Kamal, M. R., Polymeric Nanocomposites. Theory and practice. Hanser: Munich, 2008.

5. Konwarh, R.; Karak, N.; Misra, M., Electrospun cellulose acetate nanofibers: The present status and gamut of biotechnological applications. Biotechnol. Adv. 2013, 31 (4), 421-437.

6. Sakurada, I., Polyvinyl Alcohol Fibers. Taylor & Francis: 1985.

Page 29: Experimental and modeling investigation of cellulose

7

CHAPTER 2. EFFECTS OF CELLULOSE NANOCRYSTALS ON RANDOMLY ORIENTED AND ALIGNED CELLULOSE ACETATE ELECTROSPUN FIBERS

The development of random and aligned cellulose fibers with cellulose

nanocrystals as reinforcements by electrospinning is described in this chapter. Nanofibers

with CNC concentrations of up to 7.5 wt% were produced. Experimental parameters were

investigated and optimized to produce continuous, defect free fibers with consistent

diameters. The fiber surfaces were investigated using microscopy, and mechanical

properties were measured using dynamic mechanical analysis. The relationship between

CNC addition and mechanical properties is discussed.

2.1 Introduction

Cellulose nanocrystals (CNCs) are a class of newly developed and sustainable

nanomaterial that may be the beginning of a new biopolymer composites industry.

Derived from cellulose-based materials such as wood, crystalline cellulose has a greater

axial elastic modulus than Kevlar, and a greater tensile strength than steel.1 They can be

used in a wide range of applications including material packing, flexible circuits, solar

panels, and armor for defense application.2 There have been high research efforts to

utilize these materials as reinforcing agents due to their sustainability, abundance, low

cost and superior mechanical properties.1-5 However, due to their relatively new

emergence, there is a lack of fundamental understanding of their morphology, structure

Page 30: Experimental and modeling investigation of cellulose

8

and properties. To fully utilize their potential, accurate determination of the material

behavior is critical.

Cellulose nanocrystals are particles with high aspect ratio, with diameter ranging

from 2-10 nm and length from 50-500 nm, similar to stiff matchsticks. By aligning these

crystals while they are acting as reinforcing agents, the mechanical properties of the

composite in the alignment direction can be greatly increased. Alignments can be

achieved by rheological methods, magnetic fields and electric fields.6 In this study,

electrospinning, a method involving high magnetic and shear forces, will be used as the

alignment method. Electrospinning is a technique used to produce ultrafine polymer

fibers through the action of electrostatic forces. The conductive polymer solution is

pushed through a nozzle into an electric field, forming fibers in the process through

elongation, thinning, and solvent evaporation. The resulting fibers can collected as a

random or aligned mat depending on the collector type. This method generally produces

microscale and nanoscale fibers. The resulting fibers can be used in many applications

such as textiles, filtration, affinity membranes, catalyst carriers, and tissue engineering

scaffolds.7-8

During the past decade, there have been several studies where CNCs have been

used as reinforcements for composite nanofibers using the electrospinning technique.

Various synthetic polymers have been used as well as natural polymers such as alginate.9-

11 A summary of these studies can be found in Table 1.1. Successful efforts to electrospin

CNC-polymer composite fibers report typically low concentration of CNC loading (less

than 20 wt %), decrease in fiber diameter in comparison to the neat polymer fiber,

increases in mechanical strength and elastic modulus in the fiber direction. Maximum

Page 31: Experimental and modeling investigation of cellulose

9

elastic modulus improvements in comparison to the neat polymer fibers fluctuate greatly

between CNC-polymer cases and studies, from 17%10 to 3400%12.

Maximum strength values are similarly broad. However, due to difficulties of

collecting and testing a single nanoscale electrospun fiber the reported mechanical

properties are typically of fiber mats instead of single fibers. Consequently, the

mechanical properties of fibers mats are highly dependent of the number and alignment

of the electrospun fibers, thus are not representative of individual fiber, making it

difficult to link CNC additions and CNC alignment on the reinforcement potential of the

polymer system. The use of different CNC types or dispersion methods could be part of

that variation. Almost all reported strength values were below 10MPa, and improvements

were observed from CNC reinforcements up to a critical CNC loading. Further addition

of CNCs above that value decreased strength values, often attributed to the agglomeration

of the CNC fibers causing defects in the fiber. Similar behavior was observed for the

Young’s modulus of the fibers. Moreover, CNC agglomeration within the fibers is a

commonly reported issue, leading to fiber defects and decreased fiber mechanical

performance.

Dong et al 10 reporting the highest CNC concentration in a polymer at 41wt% CNC

in PMMA. The modulus was measured by nanoindentation and followed the general

trend of increasing modulus with increasing CNC concentration. Controlled orientation

of CNC in PMMA matrix was done by aligning the fibers using a rotating drum.

However, from the SEM image presented, only a vague orientation could be observed.

Park et al.13 electrospun polyethyleneoxide (PEO) fibers with 0-0.4 wt% bacterial

cellulose nanocrystals. The tensile properties of the mats were obtained using an

Page 32: Experimental and modeling investigation of cellulose

Tab

le 2

-1 S

umm

ary

of p

repa

ratio

n te

chni

ques

, com

posi

tions

, dia

met

er a

nd m

echa

nica

l pro

pert

ies

of v

ario

us p

ublis

hed

elec

tros

pun

poly

mer

-CN

C n

anoc

ompo

site

fib

er s

tudi

es

Mat

rix

CN

C ty

pe

CN

C

diam

eter

(n

m)

CN

C

leng

th

(nm

) D

ispe

rsio

n So

lven

t C

NC

ra

nge

(%w

)

Max

Mod

ulus

, MPa

M

ax S

tren

gth,

MPa

Fi

ber

Dia

met

er

[Inc

reas

e] (

% C

NC

s)

[Inc

reas

e](%

C

NC

s)

Cel

lulo

se A

ceta

te14

W

ood

NR

N

R

A

Ace

tone

, A

ceti

c A

cid

1-5

825

[+9

81%

] (

1)

NR

20

0-30

00nm

PMM

A10

So

ftw

ood

Pulp

17

19

0-66

0 B

D

MF

5-41

63

00 [

+17%

] (

17)*

N

R

100-

1000

nm

Poly

viny

l Alc

ohol

9, 1

5 R

amie

3-

10

100-

250

C

Wat

er

5-15

57

[+

270%

]

(15)

N

R

218-

235n

m

Pape

r N

R

NR

C

W

ater

3-

15

190

[+1

17%

]

(15)

10

.5 [

+96%

] (

15)

146-

292n

m

-cap

rola

cton

e)16

R

amie

10

-15

100-

250

B

DM

F 2.

5-7.

5 6.

5 [

+68

%]

(2

.5)

1.5

[+

37%

] (

2.5)

12

0-21

0nm

PEO

13

Bac

teri

al

11

420

C

Wat

er

0.2-

0.4

96.1

[+1

94%

] (

0.4)

1.

7 [+

73%

] (0

.4)

140-

300n

m

Poly

lact

ic A

cid17

-18

Bac

teri

al

20

500-

1000

A

D

MF

2.5-

7.5

NR

N

R

300n

m

W

ood

NR

N

R

A

DM

F 1-

10

5.6

[+30

%]

(1)

33

.7 [

+37%

] (

1)

300-

650n

m

Poly

acry

lic A

cid12

C

otto

n N

R

NR

A

E

than

ol

5-20

19

83 [

+340

0%]

(20)

4.

5[+1

455%

](20

) 69

-200

0nm

Poly

styr

ene19

Pa

per

10-2

0 20

0 A

T

HF

6-9

8.2

[+1

23%

] (

9)

NR

60

0-19

00nm

Dis

pers

ion

Met

hods

:

*: N

anoi

nden

tatio

n N

R: N

ot R

epor

ted

A

: Fre

eze

Dri

ed C

NC

re-

disp

erse

d us

ing

soni

cati

on

B: S

olve

nt e

xcha

nge

from

wat

er to

so

lven

t

C: S

onic

atio

n

D: M

ixin

g by

sti

rrin

g

10

Page 33: Experimental and modeling investigation of cellulose

11

universal testing machine. An average of 3-fold increase in Young's modulus (96.1 MPa

vs. 32.7 MPa) was observed for the 0.4 wt% CNC PEO composite. Additionally, TEM

images of the fibers showed both aligned CNCs along the fiber axis as well as aggregated

CNCs. Xiang et al.18 received similar results with a poly(lactic acid)/wood CNC system.

The modulus was increased when 1 wt% CNC was added, but no additional improvement

was seen with 10 wt% CNC loading. The authors attributed this to the lack of adhesion of

the CNCs to the polymer matrix. The research group at North Carolina State University

worked on incorporation of CNCs into polystyrene,19 poly( -caprolactone),16 and

poly(vinyl alcohol).15 Freeze dried CNCs were added directly into polystyrene and THF

and left to agitation overnight. Surfactants were added to minimize the beading observed

on the fibers. Storage modulus determined by Dynamic Mechanical Analysis (DMA)

showed significant increase for 6 wt% and 9wt% CNC/PS composites. An interesting

phenomena was observed for the poly( -caprolactone)/CNC composite where the initial

addition of CNCs decreased the diameter of the fibers by half while the higher dosages

increased the diameter. The authors attributed this to the lack of percolation network in

the low CNC dosage (2.5%). PCL was also chemically grafted onto the CNC surface to

improve interfacial adhesion. However, this affected the fibers negatively as they were

annealed during the electrospinning process. Poly(vinyl alcohol) was dissolved in water

and spun with up to 15 wt% of CNCs. Characterization showed no change with the

matrix structure. The elastic modulus increased with higher concentration of CNCs. The

largest increase in modulus was observed by Lu et al.12 with poly(acrylic acid)/CNC

composite. Freeze dried CNCs were used with sonication and agitation to promote

dispersion. With 20% CNC, the Young's modulus was increased by 35 times (1.983 GPa

Page 34: Experimental and modeling investigation of cellulose

12

vs. 0.056 GPa), and to an impressive 77 times (4.331 GPa) when the sample was

crosslinked by annealing. Similar to the other cases, the fiber diameter was decreased and

the fiber uniformity was increased with the addition of CNCs. Herrera et al.14 also

electrospun randomly oriented and aligned cellulose acetate fibers reinforced with

different concentration of CNCs (0-5 wt %). Similar to other literature, the diameter of

the fibers decreased with the addition. DMA was used to observe a tenfold increase in

storage modulus with only 1% CNC presence. The decrease in modulus with higher CNC

wt% was attributed to CNC aggregation in the fibers.

The focus of this study is to develop experimental techniques to produce cellulose

acetate nanofibers with cellulose nanocrystal fillers, and characterize the resulting fiber

morphology and properties with a focus on obtaining accurate mechanical strength of the

nanofibers. By obtaining accurate experimental data of how CNC affect electrospun

polymer nanofibers, this information may be used to help develop and validate models

describing the material behaviors of cellulose nanocrystals. Furthermore, these models

will allow for prediction of the alignment of the cellulose nanocrystal rods resulting from

electrospinning as well as the composite mechanical properties.

2.2 Methods and Materials

2.2.1 Materials

Cellulose Acetate (CA) was chosen as the polymer matrix due to its sustainability, and

its' existing industry market for CA fibers. It is used in apparel, diapers, filters and many

other applications, although is limited by its’ low mechanical properties.20 CA with an

acetyl content of 39.7 wt% and average molecular weight of 50,000 was purchased from

Sigma-Aldrich Chemistry. Reagent grade acetone (Sigma-Aldrich), and N, N-

Page 35: Experimental and modeling investigation of cellulose

13

Dimethylacetamide (DMAc) were used without further purification. Water-based

suspensions of cellulose nanocrystals supplied by the USDA Forest Service-Forest

Products Laboratory, Madison, WI following the procedures described by Beck-

Canandedo et al.21 were used. The morphology of individual CNC particles was

characterized via transmission electron microscopy (TEM; Philips CM-100, at 100kV,

spot 3, 200 µm condenser aperture, and 50 µm objective aperture,) of CNC particles

deposited on TEM grids (400 mesh Formvar/carbon filmed grid prepared with blow

discharge) with 2% aqueous uranyl acetate stain. The average diameter (6 ± 2nm) and

length (83 ± 43nm) of 93 individual CNC crystals were determined using within Image J

(Figure 2.1).

2.2.2 Experimental Setup

Solution of 14 wt% (CA+CNC) in a mixture of 2:1 weight ratio acetone and

DMAc were used to obtain electrospun fibers. CNCs were solvent transferred from water

to DMAc using a rotary evaporator (Yamato RE500). The cellulose acetate was added to

the solution and agitated overnight to ensure proper mixing. A high voltage power supply

(Gamma High Voltage, D-ES50PN-20W/QDPM) was used to supply both positive and

negative voltage. Strength of the positive electrode attached to the syringe were set to

12.5 kV. The negative electrode attached to the collector was kept at 1kV. The collector

was placed 20 cm away from the tip of the needle. Two types of collectors were used, a

metal plate for random non-woven mats and a parallel electrode for aligned fibers. A

gauge 16 needle was used along with an infusion pump (Cole Parmer) feeding at 0.7

mL/hr. A grounded Faraday cage was built to shield the experiment from environmental

Page 36: Experimental and modeling investigation of cellulose
Page 37: Experimental and modeling investigation of cellulose

15

interference. The electrospinning was carried out horizontally in a laboratory vent hood at

room temperature. Fibers containing 0 to 7.5 wt% CNC were successfully spun.

2.2.3 Characterization of Nanofibers

Optical Microscopy (Olympus ZX-12) was used to observe the general structure

and formation of the fibers. Flow Birefringence of CNCs was observed in a cross-

polarizer to ensure the proper dispersion of CNCs in the DMAc solution. Scanning

electron microscopy (Jeol, JSM 6400) was used at an accelerating voltage of 5 kV for

close surface observation and diameter determination. All samples were coated in

gold/palladium in a sputter (Hummer 6.2, 30 sec, 1kV, 18 mA) prior to observation.

The thermal properties of the fibers were measured by differential scanning

calorimetry (DSC) (TA Instruments, Q-200) and thermo gravimetric analysis (TGA) (TA

Instruments, Q-500). Each DSC sample contained < 5mg and were tested from -20°C; to

150°C at 10°C/min. Tg is calculated as the heat flow inflection point, or the minimum of

the heat flow derivative. The thermal stability of the fibers were determined by TGA at a

heating rate of 10°C/min from room temperature to 1000°C.

The mechanical properties of the fibers were measured using a dynamic

mechanical analysis machine (TA Instruments, Q800). Parallel aligned fiber samples

were attached to a “C” shaped holder, and mounted on the DMA. Optical microscopy

were used to estimate the number of nanofibers per cross section, and ensure there were

no obvious defects. Fibers are assumed to be cylindrical in cross section as supported by

optical imaging. Each sample was tested with a gauge length of 10 mm, a constant

applied force rate of 0.003N/min at room temperature and relative humidity of 40 to 50%

until fracture. Specimens that failed outside of the gauge length were considered

Page 38: Experimental and modeling investigation of cellulose

16

improperly loaded, and were discarded in the data analysis. The results were recalculated

using the microscopy data to represent the single fiber strength.

2.3 Discussion and Results

2.3.1 CNC Dispersion

Due to the insolubility of cellulose acetate in water, the water in the CNC

solutions needs to be removed. Two methods of water removal and CNC dispersion were

investigated: solvent transfer and freeze dried. Aqueous suspensions of CNCs (6.6 wt %)

were solvent transferred to DMAc using rotary evaporation. This method is labor

intensive and the resulting CNC% in the solution needs to be gravimetrically determined.

Additionally, there would likely be residual water in the suspension. However, one study

found the existence of small amount of water in polar organic solvents helped the

stability of the CNC suspension.22 The second method involves removing water from the

CNC suspension using a freeze drier. Determined by TGA, the final freeze dried CNCs

contained approximately 8% water. The desired amounts of CNCs were then added to

DMAc and agitated overnight. Ultra sonication were applied to solutions prepared by

both methods. Flow birefringence tests were carried out to study the dispersion of CNCs.

Vibrant colors under the crossed polarizers indicate well dispersed CNCs.

Figure 4.2 and 4.3 show the results obtained from the comparison. The solvent

transfer method was found to produce DMAc suspensions with significantly better

dispersion and stability. The freeze dried CNC suspensions showed signs of settling after

two days in cold storage where the solvent transfer solution still show no signs of settling

after 3 months. The solvent transfer method was used to add CNCs to the electrospinning

solution.

Page 39: Experimental and modeling investigation of cellulose
Page 40: Experimental and modeling investigation of cellulose
Page 41: Experimental and modeling investigation of cellulose

19

2.3.2 Surface Morphology of Electrospun CA-CNC Fibers

The electrospinning experimental parameters were initially optimized using pure

cellulose acetate fibers. Both types of collectors were used. Desirable qualities were

uniform fibers, no buildup at the tip of the needle, high number of fibers collected and

minimal defects such as surface pores and electrospray. Acetic Acid, acetone and DMAc

were investigated as potential solvents. Ambient conditions were also found to be

important. Little to no fibers were produced under high humidity conditions due to low

evaporation and charge conductivity. Pores were also found on the surface of the fibers

due to water condensation. Figure 2.4 shows the SEM images of both random and aligned

samples of cellulose acetate electrospun fibers. The images show continuous smooth

fibers with consistent diameters and no visible surface defects.

The average diameter of pure CA fibers were found to be approximately 0.75

through SEM imaging. 1, 2.5, 5 and 7.5 wt% CNCs were added to cellulose acetate and

spun. An immediate decrease in diameter was observed. 1 wt% CNC lowered the

diameter to about This behavior is also often reported in literature and accredited

to the increased conductivity of the fiber by the addition of CNCs.15, 18 Additional

additions of CNCs did not further decrease the diameter, however, led to more defects in

the fiber. One such defect is large amounts of beading in the fibers; an indication of high

surface tension (Figure 2.5). Solutions for dry CNC concentration above 7.5 wt% did not

result in fibers due to agglomeration issues.

Another issue encountered with SEM imaging is that CA is extremely heat sensitive.

Even at a low voltage setting (5kV), surface cracking of the fibers caused by the electron

Page 42: Experimental and modeling investigation of cellulose
Page 43: Experimental and modeling investigation of cellulose
Page 44: Experimental and modeling investigation of cellulose
Page 45: Experimental and modeling investigation of cellulose

23

beam was observed (Figure 2.6). This method did not produce a way to observe the

alignment of the CNCs within/on the surface of the fiber.

2.3.3 Thermal Properties of Electrospun CA-CNC Fibers

The thermal properties of the pure CA fibers and 1 wt% CNC/CA fibers were

compared to CA powders. Thermal gravimetric analysis (Figure 2.7) shows the thermal

degradation in three steps: from starting temperature to 330°C, evaporation of the

residual water and volatile matters; from 330°C to 400°C, thermal degradation of the

cellulose acetate chains; and after, carbonization of the degraded products to ash. All

sample showed similar behavior, thermal properties were not changed by electrospinning

or addition of CNCs. This behavior may change with higher dosage of CNCs. DSC

samples were run to test the effect of electrospinning on thermal stability. Each sample

contained < 5mg and were tested from -20°C; to 150°C at 10°C/min. Tg, the glass

transition temperature of the polymer, is defined as the temperature at which the

reversible transition of glassy polymer into a rubbery state happened. The Tg of the CA

electrospun fibers as well as CA-CNC fibers were similar to that of CA powders (not

pictured). The electrospinning process did not change the molecular structure, however a

slight increase of crystallinity was detected as CNC was added.

2.3.4 Mechanical Properties of Electrospun CA-CNC Fibers

To estimate the mechanical properties of a single fiber, parallel fiber mats were

collected for 10 minutes. The mechanical properties of aligned fiber mats were then

measured using dynamic mechanical analysis. Optical microscopy along with imaging

program ImageJ was used to count the number of fibers in the cross section of the fiber

mat. The fibers were assumed to be perfectly circular and of average diameter. The

Page 46: Experimental and modeling investigation of cellulose

24

Figure 2-7 TGA thermograms of the CA-CNC electrospun fibers

Page 47: Experimental and modeling investigation of cellulose

25

Figure 2-8 Mechanical Properties of CA-CNC electrospun fibers as a function of CNC concentration

Page 48: Experimental and modeling investigation of cellulose

26

modulus and ultimate tensile strength was then calculated using the new resulting cross

sectional area. The following data represent the estimated behavior of a single fiber

(Figure 2.8). Young’s modulus showed some slight improvement as CNC concentration

increases. However, further addition of CNC did not improve the modulus, and increased

the standard deviation. This suggests the CNCs are agglomeration in the fiber, thus

creating defects and leading to premature defect driven failure. The ultimate tensile

strength exhibit a similar behavior. The strength of a 7.5 wt% CNC reinforced fiber is in

fact lower than that of a pure CA electrospun fiber. The elongation of the fibers decreases

sharply as CNC is added, suggesting the presence of the CNCs are increasing the

brittleness of the fiber.

2.4 Conclusion

In this study, cellulose nanocrystals reinforced cellulose acetate nanofibers were

successfully produced using electrospinning. Continuous CNC-CA fibers with consistent

diameters were produced between varying CNC concentrations of 0 and 7.5 wt%. The

addition of CNCs did not show a significant alternation to the thermal properties of the

fibers. A maximum of 15% and 280% increase was observed in the ultimate tensile

strength and Young’s modulus of the fibers respectively. However, the addition of the

CNCs increased brittleness of the fibers, as well as the defect driven failures, leading to

an increase in deviation of the fiber performance. Overall, the electrospinning method did

not improve the mechanical performance of the fibers to the theoretical potential of CNC,

suggesting further improvement may be made. A processing method involving high shear

rates such as dry spinning may improve the agglomeration issues experienced here and

lead to further orientation of CNCs.

Page 49: Experimental and modeling investigation of cellulose

27

Bibliography

1. Moon, R. J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J., Cellulose nanomaterials review: structure, properties and nanocomposites. Chemical Society Reviews 2011, 40 (7), 3941-3994.

2. Habibi, Y.; Lucia, L. A.; Rojas, O. J., Cellulose Nanocrystals: Chemistry, Self-Assembly, and Applications. Chem Rev 2010, 110 (6), 3479-3500.

3. Klemm, D.; Schumann, D.; Kramer, F.; Hessler, N.; Hornung, M.; Schmauder, H. P.; Marsch, S., Nanocelluloses as innovative polymers in research and application. Adv Polym Sci 2006, 205, 49-96.

4. Klemm, D.; Schumann, D.; Kramer, F.; Hessler, N.; Koth, D.; Sultanova, B., Nanocellulose Materials - Different Cellulose, Different Functionality. Macromol Symp 2009, 280, 60-71.

5. Samir, M. A. S. A.; Alloin, F.; Dufresne, A., Review of recent research into cellulosic whiskers, their properties and their application in nanocomposite field. Biomacromolecules 2005, 6 (2), 612-626.

6. Aguié-Béghin, V.; Molinari, M.; Hambardzumyan, A.; Foulon, L.; Habibi, Y.; Heim, T.; Blossey, R.; Douillard, R., Preparation of Ordered Films of Cellulose Nanocrystals. In Model Cellulosic Surfaces, American Chemical Society: 2009; Vol. 1019, pp 115-136.

7. Lin, T.; Fang, J.; Niu, H. T.; Wang, X. G., Applications of electrospun nanofibers. Chinese Sci Bull 2008, 53 (15), 2265-2286.

8. Huang, Z. M.; Zhang, Y. Z.; Kotaki, M.; Ramakrishna, S., A review on polymer nanofibers by electrospinning and their applications in nanocomposites. Compos Sci Technol 2003, 63 (15), 2223-2253.

9. Lee, J.; Deng, Y. L., Increased mechanical properties of aligned and isotropic electrospun PVA nanofiber webs by cellulose nanowhisker reinforcement. Macromolecular Research 2012, 20 (1), 76-83.

10. Dong, H.; Strawhecker, K. E.; Snyder, J. F.; Orlicki, J. A.; Reiner, R. S.; Rudie, A. W., Cellulose nanocrystals as a reinforcing material for electrospun poly(methyl methacrylate) fibers: Formation, properties and nanomechanical characterization. Carbohyd Polym 2012, 87 (4), 2488-2495.

11. Habibi, Y.; Heim, T.; Douillard, R., AC electric field-assisted assembly and alignment of cellulose nanocrystals. J Polym Sci Pol Phys 2008, 46 (14), 1430-1436.

12. Lu, P.; Hsieh, Y. L., Cellulose nanocrystal-filled poly(acrylic acid) nanocomposite fibrous membranes. Nanotechnology 2009, 20 (41), -.

Page 50: Experimental and modeling investigation of cellulose

28

13. Park, W. I.; Kang, M.; Kim, H. S.; Jin, H. J., Electrospinning of poly(ethylene oxide) with bacterial cellulose whiskers. Macromol Symp 2007, 249, 289-294.

14. Herrera, N. V.; Mathew, A. P.; Wang, L. Y.; Oksman, K., Randomly oriented and aligned cellulose fibres reinforced with cellulose nanowhiskers, prepared by electrospinning. Plast Rubber Compos 2011, 40 (2), 57-64.

15. Peresin, M. S.; Habibi, Y.; Vesterinen, A. H.; Rojas, O. J.; Pawlak, J. J.; Seppala, J. V., Effect of Moisture on Electrospun Nanofiber Composites of Poly(vinyl alcohol) and Cellulose Nanocrystals. Biomacromolecules 2010, 11 (9), 2471-2477.

16. Zoppe, J. O.; Peresin, M. S.; Habibi, Y.; Venditti, R. A.; Rojas, O. J., Reinforcing Poly(epsilon-caprolactone) Nanofibers with Cellulose Nanocrystals. Acs Appl Mater Inter 2009, 1 (9), 1996-2004.

17. Liu, D. Y.; Yuan, X. W.; Bhattacharyya, D., The effects of cellulose nanowhiskers on electrospun poly (lactic acid) nanofibres. J Mater Sci 2012, 47 (7), 3159-3165.

18. Xiang, C. H.; Joo, Y. L.; Frey, M. W., Nanocomposite Fibers Electrospun from Poly(lactic acid)/Cellulose Nanocrystals. J Biobased Mater Bio 2009, 3 (2), 147-155.

19. Rojas, O. J.; Montero, G. A.; Habibi, Y., Electrospun Nanocomposites from Polystyrene Loaded with Cellulose Nanowhiskers. J Appl Polym Sci 2009, 113 (2), 927-935.

20. Frey, M. W., Electrospinning cellulose and cellulose derivatives. Polym Rev 2008, 48 (2), 378-391.

21. Beck-Candanedo, S.; Roman, M.; Gray, D. G., Effect of reaction conditions on the properties and behavior of wood cellulose nanocrystal suspensions. Biomacromolecules 2005, 6 (2), 1048-1054.

22. Viet, D.; Beck-Candanedo, S.; Gray, D. G., Dispersion of cellulose nanocrystals in polar organic solvents. Cellulose 2007, 14 (2), 109-113.

Page 51: Experimental and modeling investigation of cellulose

29

CHAPTER 3. EFFECTS OF CRYSTAL ORIENTATION ON CELLULOSE NANOCRYSTALS-CELLULOSE ACETATE NANOCOMPOSITE FIBERS

PREPARED BY DRY SPINNING

The following chapter contains content reproduced with permission from Chen, S.;

Schueneman, G.; Pipes, R. B.; Youngblood, J.; Moon, R. J., Effects of Crystal

Orientation on Cellulose Nanocrystals–Cellulose Acetate Nanocomposite Fibers Prepared

by Dry Spinning. Biomacromolecules, 15 (10), 3827-3835. Copyright 2014 American

Chemical Society

This chapter presents the development of dry spun cellulose acetate (CA) fibers

using cellulose nanocrystals (CNCs) as reinforcements. Increasing amounts of CNCs (up

to 49 wt %) were successfully dispersed into CA fibers in efforts to improve the tensile

strength and elastic modulus of the fiber. A systematic characterization of dispersion of

CNCs in the polymer fiber and their effect on the nanocomposites’ mechanical properties

is described. The birefringence, thermal properties, and degree of CNC orientation of the

fibers are discussed. 2D X-ray diffraction was used to calculate the Herman’s order

parameter, which quantified the degree of CNC alignment within the fibers. It is shown

that the CNC alignment directly correlates to the mechanical properties of the composite.

Continual improvement in the tensile strength and elastic modulus of the CNC-CA fibers

up to 34wt% of CNC was observed, which resulted in a maximum increase of 137%

tensile strength and 637% in elastic modulus. Further increases in CNC concentration

Page 52: Experimental and modeling investigation of cellulose

30

did not yield corresponding improvement and led to large variations in fiber performance,

likely due to increased defect density and/or aggregation of CNCs. Empirical

micromechanical models Halpin-Tsai equation and an orientation modified Cox model

were used to predict the fiber performance and compared with experimental results.

3.1 Introduction

Over the past two decades, cellulose nanocrystals (CNCs) have been the subject

of significant research interest as a new class of low cost, renewable and biodegradable

materials with high mechanical properties and customizable surfaces. CNCs are rod-like

nanoparticles that are produced by controlled acid hydrolysis of cellulose-based materials

such as plants, trees, algae, bacteria, and tunicates. The morphology and dimensions of

the CNCs are primarily determined by the source of cellulose, as well as hydrolysis

conditions.1-4 CNCs extracted from plants and trees typically have diameters of ~ 3 to 5

nm and lengths of ~50 to 500 nm.4-8 CNCs have a unique combination of characteristics,

such as, relatively high aspect ratio, nano size-scale fibril morphology, high mechanical

properties 9, low density, low coefficient of thermal expansion,10 and surfaces accessible

to chemical functionalization, and because of this, CNC have been studied as a

reinforcement phase in a wide variety of composite systems.2, 11-13

Much of the research on the reinforcement potential of CNCs within a polymer

matrix has been completed with films produced via solution casting 8, 14, or self-

reinforced composite where CNCs reinforce a cellulose II polymer matrix11, 15 as well as

in continuous polymer fiber fabrication16-31. CNC additions have been shown to alter

mechanical properties (stiffness, tensile strength, and strain at failure), thermal stability,

coefficient of thermal expansion, and transparency, the extent to which is dependent on

Page 53: Experimental and modeling investigation of cellulose

31

the polymer matrix, CNC concentration and the CNC surface chemistry.2, 11-13 To

maximize the effectiveness of the CNC additions in altering the polymer properties, it is

desired to have a fine CNC dispersion while maintaining CNC percolation.1 However,

agglomeration and phase separation of CNC with in the polymer matrix is common,

typically occurring at concentration well below 30 wt% CNCs, and as a result there are

limited number of studies reporting CNC-polymer composite properties for CNC

additions of 30wt% or greater concentration.1-2 Recent research has focused on CNC

surface functionalization, and surfactants as methods used to overcome problems

associated with CNC agglomeration and phase separation.8

The anisotropic arrangement of cellulose chains within CNCs results in

anisotropic properties (e.g. mechanical,32 thermal expansion 10), in which the CNC

properties are different depending on the direction of measurement. For example the

elastic moduli of CNCs in the longitudinal and transverse directions have been reported

to be within the range of 110-220 GPa1, 32 and 2-50 GPa32-33, respectively. The

ramifications of the CNC property anisotropy, is that the properties of CNC-polymer

composites will be dependent on the CNC arrangement. Thus, an area of increase interest

has been in the processing of preferred CNC alignment and its effect on composite

properties. Shear deformation rate in a suspension is one method used to facilitate high

degrees of CNC orientation.9, 34-35 Under shear, the nanocrystals naturally align along the

shearing direction due to their rod-like nature, greatly decreasing their flow resistance. By

aligning these crystals while they are acting as reinforcing agents, the stiffness of the

composite in the alignment direction can be greatly increased.

Page 54: Experimental and modeling investigation of cellulose

32

One of the processing techniques that generate high shear rates is fiber spinning (i.e. dry,

wet, electrospinning, etc.). The spinning process typically consists of a polymer dissolved

in solvent extruded through a device containing small orifices known as the spinneret.

The filaments then may be passed through air or liquids to solidify, known as dry or wet

spinning, respectively. For electrospinning high voltage is applied between the orifice

and a “target” which effectively accelerates the jet of solution to the target. Other

processing such as drawing, washing, and finishing application are also often used.

CNCs have been used in fiber spinning of various polymer systems, primarily via

electrospinning,17-20, 22-27, 36-37 whereas only a few studies have been reported using CNC

in either wet or gel spinning.28-29, 31, 38-40

Successful efforts to electrospin CNC-polymer composite fibers report typically

low concentrations of CNC loading (less than 20 wt %), decreases in fiber diameter in

comparison to the neat polymer fiber, increases in tensile strength and elastic modulus in

the fiber direction. In contrast, Dong et al 17 reported the highest CNC concentration in a

polymer at 41wt% CNC in PMMA. Maximum elastic modulus improvements in

comparison to the neat polymer fibers fluctuated greatly between CNC-polymer cases

and studies, from 17%17 to 3400%20. Maximum strength values are similarly broad.

However, due to difficulties of collecting and testing a single nanoscale electrospun fiber,

the reported mechanical properties are typically for fiber mats instead of single fibers.

Consequently, the mechanical properties of fibers mats are highly dependent of the

number and alignment of the electrospun fibers and thus, are not representative of

individual fiber, making it difficult to link CNC additions and CNC alignment on the

reinforcement potential of the polymer system. Moreover, CNC agglomeration within the

Page 55: Experimental and modeling investigation of cellulose

33

fibers is a commonly reported issue, leading to fiber defects and decreased fiber

mechanical performance.

Traditional spinning techniques (dry, wet, and gel) can circumvent some of fiber

characterization issues of electrospun fibers, however, only a few studies have been

reported28-31, 38-42. While the present study investigates dry spinning, to date, only wet

spinning of CNCs has been reported. Wet spun fibers were produced with lab-scale

syringe pump systems with large diameter needles, producing fibers with larger diameters

(60- greater CNC volume fractions (up to 50% 28) and controlled CNC

orientation based on shear rate41. This allowed for mechanical property measurements

and CNC orientation quantification of individual fibers, such that, it is now possible to

link CNC additions and CNC alignment to reinforcement potential. Liu et al. 31 wet spun

CNC reinforced artificial silk, reporting a three times increase in elastic modulus at 5wt%

CNC concentration. The same group later reported CNC/regenerated cellulose fibers with

0 to 9wt% CNC loads38, resulting in an increase in the tensile strength and Young's

modulus from 172 to 571 MPa and 2.06 to 4.15 GPa, respectively. Iwamoto et al.41

observed a direct relationship between shear rate and mechanical strength using wet-spun

fibers made from cellulose nanofibrils (CNF). X-ray diffraction analysis of the spun

fibers showed alignment of the nanofibers along the fiber direction, correlating with the

increase in mechanical strength. Hakansson et al. 42 also witnessed a similar effect in

CNF filaments formed by using a flow-focusing channel system similar to wet spinning.

By controlling the preferential fibril orientation along the filament, they were able to

increase both the modulus and tensile strength of the dried filaments by about 50%.

Urena-Benavides et al.28-30 reported a similar result with CNC-alginate composites

Page 56: Experimental and modeling investigation of cellulose

34

containing up to 50wt% CNC. They found the spinning process oriented the CNCs,

increasing the strength and the modulus of the fibers. However, above 10wt% CNC

loading, the mechanical performance of the fibers decreased with increasing CNC

concentration. This is similar to most studies with Polymer-CNC composites where a

plateau behavior is observed beyond which adding additional CNCs did not result in

further performance improvement. Despite these reported improvements in mechanical

properties by the addition of CNCs, these experimental values are significantly lower

than micromechanical model-predicted composite performance, suggesting

improvements are possible.

Figure 3.1 shows a comparison of elastic modulus vs. tensile strength of various

CNC-polymer filaments reported in literature, including this study highlighted in red, and

other common filament materials. Electrospun CNC-polymer fibers studies were grouped

together because the mechanical properties were measured for fiber mats and have other

variables that influence the measured properties and thus do not give individual fiber

properties. For the wet and dry spinning cases the effects of the CNC additions vary

significantly, which is due to the variations in starting material and, processing and

testing conditions. , It should be noted that this study reports greater values than most

recent experiments, and demonstrated the greatest percentage of improvement.

Considering the renewable and biodegradable nature of CNCs, there is a high

incentive for the development of a strong green composite. Cellulose acetate (CA) is a

semi-synthetic polymer obtained from esterification of cellulose. Since its’ invention in

1905, CA have remained a commonly used product in clothing, filters and diapers and are

typically produced by dry-spinning in fiber form. There has been some recent attempts to

Page 57: Experimental and modeling investigation of cellulose

35

Figure 3-1 Overview of elastic modulus vs. tensile strength of various CN-polymer filaments reported in literature and other common filament materials. Colored regions indicate CN studies. The red region, CA-CNC, indicates results from this study.29, 31, 38, 41,

43-45

Page 58: Experimental and modeling investigation of cellulose

36

improve the mechanical properties of CA including compounding it with clay46, modified

and unmodified nanoclays47, and organoclays48. In these studies, tensile strength and

modulus were increased 33-38% and transparencies levels were decreased. To date there

appear to be no reports on dry or wet spun CNC-CA fiber systems. Herrera et al 18

electrospun cellulose acetate fibers reinforced with wood CNC between concentrations of

1-5wt%. The maximum modulus of the electrospun fiber mat reported was 825MPa at

1% CNC, a 981% increase from the neat polymer fiber mat. No tensile strength data was

reported. For CNC-CA films, Yang et al 49 investigated the role of CNC concentration (2

to 7 wt% CNC) on the film mechanical properties. The maximum properties were

observed at 4.5 wt% CNC with a 9% increase (44MPa) in tensile strength and 39%

increase (1.5GPa) in elastic modulus respective to the neat film properties.

In the present study, CNC-CA continuous fibers were produced via dry spinning.

The role of CNC concentration and CNC alignment on the processing and resulting

mechanical properties (tensile strength, Young’s modulus, and strain to failure) were

investigated. Experimental results were compared to predictions from micromechanical

models that accounted for CNC concentration, orientation state and aspect ratio.

3.2 Experimental Section

3.2.1 Materials

Cellulose Acetate (CA) with an acetyl content of 39.7 wt% and average molecular

weight of 50,000 was purchased from Sigma-Aldrich Chemistry. Reagent grade acetone

(Sigma-Aldrich), and N,N-Dimethylacetamide (DMAc) were used without further

purification as the solvent.

Page 59: Experimental and modeling investigation of cellulose

37

Water-based suspensions of CNCs were produced and supplied by the USDA

Forest Service-Forest Products Laboratory, Madison, WI.4, 50

The morphology of individual CNC particles was characterized via transmission

electron microscopy (TEM; Philips CM-100, at 100kV, spot 3, 200 µm condenser

aperture, and 50 µm objective aperture,) of CNC particles deposited on TEM grids (400

mesh Formvar/carbon filmed grid prepared with blow discharge) with 2% aqueous uranyl

acetate stain. The average diameter (6 ± 2nm) and length (83 ± 43nm) of 93 individual

CNC crystals were determined using within Image J software.

3.2.2 Preparation of CNC-CA Solutions

Water-based suspensions of CNCs were solvent exchanged into DMAc by

vacuum-assisted rotary evaporation. Equal volumes of aqueous CNC dispersions and

DMAc were mixed together in a round bottom flask with agitation. Water and small

amounts of DMAc were evaporated using the rotary evaporator (Yamato RE500) until

the final volume was less than half of the initial volume. The final weight percentage of

CNC in the dispersion was determined gravimetrically after drying the solution under

vacuum at 80°C for 24 hours.

To prepare the dope solutions, the desired amount of CA was dissolved in a

mixture of acetone and DMAc. Varying amounts of CNC/DMAc dispersion were then

added to obtain the desired CNC/CA concentration in the solutions. The solutions were

agitated overnight and spun into fibers within 3 days of preparation.

3.2.3 Preparation of CNC-CA Fibers

The prepared solutions were dry spun using a custom built plunger style fiber

spinner based on a capillary viscometer 51 with spinneret insert having a conical V-taper

Page 60: Experimental and modeling investigation of cellulose

38

allowed processing at shear rates

similar to industrial processes (105-106/s) which is much greater than typical lab scale

setups (10-102/s). The spinner was attached to a universal tensile testing machine (MTS

Insight) to control displacement rate and measure pressure. The solutions and the

stainless steel spinneret were heated to the operating temperature (65°C to 95°C) and held

at temperature for 1 hour for degassing purposes. The spinneret was mounted onto the

MTS with counter current heated air at the spinner orifice (Figure 3.2). The dope solution

was loaded into the spinneret and the spinning process was started by pressing the

plunger into the reservoir at maximum crosshead speed (1500mm/min) until flow began.

The continual spinning rate was regulated by the minimum amount of force

needed to maintain fiber flow. The fibers were then manually collected and further dried

under vacuum at 3.4kPa and 40°C overnight. Nanocomposite fibers with CNC

concentrations between 0 and 49% were prepared. Exact experimental parameters are

listed in Table 3.1. Note that the spinning solution exhibited different drying rates

depending on the Acetone/DMAc solvent content, thus the operating temperature was

varied between 65°C to 95°C.

3.2.4 Characterization of CA-CNC Composite Fiber

3.2.4.1 Rheology

Rheology experiments were performed on all suspensions using a TA Instruments AR-

G2 rheometer. The steady state viscosity of the suspensions at room temperature was

determined using parallel plates geometry at shear rates ranging from 0.1 to 100 s-1. The

Page 61: Experimental and modeling investigation of cellulose
Page 62: Experimental and modeling investigation of cellulose

40

Table 3-1 Processing parameters for CNC-CA composite fibers

Sample

No.

CNC wt%

in fiber Weight % in solution

Operating

Temperature (°C)

CNC Acetone DMAc CA

1 0 0 78 0 22 65

2 1 <1 73 4 22 65

3 2.5 <1 68 9 22 70

4 4 <1 70 9 20 70

5 8 1 71 14 14 70

6 10 2 65 18 15 70

7 14 2 57 27 14 85

8 17 3 54 29 14 85

9 20 3 52 32 12 85

10 24 3 51 36 10 85

11 29 3 51 37 9 95

12 34 4 48 40 8 95

13 39 5 30 57 8 95

14 44 5 27 61 7 95

15 49 6 23 65 6 95

Page 63: Experimental and modeling investigation of cellulose

41

rheological behavior of the solutions was represented by the Power Law model as follows:

1n

K (3.1)

Where is the apparent viscosity, the shear rate, K the consistency index, and n is the

flow behavior index.

3.2.4.2 Microscopy

Optical Microscopy (OM) (Olympus ZX-12) was used to observe the diameter

and the surface morphology of the fibers. Optical images of each test specimen were

taken. The diameters of the fibers were measured using the measure tool in ImageJ. Flow

birefringence of CNCs within the fibers was observed in a Carl Zeiss (Axio Observer A1)

inverted microscope with cross-polarizers. Scanning electron microscopy (SEM) (FEI,

XL40, at 5 kV) was used for close surface and cross section observation. All samples

were coated in gold/palladium for 30 seconds at 1 kV, 18 mA in a sputter coater

(Hummer 6.2) prior to observation to minimize charging effects.

3.2.4.3 2D X-ray Diffraction

Two-dimensional wide-angle X-ray diffraction (2DWAXD) [Bruker D8 GADDS,

e at 40 mA, 20 kV for 600 s and a beam diameter of 500

fibers. Multiple fibers (10-15) of the same CNC content were attached in a parallel

bundle across a disk shaped fiber holder with a central opening. The holder was then

placed perpendicular to the beam to allow the X-rays to pass through only the fibers. The

WAXD patterns were recorded on a High Star 2D detector located 67mm from the

sample. The patterns were corrected for air scattering and background by subtracting a no

Page 64: Experimental and modeling investigation of cellulose

42

fiber diffraction pattern from the raw data. Furthermore, a diffraction pattern containing

only CA fibers was subtracted from all patterns of CNC-CA fibers in order to

deconvolute the two components.

According to literature29, the distribution of intensity (I) with respect to along

52, can be used to

quantify the orientation of the CNC rods along the fiber axis. This peak is determined to

-22.3°. The integrated intensity with respect to was obtained by

scanning along the arc from 0 to 180° around the (200) plane. Furthermore, Herman’s

order parameter (S) was calculated assuming axisymmetric orientation53. It is defined as:

23 cos 1

2S (3.2)

where is the angle formed between the c-axis of the cellulose unit cell and the fiber axis.

It can be calculated as follows:

2 2cos 1 2 cos (3.3)

2

2cos sin

cossin

I d

I d (3.4)

Where is the angle between the (200) plane and the fiber axis, and I( ) is the intensity

corresponding to each angle. An order parameter of S=0 indicates complete randomness

of the CNCs arrangement within the fiber, while S=1 indicates complete alignment of the

CNCs along the fiber axis.

Page 65: Experimental and modeling investigation of cellulose

43

3.2.4.4 Mechanical Testing

The tensile elastic modulus, ultimate strength and strain at failure were measured using a

dynamic mechanical analysis machine (TA Instruments, Q800). At least 10 fiber

segments were randomly selected from specimens of each CNC concentration. “C”

shaped tabs were used to attach the samples to the DMA and avoid pinching at endpoints.

Optical microscopy was used to determine the diameter of each, and ensure there were no

obvious defects. Fibers are assumed to be cylindrical in cross section as supported by

optical imaging. Single fibers were tested with a gauge length of 30mm, a constant

applied force rate of 0.3N/min at room temperature and relative humidity of 40 to 50%

until fracture. Specimens that failed outside of the gauge length were considered

improperly loaded, and were discarded in the data analysis.

3.2.4.5 Micromechanical Modeling

Micromechanical models were used to estimate the elastic modulus of CNC-CA

fibers as a function of CNC loading. Isostrain and isostress are basic micromechanical

models of idealized geometries. They assume the reinforcing filler of the composites are

continuous and perfectly collimated in one direction, and there is perfect adhesion

between the filler and the polymer matrix. Isostrain (Equation 3.5) assumes the polymer

and aligned continuous fillers experience the same axial strain. Isostress (Equation 3.6)

assumes the polymer and aligned continuous fillers experience the same stress.

f f m mcE V E V E (3.5)

f m

f mc

m f

E EE

V E V E (3.6)

Page 66: Experimental and modeling investigation of cellulose

44

Where E is the elastic modulus, and V the volume fraction. The subscripts, f and m refer

to fiber (CNC in this case), and the polymer matrix. The volume fractions are determined

from weight fractions and the material densities (1.3 g cm-3 for CA, and 1.6 g cm-3 for

CNC). The axial elastic modulus of CNC (143 GPa) is used for isostrain, and the

transverse (25 GPa) used for isostress.

The Halpin-Tsai model is an analytical form of Hill’s generalized self-consistent

model for composite materials. It takes into account short fibers and aspect ratios. The

fillers are assumed to be collimated, and there is perfect adhesion between the filler and

the polymer matrix.

11 2 wher

11

e

f

fm

mff

m

EL V EDE E

EV

E

(3.7)

Where L/D, the aspect ratio, is calculated to be length divided by diameter. Young’s

modulus values of 2GPa and 143GPa for CA and axial CNC, respectively were used.

A modified orientation Cox model including porosity and orientation effects is also used

to predict the Young’s modulus of the fiber composites.54-55

f f2

m p0 m1( )(1 )cE V E V E V (3.8)

Where E is the Young’s modulus, V the volume fraction, 0 the fiber orientation

efficiency factor, l the fiber length efficiency factor, and the subscripts c, f, m and p refer

to composite, fibers, matrix, and porosity respectively. The volume fractions are

determined from weight fractions and the material densities (1.3 g cm-3 for CA, and 1.6 g

cm-3 for CNC). In this case, the porosity term is not considered as the fibers are assumed

Page 67: Experimental and modeling investigation of cellulose

45

to be porosity free. 0 the fiber orientation efficiency factor is directly proportional to

Herman’s order parameter where random 3D gives a value of 0.5, and unidirectional

fibers gives a value of 1. l is calculated using Cox’s shear lag model55.

f

1

f

tanh2 where 1

2

22

ln

m

LGL L

DLE

V

(3.9)

Where L is the filler fiber length, D the filler fiber diameter, Gm the shear stiffness of the

matrix, and a constant controlled by the geometrical packing pattern of the fibers, and

assumed to be 0.907 in this case for hexagonal packing.

3.3 Results and Discussion

3.3.1 Suspension Properties and Fiber Spinning

The viscosities of the spinning solutions were measured to predict fiber formation

for the spinning system. It is well known that CNCs exhibit shear thinning behavior in the

dilute regime. At high shear rates, the CNCs align due to their rod-like shape and thereby

greatly decrease the shear viscosity of the solution56. However, at higher concentrations,

the suspensions transition into a gel like material 57. Considering this, the viscosity of the

spinning solution should be a balance of the concentration of the polymer, viscosity of

the polymer, the amount and orientation of CNCs, and the mixture of solvents in the

solution.

The shear rate under typical industrial dry spinning processes can reach up to

between 105 and 106 s-1. It is difficult to achieve such shear rates in a rheometer, thus

only general trends at relatively high shear was investigated. The Power Law model was

applied to the rheology behaviors of the samples. The resulting consistency coefficient,

Page 68: Experimental and modeling investigation of cellulose
Page 69: Experimental and modeling investigation of cellulose

47

K, and the flow behavior index, n, are given in Table 3.2. The solutions presented a

pseudoplastic behavior with a strong decline in apparent viscosity as shear rate increased,

as indicated by the value of n in the Power Law model. The addition of CNCs to the

solutions did not alter the value of n markedly. The change in the apparent viscosity was

largely driven by the consistency coefficient, K, which is in essence the viscosity of the

solution.

Experimental observations showed that when solutions with lower viscosities

(<10 Pa•s) were extruded from the orifice, droplets were formed rather than a

continuous/stable fiber. In contrast, stable fibers could be formed with high viscosity

solutions. However, there is a potential for the formation of porosity defects within the

fibers due to the inability for air pockets to quickly rise to the surface before

solidification. With the dry spinning configuration used in this study, typically the critical

viscosity of the dope solution for the formation of continuous fiber was ~ 10 Pa•s at 10 s-1

(Fig 3). The viscosity for pure 5.7wt% CNC in DMAc solution was too low to produce

continuous fibers and thus 100wt% CNC fiber could not be produced. However,

continuous CNC-CA fibers were able to be produced at 49wt% CNC, which is the

highest CNC concentration reported to date in CNC-CA composites.

3.3.2 Fiber Morphologies

Figure 3.4 shows OM and SEM images of the CNC-CA fiber composites. The

fibers were generally smooth on the surface, and cylindrical in cross-section. Some

defects observed were air bubbles, surface irregularities, and non-cylindrical fiber cross-

sections. Porosity within the fibers was more common with higher viscosity dope

solutions. Possible causes of the surface unevenness and deviations from cylindricality in

Page 70: Experimental and modeling investigation of cellulose

48

Table 3-2 Consistency coefficient (K), flow behavior index (n) as determined by the Power Law model for different experiment samples

Sample No. CNC% in fiber K (Pa.sn) n R2

1 0 186.2 0.36 0.98

2 1 252.6 0.35 0.99

3 2.5 45.1 0.37 0.98

4 4 57.2 0.29 0.94

10 24 35.4 0.36 0.94

14 44 70.4 0.34 0.97

5.7% CNC in DMAc Solution 2.2 0.33 0.99

Page 71: Experimental and modeling investigation of cellulose

49

the fiber cross-section may include clogging at the orifice or uneven drying conditions.

More surface defects were observed at higher CNC concentrations. This may result from

the addition of DMAc along with CNCs to the system. DMAc has a significantly higher

boiling point than acetone (165°C vs 57°C), thus greatly reducing the rate of

solidification of the fibers, allowing more time for introduction of defects into the system.

As expected, the fracture surface of the neat CA fiber showed plastic deformation (Figure

3.4b). The addition of CNCs introduced brittleness into the fibers; this is evident in

Figure 3.4d. Additionally, while smaller CNC agglomeration may still exist, at 1000x, we

do not see any large scale micron sized CNC agglomerates.

The average diameters of the fiber composites are listed in Table 3.3. The CA

-CA

fibers were generally consistent, varying between 130-

increased diameter is believed to be related, in part, from the additions of DMAc and

CNCs in the solution. The addition of CNC increased the viscosity of the solution at low

shear rates, causing the formed fiber to be more resistant to stretching due to the

gravitational forces, producing an overall larger fiber.

3.3.3 Birefringence of Fibers

The dispersion and orientation of CNCs within the polymeric fibers was

qualitatively studied with crossed polarizers. The two polarizers placed at right angles

would eliminate all light unless a birefrigent material is placed between them at an angle,

rotating the plane of linearized light. The highly crystalline CNCs are such a material,

and the effect is constructive when the crystals become locally oriented. Crystalline CA is

also capable of showing up as lighted areas. However, the majority of

Page 72: Experimental and modeling investigation of cellulose
Page 73: Experimental and modeling investigation of cellulose

51

CA in the fiber is in an amorphous state. Figure 3.5 shows optical microscopy pictures

of the fiber composites under the crossed polarizers. CA fibers show little to no

birefringence, indicating the polymer remains mostly amorphous with a little orientation

after the spinning process (Fig 3.5a). An increase in the amount of lighted area is

observed with the addition of CNCs to the system. Isotropic CNCs independent of CNC

concentration appear with some level of brightness, regardless of the orientation of the

sample with respect to the crossed polarizers, indicating randomness with some low

levels of local orientation. When CNCs are globally oriented, this effect intensifies and

the birefringence can be observed as bright colorful domains 57 whose intensity varies

with respect to the orientation of the sample. This may be due to liquid crystalline phase

ordering with shear as has been observed in CNC materials and films 10. This effect is

only achievable above a critical concentration of CNCs in the system, where there exist

enough CNCs dispersed throughout the fibers to interact with each other. At low

concentrations of CNC, the CNCs are mainly isotropic with some local orientation,

appearing as clear with dark spots under the cross polarizers (Fig 3.5b-c). As the CNC

concentration increases, birefringence also gradually increases until the entire fiber is

bright with colorful domains (Fig 3.5d-i). This is indicative of well dispersed and

ordered CNCs within the CA matrix, where CNC agglomerates would be expected to

appear as dark regions within the image. However, due to the magnification limits of the

technique, only large micron scale agglomerations would be observable.

3.3.4 CNC Orientation

The extent of CNC alignment along the fiber axis was quantified using 2D XRD

and Herman’s order parameter (S)29, 58. Figure 6 shows X-ray diffractograms and the

Page 74: Experimental and modeling investigation of cellulose
Page 75: Experimental and modeling investigation of cellulose

53

integrated intensity of two cast pure CNC films; shear and self-organized, along with

diffractograms of several fibers, increasing in CNC content. The self-organized CNC film

(Fig 3.6a) showed uniform intensity around the Debye-Scherrer ring patterns, indicating a

random orientation of CNCs in the horizontal plane of the film. The ordered CNC film

(Fig 3.6b) showed two sharp intensity peaks along the arc, indicating the CNCs are

aligned along the horizontal direction of the film-plane. For the CNC-CA fibers (Figs

3.6c-f), a clear trend of increasing sharp peaks along the (200) ring with increasing CNC

content, which is indicative of increasing CNC orientation along the fiber axis. Note if

there were no changes in CNC alignment, then with increasing CNCs additions there

would be an even/uniform increase in intensity along the entire (200) ring (i.e. for all ).

The CNC alignment along the fiber axis increases with CNC concentration,

(Table 3.3 and Fig 3.7) in agreement with the birefringence optical microscopy images.

There is an apparent plateau of the order parameter after 30 wt% CNCs, above which

adding more CNCs did not further improve CNC alignment along the fiber axis.

Furthermore, the maximum achieved level of CNC alignment within the fibers was

similar to that of the shear cast neat CNC films(S ~0.7),2,29 despite the different methods

and shear rates. This suggests there is a maximum level of orientation that can be

achieved at a high concentration of CNC through shear forces. When this concentration is

reached, the interactions between CNCs prevent further alignment. One possible

mechanism proposed for nanoparticle reinforced composites suggests the nanoparticles

aggregate together at high concentrations, forming a bundle. As a result of this behavior,

further orientation of the CNCs is limited.59

Page 76: Experimental and modeling investigation of cellulose
Page 77: Experimental and modeling investigation of cellulose
Page 78: Experimental and modeling investigation of cellulose

56

3.3.5 Fiber Mechanical Properties

The mechanical properties of the nanocomposite fiber are listed in Table 3.3 and

shown in Fig 3.8. The neat CA fiber has an elastic modulus of 1.9 GPa, a tensile strength

of 67.7 and reached 34.5% elongation before failure. Industrial grade cellulose acetate

fiber properties range in 2.7-2.9 GPa for elastic modulus, 139-163 MPa for tensile

strength and 25-45% for elongation.60 It is important to stress that the fibers produced in

the current study are not optimized by any post processing (washing, solvent extraction,

drawing and surface finishing) which may further improve properties.

In general, for increasing CNC additions, the elastic modulus and tensile strength

increased up to a critical CNC concentration after which point the properties plateau and

slightly decreases with further increases in CNC. In both cases, the maximum value was

reached at 34% CNC (elastic modulus: 1.9 to 14.0 GPa, and strength: 67.7 MPa to 160.3

MPa). The elongation at failure of the fibers decreased significantly by the addition of

CNCs. As observed from the SEM images of the fracture surfaces, the fibers become

increasingly brittle with higher CNC concentration with failure at 45o to the fiber axis,

indicating shear failure at the highest concentrations. The elongation at failure decreased

quickly from 35% to 5% with a small addition of CNC (2.5 wt %), and plateaus to 1-2%

with CNC additions up to 49wt%.

The role of CNC alignment in affecting mechanical properties is strongly

correlated. The overall curve shape of the modulus and tensile strength plots corresponds

directly to the orientation parameter curve in Fig 3.7. It can be suggested that the

increased CNC alignment in the fiber direction is one factor the improvement in

Page 79: Experimental and modeling investigation of cellulose

57

mechanical properties. As discussed before, CNCs are highly anisotropic. The axial

Young’s modulus of CNCs is 110-220 GPa while the transverse is 2-37 GPa1. As the

CNCs become more oriented along the fiber axis with increasing CNC concentration, a

higher mechanical reinforcement from the CNCs is observed. This effect describes the

exponential shape of the curve from 0 wt% to 30%. Once the maximum level of

orientation is achieved, the curve immediately plateaus. Any further addition of CNCs

does not increase mechanical properties.

The plateau and increase scatter in the measurements at higher CNC loads

(>34wt %) suggest defect mediated failure. The primary types of defects associated with

the fibers are considered to result from CNC agglomeration, porosities within the fiber,

uneven diameters and cross-section shape defects. It is also possible that the amount and

type of defects in the fibers became increasingly frequent as the CNC concentration

increases above 34wt%, as evident by SEM images. When agglomeration of CNCs

occurs, the overall surface contact area between the CNCs and the polymer is reduced.

Thus, the effective reinforcement effect is also decreased, causing the plateau. Porosities

or air bubbles within the fiber would decrease the effective cross-section, thus decreasing

the measured modulus, and creating stress concentrations within the fibers, leading to

premature failure. Uneven cross-section and diameters would similarly contribute to

stress concentrations. The fiber would prematurely fail at the weakest point of the fiber,

either the geometrical thinnest part or the structural weak point.

The mechanical property increases compares favorably to previous studies done

on CNC-CA composites 18, 49 by increasing the maximum reported modulus by 11.2 GPa

(1.5 to 12.7GPa) and the maximum reported strength by 116 MPa (44 to 160 MPa).

Page 80: Experimental and modeling investigation of cellulose
Page 81: Experimental and modeling investigation of cellulose

59

Table 3-3 Diameter, mechanical properties and order parameters of CNC-CA composite fibers

CNC% Diameter

Young’s

Modulus (GPa)

Tensile

strength (MPa)

Elongation at

break (%)

Herman’s order

parameter (S)

0 70±13 1.9±0.1 67.7±11.0 34.5±4.4 0.016

1 136±15 2.8±0.2 68.0±7.5 15.7±2.5 0.044

2.5 150±16 2.3±0.02 42.2±8.8 4.7±2.3 0.044

4 161±24 1.8±0.2 34.1±5.9 3.8±1.2 0.091

8 119±23 3.0±0.5 51.0±10.3 7.5±1.5 0.031

10 135±16 3.3±0.7 46.2±11.1 4.0±1.9 0.202

14 162±51 3.9±1.2 61.0±25.2 4.7±3.3 0.136

17 134±13 4.4±0.9 63.7±13.5 4.4±1.7 0.226

20 155±37 5.0±1.5 66.1±26.3 2.5±1.1 0.339

24 161±43 6.1±1.2 82.4±23.3 4.2±2.4 0.436

29 157±35 10.0±4.0 107.4±30.9 4.4±2.3 0.633

34 165±45 14.0±2.3 160.3±60.2 2.3±1.1 0.670

39 163±48 10.6±2.4 144±40.2 4.6±2.2 0.680

44 152±30 12.7±3.4 130±60.3 3.5±2.3 0.712

49 170±43 8.4±4.0 100±64.1 2.1±1.7 0.645

Page 82: Experimental and modeling investigation of cellulose

60

Furthermore, to the author’s knowledge, the critical concentration at which the maximum

mechanical properties were achieved is greater than all studies reported on Polymer-CNC

fiber composites.17-20, 22-29, 31, 38-39, 45, 61-64 This shows that by applying higher shear rates,

the resulting increased CNC alignment within the fiber provided a mechanism to better

utilize the CNC’s reinforcing potential for properties along the fiber direction. A

comparison of this study to other CNC fiber reports and common filament materials are

shown in Fig 3.1.

3.3.6 Micromechanical Modeling of CA-CNC Nanocomposite Fibers

Young’s modulus predictions from the isostrain and isostress model, the modified

orientation Cox model and the Halpin-Tsai equations for short fiber composites are

shown in Figure 3.9. The isostrain and isostress models assume all CNCs are continuous

fibers, which are either collimated along the fiber axis (isostrain) or perpendicular to it

(isostress). As expected, the isostrain case greatly overestimates the performance of the

composite, and the isostress case underestimates it, and thus provides the widest upper

and lower bounds for the CNC-CA composite performance. The Halpin-Tsai equations

provide refinement based on particle morphology, by considering that CNCs are discrete

short fibers, and the role of CNC aspect ratio. The Cox model provides even further

refinement, by considering both CNC aspect ratio and alignment.

Using empirical fittings, an aspect ratio (L/D) of 5 using Halpin-Tsai equations gives the

closest prediction to experimental values. An average value of 15 for aspect ratio was

calculated from TEM data collected from CNC samples. As shown in Fig 3.9, using this

aspect ratio over predicts the composites’ performance. However, TEM data presented is

a very small sample size, thus the experimentally obtained value may not necessarily be

Page 83: Experimental and modeling investigation of cellulose
Page 84: Experimental and modeling investigation of cellulose

62

representative. Nevertheless, one would expect the true aspect ratio to be greater than 5

since these equations do not take orientation and anisotropy of the fillers into

consideration, and assuming a perfect unidirectional filler orientation with all CNC

particle contributions at axial Young’s modulus values. Furthermore, the model assumes

perfect adhesion between the CNC and matrix, which may not be realistic for CNC-CA

interface. Also, it is also plausible that the wood CNCs, exhibit an axial modulus lower

than the 143GPa assumed. Any deviation from these assumptions would contribute to the

model over-predicting the composite mechanical performance.

The CNC reinforcement effects are considered to increase as the particles become

more aligned along the fiber axis. The modified orientation Cox model utilizes this

additional information to give a more accurate microstructural representative of the CNC-

CA composite, and predict elastic properties based off of this structure. The results from

the model are presented in Fig 3.9 for both aspect ratios of 15 and 5, serving as upper and

lower bounds. The model predictions and experimental data are in good agreement until

high concentrations of CNC (40-50 wt %) where composite performance is near or lower

than the lower bound. This may be an indication of the increased interactions or

agglomeration between the CNCs at a greater concentration.

Even though the models discussed here are simple approaches to predict

mechanical reinforcement by CNCs, they demonstrate the importance of CNC dispersion

and alignment. Increasing CNC past the optimum concentration could negatively impact

CNC dispersion, limiting elastic modulus improvement as CNC concentration increases.

Page 85: Experimental and modeling investigation of cellulose

63

3.4 Conclusion

Cellulose nanocrystal (CNCs) in cellulose acetate (CA) fiber nanocomposites were

successfully produced using a dry spinning process. Continuous CNC-CA fibers of

produced. Mechanical properties were found to be greatly improved by the CNC

additions up to 34 wt%, after which there the properties plateaued or decreased, possibly

due to defect formation from a loss in CNC dispersion or residual pores. The fiber

spinning process lead to shear-induced alignment of the CNCs along the fiber axis. The

degree of CNC alignment was shown to be one mechanism for altering the CNC-CA

fiber mechanical properties in the fiber axial direction. Micromechanical models suggest

that the increase in elastic modulus of the CNC-CA continuous fiber result from CNC

concentration, orientation state and aspect ratio. Maximum reinforcement of CNC

additions occurred at 34wt% CNC, in which the resulting improvement in tensile strength

and elastic modulus was of 137% and 637%, respectively.

Page 86: Experimental and modeling investigation of cellulose

64

Bibliography

1. Moon, R. J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J., Cellulose nanomaterials review: structure, properties and nanocomposites. Chemical Society Reviews 2011, 40 (7), 3941-3994. 2. Samir, M. A. S. A.; Alloin, F.; Dufresne, A., Review of recent research into cellulosic whiskers, their properties and their application in nanocomposite field. Biomacromolecules 2005, 6 (2), 612-626. 3. Siro, I.; Plackett, D., Microfibrillated cellulose and new nanocomposite materials: a review. Cellulose 2010, 17 (3), 459-494. 4. Beck-Candanedo, S.; Roman, M.; Gray, D. G., Effect of reaction conditions on the properties and behavior of wood cellulose nanocrystal suspensions. Biomacromolecules 2005, 6 (2), 1048-1054. 5. Araki, J.; Wada, M.; Kuga, S.; Okano, T., Flow properties of microcrystalline cellulose suspension prepared by acid treatment of native cellulose. Colloids and Surfaces A: Physicochemical and Engineering Aspects 1998, 142 (1), 75-82. 6. Bai, W.; Holbery, J.; Li, K., A technique for production of nanocrystalline cellulose with a narrow size distribution. Cellulose 2009, 16 (3), 455-465. 7. Bondeson, D.; Mathew, A.; Oksman, K., Optimization of the isolation of nanocrystals from microcrystalline cellulose by acid hydrolysis. Cellulose 2006, 13 (2), 171-180. 8. Habibi, Y.; Lucia, L. A.; Rojas, O. J., Cellulose Nanocrystals: Chemistry, Self-Assembly, and Applications. Chem Rev 2010, 110 (6), 3479-3500. 9. Reising, A. B.; Moon, R. J.; Youngblood, J. P., Effect of Particle Alignment on Mechanical Properties of Neat Cellulose Nanocrystal Films. J. Sci. Technol. For. Prod. Processes 2013, 2, 32-41. 10. Diaz, J. A.; Wu, X. W.; Martini, A.; Youngblood, J. P.; Moon, R. J., Thermal Expansion of Self-Organized and Shear-Oriented Cellulose Nanocrystal Films. Biomacromolecules 2013, 14 (8), 2900-2908. 11. Eichhorn, S. J.; Dufresne, A.; Aranguren, M.; Marcovich, N. E.; Capadona, J. R.; Rowan, S. J.; Weder, C.; Thielemans, W.; Roman, M.; Renneckar, S.; Gindl, W.; Veigel, S.; Keckes, J.; Yano, H.; Abe, K.; Nogi, M.; Nakagaito, A. N.; Mangalam, A.; Simonsen, J.; Benight, A. S.; Bismarck, A.; Berglund, L. A.; Peijs, T., Review: current international research into cellulose nanofibres and nanocomposites. J Mater Sci 2010, 45 (1), 1-33. 12. Siqueira, G.; Bras, J.; Dufresne, A., Cellulose Whiskers versus Microfibrils: Influence of the Nature of the Nanoparticle and its Surface Functionalization on the Thermal and Mechanical Properties of Nanocomposites. Biomacromolecules 2009, 10 (2), 425-432. 13. Lin, N.; Huang, J.; Dufresne, A., Preparation, properties and applications of polysaccharide nanocrystals in advanced functional nanomaterials: a review. Nanoscale 2012, 4 (11), 3274-3294. 14. Hubbe, M. A.; Rojas, O. J.; Lucia, L. A.; Sain, M., Cellulosic nanocomposites: a review. BioResources 2008, 3 (3), 929-980.

Page 87: Experimental and modeling investigation of cellulose

65

15. Magalhaes, W. L. E.; Cao, X. D.; Ramires, M. A.; Lucia, L. A., Novel all-cellulose composite displaying aligned cellulose nanofibers reinforced with cellulose nanocrystals. Tappi J 2011, 10 (4), 19-25. 16. Andersson, R. L.; Salajkova, M.; Mallon, P. E.; Berglund, L. A.; Hedenqvist, M. S.; Olsson, R. T., Micromechanical Tensile Testing of Cellulose-Reinforced Electrospun Fibers Using a Template Transfer Method (TTM). J. Polym. Environ. 2012, 20 (4), 967-975. 17. Dong, H.; Strawhecker, K. E.; Snyder, J. F.; Orlicki, J. A.; Reiner, R. S.; Rudie, A. W., Cellulose nanocrystals as a reinforcing material for electrospun poly(methyl methacrylate) fibers: Formation, properties and nanomechanical characterization. Carbohyd Polym 2012, 87 (4), 2488-2495. 18. Herrera, N. V.; Mathew, A. P.; Wang, L. Y.; Oksman, K., Randomly oriented and aligned cellulose fibres reinforced with cellulose nanowhiskers, prepared by electrospinning. Plast Rubber Compos 2011, 40 (2), 57-64. 19. Liu, D. Y.; Yuan, X. W.; Bhattacharyya, D., The effects of cellulose nanowhiskers on electrospun poly (lactic acid) nanofibres. J Mater Sci 2012, 47 (7), 3159-3165. 20. Lu, P.; Hsieh, Y. L., Cellulose nanocrystal-filled poly(acrylic acid) nanocomposite fibrous membranes. Nanotechnology 2009, 20 (41), -. 21. Magalhaes, W. L. E.; Cao, X. D.; Lucia, L. A., Cellulose Nanocrystals/Cellulose Core-in-Shell Nanocomposite Assemblies. Langmuir 2009, 25 (22), 13250-13257. 22. Park, W. I.; Kang, M.; Kim, H. S.; Jin, H. J., Electrospinning of poly(ethylene oxide) with bacterial cellulose whiskers. Macromol Symp 2007, 249, 289-294. 23. Peresin, M. S.; Habibi, Y.; Zoppe, J. O.; Pawlak, J. J.; Rojas, O. J., Nanofiber Composites of Polyvinyl Alcohol and Cellulose Nanocrystals: Manufacture and Characterization. Biomacromolecules 2010, 11 (3), 674-681. 24. Rojas, O. J.; Montero, G. A.; Habibi, Y., Electrospun Nanocomposites from Polystyrene Loaded with Cellulose Nanowhiskers. J Appl Polym Sci 2009, 113 (2), 927-935. 25. Xiang, C. H.; Joo, Y. L.; Frey, M. W., Nanocomposite Fibers Electrospun from Poly(lactic acid)/Cellulose Nanocrystals. J Biobased Mater Bio 2009, 3 (2), 147-155. 26. Zoppe, J. O.; Peresin, M. S.; Habibi, Y.; Venditti, R. A.; Rojas, O. J., Reinforcing Poly(epsilon-caprolactone) Nanofibers with Cellulose Nanocrystals. Acs Appl Mater Inter 2009, 1 (9), 1996-2004. 27. Lee, J.; Deng, Y. L., Increased mechanical properties of aligned and isotropic electrospun PVA nanofiber webs by cellulose nanowhisker reinforcement. Macromolecular Research 2012, 20 (1), 76-83. 28. Urena-Benavides, E. E.; Brown, P. J.; Kitchens, C. L., Effect of Jet Stretch and Particle Load on Cellulose Nanocrystal-Alginate Nanocomposite Fibers. Langmuir 2010, 26 (17), 14263-14270. 29. Urena-Benavides, E. E.; Kitchens, C. L., Wide-Angle X-ray Diffraction of Cellulose Nanocrystal-Alginate Nanocomposite Fibers. Macromolecules 2011, 44 (9), 3478-3484.

Page 88: Experimental and modeling investigation of cellulose

66

30. Urena-Benavides, E. E.; Kitchens, C. L., Cellulose Nanocrystal Reinforced Alginate Fibers-Biomimicry Meets Polymer Processing. Molecular Crystals and Liquid Crystals 2012, 556, 275-287. 31. Liu, L.; Yao, J. M., Wet-spinning of reinforced artificial silk hybrid fibres by cellulose whiskers. Adv Mater Res-Switz 2011, 175-176, 272-275. 32. Dri, F.; Hector, L., Jr.; Moon, R.; Zavattieri, P., Anisotropy of the elastic

Van der Waals interactions. Cellulose 2013, 20 (6), 2703-2718. 33. Wagner, R.; Raman, A.; Moon, R., Transverse elasticity of cellulose nanocrystals via atomic force microscopy. Cellulose 2010, 7, 27. 34. Orts, W. J.; Godbout, L.; Marchessault, R. H.; Revol, J. F., Enhanced ordering of liquid crystalline suspensions of cellulose microfibrils: A small angle neutron scattering study. Macromolecules 1998, 31 (17), 5717-5725. 35. Aguié-Béghin, V.; Molinari, M.; Hambardzumyan, A.; Foulon, L.; Habibi, Y.; Heim, T.; Blossey, R.; Douillard, R., Preparation of Ordered Films of Cellulose Nanocrystals. In Model Cellulosic Surfaces, American Chemical Society: 2009; Vol. 1019, pp 115-136. 36. Ago, M.; Jakes, J. E.; Johansson, L.-S.; Park, S.; Rojas, O. J., Interfacial Properties of Lignin-Based Electrospun Nanofibers and Films Reinforced with Cellulose Nanocrystals. Acs Appl Mater Inter 2012, 4 (12), 6849-6856. 37. Vallejos, M. E.; Peresin, M. S.; Rojas, O. J., All-Cellulose Composite Fibers Obtained by Electrospinning Dispersions of Cellulose Acetate and Cellulose Nanocrystals. J. Polym. Environ. 2012, 20 (4), 1075-1083. 38. Liu, L.; Yao, J. M., Properties of Biocomposite Fibers from Cellulose Nanowhiskers and Cellulose Matrix. In Journal of Fiber Bioengineering & Informatics Li, Y.; Takatera, M.; Kajiwara, K.; Li, J. S., Eds. 2012; pp 207-215. 39. Uddin, A.; Araki, J.; Gotoh, Y., Toward Strong Green Nanocomposites: Polyvinyl Alcohol Reinforced with Extremely Oriented Cellulose Whiskers. Biomacromolecules 2011, 12 (3), 617-624. 40. Endo, R.; Saito, T.; Isogai, A., TEMPO-oxidized cellulose nanofibril/poly(vinyl alcohol) composite drawn fibers. Polymer 2013, 54 (2), 935 - 941. 41. Iwamoto, S.; Isogai, A.; Iwata, T., Structure and Mechanical Properties of Wet-Spun Fibers Made from Natural Cellulose Nanofibers. Biomacromolecules 2011, 12 (3), 831-836. 42. Håkansson, K. M. O.; Fall, A. B.; Lundell, F.; Yu, S.; Krywka, C.; Roth, S. V.; Santoro, G.; Kvick, M.; Prahl Wittberg, L.; Wågberg, L.; Söderberg, L. D., Hydrodynamic alignment and assembly of nanofibrils resulting in strong cellulose filaments. Nat Commun 2014, 5. 43. Walther, A.; Timonen, J. V. I.; Diez, I.; Laukkanen, A.; Ikkala, O., Multifunctional High-Performance Biofibers Based on Wet-Extrusion of Renewable Native Cellulose Nanofibrils. Adv Mater 2011, 23 (26), 2924-+. 44. Chen, S.; Schueneman, G.; Pipes, R. B.; Youngblood, J.; Moon, R. J., Effects of Crystal Orientation on Cellulose Nanocrystals–Cellulose Acetate Nanocomposite Fibers Prepared by Dry Spinning. Biomacromolecules 2014, 15 (10), 3827-3835.

Page 89: Experimental and modeling investigation of cellulose

67

45. Uddin, A. J.; Araki, J.; Gotoh, Y., Characterization of the poly(vinyl alcohol)/cellulose whisker gel spun fibers. Composites Part A: Applied Science and Manufacturing 2011, 42 (7), 741 - 747. 46. Wibowo, A. C.; Misra, M.; Park, H.-M.; Drzal, L. T.; Schalek, R.; Mohanty, A. K., Biodegradable nanocomposites from cellulose acetate: Mechanical, morphological, and thermal properties. Composites Part A: Applied Science and Manufacturing 2006, 37 (9), 1428 - 1433. 47. Hassan-Nejad, M.; Ganster, J.; Bohn, A.; Pinnow, M.; Volkert, B., Bio-Based Nanocomposites of Cellulose Acetate and Nano-Clay with Superior Mechanical Properties. Macromol Symp 2009, 280 (1), 123-129. 48. Rodríguez, F. J.; Coloma, A.; Galotto, M. J.; Guarda, A.; Bruna, J. E., Effect of organoclay content and molecular weight on cellulose acetate nanocomposites properties. Polymer Degradation and Stability 2012, 97 (10), 1996-2001. 49. Yang, Z.-Y.; Wang, W.-J.; Shao, Z.-Q.; Zhu, H.-D.; Li, Y.-H.; Wang, F.-J., The transparency and mechanical properties of cellulose acetate nanocomposites using cellulose nanowhiskers as fillers. Cellulose 2013, 20 (1), 159-168. 50. Reiner, R.; Rudie, A., Process scale-up of cellulose nanocrystal production to 25 kg per batch at the Forest Products Laboratory. TAPPI Press: Peachtree Corners, GA, 2013. 51. Salyer, I. O.; Heyd, J. W.; Brodbeck, R. M.; Hartzel, L. W.; Brown, L. E., Use of a capillary extrusion rheometer to measure curing of thermosetting plastics and rubbers. Journal of Polymer Science Part A: General Papers 1965, 3 (5), 1911-1939. 52. Nishiyama, Y.; Langan, P.; Chanzy, H., Crystal structure and hydrogen-bonding system in cellulose 1 beta from synchrotron X-ray and neutron fiber diffraction. J Am Chem Soc 2002, 124 (31), 9074-9082. 53. S. Advani, Process Modeling in Composites Manufacturing, Second Edition (Manufacturing Engineering and Materials Processing); CRC Press; 2010; pp 132-134 54. Madsen, B.; Joffe, R.; Peltola, H.; Nattinen, K., Short cellulosic fiber/starch acetate composites - micromechanical modeling of Young's modulus. J Compos Mater 2011, 45 (20), 2119-2131. 55. Cox, H. L., The Elasticity and Strength of Paper and Other Fibrous Materials. British Journal of Applied Physics 1952, 3 (MAR), 72-79. 56. Liu, D.; Chen, X.; Yue, Y.; Chen, M.; Wu, Q., Structure and rheology of nanocrystalline cellulose. Carbohyd Polym 2011, 84 (1), 316 - 322. 57. Revol, J. F.; Godbout, L.; Dong, X. M.; Gray, D. G.; Chanzy, H.; Maret, G., Chiral Nematic Suspensions of Cellulose Crystallites - Phase-Separation and Magnetic-Field Orientation. Liq Cryst 1994, 16 (1), 127-134. 58. Park, S.; Baker, J. O.; Himmel, M. E.; Parilla, P. A.; Johnson, D. K., Cellulose crystallinity index: measurement techniques and their impact on interpreting cellulase performance. Biotechnol. Biofuels 2010, 3. 59. Misiego, C. R.; Pipes, R. B., Dispersion and its relation to carbon nanotube concentration in polyimide nanocomposites. Compos Sci Technol 2013, 85, 43-49. 60. Lewin, M., Handbook of Fiber Chemistry. Boca Raton : CRC/Taylor & Francis New York, 2007.

Page 90: Experimental and modeling investigation of cellulose

68

61. He, X.; Xiao, Q.; Lu, C. H.; Wang, Y. R.; Zhang, X. F.; Zhao, J. Q.; Zhang, W.; Zhang, X. M.; Deng, Y. L., Uniaxially Aligned Electrospun All-Cellulose Nanocomposite Nanofibers Reinforced with Cellulose Nanocrystals: Scaffold for Tissue Engineering. Biomacromolecules 2014, 15 (2), 618-627. 62. Li, Y. J.; Ko, F. K.; Hamad, W. Y., Effects of Emulsion Droplet Size on the Structure of Electrospun Ultrafine Biocomposite Fibers with Cellulose Nanocrystals. Biomacromolecules 2013, 14 (11), 3801-3807. 63. Liao, H. Q.; Wu, Y. Q.; Wu, M. Y.; Zhan, X. R.; Liu, H. Q., Aligned electrospun cellulose fibers reinforced epoxy resin composite films with high visible light transmittance. Cellulose 2012, 19 (1), 111-119. 64. Martinez-Sanz, M.; Olsson, R. T.; Lopez-Rubio, A.; Lagaron, J. M., Development of bacterial cellulose nanowhiskers reinforced EVOH composites by electrospinning. J Appl Polym Sci 2012, 124 (2), 1398-1408.

Page 91: Experimental and modeling investigation of cellulose

69

CHAPTER 4. DRY SPUN POLYVINYL ALCOHOL FIBERS REINFORCED WITH CELLULOSE NANOCRYSTALS

Continuous Cellulose Nanocrystal (CNC)-Polyvinyl alcohol (PVA) dry spun

composite fibers were prepared using a plunger style fiber spinner. The ternary spinning

system consists of PVA, CNCs, and water. This aqueous system was chosen due to

CNC’s optimal dispersion in water. Increasing amounts of CNCs (up to 25 wt %) were

dispersed into PVA fibers in efforts to improve the tensile strength and elastic modulus of

the fiber. Mechanical testing of the fibers show maximum modulus improvement of

602% (1.95GPa to 13.7GPa) at 15wt% CNC, and maximum tensile strength improvement

of 283% (41.16MPa to 157.5MPa). Elongation of the fibers generally decreased with

addition of CNCs. However, the variations of elongation at failure between samples of

same CNC% are large, indicating a great percentage of defect driven failures. Similar to

CNC-CA fibers, birefringence was observed in the fibers under crossed polarizers. CNC-

PVA fibers also showed significantly greater improvement in strength and modulus in

comparison to CNC-CA fibers at the same CNC wt%. This may be attributed to the

improved dispersion of CNCs in the water based spinning solution vs. the

DMAc/Acetone solution.

4.1 Introduction

Cellulose nanocrystals (CNCs) are a class of newly developed and sustainable

nanomaterial derived from cellulose-based materials such as wood. There have been high

Page 92: Experimental and modeling investigation of cellulose

70

research efforts to utilize these materials as reinforcing agents due to their sustainability,

abundance, low cost and superior mechanical properties. In order to develop CNC

nanocomposites with industrial applications, it is necessary to understand and control how

addition of CNCs leads to improvement of polymer mechanical properties. The goal of this

study is to characterize and understand the effect of CNCs on the properties of dry spun

polyvinyl alcohol-CNC fibers. To achieve this goal, we developed synthesis methods to

produce lab-scale polyvinyl alcohol (PVA) and CNC composite fibers, identified variables

that affect nanocomposite properties and proposed hypothesis that explain mechanical

property improvement by addition of CNCs to PVA, as well as comparisons of the effect

of CNCs on different polymer matrices.

Due to the highly anisotropic nature of CNCs, in which the elastic moduli in the

long axis of the particle is much greater (110-220 GPa1-2) than in the transverse directions

(2-50 GPa1, 3), preferential orientation of CNCs within the fiber composites would greatly

increase the mechanical performance of the material. In this study, shear deformation

during composite production is utilized to create orientation within the composite fiber.

For a description of how shear forces affect CNC orientation, please refer to Chapter 3.1.

Fiber spinning is one processing technique that can generate high shear rates. CNCs

have been used in fiber spinning of various polymer systems, via electrospinning,4-15 and

dry or wet spinning.16-22 The spinning process consists of a polymer dissolved in solvent

extruded through a device containing small orifices known as the spinneret. The filaments

then may be passed through air or liquids to solidify, known as dry or wet spinning,

respectively. For electrospinning, high voltage is applied between the orifice and a “target”

Page 93: Experimental and modeling investigation of cellulose

71

which effectively accelerates the jet of solution to the target. In most studies, the addition

of CNCs has resulted in an increase in mechanical properties. Electrospun CNC-polymer

composite fibers report typically low concentration of CNC loading (less than 20 wt %),

and as compared to the neat polymer fiber, there is an increase in mechanical strength and

elastic modulus in the fiber direction. However, mechanical properties were tested on fiber

mats instead of single fibers, and as a result are highly dependent of the mat configuration

(number and alignment of the fibers).

Prior to the author’s work in Chapter 3, only wet spun CNC fibers have been

produced via laboratory scale syringe pump systems, in which the resulting fibers (60-

ameter) have up to 50 vol % CNC loadings,18 and CNC alignment in the

fiber direction.23 In most cases, the addition of CNCs increased mechanical properties of

individual fibers. Liu et al.17 wet spun CNC- artificial silk fibers, reporting a 3 times

increase in modulus and tensile strength at 5 wt % CNC loading. The same group wet

spun CNC/regenerated cellulose fibers (0 to 9 wt % CNC loadings),24 resulting in an

increase in both the tensile strength (from 172 to 571 MPa) and the elastic modulus (from

2.06 to 4.15 GPa). Iwamoto et al.23 wet spun fibers made from 100 % cellulose

nanofibrils (CNF) and observed a direct relationship between shear rate, CNF alignment

and mechanical strength. Similar trends were reported by Hakansson et al.25 for CNF

fibers and by Urena-Benavides et al.18-19, 26 for CNC-alginate fibers containing up to 50

wt % CNC. Despite these improvements in mechanical properties by the addition of

CNCs, these experimental values are lower than micromechanical model prediction for

composite properties, suggesting improvements are possible.

Page 94: Experimental and modeling investigation of cellulose

72

PVA is water soluble, biodegradable, nontoxic, and biocompatible, thus has a

wide range of applications from filtration, membranes, to 3D printing. The fiber form of

PVA, sometimes known as vinylon, is widely used in textiles, ropes, and shoes. Short

fibers of PVAs are often used as reinforcement in concrete.6 Incorporating CNCs into

PVA fiber can greatly improve the mechanical properties of the fiber while still

maintaining the eco-compatibility of the polymer to create a light weight and strong

nanocomposite. There has been several reports on wet spun and electrospun PVA-CNC

fiber systems. Uddin et al.21 wet spun composite PVA fibers with CNC derived from

cotton. The study reported promising results of a 40% and 27% increase in modulus and

strength respectively. However, out of the CNC concentration studied (5 wt% to

30 wt %), the greatest results were achieved with the lowest percentage (5 wt %). Further

additions of CNC resulted in decreases in mechanical strength. Peresin et al.27

electrospun reinforced PVA nanofibers between 0 and 15 wt% CNC concentration. . The

maximum modulus of the electrospun fiber mat reported was 57 MPa at 15 % CNC, a

270 % increase from the neat polymer fiber mat. No tensile strength data was reported.

Lee et al.11 also produced electrospun CNC-PVA fibers and reported improvements in

modulus and strength. The enhancements were reported to be 117% and 96% in modulus

and strength respectively.

In this study, continuous cellulose nanocrystal (CNC)-Polyvinyl alcohol (PVA)

dry spun composite fibers were prepared using a plunge style fiber spinner. The ternary

spinning system consists of PVA, CNCs, and water. The role of CNC concentration and

CNC alignment on the processing and resulting mechanical properties were investigated.

Experimental results were compared to predictions from micromechanical models that

Page 95: Experimental and modeling investigation of cellulose

73

accounted for CNC concentration and aspect ratio. In addition, the results were

compared to a CA-CNC system that was developed on the same experimental method.

4.2 Materials and Methods

4.2.1 Materials

99+% hydrolyzed Polyvinyl Alcohol (PVA) with an average molecular weight of

89,000-98,000 was purchased from Sigma-Aldrich Chemistry. Water-based suspensions

of CNCs were produced and supplied by the USDA Forest Service-Forest Products

Laboratory, Madison, WI.28 The morphology of individual CNC particles was

characterized via transmission electron microscopy (TEM; Philips CM-100, at 100 kV,

spot 3, 200 µm condenser aperture, and 50 µm objective aperture,) of CNC particles

deposited on TEM grids (400 mesh Formvar/carbon filmed grid prepared with blow

discharge) with 2 % aqueous uranyl acetate stain. The average diameter (6 ± 2 nm) and

length (83 ± 43 nm) of 93 individual CNC crystals were determined using within Image J

software.

4.2.2 Methods

4.2.2.1 Preparation of PVA-CNC Fibers

PVA was dissolved in nanopure water by agitation overnight at 85°C, with the

desired amount of wet CNC (as received CNC in water solution) added. The spinning

system and solution was heated to 98°C for 2 hours for equilibration prior to spinning.

The prepared solutions were dry spun using a custom built plunger style fiber spinner

based on a capillary viscometer16, 29 with spinneret insert having a 30 ° conical V-taper

machined to a 0.190 ” insert for a 25 This allows processing

Page 96: Experimental and modeling investigation of cellulose

74

at shear rates similar to industrial processes (105-106 /s) that is much greater than typical

laboratory scale setups (10-102 /s). The spinner was attached to a universal tensile testing

machine (MTS Insight) to control displacement rate and measure pressure. The spinneret

was mounted onto the MTS with counter current heated air at the spinner orifice. The

dope solution was loaded into the spinneret and the spinning process was started by

pressing the plunger into the reservoir at maximum crosshead speed (1500 mm/min) until

flow began. The continual spinning rate was regulated by the minimum amount of force

needed to maintain fiber flow. The fibers were then manually collected and further dried

under vacuum at 3.4 kPa and 40 °C overnight. PVA-CNC nanocomposite fibers were

prepared with CNC concentrations of 0, 1, 2, 5, 7, 10, 15, 20, and 25 wt %. TGA

experiments were carried out to ensure the final fiber contained no more than 5 wt %

solvents. More details may be found in Chen et al.16

4.2.2.2 Microscopy of PVA-CNC Fibers

Optical Microscopy (OM) (Olympus ZX-12) was used to observe the diameter

and the surface morphology of the fibers. Optical images of each test specimen were

taken. The diameters of the fibers were measured using the measure tool in ImageJ. Flow

birefringence of CNCs within the fibers was observed in a Carl Zeiss (Axio Observer A1)

inverted microscope with cross-polarizers. Scanning electron microscopy (SEM) (FEI,

XL40, at 2 kV) was used for close surface and cross section observation. All samples

were coated in gold/palladium for 30 s at 1 kV, 18 mA in a sputter coater (Hummer 6.2)

prior to observation to minimize charging effects.

Page 97: Experimental and modeling investigation of cellulose

75

4.2.2.3 Mechanical Testing of PVA-CNC Fibers

The tensile elastic modulus, ultimate strength and strain at failure were measured

using a dynamic mechanical analysis machine (TA Instruments, Q800). At least 10 fiber

segments were randomly selected from specimens of each CNC concentration. “C”

shaped tabs with adhesives were used to attach the samples to the DMA and avoid

pinching at endpoints. Optical microscopy was used to determine the diameter of each,

and ensure there were no obvious defects. Fibers are assumed to be cylindrical in cross

section as supported by optical imaging. Single fibers were tested with a gauge length of

30 mm, a constant applied force rate of 0.3 N/min at room temperature and relative

humidity of 40 to 50 % until fracture. Specimens that failed outside of the gauge length

were considered improperly loaded, and were discarded in the data analysis.

4.2.2.4 Micromechanical Modeling of PVA-CNC Fibers

Several micromechanical models were used to estimate the elastic modulus of

CA-CNC fibers as a function of CNC loading: Isostrain, isostress, Halpin-Tsai, and a

modified Cox model. Isostrain and isostress are basic micromechanical models of

idealized geometries. They assume the reinforcing filler of the composites are continuous

and perfectly collimated in one direction, and there is perfect adhesion between the filler

and the polymer matrix. The Halpin-Tsai model is an analytical form of Hill’s

generalized self-consistent model for composite materials. It takes into account short

fibers and aspect ratios. The fillers are assumed to be collimated, and there is perfect

adhesion between the filler and the polymer matrix. The modified Cox model takes

Page 98: Experimental and modeling investigation of cellulose

76

porosity and orientation effects into account to predict the Young’s modulus of the fiber

composites. More details of the models used can be found in Chapter 3.

4.3 Results and Discussion

4.3.1 Morphologies and Surface Structure of PVA-CNC Fiber

Figure 4.1 shows the optical microscopy images of the PVA-CNC fibers. The

fibers appeared to be cylindrical in cross section with relatively smooth surfaces. Some

defects such as air voids, and surface irregularities can be observed. The fibers with

higher CNC concentrations also appear to exhibit more defects. Possible sources of these

defect include insufficient drying, air bubbles or impurities in the solution as well as

rheological changes due to the addition of CNCs to the system. A skinning effect can also

be observed on some of the fibers, where the fiber appear to have a shell where the outer

layers solidified before the inner layer, creating a visible layering effect.

Crossed polarizer optical microscopy image of PVA-CNC fibers with CNC

concentrations of 0, 1, 2, 7, 15, 25 wt% respectively are shown in Figure 4.2. Due to the

birefrigent behavior of highly crystalline CNCs, the dispersion and orientation of CNCs

within the polymeric fibers may be qualitatively studied with crossed polarizers. A

detailed explanation is included in Chapter 3. For no to low CNC content, the

micrographs show little to no birefringence, indicating the polymer remains mostly

amorphous with a low degree of orientation after the spinning process (Figure 4.2a,b). An

increase in the amount birefringence is observed with the addition of CNCs to the system.

As the CNC concentration increases, birefringence also gradually increases until the

entire fiber is bright with colorful domains (Figure 4.2c-f). This suggests that the dry

spinning process resulted in some preferential orientation of the CNCs within the fiber,

Page 99: Experimental and modeling investigation of cellulose
Page 100: Experimental and modeling investigation of cellulose
Page 101: Experimental and modeling investigation of cellulose

79

and this effect amplifies with increasing CNC concentration.30-31 Figure 4.2g shows a

PVA-CNC fiber exhibiting the skinning effect mentioned above. The change in color in

the micrograph indicate a possible change in local orientation of the fiber. This effect

may be caused by relaxation of the orientation of the CNCs in the inner core of the fiber.

While the outer layer is solidified with the CNC orientation locked in, the inner core

takes longer to solidify, thus the CNCs are given more relaxation time.

To study the surface morphology of the fibers, scanning electron microscopy

images of the fibers were taken (Figure 4.3). An increase in surface roughness is

observed as a result of CNC reinforcement. Where PVA fibers exhibit more classical

polymeric behavior such as plastic deformation, high concentration CNC-PVA fibers

exhibit brittle breakage and fibrous inner structures.

4.3.2 Mechanical Properties of PVA-CNC Fibers

The mechanical properties of the nanocomposite fiber are shown in Figure 4.4.

The neat PVA fiber has an elastic modulus of 1.95 GPa, a tensile strength of 41.16 MPa

and reached 26.4 % elongation before failure. In comparison, industrial grade PVA fiber

(KuralonTM 5900) properties are significantly higher, ranging in 46 GPa for elastic

modulus, 1.8 GPa for tensile strength.32 It is important to stress that the fibers produced in

the current study are not optimized by any post processing (washing, solvent extraction,

drawing and surface finishing) which could further improve properties.

In general, the elastic modulus and tensile strength increased up to a critical CNC

concentration after which point the properties plateau and slightly decreases with further

addition of CNC. This supports the microscopy results where more defects were visible at

higher CNC concentrations. Maximum modulus improvement of 602% (1.95GPa to

Page 102: Experimental and modeling investigation of cellulose
Page 103: Experimental and modeling investigation of cellulose

81

Figure 4-4 Mechanical properties of as spun PVA-CNC fibers, without post processing: a) Young’s modulus, b) elongation at break and c) tensile strength, as a function of

CNC %.

Page 104: Experimental and modeling investigation of cellulose

82

13.7GPa) at 15wt% CNC, and maximum tensile strength improvement of 283%

(41.16MPa to 157.5MPa). Elongation of the fibers generally decreased with addition of

CNCs. However, the variations of elongation at failure between samples of same CNC%

are very high, indicating high percentage of defect driven failures, where the fiber

prematurely fails either from the geometrical thinnest part or the structural weak point of

the fiber. Samples of static force vs. strain curves of the fibers are shown in Figure 4.5.

The graph shows elastic deformation in the initial stages of strain, the fibers undergoes

plastic deformation past 5% strain where they fail ductally at a higher strain rate. Since

the ultimate tensile strength is already reached, this results in a large deviation in

elongation, but not in ultimate tensile strength. However, the average value of elongation

does not decrease drastically, suggesting that the CNCs are not increasing the brittleness

of the fiber.

The primary types of defects associated with the fibers produced in this study are

considered to result from porosity within the fiber, CNC agglomeration, and uneven fiber

diameter and cross-sectional shape. CNC agglomeration decreases the overall CNC-

polymer interfacial area, which decreases the potential reinforcing effect of the CNCs,

and causes the plateau at higher CNC concentrations. Porosity or air voids within the

fiber not only decreases the effective cross-section, which decreases the measured

modulus, but also creates stress concentrations within the fibers, leading to premature

failure. Likewise, uneven cross-section and diameters also contribute to stress

concentrations.

The increase in mechanical property observed in this study is similar to what was

reported in the previous chapter, with few distinctions. CNC-PVA fibers showed

Page 105: Experimental and modeling investigation of cellulose

83

Figure 4-5 Examples of static force vs. strain curves of PVA-CNC fibers

Page 106: Experimental and modeling investigation of cellulose

84

higher improvement in strength and modulus in comparison to CNC-CA fibers at the

same CNC wt%. However, the maximum improvements occurred at an earlier CNC

concentration (15-20 wt %) vs the maximum for CA-CNC (34 wt %). Previously, the

standard deviation of the modulus and ultimate tensile strength increased with increasing

CNC concentration. This was attributed to agglomeration of CNCs where the overall

surface contact area between the CNCs and the polymer is reduced. Thus, the effective

reinforcement effect is also decreased locally. The large variation in performance comes

from random agglomerations across the fibers, creating uneven reinforcement. The

elongation dropped significantly upon addition of CNC from 35% to 5% with a small

addition of CNC (2.5 wt %), and plateaus to 1-2% with CNC additions up to 49wt%. In

the current study, the spread of modulus and ultimate tensile strength stay relatively

small, however, the elongation deviation is large. This suggests the stretchability of the

fibers is maintained, however, the variability is high. This difference between the two

systems may be attributed to the improved dispersion of CNCs in the water based

spinning solution vs. the DMAc/Acetone solution. CNCs are most well dispersed in

water, but have some dispersion in DMAc as well.33-34 In the CA-CNC system,

agglomeration of CNCs is the main source of fiber failure. As the dispersion of CNCs

improved in the PVA aqueous system, the level of reinforcement becomes more

consistent, and the brittleness of the fiber decreases. Instead, defects such as porosities

and uneven cross-sections becomes the dominate method of failure; the fiber would

prematurely fail at the weakest point of the fiber, either the geometrical thinnest part or

the structural weak point.

Page 107: Experimental and modeling investigation of cellulose

85

Figure 4-6 Predictions of the Young’s modulus using isostrain, isostress, Halpin-Tsai equation and the Cox model with aspect ratios of 15 plotted along with experimental data.

Page 108: Experimental and modeling investigation of cellulose

86

4.3.3 Micromechanical Modeling of PVA-CNC Fibers

The micromechanical modeling used in Chapter 3 to the CA-CNC fibers was

applied here for the PVA-CNC system. Young’s modulus predictions from the isostrain

and isostress model, the Cox model and the Halpin-Tsai equations for short fiber

composites are shown in Figure 4.5. The isostrain and isostress models assume all CNCs

are continuous fibers that are either collimated along the fiber axis (isostrain) or

perpendicular to it (isostress). Similar to the case of CA-CNC fibers, the isostrain case

greatly overestimates the performance of the composite, and the isostress case

underestimates it. Thus, these models provide the widest upper and lower bounds for the

CA-CNC composite performance. The outlier that exceeds the upper bound at 2 wt%

CNC may be attributed to experimental deviation, as well as a result of polymer chain

orientation. The spinning process aligns the molecular orientation of the polymer chain,

and consequently, the resulting fiber may exhibit a higher modulus than that of the bulk

PVA modulus.

The Halpin-Tsai equations provide refinement based on particle morphology, by

considering that CNCs are discrete short fibers, and the role of CNC aspect ratio. The

Cox model provides even further refinement, by taking into account the stress transfer

between the CNC and the matrix using the shear-lag model, as well as a filler orientation

efficiency factor. Due to the lack of orientation data, it is estimated that all CNCs are

collimated along the fiber axis.

Using the experimentally determined aspect ratio of 15, both Halpin-Tsai and Cox

model predictions are in good agreement with the experimental modulus data of the

PVA-CNC fibers up to 15 wt%. The models over predict the performance of the fibers at

Page 109: Experimental and modeling investigation of cellulose

87

20 and 25 wt%. This along with the increase of fiber performance variation at those

concentrations suggest the start of the agglomeration issue of CNCs. Note that as

previously discussed in Chapter 3, the CA-CNC fibers experiences agglomeration issues

earlier on, and consistently underperformed the predictions using an aspect ratio of 15.

This again emphasizes the importance of dispersion of CNCs in the spinning solution.

4.4 Conclusion

In this study, continuous PVA-CNC dry spun composite fibers were developed

using a plunger style fiber spinner. Increasing amounts of CNCs (up to 25 wt %) was

dispersed into PVA fibers in efforts to improve the tensile strength and elastic modulus of

the fiber. Mechanical testing of the fibers show maximum modulus improvement of

602% (1.95GPa to 13.7GPa) at 15wt% CNC, and maximum tensile strength improvement

of 283% (41.16MPa to 157.5MPa). Elongation of the fibers generally slightly decreased

with addition of CNCs, however, the stretchability of the fiber remained high. The

variations of elongation at failure between samples of same CNC% are large, indicating

great percentage of defect driven failures. Similar to CNC-CA fibers, birefringence was

observed in the fibers under crossed polarizers, and increased with respect to CNC

concentration. CNC-PVA fibers showed significantly greater improvement in strength

and modulus in comparison to CNC-CA fibers at the same CNC wt%. This may be

attributed to the improved dispersion of CNCs in the water based spinning solution.

Micromechanical data gives good predictions for the modulus performance of the

composite at CNC concentration < 20wt% using the experimentally determined aspect

ratio of 15.

Page 110: Experimental and modeling investigation of cellulose

88

Bibliography

1. Dri, F.; Hector, L., Jr.; Moon, R.; Zavattieri, P., Anisotropy of the elastic properties

Waals interactions. Cellulose 2013, 20 (6), 2703-2718. 2. Moon, R. J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J., Cellulose nanomaterials review: structure, properties and nanocomposites. Chemical Society Reviews 2011, 40 (7), 3941-3994. 3. Wagner, R.; Raman, A.; Moon, R., Transverse elasticity of cellulose nanocrystals via atomic force microscopy. Cellulose 2010, 7, 27. 4. Dong, H.; Strawhecker, K. E.; Snyder, J. F.; Orlicki, J. A.; Reiner, R. S.; Rudie, A. W., Cellulose nanocrystals as a reinforcing material for electrospun poly(methyl methacrylate) fibers: Formation, properties and nanomechanical characterization. Carbohyd Polym 2012, 87 (4), 2488-2495. 5. Herrera, N. V.; Mathew, A. P.; Wang, L. Y.; Oksman, K., Randomly oriented and aligned cellulose fibres reinforced with cellulose nanowhiskers, prepared by electrospinning. Plast Rubber Compos 2011, 40 (2), 57-64. 6. Liu, D. Y.; Yuan, X. W.; Bhattacharyya, D., The effects of cellulose nanowhiskers on electrospun poly (lactic acid) nanofibres. J Mater Sci 2012, 47 (7), 3159-3165. 7. Park, W. I.; Kang, M.; Kim, H. S.; Jin, H. J., Electrospinning of poly(ethylene oxide) with bacterial cellulose whiskers. Macromol Symp 2007, 249, 289-294. 8. Peresin, M. S.; Habibi, Y.; Zoppe, J. O.; Pawlak, J. J.; Rojas, O. J., Nanofiber Composites of Polyvinyl Alcohol and Cellulose Nanocrystals: Manufacture and Characterization. Biomacromolecules 2010, 11 (3), 674-681. 9. Rojas, O. J.; Montero, G. A.; Habibi, Y., Electrospun Nanocomposites from Polystyrene Loaded with Cellulose Nanowhiskers. J Appl Polym Sci 2009, 113 (2), 927-935. 10. Zoppe, J. O.; Peresin, M. S.; Habibi, Y.; Venditti, R. A.; Rojas, O. J., Reinforcing Poly(epsilon-caprolactone) Nanofibers with Cellulose Nanocrystals. Acs Appl Mater Inter 2009, 1 (9), 1996-2004. 11. Lee, J.; Deng, Y. L., Increased mechanical properties of aligned and isotropic electrospun PVA nanofiber webs by cellulose nanowhisker reinforcement. Macromolecular Research 2012, 20 (1), 76-83. 12. Xiang, C. H.; Joo, Y. L.; Frey, M. W., Nanocomposite Fibers Electrospun from Poly(lactic acid)/Cellulose Nanocrystals. J Biobased Mater Bio 2009, 3 (2), 147-155. 13. Ago, M.; Jakes, J. E.; Johansson, L.-S.; Park, S.; Rojas, O. J., Interfacial Properties of Lignin-Based Electrospun Nanofibers and Films Reinforced with Cellulose Nanocrystals. Acs Appl Mater Inter 2012, 4 (12), 6849-6856. 14. Vallejos, M. E.; Peresin, M. S.; Rojas, O. J., All-Cellulose Composite Fibers Obtained by Electrospinning Dispersions of Cellulose Acetate and Cellulose Nanocrystals. J. Polym. Environ. 2012, 20 (4), 1075-1083. 15. Lu, P.; Hsieh, Y. L., Cellulose nanocrystal-filled poly(acrylic acid) nanocomposite fibrous membranes. Nanotechnology 2009, 20 (41), -.

Page 111: Experimental and modeling investigation of cellulose

89

16. Chen, S.; Schueneman, G.; Pipes, R. B.; Youngblood, J.; Moon, R. J., Effects of Crystal Orientation on Cellulose Nanocrystals–Cellulose Acetate Nanocomposite Fibers Prepared by Dry Spinning. Biomacromolecules 2014, 15 (10), 3827-3835. 17. Liu, L.; Yao, J. M., Wet-spinning of reinforced artificial silk hybrid fibres by cellulose whiskers. Adv Mater Res-Switz 2011, 175-176, 272-275. 18. Urena-Benavides, E. E.; Brown, P. J.; Kitchens, C. L., Effect of Jet Stretch and Particle Load on Cellulose Nanocrystal-Alginate Nanocomposite Fibers. Langmuir 2010, 26 (17), 14263-14270. 19. Urena-Benavides, E. E.; Kitchens, C. L., Wide-Angle X-ray Diffraction of Cellulose Nanocrystal-Alginate Nanocomposite Fibers. Macromolecules 2011, 44 (9), 3478-3484. 20. Liu, L.; Yao, J. M., Properties of Biocomposite Fibers from Cellulose Nanowhiskers and Cellulose Matrix. In Journal of Fiber Bioengineering & Informatics Li, Y.; Takatera, M.; Kajiwara, K.; Li, J. S., Eds. 2012; pp 207-215. 21. Uddin, A.; Araki, J.; Gotoh, Y., Toward Strong Green Nanocomposites: Polyvinyl Alcohol Reinforced with Extremely Oriented Cellulose Whiskers. Biomacromolecules 2011, 12 (3), 617-624. 22. Endo, R.; Saito, T.; Isogai, A., TEMPO-oxidized cellulose nanofibril/poly(vinyl alcohol) composite drawn fibers. Polymer 2013, 54 (2), 935 - 941. 23. Iwamoto, S.; Isogai, A.; Iwata, T., Structure and Mechanical Properties of Wet-Spun Fibers Made from Natural Cellulose Nanofibers. Biomacromolecules 2011, 12 (3), 831-836. 24. Liu, L.; Yao, J., Properties of biocomposite fibers from cellulose nanowhiskers and cellulose matrix. Journal of Fiber Bioengineering and Informatics 2012, 5 (2), 207-215. 25. Håkansson, K. M. O.; Fall, A. B.; Lundell, F.; Yu, S.; Krywka, C.; Roth, S. V.; Santoro, G.; Kvick, M.; Prahl Wittberg, L.; Wågberg, L.; Söderberg, L. D., Hydrodynamic alignment and assembly of nanofibrils resulting in strong cellulose filaments. Nat Commun 2014, 5. 26. Urena-Benavides, E. E.; Kitchens, C. L., Cellulose Nanocrystal Reinforced Alginate Fibers-Biomimicry Meets Polymer Processing. Molecular Crystals and Liquid Crystals 2012, 556, 275-287. 27. Peresin, M. S.; Habibi, Y.; Vesterinen, A. H.; Rojas, O. J.; Pawlak, J. J.; Seppala, J. V., Effect of Moisture on Electrospun Nanofiber Composites of Poly(vinyl alcohol) and Cellulose Nanocrystals. Biomacromolecules 2010, 11 (9), 2471-2477. 28. Reiner, R.; Rudie, A., Process scale-up of cellulose nanocrystal production to 25 kg per batch at the Forest Products Laboratory. TAPPI Press: Peachtree Corners, GA, 2013. 29. Salyer, I. O.; Heyd, J. W.; Brodbeck, R. M.; Hartzel, L. W.; Brown, L. E., Use of a capillary extrusion rheometer to measure curing of thermosetting plastics and rubbers. Journal of Polymer Science Part A: General Papers 1965, 3 (5), 1911-1939. 30. Diaz, J. A.; Wu, X. W.; Martini, A.; Youngblood, J. P.; Moon, R. J., Thermal Expansion of Self-Organized and Shear-Oriented Cellulose Nanocrystal Films. Biomacromolecules 2013, 14 (8), 2900-2908.

Page 112: Experimental and modeling investigation of cellulose

90

31. Revol, J. F.; Godbout, L.; Dong, X. M.; Gray, D. G.; Chanzy, H.; Maret, G., Chiral Nematic Suspensions of Cellulose Crystallites - Phase-Separation and Magnetic-Field Orientation. Liq Cryst 1994, 16 (1), 127-134. 32. Peijs, T.; Van Vught, R.; Govaert, L., Mechanical properties of poly (vinyl alcohol) fibres and composites. Composites 1995, 26 (2), 83-90. 33. Beck, S.; Bouchard, J.; Berry, R., Dispersibility in water of dried nanocrystalline cellulose. Biomacromolecules 2012, 13 (5), 1486-1494. 34. Viet, D.; Beck-Candanedo, S.; Gray, D. G., Dispersion of cellulose nanocrystals in polar organic solvents. Cellulose 2007, 14 (2), 109-113.

Page 113: Experimental and modeling investigation of cellulose

91

CHAPTER 5. MODELING OF DRY SPUN POLYMER - CELLULOSE NANOCRYSTAL COMPOSITE FIBERS

The focus of the previous chapters has been on experimental work investigating the

effects of CNCs on the polymer-CNC composite fibers and their mechanical properties.

In this chapter, a more theoretical approach is used. A 1D model is presented using

fundamentals of transport phenomena, describing the mass, momentum and energy

transfer, combined with a viscous constitutive equation accounting for an oriented fiber

assembly.

5.1 Introduction

Over the past two decades, cellulose nanocrystals (CNCs) have been the subject of

significant research interest as a new class of low cost, renewable and biodegradable

materials with high mechanical properties and customizable surfaces. Because of their

unique combinations of properties of high strength, high aspect ratio, availability, and

nano-scale dimensions, there have been many studies investigating CNCs as the

reinforcement phase in a composite systems.1 Much of the research on the reinforcement

potential of CNCs within a polymer matrix has been focused on continuous polymer fiber

fabrication.2-18 However, to our knowledge, all of the reported literature to date are

experimental studies developing composite fibers via methods of electrospinning, dry

spinning or wet spinning. Developing a mathematical mode describing the effects of

CNCs on the spinning process is key to further understanding on the topic.

Page 114: Experimental and modeling investigation of cellulose

92

Dry spinning is an industrial continuous fiber manufacturing process. The

polymer is dissolved in a volatile solvent to form a “spin dope” then extruded through a

small orifice known as a spinneret into a drying column. A spinneret may contain one or

many openings. Typically heated air is supplied to aid the process of solvent evaporation

and fiber formation. The fiber is stretched while drying, then collected on a take-up

wheel. The stretching of the fibers promotes the alignment of the polymer chain along the

fiber direction, thus increasing the over strength of the fiber. Additional processing such

as drawing, washing, and finish applications are also often used.19 This method of

production is widely used to manufacture cellulose acetate, cellulose triacetate, acrylic,

spandex, vinyon and others.20 The addition of CNCs to a dry spinning system is

particularly interesting due to the shear-induced orientation of the CNCs by the dry

spinning process. This will have a major effect on the overall rheological behavior of the

solution, thus must be adequately described in the modeling.

In a typical dry spinning system, fundamentals of transport phenomena are used

to describe the mass, momentum and energy transfer, combined with constitutive

equations. There are large amount of parameters and variables available in such a model,

for example, the solvent/polymer concentration, temperature, axial velocity, viscosity,

diameter, and solidification point. Additional effects such as skinning and die swell are

often assumed to be negligible in the overall process.19

Previously reported modeling studies on pure polymer dry spinning systems

focused on prediction of the aforementioned parameters using various viscous and

viscoelastic constitutive equations in one-dimensional and two-dimensional models. The

earliest modeling work done by Fok and Griskey simulated the heat and mass transfer of

Page 115: Experimental and modeling investigation of cellulose

93

the system using partial differential equations solved with experimental values of radially

averaged concentration.21-22 Ohzawa et al. used the assumption of constant tension along

the filament to solve a more complex numerical one-dimensional model for multiple

polymer/solvent systems.23-24 Brazinsky et al. solved a two-dimensional mathematical

model with variable tension along the spinline and was able to show favorable

comparisons to experimental data of a cellulose acetate/acetone dry spinning system.25

Sano also investigated a cellulose acetate/acetone dry spinning system, solving equations

of momentum, heat and mass transfer simultaneously.26 In particular, the elongational

viscosity prediction was improved on by estimations from the shear viscosity data using

Krevelen theory.

The studies discussed above considered the spinning solution to be viscous

Newtonian fluids. Furthermore, these models do not take into consideration effects of

solidification on the system. More recently, a few groups published results adding in the

viscoelastic and rheological behaviors.27-29 Gou and McHugh presented a comparison of

1D Newtonian and viscoelastic models for a cellulose acetate/acetone system.27 A

modified Giesekus model was used along with a solidification mechanism. The two

different models give similar results with the exception of the Giesekus model predicting

the system to reach velocity and diameter plateaus much faster than that of the Newtonian

model. This is referred to as a “locking in” of the system by the authors. The same

authors later expanded on this model to a two-dimensional ternary system, describing a

cellulose acetate/acetone/water system.28 Deng et al. applied a similar 1D model to that of

Gou and McHugh to a polyimide fiber system.29 The White-Metzer viscoelastic

Page 116: Experimental and modeling investigation of cellulose

94

constitutive equations were used, along with considerations for a thermal cyclization

reaction, specific to polyimide dry spun fibers.

To our knowledge, there has been no published work on dry spun composite

systems. The addition of solid ellipsoidal particles, in this case CNCs, will cause

irreversible changes to the structural and rheological characteristics of the material. Due

to the anisotropic nature of the CNC particles, the orientation distribution of the particles

within the fiber will also have an effect. These changes in turn will affect the

aforementioned parameter profiles such as diameter, velocity, temperature, and overall

mass concentration.

In order to describe the flow characteristics of polymer-CNC composite, an

effective constitutive relation must be used. Batchelor gives a relation for prediction of

the effective elongational viscosities for suspensions of collimated fibers at dilute and

semi-dilute concentrations in a Newtonian fluid.30 The model assumes the shearing flow

created by the fillers moving past each other to be constant. However, at high

concentrations of filler, this assumption is invalid. In addition, it does not take maximum

possible packing fraction into consideration. Pipes developed a similar relation specific

for concentrated and highly systems derived from variable shearing stress where the

maximum packing fraction issue is also remedied.31 Though, since the model is used for

higher concentrations where the particle contribution dominates viscosity, the polymer

contribution is neglected.

The purpose of this chapter is to present a generic one-dimensional model of a dry

spun polymer system with immersed ellipsoidal particles at different concentrations, as

well as a mechanism for accounting for a system with variable orientation distribution of

Page 117: Experimental and modeling investigation of cellulose

95

the particles. A modified Pipes model is developed to take polymer contribution into

consideration as well. The model would provide insight for process optimization of dry

spinning systems involving generic ellipsoidal reinforcements.

5.2 Model Development

5.2.1 Model Assumptions

A schematic of the dry spinning process is shown in Figure 5.1. The model

developed here is a ternary system consisting of cellulose acetate, acetone, and cellulose

nanocrystals, referred to by the subscripts p, s and f respectively. Perfect volume

additivity is assumed and no specific interactions between the components unless

otherwise noted.

For simplicity purposes, a single orifice approximation is used. The system is

assumed to be at steady state. The dope is pushed through the orifice with diameter, do

into the heated column at a mass flow rate W, and temperature To. The filament is

assumed to be axisymmetric, and the cross-section circular. The traveling direction of the

filament is set as the z-axis where z=0 at the opening of the spinneret. The fiber can be

set to free fall with gravity or drawn at a speed of VL. As a one-dimensional model, any

radial variation of the system is assumed to be negligible such as temperature, density

and velocity profiles. Surface tension, die swell and crystallization is presumed negligible

for the system. The heated air at temperature Ta travels counter parallel to the fiber line at

velocity va.

For this one-dimensional model, fundamental equations of momentum, mass and

energy balances, as well as rheology and thermodynamics of the polymer composite

Page 118: Experimental and modeling investigation of cellulose
Page 119: Experimental and modeling investigation of cellulose

97

system are considered to determine the following characteristics: average axial velocity,

vz(z), spinline radius R(z), average temperature of the filament, T(z), and average

component concentration, cp(z), cs(z), and cf(z).

5.2.2 Transport Equations

The three fundamental transport phenomena equations; equations of mass, momentum

and energy for dry spinning are described below and solved simultaneously. The related

material property correlations are listed in Section 5.2.4. The subscripts p, s and f refer to

polymer, solvent, and filler respectively. In this case, the material properties for those are

cellulose acetate, acetone, and cellulose nanocrystals respectively. The detailed equation

derivations are included in the Appendix.

5.2.2.1 Equation of Continuity

It is assumed that polymer and filler do not transfer to the air during the dry spinning

process, thus the overall conservation of mass is applied to the mass of the polymer and

filler:

0pdW

dz (5.1)

0fdW

dz (5.2)

The removal of solvent from the dry spinning line is driven by the mass transfer from the

spinning line surface to the surrounding gas medium.19 As the fiber travels through the

drying column, the solvent evaporates from the gas film side at the boundary layer. The

rate of evaporation is expressed in terms of a mass transfer coefficient related to the

Nusselt number. Ideal solution behavior on the filament side and ideal gas behavior on

Page 120: Experimental and modeling investigation of cellulose

98

the air side are assumed. The overall system is at steady state with gas-liquid equilibrium

at the air-filament interface. The following equation for axial gradient of the solvent

concentration in the filament is obtained by integrating the continuity equation across the

circular fiber cross section.19, 27

,

ˆ2 1

v

sats y s ss s s

sz

k M cdc a Py

dz R P (5.3)

Where cs is the volume fraction of solvent, Ms the molecular weight of the solvent, ˆs the

partial specific volume of the solvent, and ky the mass transfer coefficient on the surface

side. vz refer to the axial velocity of the filament. satsP and P are the saturated partial

pressure of the solvent and the total pressure respectively. as is the activity coefficient of

the solvent evaluated by the Flory-Huggins theory. ,sy is the mole fraction of the solvent

The spinline radius, R is defined as

1 2

ˆ

vp p

p z

WR

c (5.4)

5.2.2.2 Equation of Momentum

The following equation is developed assuming the flow undergoes uniaxial extensional

flow.27 The axial velocity is assumed to be uniform across the cross section. As usual

practice,19 the force balance equation used for dry spinning is similar to that of wet

spinning with an additional term accounting for the air drag forces. Surface tension and

inertia due to solvent evaporation is assumed negligible.

Page 121: Experimental and modeling investigation of cellulose

99

2v

v zz zz rr f a z aA A

z z (5.5)

A is the cross

sectional area of the filament, calculated as 2. zz rr is the stress tensor terms

associated with steady-state elongational flow. Cf refers to the coefficient of friction due

a and va are the density and the velocity of the air traveling counter current to

that of the filament respectively. g is the acceleration of gravity. The term on the left

hand of the equation represents the inertia of the filament while the right hand terms refer

to the tension experienced by the filament, the air drag forces, and the gravitational forces

respectively.

5.2.2.3 Equation of Energy

The energy equation involves contributions from the latent heat of solvent vaporization,

and the heat capacity of the filament. As previously stated, any radial variations of the

temperature is neglected here. For simplicity, crystallization is also assumed to be

minimal in this system.

,

2v

sats

p z a v y s s

Pd T

d z R P (5.6)

Here Cp refer to the heat capacity of the filament, calculated by ideal mixing of

contributions from polymer, solvent, and filler. T is the temperature of the filament, h is

the convective heat transfer coefficient and Ta, the temperature of the counter current air.

v is the molar latent heat of vaporization. The rate of change in internal energy of the

Page 122: Experimental and modeling investigation of cellulose

100

system is driven by convective heat transfer between the filament and the air, as well as

the evaporation of the solvent.

5.2.3 Constitutive Model

The behavior of the spinning dope as it undergoes uniaxial elongation and

transition from fluid of solid fiber is a complicated process involving change in

temperature, composition, molecular structure, rheology, and a wide range of variables.

The majority of previous dry spinning models were based on constitutive equations of a

Newtonian, incompressible fluid.21-25 These produced valid results when compared to

experimental data for many linear bulk polymers.19 A few studies reported results from

Giesekus or White-Metzer viscoelastic constitutive models.27-29 These models included

the additional elastic properties exhibited by the spinning solutions. They produced

similar results to Netwonian models with the largest deviations observed at early stages

of spinning near the exit of the spinneret. Due to the additional complication of solid

fillers in the current system, a modified Newontian constitutive model is used.

For a Newtonian, incompressible fluids with a temperature and concentration

dependent viscosity undergoing uniaxial extension, the following velocity gradient

applies:

0 0

10 0

21

0 02

v

v

v

z

z

z

d

dzd

dzd

dz

v (5.7)

Page 123: Experimental and modeling investigation of cellulose

101

The stress tensor has the form:

0

0 0

zz

rr (5.8)

where rr and are equal.

It is known that the relation32:

11

vzzz rr

d

dz (5.9)

where 11 , the axial elongational viscosity is related to the zero shear viscosity by:

11

0

3( , )sT c

(5.10)

However, this ratio of 3 between the elongational viscosity and no shear viscosity

no longer holds true when fillers are added to the system. The fillers are assumed to be

long discontinuous rigid rods of aspect ratio of l/d, where l is the length and d is the

diameter of the fiber. The filler/polymer solution interface is assumed to be ideal, i.e., no

slip. Hexagonal packing is assumed in the transverse plane. As the filament is stretched in

elongational flow, the fillers move past each other, creating a shearing flow. This, in turn,

amplifies the overall elongational viscosity.

Pipes developed an expression for elongational viscosity for an ideal system

where the fillers are all aligned perfectly along the fiber direction.31 The variable apparent

shearing strain rate generated by the fillers is a function of aspect ratio, filler

concentration, and extensional strain rate. The resulting ratio between elongational

viscosity and zero shear viscosity is as follows:

Page 124: Experimental and modeling investigation of cellulose

102

2

11

0

(L/ D)

1ln

1

fc (5.11)

Where cf is the filler volume fraction, L and D refer to the length and the diameter of the

filler respectively. is defined as

1

1 fcF

(5.12)

Where F is 3/ 2 for hexagonal packing array and / 4 for square array.

However, Equation 5.11 only accounts for contributions to the elongational

viscosity from the filler, and not the polymer solution. While the polymer contribution is

minimal in comparison at high concentrations of filler, it is dominating at dilute

concentrations. Thus for the purpose of this model, a modified version of the Pipes’

equation is presented:

2

11

0

(L/ D)3*(1 )

1ln

1

ff

cc (5.13)

An additional consideration for the system is the effect of filler orientation on the

apparent elongational viscosity. So far, the model predicts the apparently elongational

viscosity for perfectly collimated filler systems. However, in reality, the fillers are often

at varying degrees of off axis to the fiber axis due to processing conditions. Due to the

highly anisotropic nature of filler composite systems, a small misalignment of the fillers

may result in a large deviation in the effective viscosity. Using classical micromechanical

analysis, the apparent axial elongational viscosity for an individual off-axis fiber can be

calculated as follows:31

Page 125: Experimental and modeling investigation of cellulose

103

1* 4 2 2 4

11 11 11 12 11 22/ / 1 /m m n n (5.14)

where cosm and sinn . 22 is the transverse elongational viscosity, and 12 is the

in-plane shearing viscosity. These terms are given by:

22 04 (5.15)

12 0 (5.16)

For a fiber with fillers of a normalized orientation distribution function, ,

representing an ensemble of fillers with individual orientations, a weighted average of the

axial elongational viscosity is calculated as shown below with considerations for

axisymmetry:

1/2'11 11

0

d (5.17)

Where

4 2 2 411 11 12 11 22/ 1/ 1/ /m m n n (5.18)

An accurate prediction of the zero shear viscosity of polymer system is key to the

success of the relation above. The viscosity of the spinning dope will be a function of the

concentration of the solvent in the system, as well as the temperature of the solution. The

largest increase in the viscosity will be observed as the majority of the solvent leaves the

filament, and the fiber begins to solidify. The specific calculations of the viscosity will be

discussed in the next section.

Page 126: Experimental and modeling investigation of cellulose

104

5.2.4 Material Properties and Correlations

As previously discussed, the polymer/solvent/filler system used for this model is

cellulose acetate/acetone/cellulose nanocrystals. Cellulose acetate/acetone is a well-

known dry spinning system with a large amount of existing studies with material property

data, as well as theoretical predictions. The correlations used in this study are listed

below.

5.2.4.1 Density

Density of the overall filament is assumed to be radially constant and a weight

average of the 3 components according to ideal mixing. Perfect volume additivity is

assumed and no specific interactions between the components.

p p s s f fc c c (5.19)

5.2.4.2 Heat and Mass Transfer Coefficients

The following correlations are used to estimate the heat transfer coefficient of the

filament.26

1 20.36 20.35 0.146 Re (1.03Re 0.685u )N P w (5.20)

Where Nu stands for the dimensionless Nusselt number, and Rep and Rew stands for

Reynolds numbers based on air velocity, and filament velocity respectively. They are

calculated as:

Nu2

a

Rh

k (5.21)

P

v2Re a a

a

R (5.22)

Page 127: Experimental and modeling investigation of cellulose

105

v2

Re z aw

a

R (5.23)

Where h is the heat transfer coefficient, va and vz refer to the velocity of the airflow and

the filament. ak , a and a refer to the thermal conductivity, the density and the viscosity

of the air. These are estimated by the following:27

7 0.8664.49 10a fk T (5.24)

0.351

afT

(5.25)

5 1.51.446 10

113.9f

a

f

T

T (5.26)

Where Tf refers to the average value between the filament temperature and the air

temperature in Kelvin.

The surface mass transfer coefficient, ky, is estimated using an experimentally

determined h/ky value of 11.6 cal/(gmol K) by Ohzawa et al.23

5.2.4.3 Activity Coefficient

The activity coefficient of the solvent, acetone, is estimated using Flory-Huggins

theory:27

exp 1 psss s p p ps s p

p p

dgVa c c c g c c

V dc (5.27)

Where Vs and Vp are the molar volume of acetone and cellulose acetate respectively. gps

refer to the cellulose acetate-acetone interaction parameter estimated by Yilmaz and

McHugh.33

Page 128: Experimental and modeling investigation of cellulose

106

0.535 0.11ps pg c (5.28)

For the purpose of this study, we assumed there are no interactions between acetone and

cellulose nanocrystals.

5.2.4.4 Vapor Pressure

The Antoine equation is used to describe the vapor pressure of acetone as a

function of temperature.34

1 2756.22

exp 14.314atm 5101.325 228 60C .0

satsP

T (5.29)

5.2.4.5 Air Drag Coefficient

The air drag coefficient is calculated as a function of the Reynolds number:

0.610.77 RefC (5.30)

v

Reva z a

a

2R (5.31)

Where a and a are the density and viscosity of the air, defined in a previous section. va

and vz refer to the velocity of the airflow and the filament.

5.2.4.6 Heat Capacity and Latent Heat of Evaporation

The heat capacity of the overall filament is assumed to be radially constant and a

weight average of the 3 components according to rule of mixtures.

p pp p ps s pf fC C c C c C c (5.32)

Page 129: Experimental and modeling investigation of cellulose

107

The heat capacity of cellulose acetate is set to a constant value of 0.3155 cal/g/K

due to lack of temperature correlated data.35 Similarly, the heat capacity for cellulose

nanocrystals is set to 0.2940 cal/g/K.36

The following temperature dependent relation is used to the estimate the heat

capacity of acetone.37

1

1/3 11

2.56 0.436 1R

2.91 4.28 1 0.296 1

roPs Ps

r r r

TC C

T T T (5.33)

where oPsC , the ideal gas state heat capacity, is:

2 5 2 9 31.505 26.224 10 2.992 10 4.867 10oPsC T T T (5.34)

rT , the reduced temperature, is equal to / cT T . The value of cT , the critical temperature,

and , the acentric factor are listed in Table 5.1. R is the universal gas constant.

The latent heat of evaporation of acetone is calculated as follows:37

0.354 0.456

R 7.08 1 10.95 1v c r rH T T T (5.35)

where R is the universal gas constant.

5.2.4.7 Glass Transition Temperature

The glass transition temperature is estimated as a function of the Tg of the polymer and

solvent as well as their respective concentration and densities. The Gordon-Taylor

equation is applied:38

p gp s gsg

p s

w T Kw TT

w Kw (5.36)

Page 130: Experimental and modeling investigation of cellulose

108

Where gpT and gsT refer to the glass transition temperatures of cellulose acetate and

acetone respectively. iw refer to the weight fraction of the components, converted from

volume fractions by their respective densities. K typically refer to the ratio of the

difference in thermal expansivity between the liquid and glass state between the

components. However, since these values are not known in this case, K is estimated using

the Simha-Boyer rule as ratio between the densities and the glass transition

temperatures.38

p gp

s gs

TK

T (5.37)

5.2.4.8 Solution Viscosity

The zero shear viscosity of the spinning dope is calculated as function of temperature,

concentration, degree of polymerization and glass transition temperature.39

For sT T , where the standard temperature 1.2s gT T :

712 3.76.6 10 1 exp Ro n sP w E T (5.38)

Where Pn is the degree of polymerization, E is the activation energy, R the

universal gas constant. These values are listed in Table 5.1

For s gT T T a WLF (Williams-Landel-Ferry) relation is used, 39

0 0

8.86log log

101.6s

ss

T TT T

T T (5.39)

For gT T , the above equation is used with gT T , with the assumption that the

solution has solidified.

Page 131: Experimental and modeling investigation of cellulose

109

5.2.5 Experimental Materials and Methods

For comparison to the model, cellulose acetate/cellulose nanocrystals composite

fibers were produced in a laboratory-scale fiber spinner.

5.2.5.1 Materials

Cellulose Acetate (CA) with an acetyl content of 39.7 wt% and average molecular

weight of 50,000 was purchased from Sigma-Aldrich Chemistry. Reagent grade acetone

(Sigma-Aldrich) was used without further purification as the solvent. Cellulose

nanocrystals (CNC) were produced and supplied by the USDA Forest Service-Forest

Products Laboratory, Madison, WI. The morphology of individual CNC particles was

characterized via transmission electron microscopy (TEM; Philips CM-100, at 100kV,

spot 3, 200 µm condenser aperture, and 50 µm objective aperture,) of CNC particles

deposited on TEM grids (400 mesh Formvar/carbon filmed grid prepared with blow

discharge) with 2% aqueous uranyl acetate stain. The average diameter (6 ± 2nm) and

length (83 ± 43nm) of 93 individual CNC crystals were determined using within Image J.

5.2.5.2 Preparation of CA-CNC Fibers

The prepared spinning dope was dry spun using a custom built plunger style fiber

spinner attached to an universal tensile testing machine (MTS Insight) with a single outer

orifice of 250 µm. Increasing amount of CNCs were incorporated into the CA fibers up to

49 wt% (final dry fiber concentration). Detailed description of preparation of the spinning

dope, spinning conditions and characterization is listed in Chen et al.2

Page 132: Experimental and modeling investigation of cellulose

110

5.2.6 Spinning Parameters

Table 5.1 shows values of spinning conditions used in the simulation. Table 5.2

lists the material properties used as input parameters in the simulation.. These values are

chosen to be as close to the experimental values as possible. In the case that experimental

data are unavailable, previous literature studies were consulted.

5.2.7 Numerical Method

The model described above uses a series of highly coupled simultaneous

differential equations which describe the behavior of the spinning line. Equations of

mass, momentum, and energy transfer, combined with a viscous constitutive equation

with an option for an oriented fiber assembly is used to determine the following

independent variables:

Axial velocity vz z

Axial velocity gradient vzd

dz

Temperature on the filament cross-section T z

Solvent concentration sc z

These equations are solved simultaneously using a fourth order Runga-Kutta

algorithm method. An initial value of 0 is assumed for 0

vz

z

d

dz for fibers without

drawing. For fibers with drawing, the initial value for vzd

dzis determined by iteration until

the final filament velocity matches the desired wind up velocity with a 0.1% tolerance

Page 133: Experimental and modeling investigation of cellulose

111

Table 5-1 Processing conditions for CNC-CA composite fibers

Processing Conditions Symbol Values

Initial Filament Temperature 0T 329 K

Initial mass fraction of solvent 0sc 0.80

Mass flow rate W 1.0 g/min

Diameter of Spinneret d 0.025 cm

Air velocity va 50 cm/s

Air Temperature aT 353 K

Solvent mole fraction in air ,sy 0

Ambient Pressure P 1 atm

Page 134: Experimental and modeling investigation of cellulose

112

Table 5-2 Material Properties of CA-CNC Composite Fibers

Processing Conditions Symbol Values

Density of CA p 1.310 g/cm3 35

Density of Acetone s 0.790 g/cm3 35

Density of CNC f 1.590 g/cm3 1

Molar Volume of CA pV 31147 cm3/gmol 35

Molar Volume of Acetone sV 73.53 cm3/gmol 35

Molecular Weight of Acetone sM 58.08 g/gmol

Heat Capacity of CA ppC 0.3155 cal/g/K 35

Heat Capacity of CNC pfC 0.2940 cal/g/K 36

Degree of polymerization of CA nP 200 27

Critical Temperature of Acetone cT 508.1 K 37

Glass Transition Temperature of CA gpT 468 K 27

Glass Transition Temperature of Acetone gsT 44 K 27

Acentric factor of Acetone 0.30437

Page 135: Experimental and modeling investigation of cellulose

113

error.26 To study the effect of filler orientation, the experimentally determined orientation

distribution function is used to calculate the effective elongational viscosity.

5.3 Results and Discussion

5.3.1 Newtonian Model Prediction

Figure 5.2-5.5 shows the results of the simulation with the inputs listed in Table

5.1 with a Newtonian model with no filler content. Two cases are considered, one where

the fibers are allowed to free fall without drawing, and one is drawn to a take up velocity

of 300 cm/s.

Figure 5.2 shows the diameter profile as a function of distance from the spinneret.

As the filament exits the spinneret, the diameter decreases precipitously in the case of the

fiber with drawing, whereas the free falling fiber has a much shallower profile. Once all

of the solvent leaves the filament, and the fiber solidifies, the diameter remains a constant

value until the take-up wheel. As one would expect, the drawing case resulted in a much

thinner fiber.

Figure 5.3 shows the velocity profile of the filaments. Without drawing, the

velocity of the filament remains low until it reaches a plateau, driven by the inertia of the

filament, air drag, and gravity. With drawing, the velocity rises rapidly after extrusion

until a higher plateau value. The significance of this distance at which both cases reaches

plateau can be explained by Figure 5.4 and 5.5, showing the temperature and

concentration profiles. One can see that this coincides with when the majority of the

solvent has left the system, and when the overall temperature of the system starts to rise.

Figure 5.5 shows the temperature of the filament drops rapidly as the endothermic

evaporation happens immediately after the spinneret exit. The temperature remains

Page 136: Experimental and modeling investigation of cellulose

114

Figure 5-2 Diameter profile as a function of distance from the spinneret for two cases: fibers with and without drawing

Page 137: Experimental and modeling investigation of cellulose

115

Figure 5-3Velocity profile as a function of distance from the spinneret for two cases: fibers with and without drawing

Page 138: Experimental and modeling investigation of cellulose

116

Figure 5-4 Concentration fraction profile as a function of distance from the spinneret for two cases: fibers with and without drawing. ND and D stands for non-drawing and drawing respectively.

Page 139: Experimental and modeling investigation of cellulose

117

Figure 5-5 Temperature profile as a function of distance from the spinneret for two cases: fibers with and without drawing

Page 140: Experimental and modeling investigation of cellulose

118

Figure 5-6 Viscosity profile as a function of distance from the spinneret for two cases: fibers with and without drawing

Page 141: Experimental and modeling investigation of cellulose

119

Figure 5-7 Glass transition temperature profile as a function of distance from the spinneret for two cases: fibers with and without drawing

Page 142: Experimental and modeling investigation of cellulose

120

low for a period as the acetone continually evaporates. As the solidification process

comes to an end, the temperature of the filament starts to increase, until it reaches the

temperature of the surrounding air. Similarly, Figure 5.4 shows that the majority of the

solvent has left the system around 140 cm from the spinneret. Because of the higher

velocity, the drawn fibers also maintains a slightly higher concentration of solvent, and

lower temperature compared to the non-drawn fibers at the same distance away from the

spinneret.

Figure 5.6 and Figure 5.7 also shows the zero shear viscosity profile and the glass

transition of the system. The viscosity slowly increases initially as solvent leaves the

system, at the same time, the glass transition temperature of the system increases as well.

When the glass transition begins, i.e., the temperature of the filament approaches the

glass temperature of the filament, the viscosity increases rapidly in response to the

solidification of the filament. Finally, when the solidification is compete, and all of the

solvent has left the system, the viscosity becomes constant as well as the glass transition

temperature.

5.3.2 Effect of Spinning Parameters

There are a large amount of parameters that can be changed to result in a change in the

behavior of the overall system, such as the temperature of the spinning dope and air, mass

flow rate, winding speed and the air velocity. The effect of these cases are discussed

below. Each of the parameters are changed individually as the rest of the parameters

remained the same as stated previously. The case of drawn fibers are investigated.

Page 143: Experimental and modeling investigation of cellulose

121

Figure 5-8 Diameter profile as a function of distance from the spinneret for cases with varying mass flow rate

Page 144: Experimental and modeling investigation of cellulose

122

Figure 5-9 Concentration profile as a function of distance from the spinneret for cases with varying mass flow rate

Page 145: Experimental and modeling investigation of cellulose

123

5.3.2.1 Mass Flow Rate

Because of conservation of mass, an increase in mass flow rate translates directly to an

increase in initial velocity. The effects of this can be seen below in Figure 5.8-5.9. As one

would expected, the diameter of the final solidified fiber increases due to the increase of

mass. The lack of plateau for W= 1 and 1.5 g/min cases indicate the fibers are not

completely dry at the end of 200 cm. The increase in velocity also shifts the solidification

point further downstream. In turn, the temperature profile is also shifted (not pictured).

The velocity profile is less affected due to the drawing process. The final velocity of the

fibers remain the same, however, the initial velocity is increased.

5.3.2.2 Temperature of Spinning Dope and Air

Due to the boiling point of acetone, the temperature of the spinning dope is

limited to 329K or below. A few of these cases are investigated below with the results

shown in Figure 5.10 and 5.11. It can be seen that the temperature of the spinning dope

does not have a large effect on the overall parameter profiles due to the limited amount of

mass contained in the filament. The diameter profile and the solvent concentration profile

are both slightly increased. The velocity profile shows minimal changes, and the

temperature profile shows the solidification point is slightly further downstream (not

pictured).

The effect of changing quenched air temperature is shown in Figure 5.12-5.14. As

one can expect, the surrounding air temperature has a much bigger effect on the overall

profiles than the dope temperature. As expected, the higher the air temperature, the faster

the solidification process is completed, although the overall shape is not changed. This is

Page 146: Experimental and modeling investigation of cellulose

124

Figure 5-10 Diameter profile as a function of distance from the spinneret for cases with varying spinning dope temperature

Page 147: Experimental and modeling investigation of cellulose

125

Figure 5-11 Concentration profile as a function of distance from the spinneret for cases with varying spinning dope temperature

Page 148: Experimental and modeling investigation of cellulose

126

Figure 5-12 Diameter profile as a function of distance from the spinneret for cases with varying quenched air temperature

Page 149: Experimental and modeling investigation of cellulose

127

Figure 5-13 Velocity profile as a function of distance from the spinneret for cases with varying quenched air temperature

Page 150: Experimental and modeling investigation of cellulose

128

Figure 5-14 Temperature profile as a function of distance from the spinneret for cases with varying quenched air temperature

Page 151: Experimental and modeling investigation of cellulose

129

Figure 5-15 Diameter profile as a function of distance from the spinneret for cases with varying take-up velocity.

Page 152: Experimental and modeling investigation of cellulose

130

Figure 5-16 Velocity profile as a function of distance from the spinneret for cases with varying take-up velocity.

Page 153: Experimental and modeling investigation of cellulose

131

displayed in the diameter profile in Figure 5.12, and concentration profile (not shown).

The temperature profile shows as the air temperature is increased, the overall temperature

of the filament is also increased. The distance required to evaporate all of the solvent is

shortened, and the eventual plateau is raised to the increased ambient temperature. The

velocity profile remain similar, however, the higher ambient temperature causes the

filament to reach plateau sooner (Figure 5.13).

5.3.2.3 Winding Speed and Air Velocity

The effect of winding speed is shown in Figure 5.15-5.16. Three cases are

investgiated at windup speeds of 200, 300, and 400 cm/s. One can see that a lower

winding speed result in a bigger fiber (Figure 5.15). Figure 5.16 shows the velocity

profile where the general shape is maintained while the filaments reaches the takeup

velocity pleateau at around 50cm. The effect of winding speed is minimal on both the

temperature profile and the concentration profile (Not shown).

The effect of air velocity is shown in Figure 5.17-5.19. In general, increasing air

velocity will accelerate the solidification process. This can be explained by the increase

in heat and mass transfer coefficient, as well as the increase in air drag coefficient.

Changing the air velocity has little to no effect on the velocity profile. The temperature

profile is the most affected by the change, due to its high dependence on the heat transfer

coefficient.

Page 154: Experimental and modeling investigation of cellulose

132

Figure 5-17 Diameter profile as a function of distance from the spinneret for cases with varying quenched air velocity.

Page 155: Experimental and modeling investigation of cellulose

133

Figure 5-18 Velocity profile as a function of distance from the spinneret for cases with varying quenched air velocity.

Page 156: Experimental and modeling investigation of cellulose

134

Figure 5-19 Temperature profile as a function of distance from the spinneret for cases with varying quenched air velocity.

Page 157: Experimental and modeling investigation of cellulose

135

5.3.3 Effect of Fillers

In the following chapter, the effect of fillers (CNCs) have on the dry spinning

system is discussed. The case of collimated fillers, i.e., all fillers are aligned in the fiber

direction, is discussed first, then the case of fibers filler orientation ensemble is

investigated.

5.3.3.1 Effect of Collimated Fibers

The modified Pipes model is applied to the dry spinning model in order to

describe the flow characteristics of the filler-reinforced polymer fiber. The fillers are

assumed to be long discontinuous rigid rods of aspect ratio of l/d, where l is the length

and d is the diameter of the fiber. The filler/polymer solution interface is assumed to be

ideal, i.e., no slip. Hexagonal packing is assumed in the transverse plane. As the filament

is stretched in elongational flow, the fillers move past each other, creating a shearing

flow. This, in turn, amplifies the overall elongational viscosity.

There are four regimes in composite systems, dilute, semi-dilute, concentrated,

and hyperconcentrated.31 For highly concentrated systems, it can be assumed that the

fibers will remain perfectly collimated during deformation. For fillers of aspect ratio of

15, volume fraction of filler above 10% are considered concentrated, and above 20% are

considered hyperconcentrated.

The results of the simulation are shown in Figure. 6 cases were considered,

0 wt%, 10 wt%, 20 wt%, 29 wt%, 39 wt% and 49 wt%. These translates to volume

fractions of 0%, 8.4%, 17.1%, 25.2%, 34.5%, and 44.2%. In order to compare these

results to experimental data, these fibers are free falling and not drawn. In addition, the

Page 158: Experimental and modeling investigation of cellulose

136

Figure 5-20 Diameter profile as a function of distance from the spinneret for cases with varying final CNC volume concentration.

Page 159: Experimental and modeling investigation of cellulose

137

Figure 5-21 Velocity profile as a function of distance from the spinneret for cases with varying final CNC volume concentration.

Page 160: Experimental and modeling investigation of cellulose
Page 161: Experimental and modeling investigation of cellulose

139

Figure 5-23 Temperature profile as a function of distance from the spinneret for cases with varying final CNC volume concentration.

Page 162: Experimental and modeling investigation of cellulose

140

Figure 5-24 Glass transition temperature profile as a function of distance from the spinneret for cases with varying final CNC volume concentration.

Page 163: Experimental and modeling investigation of cellulose

141

amount of acetone in the system is kept constant, and the amount of CNCs added to the

system replaces equal amount of cellulose acetate.

Figure 5.20 shows the diameter profile of the fibers with varying amount of final

CNC concentration. As the filler concentration increases, the cumulative effect on

elongational viscosity is amplified. This in turn increases the forces of tension in the

filament, thus decreasing the velocity (Figure 5.21). Since diameter is inversely related to

the velocity, this sequentially results in a thicker fiber. However, the viscosity effect

eventually reaches a diminishing limit, and additional fillers only increase the diameter

by a marginal value.

The velocity profile is shown in Figure 5.21. Inversely to the diameter profile,

increasing filler content decreases the overall velocity profile by increasing the tension.

The value of the velocity plateau decreases as well. A similar limiting effect is observed

as in the diameter case where additional fillers past 25.2% only decreases the velocity by

a marginal amount. The velocity profile is decreased significantly when the filler

concentration goes from 0 to 8.4%. Each subsequent increase results in a smaller

magnitude of reduction, approaching a limit where viscosity approaches infinity and the

filament no longer deform. In this limiting case, the velocity would be constant.

Figure 5.22(a)-(f) shows the concentration profile for each of the CNC cases and

Figure 5.23 the temperature profile. Due to similarities in values between the heat

capacities of cellulose acetate and cellulose nanocrystals, there are no large changes to

the concentration profile and the temperature profile. Increasing filler content slightly

Page 164: Experimental and modeling investigation of cellulose

142

shifts the solidification point further downstream, as well as raising the minimum

temperature of the system to some extent.

The glass transition temperature profile is shifted slightly downstream by the

addition of CNCs (Figure 5.24). The overall acetone amount remains constant while the

amount of cellulose acetate is decreased. Since the fillers do not possess a glass transition

temperature, the overall Tg transition is delayed.

5.3.3.2 Effect of fiber orientation distribution

Due to the highly anisotropic nature of the filler reinforced fiber, if the fillers’

orientation deviate from perfect alignment along the fiber axis, there will be a

considerable change to the flow behavior of the filament as well as the effective material

properties. In experimental settings, due to processing, the fillers will have a variety of

misorientations within the system. The misorientations are more likely to happen in dilute

and semi-dilute systems as there are more free space for the fillers to move. In

concentrated and hyper-concentrated systems, the fillers are naturally arranged long side

to long side for maximum packing. Furthermore, typically the shorter aspect ratios fillers

are more likely to form an off axis angle misorientation. In the case of cellulose

acetate/cellulose nanocrystals fibers, the overall orientation distribution may be

quantified using 2D x-ray diffraction. The details of the procedure are listed elsewhere.2

The normalized orientation distribution for fibers with 8.4%, 17.1%, 25.2% and

44.2% CNC are shown below in Figure 5.25 for / 2 / 2 . Because of

axisymmetry about 0 , filler at off axis angle are effectively the same off axis

angle as . As expected, the most oriented fiber is the fiber with highest amount of filler,

Page 165: Experimental and modeling investigation of cellulose

143

Figure 5-25 Normalized probability distribution for filler orientation for 4 different experimental composite fibers: 44.2%, 25.2%, 17.1% and 8.4% CNC respectively.

Page 166: Experimental and modeling investigation of cellulose

144

Figure 5-26 Effect of fiber orientation on elongational viscosity

Page 167: Experimental and modeling investigation of cellulose

145

44.2% CNC. The amplitude of the orientation parameter decrease with diminishing

amount of CNCs in the system. At the same time, the variance of the distribution function

increases as well. For the case of 8.4% CNC, the orientation distribution shows there are

large amount of fillers oriented at various off-axis angles to the fiber axis. This is far

from the assumption that all fibers are perfectly collimated, thus this effect must be

accounted for in the simulation.

Elongation at failure is shown to be highly dependent on the filler orientation

(Figure 5.26). It is also apparent that the filler concentration amplifies this effect. For a

system with 44.2% CNC, the ratio between effective elongational viscosity and

collimated elongational viscosity is 0.6 at 15 .

The above orientation distribution functions are used to calculate the effective

viscosity for each fiber system. The effective viscosity at any angle, , is calculated

using micromechanics, then a weighted average of these viscosities are represented as the

overall effective viscosity. The results of the simulations are shown in Figure 5.27-5.28.

The diameter profile shows that when the fibers are misoriented, that decreases the fiber

tension, thus results in a thinner fiber. Interestingly, there are three competing forces; one

where the misorientation effect on viscosity increases with concentration, and two,

degree of misorientation is lower at higher concentrations, and three, the diminishing

effect on diameter thickening as viscosity increases. Therefore, marginal changes are

observed both low concentration (8.4%) and high concentration (44.2%) of filler. The

larger deviations on the system are instead observed in the intermediate concentrations

(17.1%, 25.2%).

Page 168: Experimental and modeling investigation of cellulose
Page 169: Experimental and modeling investigation of cellulose
Page 170: Experimental and modeling investigation of cellulose

148

Figure 5-29 Diameter Comparison between theoretical and experimental data

Page 171: Experimental and modeling investigation of cellulose

149

A similar phenomena is observed in the velocity profile (Figure 5.28). At high

concentration (44.2%), despite the decrease of viscosity, the tension on the fiber is

sufficiently high to maintain a low velocity. For lower concentrations, the decrease in

viscosity slightly increases the velocity profile. The change in effective elongational

viscosity in the system does not have a large effect on the concentration nor the

temperature profile (not pictured).

5.3.3.3 Comparison to experimental data

The theoretically predicted diameter is compared to experimentally determined results

from CA/CNC fibers in Figure 5.29. The theoretical data consistently under predicts the

diameter of the fiber, however, the overall trend of increasing diameter with increasing

filler content is captured. This discrepancy may be explained by the existence of an

additional solvent in the experimental system. Due to the insolubility of CNCs in pure

acetone, another solvent, dimethylacetamide (DMAc), was used. DMAc is a solvent with

a significantly higher boiling point than acetone, which may explain the larger resulting

diameter. Accounting for a quaternary system should result in a more accurate model.

5.4 Conclusion

In this study, a one dimensional model was developed to describe a ternary dry

spinning system consisting of cellulose acetate, acetone and cellulose nanocrystals.

Fundamental equations of momentum, mass and energy, as well as rheology and

thermodynamics of the polymer composite was used along with a modified Pipes viscous

constitutive equation to predict the diameter, velocity, component concentration, and

temperature profile of the system. Sensitivity analysis was performed on various spinning

Page 172: Experimental and modeling investigation of cellulose

150

parameters such as mass flow rate, temperature of the spinning dope and air, the velocity

of take-up speed, and the velocity of the quenched air. In addition, the effect of

reinforcement fillers on the system was investigated. It is found that perfectly collimated

fillers increased the elongational viscosity of the system, thus resulting in a thicker fiber.

This effect is amplified with increased concentration of fillers. The more realistic case

where filler orientation consists of an ensemble of misorientation of varying degrees is

also investigated. It is found that fillers that are off-axis reduces the effective elongational

viscosity and should be considered during such simulations.

Page 173: Experimental and modeling investigation of cellulose

151

Bibliography

1. Moon, R. J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J., Cellulose nanomaterials review: structure, properties and nanocomposites. Chemical Society Reviews 2011, 40 (7), 3941-3994.

2. Chen, S.; Schueneman, G.; Pipes, R. B.; Youngblood, J.; Moon, R. J., Effects of Crystal Orientation on Cellulose Nanocrystals–Cellulose Acetate Nanocomposite Fibers Prepared by Dry Spinning. Biomacromolecules 2014, 15 (10), 3827-3835.

3. Andersson, R. L.; Salajkova, M.; Mallon, P. E.; Berglund, L. A.; Hedenqvist, M. S.; Olsson, R. T., Micromechanical Tensile Testing of Cellulose-Reinforced Electrospun Fibers Using a Template Transfer Method (TTM). J. Polym. Environ. 2012, 20 (4), 967-975.

4. Dong, H.; Strawhecker, K. E.; Snyder, J. F.; Orlicki, J. A.; Reiner, R. S.; Rudie, A. W., Cellulose nanocrystals as a reinforcing material for electrospun poly(methyl methacrylate) fibers: Formation, properties and nanomechanical characterization. Carbohyd Polym 2012, 87 (4), 2488-2495.

5. Herrera, N. V.; Mathew, A. P.; Wang, L. Y.; Oksman, K., Randomly oriented and aligned cellulose fibres reinforced with cellulose nanowhiskers, prepared by electrospinning. Plast Rubber Compos 2011, 40 (2), 57-64.

6. Liu, D. Y.; Yuan, X. W.; Bhattacharyya, D., The effects of cellulose nanowhiskers on electrospun poly (lactic acid) nanofibres. J Mater Sci 2012, 47 (7), 3159-3165.

7. Lu, P.; Hsieh, Y. L., Cellulose nanocrystal-filled poly(acrylic acid) nanocomposite fibrous membranes. Nanotechnology 2009, 20 (41), -.

8. Magalhaes, W. L. E.; Cao, X. D.; Lucia, L. A., Cellulose Nanocrystals/Cellulose Core-in-Shell Nanocomposite Assemblies. Langmuir 2009, 25 (22), 13250-13257.

9. Park, W. I.; Kang, M.; Kim, H. S.; Jin, H. J., Electrospinning of poly(ethylene oxide) with bacterial cellulose whiskers. Macromol Symp 2007, 249, 289-294.

10. Peresin, M. S.; Habibi, Y.; Zoppe, J. O.; Pawlak, J. J.; Rojas, O. J., Nanofiber Composites of Polyvinyl Alcohol and Cellulose Nanocrystals: Manufacture and Characterization. Biomacromolecules 2010, 11 (3), 674-681.

11. Rojas, O. J.; Montero, G. A.; Habibi, Y., Electrospun Nanocomposites from Polystyrene Loaded with Cellulose Nanowhiskers. J Appl Polym Sci 2009, 113 (2), 927-935.

12. Xiang, C. H.; Joo, Y. L.; Frey, M. W., Nanocomposite Fibers Electrospun from Poly(lactic acid)/Cellulose Nanocrystals. J Biobased Mater Bio 2009, 3 (2), 147-155.

Page 174: Experimental and modeling investigation of cellulose

152

13. Zoppe, J. O.; Peresin, M. S.; Habibi, Y.; Venditti, R. A.; Rojas, O. J., Reinforcing Poly(epsilon-caprolactone) Nanofibers with Cellulose Nanocrystals. Acs Appl Mater Inter 2009, 1 (9), 1996-2004.

14. Lee, J.; Deng, Y. L., Increased mechanical properties of aligned and isotropic electrospun PVA nanofiber webs by cellulose nanowhisker reinforcement. Macromolecular Research 2012, 20 (1), 76-83.

15. Urena-Benavides, E. E.; Brown, P. J.; Kitchens, C. L., Effect of Jet Stretch and Particle Load on Cellulose Nanocrystal-Alginate Nanocomposite Fibers. Langmuir 2010, 26 (17), 14263-14270.

16. Urena-Benavides, E. E.; Kitchens, C. L., Wide-Angle X-ray Diffraction of Cellulose Nanocrystal-Alginate Nanocomposite Fibers. Macromolecules 2011, 44 (9), 3478-3484.

17. Urena-Benavides, E. E.; Kitchens, C. L., Cellulose Nanocrystal Reinforced Alginate Fibers-Biomimicry Meets Polymer Processing. Molecular Crystals and Liquid Crystals 2012, 556, 275-287.

18. Liu, L.; Yao, J. M., Wet-spinning of reinforced artificial silk hybrid fibres by cellulose whiskers. Adv Mater Res-Switz 2011, 175-176, 272-275.

19. Ziabicki, A., Fundamentals of Fibre Formation: The Science of Fibre Spinning and Drawing. Wiley: 1976.

20. Lewin, M., Handbook of Fiber Chemistry, Third Edition. CRC Press: 2006.

21. Fok, S.; Griskey, R., Heat transfer during dry spinning of fibers. Appl. Sci. Res. 1966, 16 (1), 141-148.

22. Fok, S. Y.; Griskey, R. G., Mass transfer during dry spinning of fibers. J Appl Polym Sci 1967, 11 (12), 2417-2426.

23. Ohzawa, Y.; Nagano, Y., Studies on dry spinning. II. Numerical solutions for some polymer–solvent systems based on the assumption that drying is controlled by boundary-layer mass transfer. J Appl Polym Sci 1970, 14 (7), 1879-1899.

24. Ohzawa, Y.; Nagano, Y.; Matsuo, T., Studies on dry spinning. I. Fundamental equations. J Appl Polym Sci 1969, 13 (2), 257-283.

25. Brazinsky, I.; Williams, A. G.; LaNieve, H. L., The dry spinning process: Comparison of theory with experiment. Polymer Engineering & Science 1975, 15 (12), 834-841.

26. Sano, Y., Drying behavior of acetate filament in dry spinning. Dry. Technol. 2001, 19 (7), 1335-1359.

Page 175: Experimental and modeling investigation of cellulose

153

27. Gou, Z.; McHugh, A. J., A comparison of Newtonian and viscoelastic constitutive models for dry spinning of polymer fibers. J Appl Polym Sci 2003, 87 (13), 2136-2145.

28. Gou, Z.; McHugh, A. J., Two-dimensional modeling of dry spinning of polymer fibers. Journal of Non-Newtonian Fluid Mechanics 2004, 118 (2–3), 121-136.

29. Deng, G.; Xia, Q.; Xu, Y.; Zhang, Q., Simulation of dry-spinning process of polyimide fibers. J Appl Polym Sci 2009, 113 (5), 3059-3067.

30. Batchelor, G. K., The stress generated in a non-dilute suspension of elongated particles by pure straining motion. Journal of Fluid Mechanics 1971, 46 (04), 813-829.

31. Advani, S. G., Flow and Rheology in Polymer Composites Manufacturing. Elsevier: 1994.

32. Bird, R. B.; Stewart, W. E.; Lightfoot, E. N., Transport Phenomena. Wiley: 2007.

33. Yilmaz, L.; McHugh, A. J., Modelling of asymmetric membrane formation. I. Critique of evaporation models and development of a diffusion equation formalism for the quench period. J Membrane Sci 1986, 28 (3), 287-310.

34. Smith, J. M.; Van Ness, H.; Abbott, M. M., Introduction to Chemical Engineering Thermodynamics. McGraw-Hill Education: 2005.

35. Brandrup, J.; Immergut, E. H., Polymer handbook. Interscience Publishers: 1966.

36. Hatakeyama, T.; Nakamura, K.; Hatakeyama, H., Studies on heat capacity of cellulose and lignin by differential scanning calorimetry. Polymer 1982, 23 (12), 1801-1804.

37. Reid, R. C.; Prausnitz, J. M.; Poling, B. E., The Properties of Gases and Liquids. McGraw-Hill: 1987.

38. Gordon, M.; Taylor, J. S., Ideal copolymers and the second-order transitions of synthetic rubbers. i. non-crystalline copolymers. Journal of Applied Chemistry 1952, 2 (9), 493-500.

39. van Krevelen, D. W., Properties of Polymers. Elsevier Science: 2012.

Page 176: Experimental and modeling investigation of cellulose

154

CHAPTER 6. CONCLUSIONS AND FUTURE WORK

6.1 Electrospun CA-CNC Fibers

Cellulose nanocrystals reinforced cellulose acetate nanofibers were successfully

produced using electrospinning. Continuous CNC-CA fibers with consistent diameters

were produced between varying CNC concentrations of 0 and 7.5 wt%. The addition of

CNCs did not show a significant alternation to the thermal properties of the fibers. A

maximum of 15% and 280% increase was observed in the ultimate tensile strength and

Young’s modulus of the fibers respectively. However, the addition of the CNCs increased

brittleness of the fibers, as well as the defect driven failures, leading to an increase in

deviation of the fiber performance. Overall, the electrospinning method did not improve

the mechanical performance of the fibers to the theoretical potential of CNC, suggesting

further improvement may be made. A processing method involving high shear rates such

as dry spinning may improve the agglomeration issues experienced here and lead to

further orientation of CNCs.

Future work on this topic may include different polymer-CNC systems. One of

the key limitations of the system is the dispersion of CNCs in non-aqueous solvents. A

water soluble system would improve the dispersion of the CNCs within the spinning

solution, and may lead to a higher maximum CNC concentration fiber produced. In

addition, to the author’s knowledge, there has no research done to quantify neither the

Page 177: Experimental and modeling investigation of cellulose

155

behavior of the CNCs within the fibers nor the modeling of the electrospinning process

with the addition of CNCs. Orientation within the fibers has been qualitatively looked at

using methods such as TEM, however, no accurate method has been developed to

measure the global degree of alignment within the fibers.

Furthermore, literature review shows no consistent and reliable method for testing

single fiber tensile strength. In most cases, a fiber mat is tested using a DMTA. Because

of the small scale of the fibers, each fiber mat can vary dramatically in terms of number

of fibers. In this study, the single fiber properties were derived out of a fiber mat property

by optical microscopy. However, this method does include some level of uncertainty.

Further work should be done to develop a method to measure the single fiber tensile

strength as well as testing the macroscale fiber mat in a more quantifiable and accurate

way.

6.2 Effects of CNC on Dry Spun CA-CNC Fibers

Cellulose nanocrystal (CNCs) in cellulose acetate (CA) fiber nanocomposites were

successfully produced using a dry spinning process. Continuous CNC-CA fibers of

produced. Mechanical properties were found to be greatly improved by the CNC

additions up to 34 wt%, after which there the properties plateaued or decreased, possibly

due to defect formation from a loss in CNC dispersion or residual pores. The fiber

spinning process lead to shear-induced alignment of the CNCs along the fiber axis. The

degree of CNC alignment was shown to be one mechanism for altering the CNC-CA

fiber mechanical properties in the fiber axial direction. Micromechanical models suggest

Page 178: Experimental and modeling investigation of cellulose

156

that the increase in elastic modulus of the CNC-CA continuous fiber result from CNC

concentration, orientation state and aspect ratio. Maximum reinforcement of CNC

additions occurred at 34wt% CNC, in which the resulting improvement in tensile strength

and elastic modulus was of 137% and 637%, respectively.

Future work on this topic include further optimization of the fiber as well as the

spinning process. While this is a batch process, industrial level processes are continuous.

A longer drying column as well as controlled temperature drying conditions would lead

to more homogenous fibers. The fibers produced in the current study are not optimized

by any post processing (washing, solvent extraction, drawing and surface finishing)

which may further improve properties. Drawing may further improve the CNC

orientation within the fiber, as well as the polymer chain orientation.

In order to develop an industrial applicable process, CNC re-dispersion should

also be studied, as it is unrealistic to process CNCs always in wet aqueous form. As the

most common form of dried CNCs are freeze dried, studies should be done to optimize

the dispersion of freeze dried CNCs in various solvents such as water and common

organic solvents.

6.3 Dry Spun Polyvinyl Alcohol Fibers

Continuous PVA-CNC dry spun composite fibers were developed using a plunge

style fiber spinner. Increasing amounts of CNCs (up to 25 wt %) was dispersed into PVA

fibers in efforts to improve the tensile strength and elastic modulus of the fiber.

Mechanical testing of the fibers show maximum modulus improvement of 602%

(1.95GPa to 13.7GPa) at 15wt% CNC, and maximum tensile strength improvement of

Page 179: Experimental and modeling investigation of cellulose

157

283% (41.16MPa to 157.5MPa). Elongation of the fibers generally slightly decreased

with addition of CNCs, however, the stretchability of the fiber remained high. The

variations of elongation at failure between samples of same CNC% are large, indicating

great percentage of defect driven failures. Similar to CNC-CA fibers, birefringence was

observed in the fibers under crossed polarizers, and increased with respect to CNC

concentration. CNC-PVA fibers showed significantly greater improvement in strength

and modulus in comparison to CNC-CA fibers at the same CNC wt%. This may be

attributed to the improved dispersion of CNCs in the water based spinning solution.

Micromechanical data gives good predictions for the modulus performance of the

composite at CNC concentration < 20wt% using the experimentally determined aspect

ratio of 15.

The same process and fiber improvement work described in Section 6.2 would

apply here similarly for future work. In addition, development of additional polymer-

CNC systems would allow us to compare how CNC affect polymers of different

molecular structures and mechanical properties.

6.4 Modeling of Dry Spun Fibers

A one dimensional model was developed to describe a ternary dry spinning system

consisting of cellulose acetate, acetone and cellulose nanocrystals. Fundamental

equations of momentum, mass and energy, as well as rheology and thermodynamics of

the polymer composite was used along with a modified Pipes viscous constitutive

equation to predict the diameter, velocity, component concentration, and temperature

profile of the system. Sensitivity analysis was performed on various spinning parameters

such as mass flow rate, temperature of the spinning dope and air, the velocity of take-up

Page 180: Experimental and modeling investigation of cellulose

158

speed, and the velocity of the quenched air. It was found that an increase in mass flow

rate increased the final fiber diameter and the overall velocity profile. Temperature of the

spinning dope has minimal effects on the spinning process due to the small amount of

mass and heat involved in the filament. The temperature of the quenched air is directly

related to how fast the solidification process is carried out, although the overall profile

remains the same. A higher take-up speed result in a smaller fiber, but does not affect

concentration or temperature profile. Increasing air velocity will accelerate the

solidification process by increasing in heat and mass transfer coefficient, as well as the

increase in air drag coefficient. Changing the air velocity has little to no effect on the

velocity profile.

In addition, the effect of reinforcement fillers on the system was investigated. It is

found that perfectly collimated fillers increased the elongational viscosity of the system,

thus resulting in a thicker fiber. This effect is amplified with increased concentration of

fillers. The more realistic case where filler orientation consists of an ensemble of

misorientation of varying degrees is also investigated. It is found that fillers that are off-

axis reduces the effective elongational viscosity and should be considered during such

simulations.

Future work in this topic include modifications to the simplified assumptions. A

two-dimensional model including the radial changes as well as the axial changes would

give a better description of the system. The effects of solvent concentration in the air, and

surface tension, abide small, should be included in the model. Die swell is another factor

ignored in this model that should be considered and discussed. The viscoelastic model

that account for the filler addition and rheological changes would add further refinement

Page 181: Experimental and modeling investigation of cellulose

159

to the model. One key addition to the system that would be considered is the crystallinity

of the system. While this isn’t key to a cellulose acetate system, as the crystallinity of the

polymer remains relatively low, it will play a dominating effect for a crystallizable

system such as polyvinyl alcohol.

Page 182: Experimental and modeling investigation of cellulose
Page 183: Experimental and modeling investigation of cellulose

p rp p zrt r r z

rdr

R R

p p R p R rp p z p R z RR

R Rr r R R r r R

t t z z

r z RR

R R R

t t z

rp p rp r RRj

Page 184: Experimental and modeling investigation of cellulose

161

Assuming axisymmetry, the equation of continuity for the solvent using

cylindrical coordinates is:

1

v v 0ss rs s zr

t r r z (A.5)

Applying the same cross sectional integration gives:

0 0

d v v d v 0R R

s s R s R rs s z s R z RR

R Rr r

t t z z (A.6)

Using the kinematic boundary condition in Equation A.3 and rearranging terms gives:

0 0

d v d v v 0R R

s s z s R rs rR Rr r r r

t z (A.7)

Applying the steady state assumption, and simplifying the mass flux term:

rs

0

v dR

s z Rr r Rjz

(A.8)

Assuming density is constant in the radial direction, the mass flow rate of the polymer in the filament is calculated using mass balance as:

2

2 vv

ˆp z

p p zp

c RW R (A.9)

Since Wp is a constant, Equation A.10 can be arranged to give R:

1 2

1 1

1 vs z

WR

c (A.10)

The derivative of A.10 with respect to z is:

2 2 2v v 2 v v 2

ˆ ˆs sz

s z s z s z rs Rs s

dW dcdd dRc R Rc c R R Rj

dz dz dz dz dz (A.11)

The derivative of A.11 with respect to z is:

Page 185: Experimental and modeling investigation of cellulose

162

1 2ˆ v1 1 1

2 1 v 1 vp p s z

s z s z

W dc ddR

dz c c dz dz (A.12)

Using Equation A.13, A.12 becomes:

1 2

1 2

ˆvˆ2 v

ˆ1 v v1 1 1

2 v 2 1 v 1 v

p psz

rs R s s z

p ps z s z

s z s z s z

Wddj

Wdz dz d d

dz dz

cc

c c

c c c (A.13)

After simplifying, A.14 becomes:

1 2

3 2ˆ2 1 0ˆ v

srs R s s

p p z

dcj c

dz W (A.14)

The boundary condition at the interface is:

,i ,rs R y s s sj k M y y (A.15)

Where the mass flux is equal to the mass transfer coefficient multiplied by the molecular weight of the solvent, and the difference between the solvent molar fraction between the interface and the bulk gas. The last term represents the driving force for the gas phase transport.

Assuming ideal gas behavior, the interface gas side solvent molar fraction can be calculated as:

,i

sats s

s

a Py

P (A.16)

as, the activity coefficient is calculated using Flory-Huggins theory, and defined in Chapter 5.

Substituting in Equation A.16 and A.17 in Equation A.15 with some simplification gives the final equation of continuity used in the model:

,

ˆ2 1

v

sats y s ss s s

sz

k M cdc a Py

dz R P (A.17)

Page 186: Experimental and modeling investigation of cellulose

163

Appendix B Equation of Momentum

As stated in Chapter 5, for a Newtonian, incompressible fluids with a temperature

and concentration dependent viscosity undergoing uniaxial extension, the following

velocity gradient applies:

0 0

10 0

21

0 02

v

v

v

z

z

z

d

dzd

dzd

dz

v (A.18)

The stress tensor has the form:

0

0 0

zz

rr (5.19)

where rr and are equal.

The equation of motion in the z direction using cylindrical coordinates for the

system is:5

v v v T1

v vz z z zzr z rzr g

t r z r r z (A.20)

By similar treatment, integrating the above equation over the cross-sectional area, 2 rdr from r = 0 to r = R(z,t), using Leibniz’s rule, the following equation is obtained.

2 2

2v v 1v T T

2 2 2z z

z rz zz zzR R

R R RR R R g

t z z z (A.21)

The above equation assumes vz is uniform across the cross sectional area. The bar over

quantities indicate radially averaged such as:

Page 187: Experimental and modeling investigation of cellulose

164

2

0

2(T dT )zz zz

R

z r,z r rR

(A.22)

The stress boundary condition at the fiber surface is:

a

RRn T n T s n (A.23)

Where the super script refer to the air side of the filament, s is the surface tension,

and is the curvature, which can be estimated to 1/R using the slender body assumption

( R/ z << 1). n is the normal unit vector given by:

2

1

1

r z r z

R Rn e e e e

z zR

z

(A.24)

Where er and ez are unit vectors in their respective coordinate. Substituting for Ta with

aap in Equation A.24 gives:

aa RR

sp n n n T n

R (A.25)

Using Equation A.25, the z-component of Equation A.26 can be derived as:

T arz zz a rzR R R

R R s Rp

z z R z (A.26)

Similarly, the r-component is:

T aa rr zr zrR R R

R R sp

z z R (A.27)

Using Equation A.27 along with Tzz zzp , Equation A.22 becomes:

2 2

2v v 1v

2 2 2az z

z a rz zzR

R R R Rp R R s R p g

t z z z z (A.28)

Page 188: Experimental and modeling investigation of cellulose

165

The equation of motion in the r direction using cylindrical coordinates for the

system after applying axisymmetry, steady state, and radial invariance is:

1

0 zrrr

pr

r r r z r (A.29)

By similar treatment, integrating the above equation over the cross-sectional area,

22r dr from r = 0 to r = R(z,t), using Leibniz’s rule, the following equation is obtained

with some algebra.

__1 1

T r2 2

zrrr zr rrR R

Rp

z z (A.30)

Substituting Equation A.28 into Equation A.31 gives:

__1 1r

2 2a

zra zr rrR

R sp p

z R z (A.31)

Substituting the above expression for isotropic pressure into Equation A.29 and

simplifying gives:

2 __

2

v v 1 1v 2

2 2

1r

2

az zz zz rr rz R

azr rz

R

A A A R At z z

R Rs A A

z z z z (A.32)

Further simplifications is applied using rr , assuming uniaxial extension (no shear

stress), steady state and neglecting ambient shear stress derivative and surface tension,

the above equation becomes:

v

v 2 azz zz rr rz R

A A R Az z

(A.33)

Page 189: Experimental and modeling investigation of cellulose

166

Specifically, the second term on the RHS refer to the drag associated with the air. This

term is assumed to be:1, 6

2

v v

2a z aa

rz fRC (A.34)

Finally, the equation of momentum derived for the system is:

2v

v zz zz rr f a z aA A

z z (A.35)

Page 190: Experimental and modeling investigation of cellulose

167

Appendix C Equation of Energy

The equation of energy for the system assuming steady state, axisymmetry, and radial invariance is defined as:5

q :v vp z

TC

z (A.36)

Where q is the energy flux for the system:

q k T (A.37)

The viscous dissipation term, v: , is neglected in this case, and combining

the above two equations give:

vzp

TC k T

z (A.38)

Expanding the RHS gives:

v1

zp

T T TC rk k

z r r r z z (A.39)

Assuming the axial conduction is negligible gives:

v1 1

zp r

T TC rk rq

z r r r r r (A.40)

Similar to previous equations, integrating over the fiber cross section ( 2 rdr

from 0 to R) gives:

v 2zp r R

TA C R q

z (A.41)

The energy flux at the boundary condition is given as:

4 4

2r a B a v RR rq h T T T T H j (A.42)

Page 191: Experimental and modeling investigation of cellulose

168

The first term on the RHS is the heat transfer due to mass flux. The second term

on the RHS refer to the radiation term, and is neglected in this case. The last term refer to

the heat of evaporation associated with the mass flux at the interface.

Substituting A.43 with the above considerations into A.42 and simplifying gives:

2

2vp a v Rrz

TC h T T H j

z R (A.43)

Finally, using the expression for mass flux at the interface given in Equation A.16

and A.17 gives the expression for equation of energy for the model:

,

2v

sats

p z a v y s s

Pd T

d z R P (A.44)

Page 192: Experimental and modeling investigation of cellulose

169

Bibliography

1. Gou, Z.; McHugh, A. J., A comparison of Newtonian and viscoelastic constitutive models for dry spinning of polymer fibers. J Appl Polym Sci 2003, 87 (13), 2136-2145.

2. Gou, Z. Modeling of dry spinning of polymer fibers. Ph.D., University of Illinois at Urbana-Champaign, Ann Arbor, 2003.

3. Doufas, A. K.; McHugh, A. J.; Miller, C., Simulation of melt spinning including flow-induced crystallization. Part I. Model development and predictions. Journal of non-newtonian fluid mechanics 2000, 92 (1), 27-66.

4. Deng, G.; Xia, Q.; Xu, Y.; Zhang, Q., Simulation of dry-spinning process of polyimide fibers. J Appl Polym Sci 2009, 113 (5), 3059-3067.

5. Bird, R. B.; Stewart, W. E.; Lightfoot, E. N., Transport Phenomena. Wiley: 2007.

6. Shimizu, J.; Okui, N.; Tamai, K., Air Drag in High Speed Melt Spinning. Sen-I Gakkaishi 1983, 39 (10), T398-T407.

Page 193: Experimental and modeling investigation of cellulose
Page 194: Experimental and modeling investigation of cellulose