You are on page 1of 193

CHEMICAL MODIFICATION OF STARCH AND PREPARATION OF STARCH-

BASED NANOCOMPOSITES

A Dissertation

Presented to

The Graduate Faculty of The University of Akron

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

Lin Song

August, 2010
CHEMICAL MODIFICATION OF STARCH AND PREPARATION OF STARCH-

BASED NANOCOMPOSITES

Lin Song

Dissertation

Approved: Accepted:

Advisor Department Chair


Dr. Chang Dae Han Dr. Sadhan C. Jana

Committee Member Dean of the College


Dr. Mark D. Soucek Dr. Stephen Z.D. Cheng

Committee Member Dean of the Graduate School


Dr. Sadhan C. Jana Dr. George R. Newkome

Committee Member Date


Dr. Li Jia

Committee Member
Dr. Peter L. Rinaldi

ii
ABSTRACT

A series of chemically modified starches were synthesized in the present study: (1)

a starch modified with an anionic group and benzoyl group; (2) a starch modified with an

anionic group and hexanoyl group; (3) a starch modified with a cationic group and

benzoyl group; and (4) a starch modified with a cationic group and hexanoyl group. The

glass transition temperatures of the modified starches were found to be lower than that of

natural starch. The modified starches synthesized in the present study are found to be

hydrophobic and can biodegrade in composting. Blends of a modified starch and

poly(ethylene-co-vinyl alcohol) (EVOH) were prepared by melt blending. Phase

diagrams of the blends exhibited a lower critical solution temperature. The DSC

thermograms of the blends showed a single glass transition temperature and melting point

depression. The crystallization of EVOH in the blends was interrupted by the hydrogen

bonds formed between the hydroxyl groups in the modified starch and EVOH. The

crystalline structure of EVOH in the blends had an orthorhombic lattice in the quenched

specimen and evolved into a monoclinic lattice when the annealing temperature was

increased. The nanocomposites based on a modified starch and EVOH were prepared.

Natural clay (montmorillnite, MMT) and two commercial organoclays (Cloisite 30B and

Cloisite 15A) were employed to investigate how the functional groups in the modified

starch influence the dispersion characteristics of nanocomposites. Anionically modified

starch was found to be very effective in exfoliating organoclay aggregates due to the

iii
presence of ionic interaction, as determined by Fourier transmission infrared (FTIR)

spectroscopy, between the anionic group in the modified starch and positively charged N+

in the surfactant residing at the surface of an organoclay. For the nanocomposites based

on a cationically modified starch and EVOH, the preparation methods had a large

influence on the dispersion characteristics of the nanocomposites. The aggregates of

MMT have a very high degree of dispersion characteristics in the nanocomposites

prepared by solution blending but poor dispersion characteristics in the nanocomposites

prepared by melt blending. FTIR spectroscopy has indicated that the ionic interaction

between the cationic group in a modified starch and negatively charged surface of silicate

sheets of MMT could be formed in the nanocomposites prepared by solution blending but

not in the nanocomposites prepared by melt blending. Wide-angle X-ray diffraction

(WAXD) has indicated that the crystalline structure of EVOH in the nanocomposites

containing exfoliated aggregates of organoclay Cloisite 30B takes orthorhombic lattice in

both quenched and annealed specimens, which is different from that of neat EVOH. We

ascribe the difference in the crystalline structure of EVOH in the EVOH-containing

nanocomposites and in neat EVOH to the formation of hydrogen bonds between the

hydroxyl groups in EVOH and the hydroxyl groups in the surfactant residing at the

surface of Cloisite 30B. We have found that an improvement in the tensile properties of

nanocomposites can only be obtained if the nanocomposites have a very high degree of

dispersion of the aggregates of clay, and there is strong attractive interaction between the

clay and the matrix.

iv
ACKNOWLEDGEMENTS

I would like to express my sincere gratitude to my advisor, Professor Chang Dae

Han, for his advice, patience and support throughout this study. I am proud of having

worked on this dissertation under his guidance.

I would like to extend my gratitude to the committee members, Professors Mark D.

Soucek, Sadhan Jana, Li Jia and Peter L. Rinaldi, for their helpful advices and comments

on the draft of this dissertation.

This dissertation is dedicated to my parents and my wife for their understanding and

love.

v
TABLE OF CONTENTS

Page

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

LIST OF TABLES…………………………………..…………………………….….…xvi

LIST OF SCHEMES…………………………………………………………………...xvii

CHAPTER

I. INTRODUCTION .………………………………………………………….………....1

II. LITERATURE REVIEW .…………………………………………………….…..…..8

2.1 Introduction to Natural Starch …………………………………….……….…..…8

2.2 Previous Research on Chemical Modification of Starch…………...…………....17

2.3 Thermoplastic Starch (TPS)………………………………………….………..…43

2.4 Previous Research on the Nanocomposites based on Starch………………...…..47

III. EXPERIMENTAL…………...…………………………………………..……..…...51

3.1 Synthesis of Chemically Modified Starch…………………………………….....51

3.2 Preparation of Blends and Nanocomposites based on Modified Starch…….…...54

3.3 Characterization Methods……………………...…………………………….…..55

IV. CHARACTERIZATION OF CHEMICALLY MODIFIED STARCH…….….…...60

4.1 Structure Characterization and Determination of Degree of Substitution


(DS) of Chemically Modified Starch…………………………...……………......60

4.2 Moisture Resistance of Chemically Modified Starch………..…………...……...66

vi
4.3 Thermal Behavior of Chemically Modified Starch…………….………………..68

4.4 Biodegradability of Chemically Modified Starch…………..……...…………….72

4.5 Concluding Remarks……………………………………………………………..74

V. BLENDS OF MODIFIED STARCH AND EVOH……………………...………......76

5.1 Miscibility of Blends based on Modified Starch and EVOH


Characterized by Phase Diagrams………………………………………………76

5.2 Hydrogen Bonding in Blends based on Modified Starch and EVOH


Characterized by DSC and FTIR…….………………………………….………77

5.3 Crystallization of EVOH in Blends based on Modified Starch


and EVOH…………………………..…………………………………………...89

5.4 Mechanical Properties of Blends based on EVOH and Modified Starch……….93

5.5 Concluding Remarks……………………………………………………..….…..98

VI. NANOCOMPOSITES BASED ON MODIFIED STARCH..…………...………...100

6.1 Introduction………………………………………………………………….....100

6.2 Nanocomposites based on Neat EVOH………………………………….….…103

6.3 Nanocomposites containing Anionically Modified Starch….………….…..….107

6.4 Nanocomposites containing Cationically Modified Starch…...……………….116

6.5 Crystallization and Crystalline Structures of EVOH in Nanocomposites..........133

6.6 Mechanical Properties of Nanocomposites………….………………...……….141

6.7 Concluding Remarks…………………………………………………………...147

VII. CONCLUSIONS AND RECOMMENDATIONS….………………………...…..150

7.1 Conclusions………………………………………………………………….....150

7.2 Recommendations………………………………………………………….…. 153

REFERENCES ………………………………………………………………………...155

vii
APPENDICES………………………………………………………………………….170

APPENDIX A. 1H-NMR SPECTRA OF CHEMICALLY MODIFIED


STARCH…………………………….……………………………171

APPENDIX B. FTIR SPECTRA OF CHEMICALLY MODIFIED STARCH…...174

viii
LIST OF FIGURES

Figure Page

1.1 Different types of dispersion of layered silicates in polymeric matrix…………...…..3

2.1 Chemical structures of amylose and amylopectin…………………...…………...…..9

2.2 Single helix structure of amylose molecule……………………………………..…..10

2.3 Structure of branched amylopectin……………………………………………….…11

2.4 Two major types of crystalline structure in the natural starch granules:
A-type and B-type……………………………………………………………….…..13

2.5 Vh-type crystalline structure of starch……………………………………………....13

2.6 X-ray diffraction patterns of A-, B-, and Vh-type of starch……………………...…14

2.7 Schematic representation of the granule architecture of natural starch…………..…15

2.8 Hydrolysis profiles of starch treated by sulfuric acid: (▼) Waxy Maize;
(▽) Potato; (〇) Normal Maize; (□) AE Waxy Maize; and (●) Hylon VII.……...…22

2.9 DSC thermograms of waxy rice starch under acid hydrolysis (the number
next to each curve indicates the duration of hydrolysis in days) ………………...…23

2.10 Thermal stability of amylose esters……………………………………………..…32

2.11 XRD patterns of starch esters……………………………………………………...34

2.12 Crystallization of fatty acid chains within the starch esters……………………..…34

2.13 Water uptakes of starch esters versus time: (■) OCST 1.8; (◆) OCST 2.7; (▲)
DODST 2.7; (▼) OCDST 1.8; and (●) OCDST 2.7..………………….……..…..36

2.14 Stress-strain curves of potato starch with different amounts


of water content (the water contents (wt %) are indicated on the figures)………...44

ix
2.15 X-ray diffraction diagrams of (top) granular waxy maize starch and (bottom)
TPS materials stored at various temperatures and humidity. From bottom
to top: quenched and stored for 2 weeks; stored at 40% relative humidity (RH)
for 2 weeks; stored at 60% RH for 2 weeks; stored at 70% RH for 8 months;
and stored at 90% RH for 2 weeks.……………………………………………..….46

2.16 Effect of Cloisite Na+ on the XRD patterns of TPS nanocomposites: (a) Cloisite
Na+; (b) TPS/Cloisite Na+(97.5/2.5 wt %); (c) TPS/Cloisite Na+(95.0/5.0 wt %);
(d) TPS/Cloisite Na+(90.0/10.0 wt %); and (e) TPS.…………............................…48

2.17 TEM image of TPS nanocomposite containing 5 wt % of MMT…………...……..48

4.1 FTIR spectra of (a) natural starch (NS), (b) starch modified with anionic group
and benzoyl group (AS-B), and (c) starch modified with anionic group and
hexanoyl group (AS-LF).………………………………...…………………..……...61

4.2 FTIR spectra of (a) natural starch (NS), (b) starch modified with cationic group
and benzoyl group (CS-B), and (c) starch modified with cationic group and
hexanoyl group (CS-LF)………………... …………………………………..……...62

4.3 1H-NMR spectra of a starch modified by hexanoyl group……………………….…65

4.4 Moisture absorption of modified starch: (●) NS; (■) AS; (▲) CS; (▼) AS-LF;
(◆) AS-B; (□) CS-LF; and (〇) CS-B….……..…………………...………………..66

4.5 Thermal degradation temperatures of chemically modified starches…………..…...68

4.6 DSC thermograms of (a) AS, (b) AS-LF, and (c) AS-B………………………….…71

4.7 DSC thermograms of (a) CS, (b) CS-B, and (c) CS-LF………………………...…..71

4.8 Weight loss of modified starch during composting: (●) NS; (■) CS-LF;
(▲) LD-PE; (▼) CS-B; (◆) AS-LF; and (□) AS-B……...……………………..…73

5.1 Phase diagrams of the blends of a modified starch and EVOH: (●) (AS-B)/EVOH
blend; (〇) (AS-LF)/EVOH blend; (▼) (CS-B)/EVOH blend; and (▽) (CS-LF)/
EVOH blend………….…….………………..…………...………………………....77

5.2 DSC thermograms of the blends of (AS-B)/EVOH: (a) neat EVOH; (b) blend
of 20/80 (AS-B)/EVOH; (c) blend of 40/60 (AS-B)/EVOH; (d) blend of
70/30 (AS-B)/EVOH; and (e) neat AS-B…………………………….......................79

x
5.3 DSC thermograms of the blends of (AS-LF)/EVOH: (a) neat EVOH; (b) blend
of 20/80 (AS-LF)/EVOH; (c) blend of 40/60 (AS-LF)/EVOH; (d) blend
of 70/30 (AS-LF)/EVOH; and (e) neat AS-LF……………….…………..……..…..79

5.4 DSC thermograms of the blends of (CS-B)/EVOH: (a) neat EVOH; (b) blend
of 20/80 (CS-B)/EVOH; (c) blend of 40/60 (CS-B)/EVOH; (d) blend
of 70/30 (CS-B)/EVOH; and (e) neat CS-B………………….………..………..…..80

5.5 DSC thermograms of the blends of (CS-LF)/EVOH: (a) neat EVOH; (b) blend
of 20/80 (CS-LF)/EVOH; (c) blend of 40/60 (CS-LF)/EVOH; (d) blend
of 70/30 (CS-LF)/EVOH; and (e) neat CS-LF……………………………………...80

5.6 DSC thermograms in the glass transition region for the blends of
(AS-B)/EVOH: (a) neat EVOH; (b) blend of 20/80 (AS-B)/EVOH;
(c) blend of 40/60 (AS-B)/EVOH; (d) blend of 70/30 (AS-B)/EVOH;
and (e) neat AS-B……………………………………………………………………83

5.7 DSC thermograms in the glass transition region for the blends of
(AS-LF)/EVOH: (a) neat EVOH; (b) blend of 20/80 (AS-LF)/EVOH;
(c) blend of 40/60 (AS-LF)/EVOH; (d) blend of 70/30 (AS-LF)/EVOH;
and (e) neat AS-LF………………………………………………………………...…83

5.8 DSC thermograms in the glass transition region for the blends of
(CS-B)/EVOH: (a) neat EVOH; (b) blend of 20/80 (CS-B)/EVOH;
(c) blend of 40/60 (CS-B)/EVOH; (d) blend of 70/30 (CS-B)/EVOH;
and (e) neat CS-B……………………………………………………………..……..84

5.9 DSC thermograms in the glass transition region for the blends of
(CS-LF)/EVOH: (a) neat EVOH; (b) blend of 20/80 (CS-LF)/EVOH;
(c) blend of 40/60 (CS-LF)/EVOH; (d) blend of 70/30 (CS-LF)/EVOH;
and (e) neat CS-LF……………………………………………………………..........84

5.10 FTIR spectra in the hydroxyl group stretching region for: (a) neat EVOH;
(b) neat modified starch AS-B; and (c) blend of 70/30 (AS-B)/EVOH………..….87

5.11 FTIR spectra in the hydroxyl group stretching region for: (a) neat EVOH;
(b) blend of 70/30 (AS-LF)/EVOH; and (c) neat modified starch AS-LF…....…...87

5.12 FTIR spectra in the hydroxyl group stretching region for: (a) neat EVOH;
(b) blend of 70/30 (CS-B)/EVOH; and (c) neat modified starch CS-B…………....88

5.13 FTIR spectra in the hydroxyl group stretching region for: (a) neat EVOH;
(b) blend of 70/30 (CS-LF)/EVOH; and (c) neat modified starch CS-LF………....88

xi
5.14 WAXD patterns of neat EVOH at different annealing temperatures (oC):
(1) 143; (2) 123; (3) 103; (4) 83; and (5) quench. Annealing period was for 3 h,
and the specimen was quenched by liquid nitrogen………………………………..90

5.15 WAXD patterns for a 70/30 (AS-B)/EVOH blend at different


annealing temperatures (oC): (1) 145; (2) 125; (3) 105; and (4) quench……...…..91

5.16 WAXD patterns for a 70/30 (AS-LF)/EVOH blend at different


annealing temperatures (oC): (1) 130; (2) 110; (3) 90; (4) 70; and (5) quench.…...91

5.17 WAXD patterns for a 70/30 (CS-B)/EVOH blend at different


annealing temperatures (oC): (1)143; (2)123; (3)103; and (4) quench……..……...92

5.18 WAXD patterns for a 70/30 (CS-LF)/EVOH blend at different


annealing temperatures (oC): (1) 132; (2) 112; (3) 92; (4) 72; and (5) quench..…..92

5.19 Tensile properties of (AS-B)/EVOH blends: (a) tensile strength;


(b) Young’s modulus; and (c) elongation at break.…………………...……….…..94

5.20 Tensile properties of (AS-LF)/EVOH blends: (a) tensile strength;


(b) Young’s modulus; and (c) elongation at break……………………...….……...95

5.21 Tensile properties of (CS-B)/EVOH blends: (a) tensile strength;


(b) Young’s modulus; and (c) elongation at break…………………….…………..96

5.22 Tensile properties of (CS-LF)/EVOH blends: (a) tensile strength;


(b) Young’s modulus; and (c) elongation at break…………………….………..…97

6.1 XRD patterns for: (a) Cloisite 15A; (b) MMT; and (c) Cloisite 30B……………...104

6.2 XRD patterns for: (a) EVOH/Cloisite 15A nanocomposite; (b) EVOH/
Cloisite 30B nanocomposite; and (c) EVOH/MMT nanocomposite……….……...105

6.3 TEM images for: (a) EVOH/MMT nanocomposite; (b) EVOH/Cloisite 15A
nanocomposite; and (c) EVOH/Cloisite 30B nanocomposite, in which the
dark areas represent the clay, and the grey/white areas represent
the polymer matrix………………………………………………………………....106

6.4 XRD patterns for: (a) (AS-B)/EVOH/Cloisite 15A nanocomposite;


(b) (AS-B)/EVOH/Cloisite 30B nanocomposite; and (c) (AS-B)/EVOH/MMT
nanocomposite………………………………………………………………….….107

6.5 TEM images for: (a) AS-B/EVOH/Cloisite 30B nanocomposite;


(b) (AS-B)/EVOH/Cloisite 15A nanocomposite; and (c) (AS-B)/EVOH/MMT
nanocomposite………………………………………………………………….….108

xii
6.6 FTIR spectra for: (a) (AS-B)/EVOH blend; (b) (AS-B)/EVOH/Cloisite 30B
nanocomposite; (c) (AS-B)/EVOH/Cloisite 15A nanocomposite;
and (d) (AS-B)/EVOH/MMT nanocomposite…….…………………………….....109

6.7 Carboxylate group stretching region for FTIR spectra shown in Figure 6.6
for: (a) (AS-B)/EVOH blend; (b) (AS-B)/EVOH/Cloisite 30B nanocomposite;
(c) (AS-B)/EVOH/Cloisite 15A; and (d) (AS-B)/EVOH/MMT nanocomposite.....111

6.8 XRD patterns for: (1) (AS-LF)/EVOH/Cloisite 15A nanocomposite;


(2) (AS-LF)/EVOH/MMT nanocomposite; and (3) (AS-LF)/EVOH/Cloisite 30B
nanocomposite……………………………………………………………………..112

6.9 TEM images for: (a) (AS-LF)/EVOH/Cloisite 30B nanocomposite;


(b) (AS-LF)/EVOH/Cloisite 15A nanocomposite; and (c) (AS-LF)/EVOH/MMT
nanocomposite……………………………………………………………………..113

6.10 FTIR spectra at carboxyl group stretching region for: (a) (AS-LF)/EVOH blend;
(b) (AS-LF)/EVOH/Cloisite 30B nanocomposite; (c) (AS-LF)/EVOH/
Cloisite 15A; and (d) (AS-LF)/EVOH/MMT nanocomposite……………………115

6.11 XRD patterns for: (a) (CS-B)/EVOH/MMT nanocomposite; (b) (CS-B)/EVOH/


Cloisite 30B nanocomposite; and (c) (CS-B)/EVOH/Cloisite 15A
nanocomposite. The nanocomposites were prepared by melt blending…………116

6.12 TEM images for: (a) CS-B/EVOH/Cloisite 15A nanocomposite; (b) CS-B/
EVOH/Cloisite 30B nanocomposite; and (c) CS-B/EVOH/MMT
nanocomposite. The nanocomposites were prepared by melt blending……........117

6.13 XRD patterns for: (a) (CS-LF)/EVOH/MMT nanocomposite; (b) (CS-LF)/


EVOH/Cloisite 30B nanocomposite; and (c) (CS-LF)/EVOH/Cloisite 15A
nanocomposite. The nanocomposites were prepared by melt blending………....118

6.14 TEM images for: (a) (CS-LF)/EVOH/Cloisite 15A nanocomposite; (b) (CS-LF)/
EVOH/Cloisite 30B nanocomposite; and (c) (CS-LF)/EVOH/MMT
nanocomposite. The nanocomposites were prepared by melt blending…………119

6.15 XRD patterns for: (a) (CS-B)/Cloisite 30B nanocomposite; (b) (CS-B)/MMT
nanocomposite; and (c) (CS-B)/Cloisite 15A nanocomposites. The
nanocomposites were prepared by solution blending……..……………………...120

6.16 TEM images for: (a) (CS-B)/Cloisite 30B nanocomposite; (b) (CS-B)/MMT
nanocomposite; and (c) (CS-B)/Cloisite 15A nanocomposite. The
nanocomposites were prepared by solution blending……………..……………...121

xiii
6.17 XRD patterns for: (a) (CS-LF)/Cloisite 30B nanocomposite; (b) (CS-LF)/MMT
nanocomposite; and (c) (CS-LF)/Cloisite 15A nanocomposite. The
nanocomposites were prepared by solution blending……………………..……...122

6.18 TEM images for: (a) (CS-LF)/Cloisite 30B nanocomposite; (b) (CS-LF)/MMT
nanocomposite; and (c) (CS-LF)/Cloisite 15A nanocomposite. The
nanocomposites were prepared by solution blending…..……………….……..…122

6.19 FTIR spectra in the C-N+ stretching mode region of nanocomposites based on
cationically modified starch prepared by solution blending: (a) neat CS-B;
(b) (CS-B)/MMT nanocomposite; (c) (CS-B)/Cloisite 15A nanocomposite;
and (d) (CS-B)/Cloisite 30B nanocomposite…………………………..……...….124

6.20 FTIR spectrum in the C-N+ stretching mode region of nanocomposites


based on cationically modified starch prepared by melt blending:
(a) (CS-B)/EVOH blend; (b) (CS-B)/EVOH/Cloisite 15A nanocomposite;
(c) (CS-B)/EVOH/Cloisite 30B nanocomposite; and (d) (CS-B)/EVOH/MMT
nanocomposite. ………………………………………………………………......127

6.21 Structure of MMT…………………………………………………………….…..128

6.22 Structure of MMT: (a) side view of layers in MMT; (b) top view
of tetrahedral layers of MMT; (c) the ditrigonal cavity formed among 6
oxygen atoms in tetrahedral layers of MMT; and (d) coordination polyhedron
formed by ditrigonal cavities of two consecutive layers of oxygen atoms.
In (b), (c) and (d), only oxygen atoms are shown as black spots………………....130

6.23 Schematic of silicates layer in MMT with exchangeable cations:


(a) top view and (b) side view…………………………………………………….131

6.24 Schematic of electrolytic dissociation of ditrigonal cavity and the cations


on the surface of silicates in MMT…………………………………………….…132

6.25 DSC thermograms for: (a) AS-B/EVOH/Cloisite 15A nanocomposite;


(b) (AS-B)/EVOH/Cloisite 30B nanocomposite; (c) (AS-B)/EVOH/MMT
nanocomposite; (d) blend of (AS-B)/EVOH; (e) (AS-B)/Cloisite 30B
nanocomposite; (f) (AS-B)/Cloisite 15A nanocomposite; and (g) neat AS-B..….134

6.26 WAXD patterns for (AS-B)/EVOH/Cloisite 15A nanocomposite at different


annealing temperatures (oC): (1) 153; (2) 133; (3) 113; and (4) quench.…….…..137

6.27 WAXD patterns for (AS-B)/EVOH/Cloisite 30B nanocomposite at different


annealing temperatures (oC): (1) 151; (2) 131; (3) 111; and (4) quench………....137

xiv
6.28 WAXD patterns for (AS-B)/EVOH/MMT nanocomposite at different
annealing temperatures (oC): (1) 146; (2) 126; (3) 106; and (4) quench…………138

6.29 WAXD patterns for (CS-B)/EVOH/Cloisite 15A nanocomposite at different


annealing temperatures (oC): (1) 143; (2) 123; (3) 103; and (4) quench…………139

6.30 WAXD patterns for (CS-B)/EVOH/Cloisite 30B nanocomposite at different


annealing temperatures (oC): (1) 144; (2) 124; (3) 104; and (4) quench….…...…140

6.31 WAXD patterns for (CS-B)/EVOH/MMT nanocomposite at different


annealing temperatures (oC): (1) 144; (2) 124; (3) 104; and (4) quench…….…...140

6.32 Tensile properties of the nanocomposites based on AS-B and EVOH:


(a) tensile strength; (b) Young’s modulus; and (c) elongation at break.
(●) (AS-B)/EVOH blend; (□) (AS-B)/EVOH/Cloisite 30B nanocomposite;
(△) (AS-B)/EVOH/Cloisite 15A nanocomposite; and (▽) (AS-B)/
EVOH/MMT nanocomposite..…………………..………………………………..142

6.33 Tensile properties of the nanocomposites based on AS-LF and EVOH:


(a) tensile strength; (b) Young’s modulus; and (c) elongation at break.
(●) (AS-LF)/EVOH blend; (□) (AS-LF)/EVOH/Cloisite 30B nanocomposite;
(△) (AS-LF)/EVOH/Cloisite 15A nanocomposite; and (▽) (AS-LF)/
EVOH/MMT nanocomposite…....………………………………………………..143

6.34 Tensile properties of the nanocomposites based on CS-B and EVOH:


(a) tensile strength; (b) Young’s modulus; and (c) elongation at break.
(●) (CS-B)/EVOH blend; (□) (CS-B)/EVOH/Cloisite 30B nanocomposite;
(△) (CS-B)/EVOH/Cloisite 15A nanocomposite; and (▽) (CS-B)/
EVOH/MMT nanocomposite..................................................................................144

6.35 Tensile properties of the nanocomposites based on CS-LF and EVOH:


(a) tensile strength; (b) Young’s modulus; and (c) elongation at break.
(●) (CS-LF)/EVOH blend; (□) (CS-LF)/EVOH/Cloisite 30B nanocomposite;
(△) (CS-LF)/EVOH/Cloisite 15A nanocomposite; and (▽) (CS-LF)/
EVOH/MMT nanocomposite..................................................................................145

xv
LIST OF TABLES

Table Page

1.1 Bond energy and relative strength of different intermolecular forces …………....….5

2.1 The transition temperatures of starch esters…………...………………..………...…30

2.2 The tensile properties of starch esters………………………………………….....…33

2.3 XRD results of TPS/clay nanocomposites…………………..………….……..….…49

2.4 Tensile properties of TPS/Cloisite 30B and TPS/MMT nanocomposites……......…50

4.1 Peak assignments in the FTIR spectra of the chemically modified starches…….….63

4.2 Degree of substitution of chemically modified starches……………………….…....66

5.1 Melting points and the degree of crystallinity of the blends of EVOH and
a modified starch……………………………………………………………..…...…81

5.2 Glass transition temperatures of EVOH/modified starch blends……………….…...85

6.1 The changes in the thermal properties of nanocomposites….………………….….135

xvi
LIST OF SCHEMES

Scheme Page

2.1 Chemical structure of amylopectin and amylose……………………...…..……..….18

2.2 Proton-catalyzed hydrolysis of starch……………………………………………….21

2.3 Alkaline degradation of starch……………………………………………………....25

2.4 Nitration of starch………………………………………………………………...…27

2.5 Acrylation by carboxylic acid chloride…………………………………………...…30

2.6 Alkylation of starch having various side chains………………………………….....37

2.7 Reaction between starch and ethylene oxide or propylene oxide………………...…39

2.8 Quaternary ammonium starch ester……………………………………….….……..41

2.9 Iminoalkyl starch ether…………………………………………………………...…41

2.10 Aminoethylated starch ether……………………………………………...…….….42

2.11 Tertiary aminoalkyl starch ether…………………………………………..…….…42

6.1 Chemical structures of the modified starches……………………...………........…102

6.2 Chemical structures of the surfactant MT2EtOH residing at the surface of


Cloisite 30B and the surfactant 2M2HT residing at the surface of Cloisite 15A ....103

xvii
CHAPTER I

INTRODUCTION

Increasing interest has been cast upon the research of biopolymers because of the

environmental pressure of using lesser amounts of synthetic petroleum-based polymers.

Most petroleum-based polymeric products are not biodegradable. It usually takes several

decades, even centuries, for nature to decompose these polymeric materials. Much

energy and effort spent on the landfill of polymeric wastes could be saved by using

biodegradable polymer products. Besides environmental pressure, exhausting petroleum

resources also drive the researchers to look for alternative resources for polymer

materials. Biopolymeric products manufactured from natural resources, such as starch

and cellulose, can offer the polymer industry inexhaustible and green resources. Natural

starch, as one of the most abundant natural resources for polymer materials, is

inexpensive and biodegradable. With necessary modification and processing, starch can

be used as ideal “green” polymer material that comes from nature and readily goes to

nature. Judicious chemical modifications can be made on starch because of its abundant

hydroxyl groups.

Research showed that all natural starches have granular structure containing

semicrystalline lamellae.1 Both the glass transition temperature and the melting

temperature of pure, dry starch are higher than its thermal decomposition temperature.

Thus, the starch can only be processed as thermoplastic materials in the presence of
1
plasticizers and under the action of heat and shear. Many low-molecular-weight

materials have been used as plasticizers, such as water, glycerol, sorbitol, amine, etc. to

produce thermoplastic starch (TPS).

Water, among all the plasticizers, is the most effective. However, TPS containing

low-molecular-weight plasticizers has inevitable shortcomings. The abundant hydroxyl

groups on the molecules render the starch hydrophilic. The low moisture resistance of

starch limits its use for thermoplastic processing. Additionally, the content of plasticizers,

such as water, in the starch continuously changes because of the plasticizer’s migration,

evaporation, and the starch’s moisture absorption. The physical and mechanical

properties of starch are very sensitive to the content of plasticizers.

Chemical modification is an alternative approach to making starch thermoplastic.

Side chains can be introduced into the starch molecules through chemical reactions

between the hydroxyl groups and functional groups. The side chains interrupt the

hydrogen bonds between the hydroxyl groups of starch and destroy the granular

semicrystalline structure of natural starch. The mobility of starch molecules is enhanced

so that the glass transition temperature and the melting temperature can be lowered.

Furthermore, the decreasing number of available hydroxyl groups and the introduction of

hydrophobic functional groups make the starch hydrophobic.

When a thermoplastic polymer is mixed with layered silicates, the state of

dispersion of aggregation of layered silicates in nanocomposites is schematically shown

in Figure 1.1.

2
Figure 1.1. Different types of dispersion of layered silicates in polymeric matrix.

In the past two decades, the preparation of polymer-based nanocomposites, using

nanoclay, has been a very active research area.2 For the majority of industrial

applications, the ultimate goal for the preparation of nanocomposites is to achieve

exfoliation of the aggregates of layered silicates. The layered silicates consist of platelets

of approximately 1 nm thick and a large aspect ratio (about 1000). From the point of

view of obtaining enhanced physical and mechanical properties of nanocomposites, large

surface areas of exfoliated nanocomposites are highly desirable. Obviously, exfoliated

layered silicates in a polymeric matrix will enhance the available surface areas. Thus,

exfoliation is preferred to intercalation.

It is known that the layered silicates of natural clay are hydrophilic3,4 and do not

have any chemical affinity with hydrophobic synthetic polymers. This is the reason why

so much effort has been placed on providing organophilic surfaces of layered silicates

through chemical modification. The chemical affinity between a surfactant residing at

the surface of an organoclay and the polymeric matrix can give rise to highly dispersed

3
(exfoliated) organoclay nanocomposites. The extent of dispersion of the aggregates of

layered silicates in a polymeric matrix is usually determined by: (1) the chemical

structure of a surfactant residing at the surface of layered silicates; (2) the chemical

structure and the architecture of polymer matrix; (3) the molecular weight of polymer

matrix; (4) the extent of functionality in polymer matrix; (5) the location(s) of the

functionalization; and (6) the gallery distance of layered silicates. Such factors affecting

the dispersion of the aggregates of layered silicates in a polymer are determined by the

nature of chemical interaction between the surface of the layered silicates and polymer.

The preparation of starch-based nanocomposites using montmorillonite (MMT) or

organoclay began several years ago.5,6 It was assumed that the hydrophilic starch would

be compatible with hydrophilic MMT. However, so far researches have only observed a

rather poor dispersion of the aggregates of MMT or organoclay in starch, as determined

by X-ray diffraction (XRD) and/or transmission electron microscopy (TEM). The details

of this research will be discussed in the next chapter. The reasons for the poor dispersion

of aggregates of clay in starch lie in the fact that the assumed attractive interaction

between starch and clay is very weak, and the mobility of amylose and amylopectin

molecules in starch is very low, owing to very high molecular weights and intense

hydrogen bonds, making them difficult to approach the gallery of the layered silicates.

Strong attractive interaction between the layered silicates and the polymer matrix

must be provided in order to achieve a high degree of dispersion (exfoliation) of

aggregates of clay in nanocomposites. In this regard, strong intermolecular forces must

be introduced. Some well-known specific interactions include ionic interactions, ion-

dipole interactions, hydrogen bonding, and dipole-dipole interactions. Table 1.1 gives a

4
summary of bond energy and the relative strength of such intermolecular forces. It can

be concluded that ionic interaction provides the strongest intermolecular forces.

Table 1.1. Bond energy and relative strength of different intermolecular forces7

Type of interaction Bond energy (kJ/mol) Relative strength


Ionic attraction 850-1700 1000
Ion-dipole interaction 500-700 600
Hydrogen bonding 50-170 100
Dipole-dipole interaction 2-8 10
van der Waals interaction less than 1 1

Once a strong interaction, such as ionic interaction, is introduced between starch

and clay, a high degree of dispersion (exfoliation) of aggregates of clay can be expected.

On the other hand, the substitution of hydroxyl groups by other functional groups can

also effectively enhance the mobility of starch molecular chains.

A biodegradable polymeric product based on modified starch, having comparable

physical and mechanical properties, can potentially replace the synthetic petroleum-based

polymeric materials, such as low-density polyethylene (LDPE), which is widely used in

the current market in such applications as grocery bags, garbage bags, baby diapers, and

food packaging materials. The present research was to prepare polymer blends and

nanocomposites based on chemically modified starch. Through our research, we can find

a way of making use of natural starch as thermoplastic polymer materials. Moreover,

from a fundamental point of view, the research of blends and nanocomposites based on

modified starch is still a new field. Few previous researchers have ever done such a

study, and there was no good dispersion of nanoclay (intercalation at best) ever reported.

5
In the first part of this dissertation, we present the synthesis of modified starches

having different functional groups, such as acetate sodium, trimethylammonium chloride,

hexanoyl group, and benzoyl group. The role of hexanoyl group and benzoyl group on

the starch is to lower the number of hydroxyl group so that the hydrogen bonds present in

starch can be destroyed. Then the mobility of starch molecules is enhanced, and the glass

transition temperature is lowered. Furthermore, the hydrophobic functional group can

make a chemically modified starch hydrophobic. The role of sodium acetate and

trimethylammonium chloride is to introduce ionic groups into the starch molecules. The

ionic group in starch molecules can form ionic interaction with the aggregates of

nanoclay so that a high degree of dispersion of silicate sheets can be obtained in

nanocomposites.

In the second part of this dissertation, we present experimental results for the

blends of modified starch and poly(ethylene-co-vinyl alcohol) (EVOH). The miscibility

between a modified starch and EVOH was investigated. Miscible blends of a modified

starch and EVOH were obtained at specific compositions. The FTIR spectra show the

presence of the hydrogen bonds between a modified starch and EVOH. We ascribed the

hydrogen bonds between modified starch and EVOH to the reason of high miscibility in

blends. In addition, the crystallization and crystalline structure of EVOH in the blends

with the interference of hydrogen bonds were investigated.

In the third part of this dissertation, we present the dispersion characteristics of the

nanocomposites based on miscible blends of a modified starch and EVOH. We

investigated the dispersion characteristics of the nanocomposites, using X-ray diffraction

(XRD) and transmission electron microscopy (TEM). Natural clay (MMT) and

6
organoclays (Cloisite 30B and Cloisite 15A) that were treated by surfactants were

employed in the present study. We also took effort to find out how the degree of

dispersion of clay aggregates depends on the functional groups in a modified starch

molecules and the surfactant residing at the surface of organoclay.

In the fourth part of this dissertation, we present the crystallization and crystalline

structure of EVOH in the nanocomposites prepared. We found that the degree of

dispersion of clay aggregates and the surfactant residing at the surface of clay have a

profound influence on the crystallization and crystalline structure of EVOH. The tensile

properties were investigated, including tensile strength, Young’s modulus, and elongation

at break, of miscible blends of a modified starch and EVOH, and nanocomposites based

on such blends.

7
CHAPTER II

LITERATURE REVIEW

2.1 Introduction to Natural Starch

Starch is a carbohydrate material that can be hydrolyzed into small molecules by

specific enzymes during animal metastasis, generating biological energy for animals.

Many plants, in their fruits, seeds, tubers or rhizomes, grow with starch. Potato, corn,

wheat and rice are major resources for production of starch worldwide.

2.1.1 Compositions of Natural Starch

Starch contains two major components: one is amylose, which is a linear polymer,

and the other is amylopectin, which is a highly branched polymer. The repeat units in

both amylose and amylopectin are identical (so-called α-D-glucosyl residues) but

connected in different ways, as given in Figure 2.1. Most starch contain 20―30 wt % of

amylose, although some waxy starches contain very little amylose.

In amylose, α-D-glucosyl residues are linked by an α[1→4] bond. Typically, the

molecular weights of extracted amylose are in the order of 105―106 g/mol.8 Amylose

generally is recognized as linear macromolecule. However, it has been reported that

there is a slight degree of branching (9―20 branch points per molecule) in amylose from

various botanical sources.9 The length of side chains ranges from 4 to over 100 repeat
8
OH
OH O
O OH
OH
O HO OH
OH
O O OH

OH O OH OH OH
OH
acetal OH 6
O O O 5 O O
HO 4 1 OH
O O O O
3
2
HO OH HO OH HO OH HO OH HO OH
anomeric carbon

amylopectin n

OH OH OH
6
O 5 O O hemiacetal
HO 4 1 OH
O O
3
2
HO OH HO OH HO OH

amylose

Figure 2.1. Chemical structures of amylose and amylopectin.

units.9,10 There is no effective way of separating the branched amylose from the linear

amylose.

Side chains on branched amylose form small clusters rather than a random comb

structure. Note that amylopectin also forms clusters by its branched side chains, which

will be elaborated on later. Branched amyloses still have properties, such as iodine

binding capability, which amylopectin does not have.9 In aqueous solutions, amylose

molecules can form extended chains with a hydrodynamic radius of 7―22 nm.11 In

natural starch, the amylose molecules tend to have a rather stiff left-handed single helical

9
structure or form an even stiffer parallel double left-handed helix.11 The helix consists of

six α-D-glucosyl residues per turn with a pitch of 0.8 nm. Figure 2.2 shows a single helix

structure of amylose molecule.

On each of the glucose residues, hydroxyl groups connecting with carbons on

position 2 and 6 (C2 and C6) point outward of the helix, while the hydroxyl group

connecting with carbon on position 3 (C3) points inward of the helix, which makes the

inner surface of the helix more hydrophobic than the outer surface. Thus, amylose with a

helix structure can hold the hydrocarbon portion of glycerides and fatty acid in its helix

cavity to form a V-helix complex. The distinct capability of amylose to bind with iodine

is also by holding iodine ions in its hydrophobic helix cavity.

Figure 2.2. Single helix structure of amylose molecule.

10
Amylopectin is a highly branched component of starch. The α-D-glucosyl residues

are mainly linked by α[1→4] bonds but with 5-6% of α[1→6] bonds at the branch

points. Its molecular weight ranges 107―109 g/mol. The amylopectin molecular chains

can be classified into three categories by their ways of connecting with each other: the C

chain carries the sole reducing end of the molecule; the A chain (unbranched) is only

linked to the rest of the molecule through its potential reducing end; and the B chains

carry one or more A chains and/or B chains. Figure 2.3 shows a schematic describing

this classification.

Figure 2.3. Structure of branched amylopectin.

In most experiments, the C chain is not distinguished from B chains. According to

their positions in the cluster structure, B chains are further subdivided: B1 is a short chain,

being components of only one cluster while B2, B3, etc. are long chains that span over

two, three, or more clusters, thereby interconnecting them.

11
The branching points on the amylopectin molecules are not randomly distributed

but periodically clustered. The distance between two clusters on the same B chains is 22

α-D-glucosyl residues on average. Most short chains (the average degree of

polymerization ranges from 14 to 18) are organized within clusters, while long chains

(the average degree of polymerization is more than 55) usually span more than one

cluster. The molar ratio of short to long chains, influenced by the source of starch, varies

from 5 in potato starch amylopectin to 8―10 in cereal amylopectin. The distribution of

chain length in amylopectin critically influences the inner assembly structure of starch

granules and the interaction with the water or solvent.

2.1.2 Structures of Natural Starch

As described above, the linear polysaccharide chains tend to have helix structure

either left-handed or right-handed. The left-handed form is energetically preferred to the

right-handed form.11 There are two ways the helical molecule is packed in the crystalline

phases of starch: A-type and B-type, as shown in Figure 2.4.

In A-type crystalline structure, these helices are packed in a monoclinic unit cell (a

= 2.124 nm, b = 1.172 nm, c = 1.069 nm, γ= 123O);11,12 in the B-type structure, the

helices are packed in a hexagonal unit cell (a = b = 1.85 nm, c = 1.04 nm).12 In both

A- and B-type crystals, the helices are packed parallel.

Besides A-type and B-type crystalline structures in natural starches, previous

researchers12 also found Vh-type crystalline structure in starch complex formed between

the single helix of amylose molecules with compounds, such as iodine, dimethyl

12
Figure 2.4. Two major types of crystalline structure in the natural starch granules: A-type
and B-type.13 (Reprinted with permission from Elsevier.)

sulfoxide (DMSO), alcohol, and fatty acid. In such complexes, iodine atoms or the

hydrocarbon part of alcohol or fatty acid are held by the hydrophobic inner surface in the

helix, as given in Figure 2.5. Then these complexes are packed parallel in an

orthorhombic unit cell.

Figure 2.5. Vh-type crystalline structure of starch.13 (Reprinted with permission from
Elsevier.)
13
Figure 2.6 shows the typical X-ray diffraction patterns for the three crystalline

structures described above. In the natural starch source, A-type crystalline is found

mainly in cereal starch, and B-type is in tuber and amylose rich starches.14,15 The Vh-

type crystal is rarely observed in natural starch.16 X-ray diffraction results also showed

that in some natural starch there is a mixture of A-type and B-type crystalline phases,

called C-type, mostly found in legume starches and cereal grown under specific

conditions of temperature and hydration.

In natural starch, polysaccharide chains are packed in granules. Their size varies

1 to 100 μm. The organization of crystalline and amorphous phases in granules can be

observed by acid hydrolysis. Crystalline phase is believed to be more resistant to acid

hydrolysis than amorphous phase. A schematic representation of the granule architecture

is given in Figure 2.7. The layered concentric shell, the so-called “growth ring”, was

observed. These shells, with the thickness of 120-400 nm,17,18 consist of alternating

amorphous and semi-crystalline layers.

Figure 2.6. X-ray diffraction patterns of A-, B-, and Vh-type of starch.13 (Reprinted with
permission from Elsevier.)

14
Figure 2.7. Schematic representation of the granule architecture of natural starch.18
(Reprinted with permission from Elsevier.)

2.1.3 Gelatinization, Glass Transition, and Retrogradation of Starch

Gelatinization of starch is usually known as the disruption of the crystalline

structure and ultimately the destruction of the starch granular structure when heat and

water (or other effective solvents) are applied. In the presence of excessive water

(usually > 90 wt %), starch granules start to swell because the amylose molecules are

preferentially solubilised in the water. Below a characteristic temperature known as

gelatinization temperature, the crystalline region, mainly composed of high molecular

weight and branched amylopectin molecules, still maintains its integrity. Above the

gelatinization temperature, the crystalline region starts to lose its order and swell

irreversibly.19 If approximately treating starch granules as a polymer network, the

swelling of the starch before the occurrence of gelatinization can be considered as a

reversible equilibrium. In this equilibrium, the affinity between the polymer network and

the solvent generates osmotic pressure (favoring the swelling), which is equal to the

restoring force (disfavoring the swelling), from the stiffness of polymer network.

Obviously, enhancing the temperature or choosing the solvent that has a high affinity
15
with starch will increase swelling. Swelling above the gelatinization temperature is

profoundly irreversible.

The glass transition temperature (Tg) of dry starch is experimentally inaccessible,

because its degradation temperature is higher than Tg. The addition of water has a very

strong plasticizing effect, causing a remarkable depression in Tg.20 Since the Tg of starch

is considerably influenced by the water content, the mechanical properties are also

affected by it. Due to the evaporation and migration of water, the water content in starch

keeps changing, causing difficulty in maintaining steady mechanical properties of the

starch materials. Other non-volatile plasticizers have been used. Many studies were

carried out on low molecular-weight compounds, such as glycerol and sorbitol.21-25

Retrogradation is another characteristic behavior of natural starch granules while

interacting with water. If the gelatinized starch/water mixture is cooled to room

temperature, there will be a strong driving force for starch to crystallize. Then water will

normally be driven out of the starch (phase separation). This process, which occurs upon

cooling, is known as retrogradation. Obviously, the thermodynamic force for

crystallization will be different between amylose and amylopectin because of their highly

different conformations of molecular structure. The former has very long molecular

chains, which make the formation of crystalline phase harder than the branched

amylopectin having abundant short linear molecular chains.26 If a concentrated aqueous

amylose solution is cooled to room temperature, a very rapid phase separation will be

observed. The clear solution turns to opaque very fast, indicating the appearance of the

polymer aggregates. The size of the aggregates is larger than the wavelength of the light.

When the concentration of amylose solution reaches to a certain level (greater than 1―2

16
wt %), an elastic gel is observed after a sudden temperature drop. Double helices were

observed in the initial aggregates.27 It was reported that the rate of crystallization of

amylose was very low, as observed by X-ray diffraction.28,29

When a concentrated amylopectin/water mixture was quenched to room

temperature or below, a slow crystallization of amylopectin was observed by X-ray

diffraction.30,31 This crystallization is associated with the development of stiffness of the

material. The crystallinity could reach up to 30%, comparable to that found in the natural

starch granules. It was also reported that the longer the chain of amylopectin and the

more abundance of short chain fraction, the greater tendency the amylopectin will have to

retrograde and crystallize from aqueous solution.30-34

2.2 Previous Research on the Chemical Modification of Starch

Various types of chemical reactions can occur on different sites in the molecular

chains of starch, resulting in different properties and molecular assembly.

2.2.1 Reactive Positions on the Molecular Chains of Starch

The chemical reactivity of starch is controlled by the reactivity of its glucose

residues. Scheme 2.1 shows the molecular structure and reactive sites of amylose and

amylopectin. On every glucose residue, there are six carbons (referred as C1, C2, C3 . . . ,

etc.). Generally, there are two terminal hydroxyl groups at the ends of the molecular

chains at C1 and C4, as shown in Scheme 2.1. In each glucose residue of amylose, there

are two secondary hydroxyl groups at C2 and C3, as well as one primary hydroxyl group

at C6. In amylopectin, each branch point decreases the number of primary hydroxyl

17
groups by one and simultaneously increases the number of secondary hydroxyl groups at

C4 by one.

Scheme 2.1. Chemical structure of amylopectin and amylose.

OH
OH O
O OH
OH
O HO OH
OH
O O OH

OH O OH OH OH
OH
acetal OH 6
O O O 5 O O
HO 4 1 OH
O O O O
3
2
HO OH HO OH HO OH HO OH HO OH
anomeric carbon

amylopectin n

OH OH OH
6
O 5 O O hemiacetal
HO 4 1 OH
O O
3
2
HO OH HO OH HO OH

amylose

All available hydroxyl groups of amylose and amylopectin are capable of being

oxidized and reduced, as well as participating in the formation of hydrogen bonds. These

hydroxyl groups may form salts and participate in the formation of ethers and esters. The

starch esters may be formed with inorganic, as well as organic acids.

18
The macrostructure of starch also has an impact on its reactivity. Inter- and

intramolecular hydrogen bonding decrease the solubility of the material and limit the

accessibility of potential reaction sites to the reagents used.35 Entanglement of amylose

and amylopectin chains may also be a factor. These circumstances do not change the

reactivity of starch in the qualitative sense but may have serious quantitative impact.

Hence, the selection of the reagents and reaction conditions require special consideration.

The granular character of native starch also influences the reactivity of starch.

Depending on the penetrating ability of the reagent, the reaction of starch proceeds either

on the granule surface or in its interior.35,36 Moreover, the capillaries between granules,

together with the polar character of the granule surface, make starch a good, porous

absorbent. Many reagents undergo inclusion in the capillaries, forming starch capillary

complexes.36

Since humans started using starch, not only as a food but also as various applied

materials, chemical modification has been a major way of obtaining desirable properties

and adding special value to the natural starch. We do not intend to cover all types of the

chemical modifications reported. Instead, several major chemical modifications, mainly

involving hydroxyl reaction and acidic/basic hydrolysis, will be described below. It is

interesting to note that some chemical modifications that were mainly used in food

related practices may also be used to obtain thermoplastic materials.

2.2.2 Acidic Hydrolysis of Natural Starch

Because of its polyacetal structure, starch is very sensitive to protonic acids that

readily catalyze the hydrolysis of starch. Even very small amounts of weak organic acids,

19
in the form of complexes within the amylose helix, produce visible decreases in the

viscosity of starch gels.37,38 Potato starch contains some phosphate ester groups as metal

salts. These phosphate esters can be decationized and turned into acids that finally

catalyze the hydrolysis, producing the so-called hydrogenated starch. Although this

acidic group occurs on only one of every 30-200 glucose units, it is sufficient to make

this starch unstable during storage. Slightly hydrolyzed starch has low viscosity and

good gel-forming ability. More extended hydrolysis has been used to produce adhesives.

Finally, oligo- and monosaccharides were achieved with further extended hydrolysis.39

Scheme 2.2 shows the hydrolysis pathway that starts with a fast, reversible

protonation at either the oxygen atoms on the pyranose ring and/or the oxygen atoms of

the glycosidic bonds. This step is followed by the slow breakage of the respective C-O

bond. The resulting carbocations rapidly accept a water molecule from the solution. It

was shown that both mechanisms operate in parallel, and that particular reaction

conditions give priority to one of them.

Acid-catalyzed hydrolysis of starch was reported as a unimolecular process

resembling the hydrolysis of simple glycosides and led to glucoses as a single, acid-stable

product.40 The hydrolysis of starch catalyzed by acid was usually accompanied by

reversion. The primary hydrolysis products could be repolymerized. In addition, both

this repolymerization and decomposition depended on the reaction time, temperature,

concentration, and pH.

Hydrolysis reaction follows the slow penetration of acid into starch granules, which

may differentiate the hydrolyzing ability of various acids. Hydrochloric acid has always

20
Scheme 2.2. Proton-catalyzed hydrolysis of starch.
H
HO O HO O HO O

O H O
H O
O fast O fast O
OH OH OH
H
HO HO HO

H
HO O HO O
HO O
O O
O OH
O OH OH
OH
HO HO
HO

proven to be the most effective catalyst.41 Electron microscopy of acid-hydrolyzed starch

granules indicates that the amorphous regions of granules accept the catalyzing acid

first.42 In contrast, the crystalline regions hydrolyze much more slowly. It is reported

that all starches exhibit a two-stage hydrolysis pattern.40 Figure 2.8 gives the hydrolysis

profiles of various starches treated by sulfuric acid. The two-stage process is evident.

The faster stage corresponds to the hydrolysis of the more amorphous parts of the starch

granule. During the second stage, the crystalline material is slowly degraded. This is

consistent with the observation that hydrolysis increases the crystallinity of the starch

granules.43 There are two hypotheses proposed to account for the slower hydrolysis rate

of the crystalline parts of the starch.44,45 First, the dense packing of starch chains within

the starch crystallites does not readily allow the penetration of protonic cation into the

regions. Second, acid hydrolysis of a glucosidic bond may require a change in

21
Figure 2.8. Hydrolysis profiles of starch treated by sulfuric acid: (▼) Waxy Maize; (▽)
Potato; (〇) Normal Maize; (□) AE Waxy Maize; and (●) Hylon VII.46 (Reprinted with
permission from Elsevier.)

conformation (from chair to half-chair) of the D-glucose unit, and the crystalline structure

immobilizes the conformation, making this transition sterically difficult. Furthermore,

the differences of the rate and the extent of hydrolysis have been attributed to differences

in granular size, the extent of starch chain interactions (within the amorphous and

crystalline regions of the granule), and starch composition (the extent of phosphorylation,

amylose content, and lipid-complexed amylose chains).

The hydrolysis of starch initially results in the formation of amylodextrin paste that

still develops a strong blue stain with iodine.47 Then it turns into a complex of dextrin

and maltose. The gelatinization temperature and the breadth of the gelatinization

endotherm have been shown to increase on acid hydrolysis as shown in Figure 2.9.30

22
Figure 2.9. DSC thermograms of waxy rice starch under acid hydrolysis (the number
next to each curve indicates the duration of hydrolysis in days).30 (Reprinted with
permission from Elsevier.)

Since acid hydrolysis preferentially attacks the amorphous regions in the granule, the

crystalline regions are decoupled from and no longer destabilized by the amorphous

parts.18 Consequently, the starch crystallites of the acid-treated starches melt at a higher

temperature, and the transition temperature range is broader.48,49 The studies on normal

and waxy barley starch lintner residues suggested that the higher transition temperature

might be due to longer amylopectin double helices than in the unhydrolyzed amylopectin

molecules, whereas the branch points might reduce the length of the helix.50,51

23
As described above, the hydrolysis of starch resulting from protonic acid mainly

involves the oxygen atoms on the pyranose ring or glycosidic bonds. The acidic

hydrolysis results in more primary hydroxyl groups and short molecular chains. The

three hydroxyl groups on the pyranose are not involved in the acidic hydrolysis. This is

in contrast to the chemical reaction between starch and acryl acids (described later) and is

a very important way to functionalize starch by adding functional group onto starch

molecular chains.

2.2.3 Reactions of Natural Starch Under Basic Conditions

The behavior of starch with base depends on the type of base, its concentration,

either the presence or absence of oxygen, and the temperature and time of the reaction.

The term “degradation” is a better word to describe the reaction of starch under basic

conditions.

Diluent aqueous alkali solution (0.1-0.4 wt %) at room temperature caused the

swelling of starch granules and the liberation of proteins and lipids present in the starch.52

It was reported that starch gelatinization was retarded by hydroxyl anions, especially in

the presence of salts and that the gelatinization temperature increases with anion

concentration.53

Scheme 2.3 shows reaction pathways in the alkaline degradation of starch. Various

carboxylic acids (formic, acetic, glycolic, lactic, 2-hydroxybutanoic, 2-hydroxyl-2-

methylpropanoic, and 2-hydroxypentanoic acids) were isolated from alkali-degraded

starch.54 They constituted 41-46 wt % of the starting material. The alkaline

24
Scheme 2.3. Alkaline degradation of starch.

OH H
O OH H OH
O O
R H H
O R R H
OH O O O
HO
HO OH O HO OH O

OH H OH H
O O
R CH 2 OH CH 2 OH
O
HO O HO O

OH H OH H
Benzylic acid
O O
rearrangement
COO
CH 2 OH CH 2 OH
O O HO

COO
COO
H C H
CH 2 OH
H C OH
CH 2 OH

degradation of starch under anaerobic conditions is a slow process that increases with the

concentration of alkali.52 Part of the starch undergoes degradation with diluted aqueous

alkali, whereas other parts are resistant, even to concentrated alkali solutions. It has been

reported that these proportions of the alkali-resistant part may be changed by treatment of

starch with acids, heat, or by grinding.55 Moreover, viscosity measurements showed that

the rate of alkali degradation was faster during the initial stages.44,55 These findings

suggest that starch undergoes selective degradation and that these reaction sites are

hindered inside the starch matrix. Since the degradation of starch is considered to

25
progress from the reducing end of the macromolecular chains, the reducing-terminal

glucose units, even in branched structures are split off, indicating that, not only amylose

but also amylopectin is alkali labile.

In air, the reaction retains its unimolecular character, but it proceeds faster and is

accelerated by phosphate and acetate ions.56 The degradation just described is further

assisted by oxidation. Starch can also be degraded by treatment with hydrolyzing salts

formed from strong bases and weak acids, i.e., sodium carbonate, sodium

hydrogencarbonate, sodium borate, hydrogenphosphates, and alumina blended with

sodium oxide.57

2.2.4 Esterification

Starch readily forms esters with organic and inorganic reagents because of its

hydroxyl groups on each glucose residue. One, two, or all three hydroxyl groups of the

starch glucose units may be involved in esterification. The selectivity of these groups in

esterification is affected by the reagent, reaction conditions, and the degree of chain

branching in the amylopectin component.36,58 The reactions usually proceed in

suspension. With granular starch, such reactions take place, first on the granular surface.

After the reagents penetrate the granules, the reaction occurs in both the amorphous

region and crystalline region of the granule interiors.

(a) Nitration

The study of nitration of starch initially was inspired by the great success of

cellulose nitrates. The chemical similarity of cellulose and starch motivated the studies

of starch esterified with nitric acid. Scheme 2.4 shows the starch of nitration.

26
Scheme 2.4. Nitration of starch.

NO2
OH O
O O
O -
O O
+ H N
O O O O
OH O
OH O
O 2N NO2

starch nitrate

O
N
OH O
O O
- O
O + H O N O
O O
OH O
OH O N
N
O
O

starch nitrite

Significant differences were reported in the viscosities of solutions of esterified

cellulose and starch having the same degree of nitration.59 Increasing the concentration

of nitric acid and maintaining a low reaction temperature was crucial to achieving a high

degree of esterification and a high overall yield.60 Impurities in native starch, such as

proteins and lipids, should be removed prior to nitration to avoid contamination by their

nitration. This can be achieved by the preliminary treatment of starch with a solution of

1.5-2 wt % aqueous sodium hydroxide.61

Nitration also occurred with a mixture of nitric and phosphoric acid to produce a

material having 13.0% nitrogen content, which means an almost completely trinitrated

starch (fully substituted starch has 14.14% of nitrogen content).62 This method of

nitration inhibits acid-hydrolytic degradation of the starch structure.

27
It was shown that nitration with nitric acid alone gives a methanol-soluble product.

Nitration with mixtures of nitric and sulfuric acids gave products partially soluble in

methanol.63 The insoluble portion consisted of amylopectin nitrates. The highest degree

of esterification occurs when either gaseous or liquid nitric anhydride was used and also

in the presence of sodium fluoride in chloroform, which allows the reaction to proceed at

a temperature close to ambient.

Amylopectin nitrate is more viscous than amylose nitrate.64 Solid esters are pure

white, soluble in concentrated sulfuric acid, and also more hygroscopic than esterified

cellulose. Their stability was high, and their density exceeded that of the cellulose

esters.62,64 Amylose nitrate is more stable than amylopectin nitrate, and the latter is more

stable than starch nitrate. Isolated starch nitrate could also be protected from hydrolysis

by washing the crude isolated product with a solution of 50-90 wt % ethanol or by

washing reacting acids out with a stream of water. Isolated starch nitrate could also be

protected by boiling crude starch nitrate in water under pressure and eventually either by

the addition of sodium carbonate or by boiling in ethanol. Nevertheless, the extent of

acid hydrolysis upon nitration is significant. Starch nitrate is actually an ester of an

oligosaccharide of 6-7 glucose units and even maltose octanitrate.35 The nitrated starch

was mainly used as an explosive when the nitrogen content was higher than 9%.62

(b) Acylation by carboxylic acid chloride

In the acylation of starch, it was seen that anhydrides, carboxylic acid, and

carboxylic acid chloride could all be used to react with hydroxyl groups. However,

previous research showed that the reactivity of carboxylic acid and its anhydride usually

28
are very low.35,65 The reaction has to be catalyzed by mineral acids, such as sulfuric acid

and perchlorate acid, but the results of acylation were influenced by hydrolysis of the

esters in the acidic reaction medium. Thus, some metal oxide or metal hydroxide, for

example MgO or Mg(OH)2, must also be added to control the acidity of the reaction

mixture.63

Carboxylic acid chlorides were more often used in the acylation of starch as its

reactivity is stronger than carboxylic acid. Thus, when starch with higher fatty acid esters

is desired, especially with a high degree of substitution, the acyl chlorides were used.

Scheme 2.5 shows the reaction. Pyridine was proven to be a very effective catalyst in the

acylation starch with acyl chlorides.66,67 The first reason for its effectiveness lies in the

fact that the long pair electron on the nitrogen atom can accelerate the breakage of acetyl

chloride bond. Moreover, pyridine, acting like tertiary amines, is a good acceptor of

hydrochloric acid produced during the esterfication. In some research, esterification was

even conducted using pyridine as a suspension medium without any other solvents.67-69

Other catalysts, such as quioline and triethyamine, were also used. Starches having long-

chain fatty acid esters with various chain lengths have been obtained, ranging from starch

having butanoic acid ester (four carbons on the fatty acid chain) to starch having

octadecanoic acid ester (eighteen carbons on the fatty acid chain).66,67,70,71 The thermal

and mechanical properties of these starch esters were examined. The relationship

between the properties of starch esters and the side-chain length and degree of

substitution was researched. The resulting starch esters are hydrophobic, thermoplastic

materials that can be shaped, handled, and processed with much greater ease than the

native starch.

29
Scheme 2.5. Acrylation by carboxylic acid chloride.

OH
OH
O
O
O O
O Pyridine
+ Cl C(CH 2 ) n CH 3 O
O HO O
HO OH
(CH 2 ) n CH 3
O

Table 2.1 lists transition temperatures of some starch long-chain fatty acid

esters.66,68,72 It is known that the Tg and Tm of dry native starch cannot be determined,

because they lie above the decomposition temperature. The transition temperatures of

starch esters were dramatically lowered through chemical modification. This can be

explained by the plasticization introduced by ester groups. On the other hand,

esterification of long-chain fatty acid reduces the number of hydroxyl groups and

interferes with the hydrogen bonding. The crystalline phase presented in the native starch

was destroyed by the side chains. The mobility of amylose and amylopectin molecules

was enhanced. Previous research showed that the increase of side-chain length has

resulted in a depression of glass transition temperature.67 This behavior can be explained

by the increase of free volume of the polymer resulting from the increasing size of bulky

Table 2.1. The transition temperatures of starch esters66,68,72

Starch esters Tg (oC) Tm (oC)


starch having butanoic acid ester 65 ―
starch having pentanoic acid ester 63 124
starch having hexanoic acid ester 50 91
starch having octanoic acid ester (OCST) − 50/40 ―
starch having dodecanoic acid ester (DODST) − 56/25 ―

30
flexible side-chain groups. It should be noted that some starch esters have two glass

transition temperatures, and this behavior can only be found during the first scan in the

DSC. In the second scan, only one glass transition was observed at a temperature close to

the larger transition observed in the first scan. So far, there is no systematic research or

conclusive explanation on this behavior. A possible suggestion is that these two glass

transition temperatures correspond to two components of starch esters, amylose ester and

amylopectin ester.66 As the unmodified two components in native starch (amylose and

amylopectin are incompatible, located in different granule regions), they remain to be

incompatible after esterification. Since amylopectin has a branched structure, it must

have a higher glass transition temperature than amylose, which applies accordingly to

their esters.

Previous studies also showed that decomposition temperature was increased after

esterification.67,68 Figure 2.10 shows the thermal stability of starch esters.67 This greater

thermal stability of starch ester is due to the reduced amount of the remaining hydroxyl

groups after esterification. The amylose esters have similar results in thermal stability.70

The study also showed that the main mechanism of starch thermal decomposition is the

dehydration among the hydroxyl groups on the starch molecules.73 The internal

plasticization provided by the bulky flexible side groups also profoundly affects the

mechanical properties of starch esters. The starch esters with a high degree of

substitution behave like typical thermoplastic materials, showing tensile strength and

high elongation at break. The side-chain length and the degree of substitution both play

important roles in the tensile strength and elongation at break.

31
Figure 2.10. Thermal stability of amylose esters.67 (Reprinted with permission from
John Wiley & Sons.)

Table 2.2 lists the mechanical properties of some starch esters.66,70,72 First, it is

shown that the tensile strength decreases as DS gets higher, whereas the elongation at

break increases with the DS. On the other hand, the increase of the side chain length

results in lower tensile strength and higher elongation at break once the DS remains

consistent. It seems that the longer side chain provides more intensive plasticization to

the final materials. However, there is a reversal trend on the mechanical properties.

When the side chain length increases to 18 carbons, the starch ester has higher tensile

strength and lower elongation at break than starch esters with shorter side chains (such as

8 carbons and 12 carbons). This behavior must be attributed to the partial crystallization

of 18-carbon side chain. The crystalline phase in the material usually contributes to the

32
Table 2.2. The tensile properties of starch esters66,70,72

Starch esters (DS) Tensile strength Elongation at break


(MPa) (%)

starch having butanoic acid ester (2.7) 15.9 36.7


starch having pentanoic acid ester (2.7) 12.7 20
starch having hexanoic acid ester (2.7) 5.48 74.9
starch having octanoic acid ester (2.7) 0.7 380
starch having octanoic acid ester (1.8) 3.1 9
starch having dodecanoic acid ester (2.7) 0.7 1500
starch having octadecanoic acid (2.7) 1.9 10
starch having octadecanoic acid (1.8) 3.7 9

high tensile strength and low elongation at break. The results of DSC and X-ray

diffraction already proved the presence of the side chain crystalline in starch esters.67,71

A complete loss of native crystallinity after substitution of hydroxyl groups by a fatty

acid chain is observed by XRD as Figure 2.11 shows. The large diffraction peak between

9.9 and 10.5 increases when the length of fatty acid chains grafted on the starch is

increased up to 18 carbons. These two diffraction peaks were attributed to the fatty acid-

amylose complex as shown in Figure 2.11. It is seen that there is an optimum side-chain

length, which can provide effective plasticization and desirable mechanical properties.

Figure 2.12 schematically displays the crystallization of fatty acid groups on the starch.

Sagar et al.66 have reported on the long-chain fatty acid starch esters, and the DSC

thermograms also indicated the presence of some crystallinity. It was also found that the

crystalline structure did not reappear in the second scan during the measurement of DSC.

However, annealing at a temperature above the glass transition temperature and below

the melting point led to the reappearance of endotherms of starch esters.

33
Figure 2.11. XRD patterns of starch esters.71 (Reprinted with permission from
Wiley-VCH.)

Figure 2.12. Crystallization of fatty acid chains within the starch esters.71 (Reprinted
with permission from Wiley-VCH.)

34
Since the long-chain fatty acid groups lead to the loss of the hydrophilic character

of starch, the water uptake of starch esters will reflect the hydrophobicity. Figure 2.13

shows the diagram of water uptake of starch esters versus time.70 The DS is the main

factor affecting the water uptake. It shows that starch esters, such as the octanoated and

the octadecanoated with a degree of substitution 1.8, have significantly higher water

absorption compared to the corresponding esters with a degree of substitution of 2.7.

Even so, the maximum increase does not exceed 4 wt %, indicating essentially

hydrophobic materials. The length of fatty acid groups has a minor effect on the water

uptake. Dodecanoated and octadecanoated starch esters having DS = 2.7 demonstrate a

similar behavior, which is different from that of octanoated esters. The latter has a very

small water absorption, which seems to grow steadily without reaching a plateau as in the

case with esters having DS = 1.8. This behavior could be attributed to the lower chain

length of the ester, which cannot effectively hinder the unreacted hydroxyl groups to

absorb water. Besides being catalyzed by pyridine in organic solvents, such as dioxane

and toluene, the same long side-chain starch esters could be produced in a organic

solvent-absence method.67,68,70 In this method, starch was first treated by formic acid

prior to reacting with long-chain fatty acid chloride. The highly reactive formic acid

reacts with hydroxyl groups on the starch, producing formate starch esters. The initial

reaction decreases the number of hydroxyl groups and interferes with the strong hydrogen

bonding among the starch molecules. The remaining hydroxyl groups in formate starch

ester are more accessible to the fatty acid chloride that was added immediately after the

initial reaction, without interruption. In each of the steps, the reagents formic acid and

fatty acid chloride were in excess. This one-pot process was conducted under a stream of

35
Figure 2.13. Water uptakes of starch esters versus time: (■) OCST 1.8; (◆) OCST 2.7;
(▲) DODST 2.7; (▼) OCDST 1.8; and (●) OCDST 2.7.70 (Reprinted with permission
from John Wiley & Sons.)

nitrogen gas that drew the hydrochloric acid away from the reaction system. The formate

group is inherently unstable and removed at the end of the reaction, giving rise to pure

starch long-chain fatty acid ester. Starch esters can be obtained with minimum

degradation during the reaction. However, compared with those of the catalyzed

esterification, the degree of substitution (DS) of starch esters cannot be higher than 2.

Moreover, the results indicated that the DS of starch ester dramatically diminished with

an increase in the fatty acid chain length, because the longer fatty acid has less

accessibility to the hydroxyl groups in the absence of solvent.

36
2.2.5 Etherification

(a) Alkylation with alkyl and benzyl halides and sulfates

The hydroxyl groups of the glucose residues in starch, exhibiting reactivity typical

for alcohols, could undergo conversion into ethers by various alkylation agents. Methyl

and ethyl starch ethers were produced by using methylating and ethylating agents,

dimethyl sulfate, and diethyl sulfate.74-76 The reactions took place in dimethyl sulfoxide

(DMSO) with the addition of aqueous NaOH. The degree of substitution of the reaction

products could be very high. The infrared spectra of the products had little absorption at

3400-3600 cm−1, indicating that nearly all of the hydroxyl groups had been converted

into ether groups.77

More often, starch ethers with various side chains were produced through

reactions with alkyl halides. Scheme 2.6 shows the reaction of alkylation of starch with

various side chains. Early reports78 showed that starch ethers could be produced by

blending starch with alkyl halide in the presence of such desiccating agents as CaCl2.

Higher alkyl halides react with starch in aqueous alkali, but the use of high pressure is

recommended.63 Research indicated that methylation of starch may retain the granular

Scheme 2.6. Alkylation of starch having various side chains.

OH OH
O O
O A queous N aO H O
+ Cl CH2 R
O O
HO OH HO O

R= H CH3 ,
OH , ,

37
structure of the product, although the alkylation was uniform throughout the granule

interior.79,80 The addition of sodium phosphate inhibits gelatinization of starch granules

and has been used in the preparation of granular methyl starch ether with a degree of

substitution above 2.77

So far, there has been no report regarding alkyl starch ethers as thermoplastics.

This is partly because long-chain alkyl groups can hardly be attached to the starch by

alkyl halide, in contrast to the long-chain fatty acid chloride esterification of starch. The

reactivity of alkyl halide was highly limited by its length.65

Besides the alkyl starch ethers, the benzyl group could also be added onto starch

through the reaction between starches with benzyl chloride.81,82 Presently, benzyl starch

ether is not widely used, but there are possibilities for future use in the industry. The

benzyl group provides hydrophobicity to the products that can be used as molding

powder, resin additive, and detergent component. Moreover, research showed good

miscibility between benzyl starch and certain synthetic polymers, such as poly(vinyl

acetate), polystyrene, polyurethane, etc.83-85 Such polymer composites, copolymerized

or blended, show good tensile strength, high softening temperature, increased flexibility,

and more importantly, biodegradability. Benzyl starch could be a potential choice for

TPS, because the benzene ring could effectively provide hydrophobicity to the materials

and interfere with the hydrogen bonding between hydroxyl groups to mobilize the starch

molecules. The benzene ring could also act as an internal plasticizer, similar to long-

chain fatty acid groups. So far, the reported benzyl starch had only low DS (less than

0.5).81 In the future, benzyl starch having high DS should be synthesized to obtain

thermoplastic characters.

38
(b) Alkylation with epoxide compounds

Etherification of starch by epoxides (also referred to as alkylene oxides) can occur

without the addition of alkali. Alkylene oxides of complicated structures provided novel

opportunities for starch modifications. The asymmetrical epoxide group is very reactive

because of its highly strained three-membered oxide ring with bond angles of 60o.36

Opening of the asymmetrical epoxide ring occurs by attacking at the sterically less

hindered end of epoxide to produce 2-hydroxylalkyl starch ether.86 Scheme 2.7 shows

the reaction between hydroxyl groups on the starch and the expoxides.

Scheme 2.7. Reaction between starch and ethylene oxide or propylene oxide.

OH
O
O
OH O
O HO O
O Aqueous NaOH
+
O or O OH
HO OH O or
OH
O
O
O O
HO

OH

Ethylene oxide could etherify starch in granular form by pressurized hot air

through dried starch containing 10 wt % or less of moisture.58 The reaction rate

increased with temperature and with the concentration of absorbed ethylene oxide.

Hydroxylethylation of starch increases the water-binding capacity when compared with

the untreated starch.87 Hydroxyethyl starch is one of the most frequently used blood

plasma substitutes and is also used in oil drilling. Studies about the substitution pattern

39
of hydroxyethyl starch showed that most of the substituents (above 80%) were located on

the oxygen atom at 2-position (O-2), and the remaining hydroxyethyl groups were found

almost exclusively on the oxygen atom at 6-position (O-6).58,88 Only traces of

etherification at 3-position (O-3) took place.

The reaction of propylene oxide with starch under alkaline conditions is a

nucleophilic substitution, as Scheme 2.7 shows. The hydroxypropyl groups are

predominantly located at O-2 position on the glucose residues. The possible reason for

the higher reactivity of the oxygen atom at 2 position over the other two positions

(oxygen at 6 and 3) is the higher relative acidity of the O-2 due to its proximity to the

anomeric center.89 There are two ways of hydroxypropylation of starch, depending on

the desirable DS. When low DS (lower than 0.25) is desirable, starch is slurried in

aqueous NaOH.90 Propylene oxide is added based on the starch dry weight. The

preparation of hydroxypropyl starch with a high DS is industrially carried out in a

semidry state. Starch is first sprayed with aqueous NaOH until it contains 20% moisture.

Then the alkali starch is stirred in a closed reactor with gaseous propylene oxide of an

excess amount.89

Introduction of hydroxypropyl groups has affected the biodegradability of starch.

Research on the hydroxypropyl starch showed that the substitution of hydroxypropyl

groups decreased the extent of enzymatic hydrolysis of modified starch with increasing

DS.90

40
(c) Cationization

Cationic starches are important industrial derivatives in which starch is given a

positive ion by introducing ammonium, amino, and imino groups. Etherification by

different aminoalkyl agents, such as 2-diethylaminoethyl chloride, 2,3-(epoxypropyl)

trimethylammonuim chloride, (4-chlorobutene-2)-trimethylammonium chloride, etc., is a

major way of producing cationic starches.36 Schemes 2.8-2.11 show their structures

and the reactions with starch.

Scheme 2.8. Quaternary ammonium starch ester.

+ +
OH
H2C CHCH2N(CH3)3 Cl + NaCl
Cl CH2CH CH2N(CH3)3 Cl + NaOH
O

2,3-(epoxypropyl)trimethylammonuim chloride

+ +
H2C CHCH2N(CH3)3 Cl 50 oC
Starch OH + Starch O CH2CHCH2N(CH3)3 Cl
O Aqueous NaOH
OH

Quaternary ammonium starch ether

Scheme 2.9. Iminoalkyl starch ether.

R NH R
OH
Starch OH + N C N Starch O C N
R R

Iminoalkyl starch ether

41
Scheme 2.10. Aminoethylated starch ether.

H2C CH2
Starch OH + N CCl4 and benzene
Starch O CH2CH2NH2
H

Aminoethylated starch ether

Scheme 2.11. Tertiary aminoalkyl starch ether.

+
OH (C2H5)N CH2
ClCH2CH2N(CH2CH3)2 Cl
H 2C
2-diethylaminoethyl chloride

+
(C2H5)N CH2
Starch OH + Cl Starch O CH2CH2N(CH2CH3)2
H 2C

H +
O HCl
Starch CH2CH2N(CH2CH3)2 Starch O CH2CH2N(CH2CH3)2 Cl

Tertiary aminoalkyl starch ether

So far, all of the research on cationic starch is still concentrated on its

gelatinization behavior, pasting properties, and solubility. There are no reports about the

thermal or mechanical properties of cationic starch, except one report about cationic

starch that was used to modify clay as compatibilizer in the production of

nanocomposites (described later). Cationic starches are mainly used in the manufacture

of paper. They are employed as wet-end additives during the formation of paper to

improve the retention and drainage rate of the pulp and strength of the finished sheets.

42
2.3 Thermoplastic Starch (TPS)

Recently, there has been much interest in the research of thermoplastic starch

because of its biodegradability and bio-sustainable source. Both the glass transition

temperature (Tg) and the melting temperature (Tm) of natural, dry starch are higher than

its thermal decomposition temperature.91 The obstacle of using starch in plastics is its

brittle nature due to its relatively high Tg and lack of a sub-Tg main-chain relaxation.

Moreover, brittleness increases with time because of the free volume relaxation and

retrogradation.91 Thus, unlike other synthetic thermoplastic polymers, starch can only be

processed into a thermoplastic material by providing plasticization to starch molecules.

In the previous section, several basic chemical modifications of starch were

described. Various novel functional groups could be attached onto the starch through

such chemical modifications, depending on the desired properties. However, so far only

starch modified with long-chain fatty acid esters groups were systematically studied as a

thermoplastic material and showed some promising results (reduced Tg and Tm, increased

elongation at break and flexibility). Currently, most thermoplastic starches are produced

by the addition of low-molecular-weight plasticizers, such as water, glycerol, sorbitol,

urea, amide, sugars, quaternary amine, etc.15,23,92-94

Water, due to its strong affinity with hydroxyl groups on starch, is a very effective

plasticizer for thermoplastic starch. Compared with the preparation of gels or food

related products, the preparation of thermoplastic starch products require a relatively low

amount of water.95 In most cases, the water content is less than 20 wt % during

processing.15,96 Usually, a small amount of glycerol is used together with water.

Because glycerol enhances the boiling point of water, the migration and evaporation of

43
water inside the starch could be hindered.23,97 The thermoplastic properties of potato

starch with water and glycerol prepared by extrusion were studied.94 The samples

containing less than 9 wt % of water were glasslike and had an elastic modulus of 400

MPa.98 The viscoelastic behavior of thermoplastic starch highly depends on the water

content and typically shows three regions: a glassy region when the water content is

lower than 9 wt %, followed by a glass-to-rubber transition, then a rubbery region when

water is between 9 and 15 wt %, and finally, the viscous gel-like region with the water

content higher than 15 wt %.94 Figure 2.14 shows the stress-strain curve of TPS with

different amounts of water as a plasticizer.94

Figure 2.14. Stress-strain curves of potato starch with different amounts of water content
(the water contents (wt %) are indicated on the figures).94 (Reprinted with permission
from Elsevier.)

The structure of TPS plasticized by water was also studied.98 The extrudates of

TPS with water seemed to be translucent, suggesting that the granular structure of starch

44
was destroyed during extrusion. X-ray diffraction results proved the disappearance of the

native crystallinity of starch. Granular waxy maize starch showed A-type crystalline

structure on the X-ray diffraction curve. After being plasticized by water and extruded,

the X-ray diffraction results showed the amorphous morphology and a small amount of

B-type crystallinity. Thermal analysis indicated that the glass transition temperature of

the starch decreased appreciably with an increase in water content. The Tg of TPS (potato

starch) was decreased from 59 oC with 5 wt % of water content to –10 oC with 20 wt %

of water content.94 Other research on the starch/glycerol/water blends found that the Tg

of the TPS depends on the water contents almost linearly when the glycerol content is

fixed.99 Moreover, the amount of glycerol (higher than 14 wt %) only slightly affects the

Tg.

Since the thermal behavior and mechanical properties of TPS are very sensitive to

the water content, the major drawback of the TPS-based materials is the decline of the

material properties with time because of the migration and evaporation of the water

inside the starch, or the absorbing water by hydrophilic starch from the

atmosphere.93,95,98 Furthermore, studies also found the time-dependent post-

crystallization of starch in the presence of water and glycerol during storage. Figure 2.15

shows X-ray diffraction curves of TPS containing water stored at various periods and

under different humidity values.98 The problem of the decline in material properties with

time is at least in part due to the reorganization and crystallization of the amylose and

amylopectin molecules. Naturally, it is limited to the use of materials manufactured only

from TPS because of its poor water resistance. Thus, blending TPS with other polymers

would be helpful to enhance the overall water resistance.

45
Figure 2.15. X-ray diffraction patterns of (top) granular waxy maize starch and (bottom)
TPS materials stored at various temperatures and humidity. From bottom to top:
quenched and stored for 2 weeks; stored at 40% relative humidity (RH) for 2 weeks;
stored at 60% RH for 2 weeks; stored at 70% RH for 8 months; and stored at 90% RH for
2 weeks.98 (Reprinted with permission from John Wiley & Sons.)

46
2.4 Previous Research on the Nanocomposites based on Starch

The research on the starch-based nanocomposites began several years ago. The

most intensive research has been focused on organically modified silicates. Improved

properties, such as gas barrier, mechanical stiffness, and thermal stability, could be

achieved. Many reports have been published on starch-base nanocomposites.5,6,100-103

However, none of them achieved a high degree of exfoliation. Because both starch and

natural clay (montmorillnite, MMT) are hydrophilic, it was assumed that these

compounds would have high affinity with each other. But according to the X-ray

diffraction studies,6,104 it seems that mainly those low molecular-weight plasticizers such

as glycerol are intercalated, resulting in a low degree of exfoliation.102,105 Figure 2.16

shows X-ray diffraction (XRD) patterns, and Figure 2.17 shows TEM images of

TPS/MMT (the counter ion on the surface of layered silicates is Na+) nanocomposites

with various amounts of MMT in which TPS was plasticized by glycerol and water.103
o
The XRD patterns indicate that the peak around 7.85 of MMT shifted to

4.5o-5.0o regardless of the clay content, suggesting the intercalated nanocomposites.

One can observe agglomerates of MMT from the TEM images given in Figure 2.17.

Other research groups also showed similar results.5,104,106-109 So far, there is no

publication reported on exfoliated nanocomposites based on starch.

Table 2.3 lists the XRD results of various TPS/clay (treated and untreated)

nanocomposites. It is seen in Table 2.3 that only intercalation was achieved in TPS/clay

nanocomposites regardless of whether MMT or organoclay was used.

47
Figure 2.16. Effect of Cloisite Na+ on the XRD patterns for TPS nanocomposites: (a)
Cloisite Na+; (b) TPS/Cloisite Na+(97.5/2.5 wt %); (c) TPS/Cloisite Na+(95.0/5.0 wt %);
(d) TPS/Cloisite Na+(90.0/10.0 wt %); and (e) TPS.103 (Reprinted with permission from
Kluwer Academic Publishers.)

Figure 2.17. TEM image of TPS nanocomposite containing 5 wt % of MMT.103


(Reprinted with permission from Kluwer Academic Publishers.)

48
Table 2.3. XRD results of TPS/clay nanocomposites5,104,106-109

XRD peak XRD peak XRD peak position


position (2θ) of position (2θ) (2θ) layered
layered silicates layered silicates silicates in the
Sample code before treatment after treatment nanocomposites
TPS/Cloisite-Na+ 7.85 N/A 4.68
TPS/Cloisite-30B 7.85 4.73 4.30
TPS/MMT(treated by 8.75 7.06 4.24
ethanolamine)
TPS/MMT (treated by 5.975 5.590 5.015
chitosan)
TPS/MMT(treated by 7.3 3.6 3.6
cationic starch

Another way of improving the chemical affinity between starch and clay is to

modify the starch by the introduction of functional groups. Presently, there are very few

studies reported on such an approach. Two studies reported on chemically modified

starch-based nanocomposites.101,110 Oxidized starch by hydrogen peroxide and

acetylated starch were produced. However, neither of them exhibited good dispersion of

clay. Only a low degree of intercalation of layered silicates was observed by XRD, and

the experimental results can be explained as follows. (1) The amylose and amylopectin

in starch were stiff molecules, and such modifications could not effectively enhance the

mobility of molecules. It was very difficult for them to diffuse into the gallery of the

layered silicates exfoliating the aggregates of MMT. (2) Both modifications have mainly

introduced carbonyl or carboxyl groups into starch, which could have formed hydrogen

bonding, that apparently was not strong enough to exfoliate the aggregates of MMT.

Efforts on dispersion of clay in TPS nanocomposites are still underway, and

enhanced mechanical properties, gas barrier, and improved thermal stability have been

reported in various publications.5,6,103,104,107,108 Table 2.4 gives a summary of the

49
Table 2.4. Tensile properties of TPS/Cloisite 30B and TPS/MMT nanocomposites103

Blend composition Tensile strength Elongation at break


(MPa) (%)
TPS 97.5/MMT 2.5 wt % 2.79 48.9
TPS 95/MMT 5.0 wt % 3.32 57.2
TPS 90/MMT 10.0 wt% 3.20 52.0
TPS 97.5/Cloisite 30B 2.5 wt % 2.75 45.7
TPS 95/Cloisite 30B 5.0 wt % 2.80 44.5
TPS 90/Cloisite 30B 10.0 wt % 3.00 45.7
TPS 100 wt % 2.61 47.0

mechanical properties of TPS/MMT and TPS/Cloisite 30B nanocomposites.103 It is seen

that the tensile properties of TPS/MMT are increased with increasing content of MMT. It

should be noted that the TPS/MMT nanocomposite showed higher tensile strength than

the TPS/Cloisite 30B nanocomposite. As far as the dispersion of clay in TPS is

concerned, Cloisite 30B would be a poor example of modification of MMT. Because the

hydrophobic functional groups introduced into the Cloisite 30B decreased the chemical

affinity between Cloisite 30B and TPS, it resulted in lower intercalation of layered

silicates in comparison with TPS/MMT nanocomposites.

50
CHAPTER III

EXPERIMENTAL

3.1 Synthesis of Chemically Modified Starch

Starch employed in this study is a cornstarch purchased from Sigma-Aldrich

Chemical Company. Other chemicals used for the synthesis were purchased from Sigma-

Aldrich Chemical Company, and they were used as received without further purification.

Natural clay (montmorillonite, MMT) and two commercial organoclays (Cloisite 30B

and Cloisite 15A) used in the preparation of nanocomposites were provided by Southern

Clay Products. EVOH contained 27 mol % of ethylene units.

3.1.1 Synthesis of Anionically Modified Starch with Hexanoyl Group (AS-LF)

5 g of natural starch (NS) was added to a flask containing 150 ML isopropyl

alcohol. The reaction mixture was vigorously stirred at room temperature for 60 min

upon the addition of aqueous NaOH solution (10 ML, 5%). Sodium chloroacetate with a

predetermined amount (based on the desired DS) was then added into the reaction

mixture, and the temperature of the reactor was raised to 55 oC. The reaction was

stopped after 6 h by pouring the reaction mixture into a beaker containing ethanol. The

modified starch was precipitated from the ethanol. Then the product was filtered off.

The anionically modified starch (AS) was stirred in anhydrous ethanol for at least 24 h

and dried in a vacuum oven at 50 oC for 24 h.

51
5 g of dried AS was added to a flask containing 80 ML pyridine. The suspension

was vigorously stirred at 115 oC for 60 min. The reaction system was connected to a

condenser cooled by water. Hexanoyl chloride with a predetermined amount (based on

desired DS) was added dropwise into the reaction mixture, and the temperature of the

reactor was kept at 115 oC for 6 h. The reaction mixture was poured into a beaker

containing ethanol. The product (AS-LF) was precipitated from the ethanol and filtered

off. The product was stirred in anhydrous ethanol for at least 24 h and dried in a vacuum

oven at 60 oC for 24 h.

3.1.2 Synthesis of Anionically Modified Starch with Benzoyl Group (AS-B)

5 g of dried AS was added to a flask containing 80 ML pyridine. The suspension

was vigorously stirred at 115 oC for 60 min. The reaction system was connected with a

condenser cooled by water. Benzoyl chloride with a predetermined amount (based on

desired DS) was added dropwise into the reaction mixture, and the temperature of the

reaction was kept at 115 oC for 6 h. The reaction mixture was poured into a beaker

containing ethanol. The product (AS-B) was precipitated from the ethanol and filtered

off. The product was stirred in anhydrous ethanol for at least 24 h and dried in a vacuum

oven at 100 oC for 12 h.

3.1.3 Synthesis of Cationically Modified Starch with Hexanoyl Group (CS-LF)

5 g of natural starch was added to a flask containing 100 ML aqueous NaOH

solution (5%). The reaction mixture was vigorously stirred at room temperature for 60

min. Then glycidyltrimethylammonium chloride with a predetermined amount (based on

52
the desired DS) was added dropwise into the reaction mixture, and the temperature of the

reaction was raised to 60 oC. The reaction was stopped after 6 h by pouring the reaction

mixture into a beaker containing ethanol. The product (CS) was precipitated from the

ethanol and filtered off. The CS was stirred in anhydrous ethanol for at least 24 h and

dried in a vacuum oven at 50 oC for at least 24 h.

5 g of CS was added to a flask containing 80 ML pyridine. The suspension was

vigorously stirred at 115 oC for 60 min. The reaction system was connected with a

condenser cooled by water. Hexanoyl chloride with a predetermined amount (based on

desired DS) was added dropwise into the reaction mixture, and the temperature of the

reaction was kept at 115 oC for 6 h. The reaction mixture was poured into a beaker

containing ethanol. The product (CS-LF) was precipitated from the ethanol and filtered

off. The product was stirred in anhydrous ethanol for at least 24 h and dried in a vacuum

oven at 60 oC for 24 h.

3.1.4 Synthesis of Cationically Modified Starch with Benzoyl Group (CS-B)

5 g of CS was added to a flask containing 80 ML pyridine. The suspension was

vigorously stirred at 115 oC for 60 min. The reaction system was connected with a

condenser cooled by water. Benzoyl chloride with a predetermined amount (based on

desired DS) was added dropwise into the reaction mixture, and the temperature of the

reaction was kept at 115 oC for 6 h. The reaction mixture was poured into a beaker

containing ethanol. The product (CS-B) was precipitated from the ethanol and filtered

off. The product was stirred in anhydrous ethanol for 24 h and dried in a vacuum oven at

100 oC for 12 h.

53
3.2 Preparation of Blends and Nanocomposites based on Modified Starch and

EVOH

Solution Blending of Blends of Modified Starch and EVOH A predetermined amount

of modified starch and EVOH were dissolved into dimethyl sulfoxide(DMSO) at a mildly

elevated temperature (around 60 oC). A magnetic stirrer was used to facilitate the

dissolution. A transparent and yellowish solution was obtained after total dissolution of

polymers. The solution was put into a hood, allowing the evaporation of the solvent.

Then the mixture was put into a vacuum oven at 100 oC until no weight changes were

detected.

Solution Blending of Nanocomposites based on Modified Starch A predetermined

amount of modified starch was dissolved in tetrahydrofuran (THF) upon heating. 5 wt %

of organoclay (or MMT) was pre-dispersed in THF (a co-solvent THF/water, 10/1 v/v

was used to disperse MMT) by ultrasonic bath for 30 min, and then it was added slowly,

while stirring, into the polymer solution. The mixture of modified starch and organoclay

was stirred constantly for 2 days at an ambient temperature, allowing the evaporation of

the solvent. Then the mixture were put into a vacuum oven at 100 oC until no weight

changes were detected.

Melt Blending of Blends and Nanocomposites Melt blending was performed by a

twin-screw mini compounder (Haake Instruments). 5.5 g of polymer mixture, including a

modified starch, EVOH, and 5 wt % of nanoclay, were fed to the compounder, which had

been preheated to 185 oC. The compounder was run in a closed-loop mode, and the melt

54
blending was performed by the twin-screw extruding in the compounder before discharge

into a circular die. An external nitrogen supply prevented possible oxidation. For other

characterization purposes, the nanocomposites were hot-compression molded when

needed.

3.3 Characterization Methods

Proton Nuclear Magnetic Resonance (1H-NMR) Spectroscopy 1H-NMR spectra were

obtained with a Varian Gemini 300 MHz NMR spectrometer. Solid samples were

solubilised in deuterated dimethyl sulfoxide (DMSO-d6) and placed in 5 mm NMR tube.

The spectra were collected at 25 oC over a spectra width of 4796 Hz using a 90o pulse.

The acquisition time was 1.98 s, the relaxation delay (d1) was 5 s (the relaxation time (T1)

of anomeric proton was measured to be 0.9 s), and 16 scans were accumulated. Data

were processed with 0.5 Hz exponential line broadening and zero filled to 131,072 points

before Fourier transformation.

Differential Scanning Calorimetry (DSC) The thermal transition temperatures of a

modified starch, EVOH, blends, and nanocomposites prepared in the present research

were determined by differential scanning calorimetry (DSC) (TA instruments). All DSC

runs were made under a nitrogen atmosphere with heating rates of 10 oC/min. Indium

and zinc were used to calibrate the temperature and the heat of fusion.

Optical Microscopy The phase separation of blends of modified starch and EVOH was

observed by an optical microscope equipped with a hot stage (Instec) and a digital camera

55
(Spot insight 2, Diagnostic Instrument). Specimens were cast from 1 wt % solution of

polymer mixtures on a slide glass to obtain a film of about 2-3 μm in thickness, which

was then first dried in a fume hood and subsequently in a vacuum oven. The heating rates

employed were 0.5 oC/min. Images of single phase and separated phase were obtained by

digital camera. At each weight fraction of EVOH, at least 5 samples were measured for

determining the cloud point of the mixture, and an average value was obtained from the

repeated measurements.

Fourier Transform Infrared (FTIR) Spectroscopy Using a Fourier transform infrared

spectrometer (Perkin Elmer), FTIR spectra were obtained at room temperature for

modified starches, blends of a modified starch and EVOH, and nanocomposites based on

a modified starch with mode of attenuated total reflectance (ATR). Spectral resolution

was maintained at 4 cm−1. Specimens were grinded into fine powders and slowly dried

for 24 h in a fume hood until the solvent and moisture evaporated, and they were then

stored in a vacuum oven until use.

Wide Angle X-ray Diffraction (WAXD) WAXD experiments were conducted at room

temperature on the films, which were hot pressed by compression molding, using a

General Electric X-ray generator (Model XRD-6) operated at 30 kV and 30 mA (Ni-

filtered Cu Kα radiation). For annealed specimens, the films were first heated at 200 oC

for 5 min and then immediately transferred to a vacuum oven in which the temperature

was regulated at an annealing temperature. The samples were annealed for at least 3 h

and quenched by liquid nitrogen before the WAXD measurements began.


56
X-ray Diffraction (XRD) Using a Rigaku X-ray generator operated at 40 kV and 150

mA, X-ray diffraction patterns were obtained to determine the mean interlayer spacing of

the (001) plane (d001) for the natural clay or organoclays, and their nanocomposites. The

X-ray beam was monochromatized to Cu Kα with a graphite crystal. The range of 2θ

scanning of X-ray intensity employed was 1.5–10 degrees.

Transmission Electron Microscopy (TEM) TEM images of specimens were taken at

room temperature. The ultrathin sectioning (around 100 nm) was performed by

ultramicrotomy at room temperature for the clay (or organoclay) nanocomposites. A

transmission electron microscope (JEM1200EX 11, JEOL) operated at 120 kV was used

to obtain images of nanocomposite specimens.

Thermogravometric Analysis (TGA) The thermal degradation temperature of

polymers was determined by a thermogravometric analyzer (TGA 2050, TA Instrument).

All TGA data was obtained under a nitrongen atmosphere at a heating rate of 10 oC/min,

and the thermal degradation temperature was determined at the point of 90 wt % of the

original weight.

Tensile Test The tensile test of the blends and nanocomposites based on a modified

starch was performed by a tensile tester (Instron, Model 5567), following the ASTM

standard, D638-00. The load cell used for the tensile test was an Instron static load cell

with a 10 kN capacity. A cross-head speed of 5 mm/min was used for the test. The
57
upper yield stress on the stress-strain curve was considered as the tensile strength. For all

the results reported in this dissertation, at least 3 replicates of specimens were tested, and

an average value of the measured values was used. All specimens for tensile testing were

compounded by a twin-screw mini compounder (Haake Instruments) and hot compressed

into films with uniform thickness. Then the films were cut into dumbbell shape for the

tensile test. The thickness and length of the specimens were accurately measured before

use.

Hydrophilicity Test Saturated aqueous sodium chloride solution was prepared, and it

was put into a sealable chamber. The ambient temperature of this chamber was kept

around 20 °C. In this way, the constant relative humidity (74%) was obtained in the

chamber. Dried, chemically modified starch samples were put into the chamber after

being accurately weighed. Then the samples were weighed every two days until the

weight did not change further.

Biodegradability Test Materials that were to be characterized for their biodegradability

include NS (the positive control), modified starches employed in the present study, and

LDPE (the negative control). Since dry natural starch is not processable, it was premixed

with 10 wt % of glycerol (acting as plasticizer) by mini twin-screw compounder. All

materials were molded between two Teflon sheets into thin films (25×50×1 mm ) by

applying a 250 kg/cm2 pressure using a preheated hydraulic press, and accurately

weighted by digital balance. Every sample had 10 duplicates with the same size and

weight. The compost was offered by KB Compost Services, Inc. The composition of the

compost was (approximate dry weight): 5% animal dung; 10% sawdust; 10% shredded
58
paper; 10% food waste; 20% shredded leaves; and 45% garden soil. The compost was

fully shredded and sifted to remove large clumps. The moisture content of the compost

was maintained by periodic spraying of water. To avoid anaerobic condition, the

compost was aerated by turning at least 3 times a week. All samples were buried inside

the compost at a depth of at least 1.5 feet. One duplicate of each sample was dug out

every two weeks and weighed after thorough washing with water and drying in a vacuum

oven until it reached a constant weight. The weight loss (Wloss) was calculated by:

Winitial − Wweighted
Wloss = × 100%
Winitial

59
CHAPTER IV

CHARACTERIZATION OF CHEMICALLY MODIFIED STARCH

4.1 Structure Characterization and Determination of Degree of Substitution (DS)

of Chemically Modified Starch

As described in Chapter II, there are three hydroxyl groups in each repeat unit of

starch molecules. Chemical modification explored in the present study is to substitute

these hydroxyl groups with functional groups, such as anionic groups, cationic groups,

hexanoyl groups, or benzoyl groups. The rationale behind this modification was as

follows. (1) Because of ionic interaction between the modified starch and clay, starch

having anionic or cationic groups can have higher compatibility with organoclay or

natural clay. (2) Because the granular semicrystalline structure of natural starch was

destroyed by the interruption of functional groups, starch having hexanoyl groups or

benzoyl groups can have higher molecular mobility than natural starch. (3) Because the

hexanoyl groups or benzoyl groups are hydrophobic, starch having hexanoyl groups or

benzoyl groups can have lower affinity with water than natural starch.

Figures 4.1 and 4.2 give the FTIR spectra of the modified starches synthesized in

the present study. Following the FTIR spectra, Table 4.1 lists the major absorbance

bands shown in the FTIR spectra for the functional groups in the modified starches.

The FTIR spectra of natural starch (NS) show a very broad and strong band at

3263 cm−1, suggesting strong hydrogen bonds in natural starch. These hydrogen bonds

60
Absorbance (arb. unit)

(1)

(a) (2)
(1)
(b)

(c)

(2) (1)

4000 3000 2000 1000


−1
Wavenumber (cm )
Figure 4.1. FTIR spectra of (a) natural starch (NS), (b) starch modified with anionic
group and benzoyl group (AS-B), and (c) starch modified with anionic group and
hexanoyl group (AS-LF). (Arrow (1) points to the absorbance of the hydrogen-bonded
−OH group, and arrow (2) points to the absorbance of the free −OH group.)

61
Absorbance (arb. unit)

(1)

(a)
(2) (1)
(b)
(c)
(2) (1)

4000 3000 2000 1000


−1
Wavenumber (cm )
Figure 4.2. FTIR spectra of (a) natural starch (NS), (b) starch modified with cationic
group and benzoyl group (CS-B), and (c) starch modified with cationic group and
hexanoyl group (CS-LF). (Arrow (1) points to the absorbance of the hydrogen-bonded
−OH group, and arrow (2) points to the absorbance of the free −OH group.)

62
Table 4.1. Peak assignments in the FTIR spectra of the chemically modified starches

Bond Peak

hydrogen-bonded v (O−H) 3263 cm−1 in NS, 3410 cm−1 in AS-B, 3456 cm−1 in
AS-LF, 3401 cm−1 in CS-B, and 3422 cm−1 in CS-LF

free v (O−H) 3566 cm−1 in AS-B, 3589 cm−1 in AS-LF, 3559 cm−1
in CS-B, and 3577 cm−1 in CS-LF

v (C=O) 1737 cm−1 in AS-B and CS-B, 1743 cm−1 in AS-LF


and CS-LF

v (COO−Na+) 1595 cm−1 in AS-B and AS-LF

v (C−H) in CH2 or CH3 2854 cm−1 and 2956 cm−1 in AS-LF and CS-LF

δ (C−H) in −N+(CH3)3 1470 cm−1 in CS-B and CS-LF

v (C−N+) in −N+(CH3)3 957 cm−1, 911 cm−1, and 872 cm−1 in CS-B and
CS-LF

are present in the granular structure of natural starch, in both the amorphous and

crystalline phases. As described in Chapter II, the mobility of starch molecular chains is

very low (its glass transition temperature is higher than its thermal degradation

temperature), which was partly resulting from the strong hydrogen bonds present among

molecules. (The other reasons could include high molecular weight and microscopically

granular structure of natural starch.)

The FTIR spectra of the modified starches synthesized in the present study include

a starch modified with anionic group and benzoyl group (AS-B), a starch modified with

anionic group and hexanoyl group (AS-LF), a starch modified with cationic group and

benzoyl group (CS-B), and a starch modified with cationic group and hexanoyl group

(CS-LF). They show that the absorbance band of hydrogen-bonded −OH group was

63
lowered and shifted to a higher wavenumber, suggesting that the hydrogen bonds in a

modified starch were drastically decreased as compared with natural starch. Furthermore,

the absorbance band of free −OH group was observed in the FTIR spectra of a modified

starch. The substitution of hydroxyl groups in a modified starch by other functional

groups decreased the total number of hydroxyl group so that the strong hydrogen bonds

were weakened. Additionally, the presence of functional groups, such as hexanoyl

groups or benzoyl groups, interfere with the hydrogen bonds present between the

available hydroxyl groups. That is the reason why we can observe the absorbance peaks

of free −OH groups in a modified starch.

Qualitatively, we have shown above that chemical modification decreased the total

number of hydroxyl groups in the starch molecules. The hydrogen bonds present in

starch molecules were decreased, upon chemical modification, because of the lowered

number of available hydroxyl groups. However, the number of hydroxyl groups

substituted by functional groups and the number of hydroxyl groups available in a starch

should be determined quantitatively. The values of degree of substitution (DS) indicate

that the number of hydroxyl groups substituted in each repeat unit of a starch molecule

ranges from 0 to 3. The DS of modified starches were determined from the 1H-NMR

spectra of a modified starch by equations 4.1 to 4.4:

DSanionic = Amethylene / 2Aanomeric (4.1)

DScationic = Amethyl/ 9Aanomeric (4.2)

DShexanoyl = Amethyl / 3Aanomeric (4.3)

DSbenzoyl = Abenzoyl / 5Aanomeric (4.4)

64
in which A stands for the integrated area of chemical shift found from the 1H-NMR

spectra for specific functional groups. Figure 4.3 gives the 1H-NMR spectra, showing

how the DS of hexanoyl group was determined. The peak locating at 5.1 ppm is the

chemical shift of anomeric proton in the glycosidic ring.111 The peaks locating around

1.1 ppm are the chemical shift of methyl protons in hexanoyl group. The numbers under

both of those two peaks denote the integrated peak area (Aanomeric and Amethyl). The value

of DS for all modified starches were determined in the same way. Table 4.2 lists the

values of DS for the four chemically modified starches.

H OH
H
a
O O H
H H b
Intensity (arb. unit)

H O b
H
OH O
C Hb

O Hb

1.0 3.67

10 9 8 7 6 5 4 3 2 1 0

Chemical shift (ppm)

Figure 4.3. 1H-NMR spectra of a starch modified by hexanoyl group.

65
Table 4.2. Degree of substitution of chemically modified starches

Sample code DSa1 DSa2

AS-B 0.03 1.3


AS-LF 0.03 1.2
CS-B 0.1 1.7
CS-LF 0.03 1.6

a DS
1 is the degree of substitution of ionic group in modified starch, and DS2 is the degree
of substitution of hexanoyl groups or benzoyl groups in modified starch.

4.2 Moisture Resistance of Chemically Modified Starch

Figure 4.4 shows how much moisture was absorbed by a modified starch over time

in an environment having a constant humidity (75%). It is seen that natural starch and

60

50
Absorbed Moisture (wt %)

40

30

20

10

0 2 4 6 8 10 12 14 16 18 20 22 24
Period (day)

Figure 4.4. Moisture absorption of modified starches: (●) NS; (■) AS; (▲) CS;
(▼) AS-LF; (◆) AS-B; (□) CS-LF; and (〇) CS-B.

66
the starches modified with ionic groups (AS or CS) are very hydrophilic. Their moisture

absorption could reach up to an equilibrium value of 50 wt %. Natural starch has a strong

affinity with water due to its abundant hydroxyl groups. As we described in Chapter II,

the mechanical properties of natural starch-based products keep changing because the

product absorbs water from its ambient environment due to a strong affinity with water.

Furthermore, its high water content limits the use of starch as a conventional

thermoplastic material. Figure 4.4 shows that the starches modified with ionic groups

(AS or CS) have even higher moisture absorption than natural starch, suggesting that both

AS and CS are more hydrophilic than natural starch. It should be noted in Table 4.2 that

the values of DS of AS and CS are around 0.1, which means that only a very small

amount of hydroxyl groups in starch was substituted by ionic groups. Considering that

the ionic groups are more hydrophilic than hydroxyl groups, it is understandable that the

starch modified with ionic groups have a higher affinity with water than natural starch.

Besides moisture absorption, we have found that both AS and CS are soluble in cold

water, while natural starch is not, which also suggests a higher affinity of AS and CS with

water.

On the other hand, Figure 4.4 shows that the starch modified with hexanoyl groups

or benzoyl groups have very low moisture absorption. We can see that starch was

changed from hydrophilic to hydrophobic material through the chemical modification in

the present study. The substitution of functional groups decreased the number of

hydroxyl groups. It was well known that hydroxyl groups have a strong affinity with

water. Additionally, the hydrophobic substituents limit the access of available hydroxyl

group to moisture.

67
4.3 Thermal Behavior of Chemically Modified Starch

Thermal Stability of Chemically Modified Starch

The thermal stability of chemically modified starch was characterized by

thermogravimetric analysis (TGA) in a nitrogen atmosphere. Figure 4.5 shows how the

thermal degradation temperature of a starch was changed by chemical modification. The

thermal degradation temperature is defined by the temperature at which the weight loss of

a sample reaches 10 wt % in the weight loss versus temperature curves obtained from

TGA measurements.
Thermal Degradation Temperature (oC)

400

NS NS
AS-LF AS-B CS-LF
CS-B
300

CS
200 AS

100

Starch samples

Figure 4.5. Thermal degradation temperatures of chemically modified starches.

It is worth noting that the thermal degradation temperature of starch greatly

decreased from 330 oC (natural starch) to around 200 oC after modification with ionic

groups (AS or CS). Previous studies73 on the thermal stability of starch showed that the

68
thermal weight loss of starch started with the intra- and intermolecular dehydration

among hydroxyl groups and then was followed by decomposition and depolymerization

at a higher temperature. In both AS and CS, only a very small amount of hydroxyl

groups were substituted by ionic groups. The intra- and intermolecular dehydration

among hydroxyl groups would not be hindered by the substitution of ionic groups.

Moreover, the substitution of ionic groups destroyed the granular structure of natural

starch and made abundant hydroxyl groups more easily accessible to each other for the

dehydration.

On the other hand, Figure 4.5 shows that the thermal degradation temperatures of

the starches modified with hexanoyl groups or benzoyl groups (AS-B, AS-LF, CS-B, or

CS-LF) were almost not changed. In these modified starches, the DS of hexanoyl groups

or benzoyl groups is around 2, suggesting that out of three hydroxyl groups on each

repeat unit, two hydroxyl groups were substituted. The decreased total number of

hydroxyl groups in the modified starch has lowered the possibility of having dehydration

among the available hydroxyl groups. Furthermore, the hexanoyl groups or the benzoyl

groups interfered with the accessibility of hydroxyl groups.

Glass Transition Temperature

As described in the previous chapter, the glass transition temperature of dried

natural starch cannot be practically measured, because it is higher than the thermal

decomposition temperature. It is necessary to lower the glass transition temperature of a

starch below its thermal decomposition temperature so that we can possibly process

starch as a conventional thermoplastic material. Thermoplastic starch can be prepared in

two ways: one is by the addition of low-molecular-weight plasticizer, and the other is by

69
chemical modification. In this dissertation, side groups such as hexanoyl groups and

benzoyl groups were attached into a starch through chemical reaction with hydroxyl

groups. Figures 4.6 and 4.7 give the DSC thermograms of the modified starches

synthesized in this study. For starches modified only with ionic groups (AS or CS), there

was no discernible glass transition temperature found in the DSC thermograms. For

starches modified with multiple functional groups (AS-B, AS-LF, CS-B, or CS-LF), we

can clearly observe the glass transition temperature in the DSC thermograms. The

introduction of hexanoyl groups or benzoyl groups into a starch destroyed the

semicrystalline granular structure and the hydrogen bonds formed between the hydroxyl

groups. Hence, the mobility of starch was enhanced. In addition, the introduction of

functional groups into a starch have increased the free volume of the modified starch.

Thus, the glass transition temperature of a starch was lowered and it became discernible

in the DSC measurements. On the other hand, for a starch modified only with ionic

groups (AS or CS), the DS of ionic groups was around 0.1. Only a very small amount of

hydroxyl groups was substituted by the ionic groups. There still are very strong

hydrogen bonds in both AS and CS, which limit the mobility of the starch molecules.

Thus, the glass transition temperature of AS or CS was not discernible in the DSC

thermograms.

70
90 (c)

45 (b)
Endotherm

(a)

0 50 100 150 200


Temperature (oC)

Figure 4.6. DSC thermograms of (a) AS, (b) AS-LF, and (c) AS-B.

(c)
44

(b)
Endotherm

92

(a)

0 50 100 150 200 250


Temperature (oC)
Figure 4.7. DSC thermograms of (a) CS, (b) CS-B, and (c) CS-LF.

71
It should be noted that there was no melting peak on the DSC thermograms of

modified starches, indicating an amorphous morphology in the modified starches. The

modification by functional groups destroyed the semicrystalline granular structure of

natural starch. Furthermore, the bulky side chains in the starch molecules after

modification decreased the irregularity of starch molecules so that it was hard for

modified starch to crystallize.

4.4 Biodegradability of Chemically Modified Starch

Biodegradable material can be defined as a material that has a proven capability to

decompose in a specific environment within a period through natural biological processes

into non-toxic carbonaceous soil, water, carbon dioxide, or methane. Apparently, natural

starch is a biodegradable material. In this dissertation, natural starch has been chemically

modified by various functional groups, including ionic groups, hexanoyl groups, and

benzoyl groups. Whether or not these chemically modified starches are still

biodegradable crucially dictates their potential applications. Biodegradation represents a

complex phenomenon brought about by the actions of naturally occurring

microorganisms, such as bacteria, fungi, and algae. As biodegradation proceeds, it

produces carbon dioxide, water, and/or methane. Biodegradability can be characterized

by monitoring the change of such properties as thickness, transparency, weight, tensile

strength, and molecular weight/distribution of molecular weight of a material or the

production of by-products (e.g., carbon dioxide and methane) during the biodegradation.

It must be noted that a specific time frame should be added to any characterization of

biodegradability.

72
In this dissertation, composting environment was built for the characterization of

biodegradability of the modified starches. Composting is one of the most important

waste management strategies, and natural starch is readily biodegradable in a composting

environment. The weight loss of samples during composting will indicate biodegradation.

Thus, the biodegradability of natural starch and the modified starches can be

quantitatively evaluated by measuring the weight loss of samples against time. Figure

4.8 shows the weight loss of the modified starches during composting against time. For

natural starch (the positive control in our research), we can see from Figure 4.8 that it

degraded very fast in composting. After 6 weeks, natural starch completely degraded.

For LDPE (the negative control in our research), we cannot find any weight loss during

the whole testing period, suggesting no degradation during composting.

100

80
Weight loss (%)

60

40

20

0 2 4 6 8 10 12 14 16 18
Composting period (week)

Figure 4.8. Weight loss of the modified starch during composting: (●) NS; (■) CS-LF;
(▲) LDPE; (▼) CS-B; (◆) AS-LF; and (□) AS-B.
73
For the modified starches synthesized in the present study, we can clearly find

weight loss during composting, suggesting biodegradability of a modified starch. In

Figure 4.8, we can observe that the degradation of natural starch is faster than the

modified starches, suggesting that the biodegradability of natural starch is higher than

modified starches. It is interesting to observe that the modified starches having cationic

functional groups (CS-LF and CS-B) have different degradation patterns from the

modified starches having anionic functional groups (AS-LF and AS-B).

4.5 Concluding Remarks

A series of chemically modified starches has been synthesized in the present study.

The hydroxyl groups in starch molecules were substituted by ionic groups (acetate

sodium or trimethylammonium chloride) and hydrophobic groups (hexanoyl group or

benzoyl group). The FTIR spectra showed that the hydrogen bonds present in natural

starch were weakened by the chemical modification. Through the calculation from the
1
H-NMR spectra, the degree of substitution of the modified starch prepared was

determined.

The modified starches, including AS-B, AS-LF, CS-B, and CS-LF, were

characterized by DSC. The DSC thermograms showed that the chemical modification

has lowered the glass transition temperature of starch. Additionally, there was no melting

peak in the DSC thermograms of modified starches, suggesting an amorphous

morphology in the modified starches.

The thermal stability of the modified starches was characterized by TGA. When

starch was modified only with ionic groups, the thermal degradation temperature of

74
starch was largely decreased from over 300 oC to around 200 oC. On the other hand, the

starches modified with both ionic groups and hydrophobic groups have a higher thermal

degradation temperature than the starch modified only with ionic groups.

Natural starch is known to be very hydrophilic, which limits its use in conventional

thermoplastic processing. The substitution of hydrophobic groups (hexanoyl groups or

benzoyl groups) decreased the hydrophilicity of starch. We characterized the moisture

absorbance of the modified starches in an environment having constant humidity over

time. The results showed that the chemical modification has changed the starch from

hydrophilic to hydrophobic.

We also investigated whether or not the biodegradability of a starch was influenced

by the chemical modification. The weight loss of a modified starch during composting in

a given period clearly has shown biodegradability of the modified starch. However, the

weight loss of a modified starch is slower than natural starch, suggesting that the

biodegradability of modified starches is lower than natural starch.

75
CHAPTER V

BLENDS OF MODIFIED STARCH AND EVOH

5.1 Miscibility of Blends Based on Modified Starch and EVOH Characterized by

Phase Diagrams

In this dissertation, blends of a chemically modified starch (including AS-LF, AS-B,

CS-LF, or CS-B) and EVOH were prepared. There is plenty of literature reporting on the

mechanical properties, biodegradability, processability, thermal behavior of blends of

natural starch and EVOH. However, few of them discussed the phase behavior of blends

of starch, and EVOH. So far, there is no literature reporting on the blends of chemically

modified starch and EVOH. The purpose of this study was to prepare miscible blends of

a modified starch and EVOH so that exfoliated organoclay nanocomposites could be

prepared. Moreover, miscible blends can effectively improve their mechanical properties

of blend. Using an optical microscope equipped with a programmable heating system,

the phase behavior of a blend can be observed at a specific temperature and composition.

Figure 5.1 gives phase diagrams of blends based on a modified starch and EVOH.

It is seen that all blend specimens exhibit a lower critical solution temperature (LCST).

The blends containing a starch modified by hexanoyl groups (AS-LF or CS-LF) and a

starch modified by benzoyl groups (AS-B or CS-B) have a critical temperature at the

same blend composition (0.6 weight fraction of EVOH). The former (AS-LF or CS-LF)

has a higher critical temperature than the latter (AS-B or CS-B).

76
Temperature (oC) 300

200

100

0
0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction)

Figure 5.1. Phase diagrams of the blends of a modified starch and EVOH: (●) (AS-B)/
EVOH blend; (〇) (AS-LF)/EVOH blend; (▼) (CS-B)/EVOH blend; and (▽) (CS-LF)/
EVOH blend.

5.2 Hydrogen Bonding in Blends based on Modified Starch and EVOH

Characterized by DSC and FTIR

The blends based on a modified starch and EVOH were prepared by melt blending.

The modified starches include AS-B, AS-LF, CS-B, and CS-LF, the thermal behavior and

the biodegradability of which were summarized in Chapter IV. EVOH used in the

present study contains 27 mol % ethylene.

It should be noted that EVOH is a semicrystalline polymer. Previous results112,113

showed that the vinyl alcohol content in the crystalline phase is essentially equal to that

77
in the bulk sample. EVOH employed in the present study contains 73 mol % of vinyl

alcohol, suggesting that most of the crystalline phase was contributed by the vinyl alcohol

component. Because the vinyl alcohol groups in EVOH can form hydrogen bonds with

the hydroxyl groups in a modified starch, we might expect to find that the crystallization

of EVOH will be interrupted by the presence of the starch in the blend of EVOH and the

modified starches.

Figures 5.2 to 5.5 give the DSC thermograms of the blends of EVOH and a

modified starch for different compositions. Several common features in the DSC

thermograms are worth noting as follows. (1) Endothermic peaks were found on every

DSC thermogram of the blends of EVOH and a modified starch, which indicate that the

crystalline phase was present in the blends. Because the modified starch was an

amorphous material, we ascribed the crystallization in the blends to the presence of

EVOH. (2) The melting point of the crystalline phase in the blends decreases with an

increasing amount of modified starch. (3) The melting enthalpy of the crystalline phase

in the blends decreases with increasing amount of modified starch. The melting point

depression in the blends suggests that the crystallization of EVOH was interrupted. The

imperfection of the crystalline phase in the blend was increased because of the

interruption induced by the modified starch. We ascribe this interruption to the formation

of hydrogen bonds between the hydroxyl groups in a modified starch and the vinyl

alcohol groups of EVOH. On the other hand, the decreased melting enthalpy of the

crystalline phase with increasing amount of modified starch in a blend, shown in Figures

5.2 to 5.5, clearly indicates that the crystallinity of the blend is decreased as the weight

fraction of modified starch is increased.

78
(a)
Endotherm (b)
(c)
(d)

(e)

0 50 100 150 200


o
Temperature ( C)
Figure 5.2. DSC thermograms of the blends of (AS-B)/EVOH: (a) neat EVOH; (b) blend
of 20/80 (weight ratio) (AS-B)/EVOH; (c) blend of 40/60 (AS-B)/EVOH; (d) blend of
70/30 (AS-B)/EVOH; and (e) neat AS-B.

(a)
Endotherm

(b)
(c)
(d)
(e)

0 50 100 150 200


o
Temperature ( C)
Figure 5.3. DSC thermograms of the blends of (AS-LF)/EVOH: (a) neat EVOH; (b)
blend of 20/80 (AS-LF)/EVOH; (c) blend of 40/60 (AS-LF)/EVOH; (d) blend of 70/30
(AS-LF)/EVOH; and (e) neat AS-LF.

79
(a)
(b)
Endotherm
(c)
(d)
(e)

0 50 100 150 200


Temperature (o C)
Figure 5.4. DSC thermograms of the blends of (CS-B)/EVOH: (a) neat EVOH; (b) blend
of 20/80 (CS-B)/EVOH; (c) blend of 40/60 (CS-B)/EVOH; (d) blend of 70/30 (CS-B)/
EVOH; and (e) neat CS-B.

(a)
Endotherm

(b)
(c)
(d)
(e)

0 50 100 150 200


Temperature (oC)

Figure 5.5. DSC thermograms of the blends of (CS-LF)/EVOH: (a) neat EVOH; (b)
blend of 20/80 (CS-LF)/EVOH; (c) blend of 40/60 (CS-LF)/EVOH; (d) blend of 70/30
(CS-LF)/EVOH; and (e) neat CS-LF.
80
The values of melting point Tm and the degree of crystallinity of the blends are

summarized in Table 5.1. Two points should be addressed in the calculation of the

degree of crystallinity in the blends. First, since we calculated the degree of crystallinity

of a blend using melting enthalpy determined from the DSC thermograms, we had to

determine the melting enthalpy of neat EVOH ( ΔH m0 ). However, since ΔH m0 could not

be determined experimentally, we estimated ΔH m0 from the enthalpies of completely

crystallized poly(vinyl alcohol) (PVA), ΔH PVA


0
(156.14 J/g)114, and polyethylene (PE),

ΔH PE
0
(290 J/g)115, by

ΔH m0 = αΔH PVA
0
+ βΔH PE
0
(5.1)

where α and β are the weight fractions of the vinyl alcohol component and the ethylene

Table 5.1. Melting points and the degree of crystallinity of blends of EVOH and
modified starch

Weight ratio Melting Point Tm Degree of crystallinity


(oC) (wt %)
EVOH n/a 184 58.3
AS-B/EVOH 2/8 179 49.7
4/6 172 40.1
7/3 162 23.2
AS-LF/EVOH 2/8 180 45.2
4/6 174 36.8
7/3 166 20.7
CS-B/EVOH 2/8 178 50.2
4/6 173 41.3
7/3 164 24.3
CS-LF/EVOH 2/8 177 43.2
4/6 172 32.7
7/3 163 21.6

81
component in EVOH, respectively. In this study EVOH had 73 mol % (α = 0.81) of

vinyl alcohol and 27 mol % (β = 0.19) of ethylene. Thus, we have calculated the value of

ΔH m0 to be 181.6 J/g. The second point that should be addressed is that the degree of

crystallinity calculated is the percentage of the crystalline phase in a crystallizable

material (EVOH) instead of the entire sample (a blend of a modified starch and EVOH).

Thus, the degree of crystallinity (%) of a blend was calculated from:

C = 100ΔH m (ΔH m0 WEVOH ) (5.2)

where ΔH m is the melting enthalpy of the crystalline phase in a blend determined from

the DSC thermograms, and WEVOH is the weight fraction of EVOH in a blend sample.

The increasing imperfection of the crystalline phase (indicated by melting point

depression) and the decreasing degree of crystallinity (indicated by decreasing melting

enthalpy) in a blend with increasing weight fraction of a modified starch suggest the

presence of strong interactions between the modified starch and EVOH. We ascribe this

interaction to the formation of hydrogen bonds between the two components in a blend.

This strong interaction not only interrupts the crystallization of EVOH in the blend, but

also enhances the miscibility of the blend. It is seen in the DSC thermograms that only

one glass transition temperature is observed on each DSC thermogram over the entire

composition of a blend.

Figures 5.6 to 5.9 give the DSC thermograms in the glass transition region for all

blends. It should be noted in Figures 5.6 to 5.9 that the vertical bar denotes glass

82
(a)

(b)

Endotherm (c)

(d)

(e)

0 50 100 150
Temperature (o C)
Figure 5.6. DSC thermograms in the glass transition region for the blends of (AS-B)/
EVOH: (a) neat EVOH; (b) blend of 20/80 (AS-B)/EVOH; (c) blend of 40/60 (AS-B)/
EVOH; (d) blend of 70/30 (AS-B)/EVOH; and (e) neat AS-B.

(a)

(b)
Endotherm

(c)
(d)

(e)

0 50 100 150
Temperature (oC)

Figure 5.7. DSC thermograms in the glass transition region for the blends of (AS-LF)/
EVOH: (a) neat EVOH; (b) blend of 20/80 (AS-LF)/EVOH; (c) blend of 40/60 (AS-LF)/
EVOH; (d) blend of 70/30 (AS-LF)/EVOH; and (e) neat AS-LF.

83
(a)

(b)

Endotherm
(c)
(d)

(e)

0 50 100 150
Temperature ( o C)

Figure 5.8. DSC thermograms in the glass transition region for the blends of (CS-B)/
EVOH: (a) neat EVOH; (b) blend of 20/80 (CS-B)/EVOH; (c) blend of 40/60 (CS-B)/
EVOH; (d) blend of 70/30 (CS-B)/EVOH; and (e) neat CS-B.

(a)

(b)
Endotherm

(c)

(d)

(e)

0 50 100 150
Temperature (oC)

Figure 5.9. DSC thermograms in the glass transition region for the blends of (CS-LF)/
EVOH: (a) neat EVOH; (b) blend of 20/80 (CS-LF)/EVOH; (c) blend of 40/60 (CS-LF)/
EVOH; (d) blend of 70/30 (CS-LF)/EVOH; and (e) neat CS-LF.
84
transition temperature Tg, the arrow pointing upward denotes the onset of glass transition

Tgi, and the arrow pointing downward denotes the final point of the glass transition

temperature Tgf. All the values were determined by Universal Analysis 2000 (TA

Instruments). The following observations are worth noting in the DSC thermograms.

First, only one glass transition temperature can be found on each DSC thermogram of a

blend, suggesting that the modified starches (including AS-B, AS-LF, CS-B, and CS-LF)

are miscible with EVOH (within the size scale, which DSC can detect). The values of Tg

of the four blend systems are summarized in Table 5.2. Second, the blend composition

has a significant influence on the width of the glass transition of each blend system,

namely, the width of glass transition (ΔwTg = Tgf − Tgi) decreases with increasing weight

fraction of a modified starch, as summarized in Table 5.2. Note that the value of ΔwTg of

Table 5.2. Glass transition temperatures of EVOH/modified starch blends

Weight ratio Tg (oC) ΔwTg


EVOH n/a 63 41
(AS-B)/EVOH 2/8 71 44
4/6 79 40
7/3 85 26
10/0 90 15
(AS-LF)/EVOH 2/8 60 42
4/6 57 40
7/3 50 25
10/0 45 16
(CS-B)/EVOH 2/8 72 43
4/6 80 27
7/3 83 20
10/0 92 18
(CS-LF)/EVOH 2/8 61 43
4/6 58 30
7/3 52 25
10/0 44 18

85
neat EVOH is higher than that of neat modified starch, which might be explained by the

fact that the self-association in EVOH is stronger than that in a modified starch. When a

modified starch was introduced into neat EVOH, the hydrogen bonds between EVOH

and a modified starch decreased the self-association of EVOH. In other words, the

hydrogen bonds between a modified starch and EVOH in a blend is favored over the

hydrogen bonds in self-association of neat EVOH.

FTIR spectroscopy was used to evaluate the strength of hydrogen bonding in a

modified starch, neat EVOH, and between a modified starch and EVOH in a given blend.

Figures 5.10 to 5.13 give the FTIR spectra of the blends based on a modified starch and

EVOH. In the FTIR spectra of neat EVOH, a broad and strong absorption band centered

at 3349 cm−1 is ascribed to the stretching vibration of the hydrogen-bonded hydroxyl

group, suggesting strong self-association present in EVOH. From the FTIR spectra of

four modified starches (AS-B, AS-LF, CS-B, and CS-LF) we made the following

observations. (1) The absorption band located at 3550 cm−1 is ascribed to the free

hydroxyl group in a modified starch. (2) A weak but broad absorption ranging from

3220 cm−1 to 3500 cm−1 (centered at 3470 cm−1) is ascribed to the hydrogen-bonded

hydroxyl groups in a modified starch. Some investigations116 reported that the difference

in wavenumbers between the absorption bands of hydrogen-bonded and free hydroxyl

groups could be used as a measure of the relative strength of the hydrogen bonds. Thus,

it is clearly seen in Figures 5.10 to 5.13 that the hydrogen bonding in neat EVOH is

stronger than that in a modified starch. In the FTIR spectra of the blends of EVOH and a

modified starch, we find that the absorption band of the hydrogen-bonded hydroxyl

86
Absorbance (arb. unit)
(a) −OH −OH

(b)

free −OH −OH −OH

(c)

3800 3600 3400 3200 3000 2800


−1
Wavenumber (cm )

Figure 5.10. FTIR spectra in the hydroxyl group stretching region for: (a) neat EVOH; (b)
neat modified starch AS-B; and (c) blend of 70/30 (AS-B)/EVOH.
Absorbance (arb. unit)

−OH −OH
(a)
free −OH
(b)

(c)
free −OH

3800 3600 3400 3200 3000 2800


−1
Wavenumber (cm )
Figure 5.11. FTIR spectra in the hydroxyl group stretching region for: (a) neat EVOH; (b)
blend of 70/30 (AS-LF)/EVOH; and (c) neat modified starch AS-LF.

87
Absorbance (arb. unit)
(a) −OH −OH

(b)
free −OH
(c)

3800 3600 3400 3200 3000 2800


−1
Wavenumber (cm )
Figure 5.12. FTIR spectra in the hydroxyl group stretching region for: (a) neat EVOH; (b)
blend of 70/30 (CS-B)/EVOH; and (c) neat modified starch CS-B.
Absorbance (arb. unit)

(a)
−OH −OH

(b)
free −OH
(c)

3800 3600 3400 3200 3000 2800


−1
Wavenumber (cm )

Figure 5.13. FTIR spectra in the hydroxyl group stretching region for: (a) neat EVOH; (b)
blend of 70/30 (CS-LF)/EVOH; and (c) neat modified starch CS-LF.

88
groups is shifted to around 3220 cm−1. The difference in wavenumbers between the

absorption bands of the hydrogen-bonded hydroxyl groups and the free-hydroxyl groups

is found to be larger in the blends of EVOH and a modified starch than in a modified

starch or in neat EVOH. Thus, the hydrogen bonding between EVOH and a modified

starch in a blend is stronger than that in self-association of neat EVOH.

5.3 Crystallization of EVOH in Blends based on Modified Starch and EVOH

Wide-angle X-ray diffraction (WAXD) was used to investigate whether or not the

crystalline structure of EVOH was influenced by the hydrogen bonding between EVOH

and a modified starch in a blend. Previous research113,117 reported that the crystalline

structure was largely affected by the amount of vinyl alcohol groups in EVOH. The

crystalline structure takes orthorhombic lattice when the amount of vinyl alcohol groups

is lower than 20 mol %, and takes monoclinic lattice when the amount of vinyl alcohol

groups is higher than 20 mol %. It was also found that the composition-dependence of

crystalline structure described above is only valid for slow crystallized samples. Thermal

history also has an influence on the crystalline structures of EVOH. The literature112,118

showed that various cooling rates or crystallization temperatures could induce a change in

crystalline lattice between monoclinic lattice and orthorhombic lattice.

Figure 5.14 gives WAXD patterns of neat EVOH samples at various annealing

temperatures. It is clearly seen that the diffraction patterns of EVOH are largely

influenced by crystallization temperature. The quenched sample has a pattern

characteristic of an orthorhombic lattice, with the (110) reflection peak being located at

89
monoclinic lattice group

Intensity (arb. unit)


(1)

(2)

(3)

(4)

(5)
orthohombic lattice group

16 17 18 19 20 21 22 23 24 25
2θ (degree)
Figure 5.14. WAXD patterns of neat EVOH at different annealing temperatures (oC):
(1) 143; (2) 123; (3) 103; (4) 83; and (5) quench. Annealing period was for 3 h, and the
specimen was quenched by liquid nitrogen.

20.5o (2θ). With increasing crystallization temperature, the (110) reflection peak is

broadened and split into two monoclinic reflection (110) and (110) peaks. This change in

the crystallization morphology was not observed during a typical temperature scan of the

samples, even at the relatively slow ramp speed of 5 oC/min, because of its rather slow

crystallization kinetics.

The change in crystalline morphology described above in neat EVOH has been

reported.112,118 Figures 5.15 to 5.18 give WAXD patterns of the blends based on EVOH

and a modified starch, crystallizing at various temperatures. The reflection peaks shown

in the WAXD patterns of the blends are ascribed to the crystalline phase of EVOH in the

blend (the modified starch is amorphous). Referring to Figures 5.15 to 5.18, it is shown

90
Intensity (arb. unit) (1)
(2)
(3)
(4)

16 17 18 19 20 21 22 23 24 25
2θ (degree)

Figure 5.15. WAXD patterns for a 70/30 (AS-B)/EVOH blend at different annealing
temperatures (oC): (1) 145; (2) 125; (3) 105; and (4) quench.
Intensity (arb. unit)

(1)
(2)
(3)
(4)
(5)

16 17 18 19 20 21 22 23 24 25
2θ (degree)

Figure 5.16. WAXD patterns for a 70/30 (AS-LF)/EVOH blend at different annealing
temperatures (oC): (1) 130; (2) 110; (3) 90; (4) 70; and (5) quench.

91
Intensity (arb. unit)
(1)

(2)

(3)
(4)

16 17 18 19 20 21 22 23 24 25
2θ (degree)

Figure 5.17. WAXD patterns for a 70/30 (CS-B)/EVOH blend at different annealing
temperatures (oC): (1) 143; (2) 123; (3) 103; and (4) quench.
Intensity (arb. unit)

(1)

(2)

(3)
(4)

(5)

16 17 18 19 20 21 22 23 24 25
2θ (degree)

Figure 5.18. WAXD patterns for a 70/30 (CS-LF)/EVOH blend at different annealing
temperatures (oC): (1) 132; (2) 112; (3) 92; (4) 72; and (5) quench.

92
that the change in the crystalline morphology in a blend with increasing crystallization

temperature is similar to that in neat EVOH (see Figure 5.14).

5.4 Mechanical Properties of Blends based on EVOH and Modified Starch

Figures 5.19 to 5.22 give the tensile properties (tensile strength, Young’s modulus,

and elongation at break) of the blends based on a modified starch and EVOH. The

dashed line in the figures denotes a linear additive rule as a reference. For the tensile

properties of the blend based on modified starch and EVOH, it is worth noting that all

(AS-B)/EVOH blends, (AS-LF)/EVOH blends, (CS-B)/EVOH blends, and (CS-LF)/

EVOH blends exhibit a positive deviation from linearity for the tensile properties versus

composition plots. Then the tensile properties of blends are enhanced over those of the

constituent components. These synergistic effects obtained in the blends suggest an

enhanced miscibility between the modified starch and EVOH. Usually, the synergistic

effects for tensile properties have been found in miscible blends in which the

densification of the blends occurred due to the specific interaction. We ascribed the

apparent miscibility of the blends to the formation of hydrogen bonds between the

hydroxyl groups in a modified starch and the hydroxyl group in vinyl alcohol units of

EVOH. In the previous section, the attractive interaction between a modified starch and

EVOH in the blends have been characterized by FTIR spectroscopy and DSC.

93
140 7.0
(a) (b)
Tensile strength (MPa)

Young's modulus (GPa)


6.5
120

6.0
100
5.5

80
5.0

60 4.5
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction) EVOH (wt fraction)

60
(c)
50
Elongation at break (%)

40

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction)

Figure 5.19. Tensile properties of (AS-B)/EVOH blends: (a) tensile strength; (b)

Young’s modulus; and (c) elongation at break.

94
140 7
(a) (b)
120

Young's Modulus (GPa)


Tensile strength (MPa)

100
5
80

4
60

40 3
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction) EVOH (wt fraction)

60
(c)
50
Elongation at Break (%)

40

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction)

Figure 5.20. Tensile properties of (AS-LF)/EVOH blends: (a) tensile strength; (b)
Young’s modulus; and (c) elongation at break.

95
140 7.0
(a) (b)

Young's Modulus (GPa)


120 6.5
Tensile strength (MPa)

100 6.0

80 5.5

60 5.0

40 4.5
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction) EVOH (wt fraction)

60
(c)
50
Elongation at break (%)

40

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction)

Figure 5.21. Tensile properties of (CS-B)/EVOH blends: (a) tensile strength; (b)
Young’s modulus; and (c) elongation at break.

96
140 7.0
(a) (b)
6.5

Young's Modulus (GPa)


120
Tensile strength (MPa)

6.0
100 5.5

5.0
80
4.5
60
4.0

40 3.5
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction) EVOH (wt fraction)

60
(c)
50
Elongation at break (%)

40

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction)

Figure 5.22. Tensile properties of (CS-LF)/EVOH blends: (a) tensile strength; (b)
Young’s modulus; and (c) elongation at break

97
5.5 Concluding Remarks

A series of chemically modified starch synthesized in the present study was blended

with EVOH (having 27 mol % of ethylene component). The miscibility between the

modified starch and EVOH was characterized by phase diagrams. The blends of a

modified starch and EVOH was characterized by DSC. The DSC thermograms have

shown a single glass transition temperature, suggesting a good miscibility between the

modified starch and EVOH. We ascribed the miscibility in the blends to the formation of

the hydrogen bonds between the hydroxyl groups in EVOH and a modified starch.

Additionally, the DSC thermograms have shown a melting point depression and a

decreasing melting enthalpy with the increase of the composition of the modified starch

in the blends, suggesting that the crystallization of EVOH was interrupted by the

modified starch. We also ascribed this interruption to the formation of the hydrogen

bonds between a modified starch and EVOH. The FTIR spectra of the blends have

shown a broad absorbance band of the hydrogen-bonded hydroxyl groups, which was

located at a lower wavenumber than the absorbance band of hydrogen-bonded hydroxyl

groups in neat EVOH, suggesting that the hydrogen bonds formed between a modified

starch and EVOH in the binary blends are favored over the hydrogen bonds due to self-

association in neat EVOH.

Furthermore, the crystalline structure of EVOH in the blends of a modified starch

and EVOH was characterized by WAXD. In neat EVOH, the crystalline structure of

EVOH takes orthorhombic lattice in the quenched specimen and evolves into monoclinic

lattice as the crystallization temperature is increased. The WAXD patterns have shown

that the crystalline structure of EVOH in the blends of a modified starch and EVOH did

98
not change from that of neat EVOH. The tensile properties of blends of a modified starch

and EVOH are found to have a positive deviation from linearity over the entire blend

composition, indicating that the blends are miscible.

99
CHAPTER VI

NANOCOMPOSITES BASED ON MODIFIED STARCH

6.1 Introduction

The research on nanocomposites started in the early 1990s by the publication about

nylon 6 of Toyota scientists.119-121 A very high degree of dispersion of organoclay

aggregates will generate large surface areas, which is supposed to significantly enhance

the mechanical properties of nanocomposites. Furthermore, a high degree of dispersion

of nanoclay aggregates also was reported to have a profound influence on the

morphology of the polymer matrix due to its enhanced surface areas.122 On the other

hand, highly dispersed organoclay nanocomposites have been found to increase heat

distortion temperature and the gas barrier properties.119,123-127

To achieve a high degree of dispersion of nanoclay aggregates, a fundamental issue

should be addressed about compatibility between a polymer matrix and clay. During the

last two decades, much effort has been spent on the modification of the surface of natural

clay by developing surfactants. However, there is no universal surfactant that enables

one to achieve a high degree of dispersion of nanocomposites based on any polymer

matrix. One has to modify either the chemical structure of the polymer matrix or the

chemical structure of a surfactant residing at the surface of the nanoclay to obtain the

desired compatibility between the polymer matrix and nanoclay.

100
Han et al.128-134 have reported on successful research results on the preparation of a

high degree of dispersion of nanocomposites. They found exfoliated nanoclay

(evidenced by TEM and XRD) in nanocomposites by introducing functional group(s) into

a polymer matrix. In those studies, natural clay (MMT) and two commercial organoclays

(Cloisite 30B and Cloisite 15A) were employed. The specific interactions, including

hydrogen bonding, ionic interaction, or ion-dipole interaction, between the functional

group(s) in a polymer matrix and the layered silicates of nanoclay were confirmed by

FTIR spectroscopy, UV-vis spectroscopy, and solid-state NMR spectroscopy. They

established guidelines for the preparation of nanocomposites with a high degree of

dispersion.

The research on the starch-based nanocomposites began several years ago.5,6 Since

then, many reports have been published on starch-based nanocomposites.100-110 However,

all XRD patterns and TEM images have shown quite a low degree of dispersion of

nanoclay aggregates.

In this dissertation, we prepared and characterized the nanocomposites based on the

chemically modified starch. To investigate how functional groups in a modified starch

affect the dispersion characteristics of nanocomposites, a series of chemically modified

starches has been synthesized with the chemical structures, as shown in Scheme 6.1. In

Chapter IV we presented the chemical structures, thermal behavior, biodegradability, and

hydrophobicity of these chemically modified starches, and in Chapter V we presented the

miscibility of the blends based on a chemically modified starch and EVOH. As described

in Chapter IV, we attached the hexanoyl groups or benzoyl groups into starch molecules

to lower the glass transition temperature and enhance the hydrophobicity of starch.

101
Scheme 6.1. Chemical structures of the modified starches.

(CH2)4CH3
O
O
O
(CH2)4CH3 O
O O
O
O
O
O O
OH O
O O O
O
O O O OH
O O
OH OH
O
O OH
O
Cl OH
O O H3C N CH3 Cl
ONa ONa CH3 H3C N CH3
CH3
AS-LF AS-B CS-LF
CS-B

On the other hand, from the point of view of preparing nanocomposites, we

attached ionic groups (acetate sodium for anionic group or trimethylammonium chloride

for cationic group) into the starch molecules with the expectation that the attached ionic

groups may have ionic interactions with layered silicates of clay. Specifically, the

cationic groups attached onto a starch are expected to have ionic interaction with the

negatively charged surface of natural clay (MMT). Note that the counter ion residing at

the surface of layered silicates in MMT is Na+. And the anionic groups are expected to

have ionic interaction with the positively charged N+ in the surfactant (MT2EtOH for

Cloisite 30B and 2M2HT for Cloisite 15A) residing at the surface of organoclay. The

chemical structures of the surfactants are shown in Scheme 6.2. In the chemical structure

of MT2EtOH, N+ denotes quaternary ammonium chloride, and T denotes tallow

102
consisting of ca. 65% C18, ca. 30% C16, and ca. 5% C14. In the chemical structure of

2M2HT, N+ denotes quaternary ammonium chloride, and HT denotes hydrogenated

tallow consisting of ca. 65% C18, ca. 30% C16, and ca. 5% C14.

Scheme 6.2. Chemical structures of the surfactant MT2EtOH residing at the surface of
Cloisite 30B and the surfactant 2M2HT residing at the surface of Cloisite 15A.

CH2CH2OH CH3
CH3 N T CH3 N HT
CH2CH2OH HT

MT2EtOH 2M2HT

The dispersion characteristics of the nanocomposites prepared in the present study

were investigated using X-ray diffraction (XRD) and transmission electron microscopy

(TEM). The presence of ionic interactions between a modified starch and nanoclay was

investigated using Fourier transform infrared (FTIR) spectroscopy. The mechanical

properties of nanocomposites prepared were also measured.

6.2 Nanocomposites based on Neat EVOH

Figure 6.1 gives the XRD patterns for nanoclays used in this study, which include

MMT, Cloisite 30B, and Cloisite 15A. Note in Figure 6.1 that MMT has a gallery

distance (d001 spacing) of 1.1 nm, while Cloisite 30B and Cloisite 15A, due to the

presence of a surfactant, have a larger gallery distance, 1.9 nm and 3.1 nm, respectively.

103
3.1 nm

Intensity (arb. unit)


(a)
1.1 nm

1.9 nm (b)
(c)

2 3 4 5 6 7 8 9 10
2θ (degree)

Figure 6.1. XRD patterns for: (a) Cloisite 15A; (b) MMT; and (c) Cloisite 30B.

The goal of this research was to investigate how the functional group(s) attached

onto starch might affect the dispersion characteristics of organoclay (or natural clay)

aggregates by specific interactions. Thus, before characterizing the dispersion

characteristics of nanocomposites based on a mixture of a modified starch and EVOH,

the dispersion of nanocomposites based on neat EVOH should be investigated. Figure

6.2 gives XRD patterns for nanocomposites based on neat EVOH. It is worth noting that

the XRD patterns for EVOH/MMT nanocomposite and EVOH/Cloisite 15A

nanocomposite show little difference from the XRD patterns for Cloisite 15A and MMT,

suggesting that both nanocomposites have a very poor dispersion of nanoclay aggregates.

This is not surprising, because no attractive interaction can be expected between neat

EVOH and the surface of MMT or a surfactant residing at the surface of Cloisite 15A.

104
Intensity (arb. unit)
3.2 nm

(a)
2.3 nm

(b)
1.2 nm
(c)

2 3 4 5 6 7 8 9 10
2θ (degree)
Figure 6.2. XRD patterns for: (a) EVOH/Cloisite 15A nanocomposite; (b) EVOH/
Cloisite 30B nanocomposite; and (c) EVOH/MMT nanocomposite.

On the other hand, the EVOH/Cloisite 30B nanocomposite has a d001 spacing of 2.3 nm,

an increase of 0.4 nm over the d001 spacing of Cloisite 30B, suggesting that the

nanocomposite might have intercalated Cloisite 30B aggregates.

Figure 6.3 shows TEM images for the nanocomposites based on neat EVOH.

Both XRD patterns and TEM images indicated that there is no high degree of dispersion

of organoclay aggregates in neat EVOH. The preparation of nanocomposites based on

EVOH has previously been reported.132 In that study, EVOH had 53 mol % of vinyl

alcohol, while in the present study EVOH has 79 mol % of vinyl alcohol. The XRD

results showed that the d001 spacing of EVOH/Cloisite 30B nanocomposite has increased

by 1.69 nm over the d001 spacing of Cloisite 30B. Although there are hydroxyl groups on

both EVOH and the surfactant residing at the surface of Cloisite 30B, the hydrogen

105
(a) (b)

200 nm 200 nm

(c)

200 nm

Figure 6.3. TEM images for: (a) EVOH/MMT nanocomposite; (b) EVOH/Cloisite 15A
nanocomposite; and (c) EVOH/Cloisite 30B nanocomposite, in which the dark areas
represent the clay, and the grey/white areas represent the polymer matrix.

bonding between them was believed to be hindered, because the attractive interaction

within EVOH molecules was very strong.132 The above observations led us to conclude

that the EVOH employed in the present study could not have a high degree of dispersion

of organoclay aggregates.

106
6.3 Nanocomposites containing Anionically Modified Starch

Figure 6.4 gives XRD patterns for the nanocomposites based on the (AS-B)/EVOH

blend. It is worth noting that the nanocomposites containing organoclay (Cloisite 30B or

Cloisite 15A) show featureless XRD patterns, suggesting that organoclay aggregates

might have been highly dispersed. On the other hand, the nanocomposite containing

MMT has a d001 spacing of 1.3 nm, which is slightly higher than the d001 spacing of MMT

(1.1 nm). This is not surprising because there is no attractive interaction expected

between AS-B and the surface of MMT.

Figure 6.5 shows TEM images for the nanocomposites based on the (AS-B)/EVOH

blend, and Cloisite 15A, Cloisite 30B, or MMT. It is clearly seen in Figure 6.5 that both

the (AS-B)/EVOH/Cloisite 30B and (AS-B)/EVOH/Cloisite 15A nanocomposite have a


Intensity (arb. unit)

(a)

(b)

1.3 nm

(c)

2 3 4 5 6 7 8 9 10
2θ (degree)

Figure 6.4. XRD patterns for: (a) (AS-B)/EVOH/Cloisite 15A nanocomposite;


(b) (AS-B)/EVOH/Cloisite 30B nanocomposite; and (c) (AS-B)/EVOH/MMT
nanocomposite.

107
(a) (b)

200 nm 200 nm

(c)

200 nm

Figure 6.5. TEM images for: (a) (AS-B)/EVOH/Cloisite 30B nanocomposite;


(b) (AS-B)/EVOH/Cloisite 15A nanocomposite; and (c) (AS-B)/EVOH/MMT
nanocomposite.

very high degree of dispersion of organoclay aggregates in the matrix, whereas the

(AS-B)/EVOH/MMT nanocomposite has large MMT aggregates, indicating a poor

dispersion of the nanocomposites.

Both XRD patterns and TEM images shown above indicate that both the (AS-B)/

EVOH/Cloisite 30B and (AS-B)/EVOH/Cloisite 15A nanocomposites have a high degree

of dispersion of organoclay aggregates, whereas the (AS-B)/EVOH/MMT nanocomposite

108
has a very poor dispersion of MMT aggregates. Note that the surfactant residing at the

surface of Cloisite 30B or Cloisite 15A have positively charged N+, and AS-B has

anionic functional group −COO−. Thus, the high degree of dispersion of organoclay

aggregates (so-called exfoliation of layered silicates) might have resulted from the ionic

interaction between N+ at the surface of the organoclay and −COO− in the modified

starch. It is well established that the ionic interaction is much stronger, although not as

strong as covalent bonding, than the hydrogen bonding.

To directly probe evidence of the ionic interaction between −COO− and N+, FTIR

spectroscopy was employed. Figure 6.6 gives the FTIR spectra for (a) (AS-B)/EVOH

blend, (b) (AS-B)/EVOH/Cloisite 30B nanocomposite, (c) (AS-B)/EVOH/Cloisite 15A

nanocomposite, and (d) (AS-B)/EVOH/MMT nanocomposite.


Absorbance (arb. unit)

(d)

(c)

(b)

(a)

4000 3000 2000 1000


−1
Wavenumber (cm )
Figure 6.6. FTIR spectra for: (a) (AS-B)/EVOH blend; (b) (AS-B)/EVOH/Cloisite 30B
nanocomposite; (c) (AS-B)/EVOH/Cloisite 15A nanocomposite; and (d) (AS-B)/EVOH/
MMT nanocomposite.

109
Figure 6.7 gives the carboxyl group stretching region for the same FTIR spectra

shown in Figure 6.6. In reference to Figure 6.7, the absorbance band at 1737 cm−1 is

attributed to the stretching of carbonyl groups (C=O) in the benzoyl ester group in a

modified starch. The absorbance band at 1594 cm−1 in the spectra of the AS-B/EVOH


blend is attributed to asymmetric stretching carboxylate group (−COO ) in the AS-B. It

is seen that the absorbance band of carboxylate group is shifted to 1572 cm−1 for the

nanocomposites containing a high degree of dispersion of Cloisite 30B or Cloisite 15A

aggregates. There is no shift in the absorbance band of carboxylate group for the

nanocomposite containing MMT aggregates. Previous studies133,135,136 reported that the

absorbance band of carboxylate group was shifted to a lower wavenumber when ionic

interactions were present between the carboxylate groups and other positively charged

groups. Yano and coworkers135,136 reported that in the binary blends of sodium and zinc

salt ionomers of poly(ethylene-co-methacrylate) (PEMMA), the absorbance band of

asymmetric stretching of carboxylate groups was found at 1569 cm−1 when carboxylate

groups bridged sodium and zinc cations. Zha et al.133 reported that in the preparation of

nanocomposites based on synthetic polymers (homopolymer or block polymer)

end-functionalized by carboxylate sodium salt, the absorbance band of asymmetric

stretching of carboxylate groups was found at 1570 cm−1 because of the ionic interaction

between the carboxylate groups and the N+ in the surfactant residing at the surface of

Cloisite 30B. Thus, it is reasonable to state that the absorbance band at 1572 cm−1 found

in Figure 6.7 is asymmetric stretching of carboxylate group, while having ionic

interaction with N+ in the surfactant residing at the surface of Cloisite 30B or Cloisite

110
O
C
(1)
(1) O
Absorbance (arb. unit)
(3)
(d)
O (2)
(c) O C
O Na
(2)
(b) N

(3) O
(a) (3)
O C
O Na

1800 1700 1600 1500 1400


−1
Wavenumber (cm )
Figure 6.7. Carboxylate group stretching region for the FTIR spectra shown in Figure 6.6
for: (a) (AS-B)/EVOH blend; (b) (AS-B)/EVOH/Cloisite 30B nanocomposite; (c)
(AS-B)/EVOH/Cloisite 15A; and (d) (AS-B)/EVOH/MMT nanocomposite.

15A. On the other hand, since there are no functional groups on the surface of MMT, the

absorbance band of asymmetric stretching of carboxylate groups on a modified starch did

not shift along the spectra of the (AS-B)/EVOH/MMT nanocomposite, because there was

no ionic interaction present.

The XRD patterns and TEM images indicated that both (AS-B)/EVOH/Cloisite 30B

and (AS-B)/EVOH/Cloisite 15A nanocomposites have a high degree of dispersion of

organoclay aggregates. We ascribed this observation to the presence of the ionic

interaction between the positively charged N+ residing at the surface of organoclay,

Cloisite 30B or Cloisite 15A, and the negatively charged carboxylate groups attached

onto the main chain of a modified starch. The FTIR spectra showed that the absorbance

111
band of carboxylate groups was shifted to a lower wavenumber, indicative of the

presence of the ionic interaction occurring on the carboxylate groups. On the other hand,

the (AS-B)/EVOH/MMT nanocomposite has a poor dispersion of MMT aggregates.

There is no attractive interaction formed between the AS-B and MMT in the

nanocomposite.

The dispersion of nanoclay (Cloisite 30B, Cloisite 15A, or MMT) aggregates in the

nanocomposites based on the blends of (AS-LF)/EVOH was also characterized. It should

be noted that the anionic moiety in the AS-LF also has a carboxylate group, which is the

same as in the AS-B. Figure 6.8 shows XRD patterns for the nanocomposites based on

the blends of (AS-LF)/EVOH. It can clearly be seen that the aggregates of clay have

similar dispersion characteristics in the (AS-LF)/EVOH blend as in the (AS-B)/EVOH


Intensity (arb. unit)

(1)

1.4 nm
(2)

(3)

2 3 4 5 6 7 8 9 10
2θ (degree)
Figure 6.8. XRD patterns for: (1) (AS-LF)/EVOH/Cloisite 15A nanocomposite;
(2) (AS-LF)/EVOH/MMT nanocomposite; and (3) (AS-LF)/EVOH/Cloisite 30B
nanocomposite.

112
blend. Featureless XRD patterns for (AS-LF)/EVOH/Cloisite 30B and (AS-LF)/EVOH/

Cloisite 15A nanocomposite indicate a high degree of dispersion of organoclay

aggregates in the respective blends. On the other hand, the (AS-LF)/EVOH/MMT

nanocomposite has a d001 spacing of 1.4 nm, which is 0.3 nm higher than the d001 spacing

of MMT (1.1 nm) (see Figure 6.1) and only 0.1 nm higher than the d001 spacing of the

(AS-B)/EVOH/MMT nanocomposite (1.3 nm), given in Figure 6.4.

Figure 6.9 gives TEM images of the nanocomposites based on (AS-LF)/EVOH

blends, and Cloisite 30B, Cloisite 15A, or MMT. It is clearly seen that both (AS-LF)/

(a) (b)

200 nm 200 nm

(c)

200 nm

Figure 6.9. TEM images for: (a) (AS-LF)/EVOH/Cloisite 30B nanocomposite;


(b) (AS-LF)/EVOH/Cloisite 15A nanocomposite; and (c) (AS-LF)/EVOH/MMT
nanocomposite.

113
EVOH/Cloisite 30B and AS-LF/EVOH/Cloisite 15A nanocomposites have a very high

degree of dispersion of organoclay aggregates in the matrix, whereas the (AS-LF)/

EVOH/MMT nanocomposite has large clay aggregates, indicating a poor dispersion of

MMT aggregates.

The XRD patterns and TEM images shown above suggest that organoclay

aggregates (Cloisite 30B or Cloisite 15) have a high degree of dispersion in the

nanocomposites based on the (AS-LF)/EVOH blend, whereas MMT aggregates have a

very poor dispersion in the (AS-LF)/EVOH/MMT nanocomposite.

Figure 6.10 gives the FTIR spectra in carboxylate group stretching region for the

nanocomposites based on the (AS-LF)/EVOH blend, and Cloisite 15A, Cloisite 30B, or

MMT. The absorbance band at 1743 cm−1 was designated to the stretching of carbonyl

group (C=O) in the hexanoyl group, contrasting to the absorbance band at 1737 cm−1

designated to the same carbonyl group (C=O) in the benzoyl ester group, shown in Figure

6.7. In the FTIR spectra of the (AS-LF)/EVOH blend shown in Figure 6.10, the

absorbance band at 1594 cm−1 was designated to asymmetric stretching of the

carboxylate group (−COO−) in the AS-LF. Similar to that observed in Figure 6.7, the

absorbance band of carboxylate group shifted to a lower wavenumber in the

nanocomposites containing Cloisite 30B or Cloisite 15A with a high degree of dispersion,

indicating the presence of the ionic interaction between the carboxylate groups (−COO−)

and N+ residing at the surface of Cloisite 30B or Cloisite 15A. On the other hand, the

spectra of the (AS-LF)/EVOH/MMT nanocomposite show that the absorbance band of

carboxylate group has not changed at all, suggesting no ionic interaction in the

nanocomposite.
114
O C C C
(1) C C C
(1)
Absorbance (arb. unit) O

(3) (d)
O
(2)
(c) O C
O Na

(2) (b) N

(a)
O
(3)
O C
(3) O Na

1800 1700 1600 1500 1400


−1
Wavenumber (cm )

Figure 6.10. FTIR spectra at carboxyl group stretching region for: (a) (AS-LF)/EVOH
blend; (b) (AS-LF)/EVOH/Cloisite 30B nanocomposite; (c) (AS-LF)/EVOH/Cloisite
15A nanocomposite; and (d) (AS-LF)/EVOH/MMT nanocomposite.

The dispersion characteristics of the nanoclay aggregates in the blends based on the

anionically modified starch and EVOH have been characterized by XRD and TEM. It

was found that the aggregates of organoclay (Cloisite 30B or Cloisite 15A) have a high

degree of dispersion, whereas the aggregates of MMT have a very poor dispersion in the

anionically modified starch. We ascribed the high degree of dispersion of organoclay

aggregates in nanocomposites based on anionically modified starch to the presence of the


ionic interaction between the negatively charged carboxylate group (−COO ) in the

modified starch and the positively charged N+ in the surfactant residing at the surface of

organoclay. This ionic interaction was confirmed by FTIR spectroscopy.

115
6.4 Nanocomposites containing Cationically Modified Starch

Figure 6.11 gives XRD patterns for the nanocomposites based on the blends of

cationically modified starch (CS-B) and EVOH, and Cloisite 15A, Cloisite 30B, or MMT.

It is seen in Figure 6.11 that the nanocomposite containing Cloisite 15A has the d001

spacing of 3.2 nm, and the nanocomposite containing Cloisite 30B has the d001 spacing of

2.2 nm. It is worth noting that both nanocomposites containing either Cloisite 15A or

Cloisite 30B have a d001 spacing only slightly higher than the d001 spacing of organoclay

shown in Figure 6.1, suggesting that the dispersion characteristics of the organoclay

aggregates in the polymer matrix would be low. This observation is not surprising

because sufficiently strong attractive interaction is not expected between the cationically

modified starch and the organoclay.


Intensity (arb. unit)

3.2 nm

(c)
2.2 nm
(b)
1.2 nm (a)

2 3 4 5 6 7 8 9 10
2θ (degree)
Figure 6.11. XRD patterns for: (a) (CS-B)/EVOH/MMT nanocomposite; (b) (CS-B)/
EVOH/Cloisite 30B nanocomposite; and (c) (CS-B)/EVOH/Cloisite 15A nanocomposite.
The nanocomposites were prepared by melt blending.

116
The XRD patterns further indicate that the nanocomposite containing MMT has a

d001 spacing of 1.2 nm, only 0.1 nm higher than the d001 spacing of MMT, suggesting that

the dispersion characteristics of MMT aggregates in the cationically modified starch

would also be very low.

Figure 6.12 gives TEM images for the nanocomposites based on the (CS-B)/EVOH

blend and Cloisite 15A, Cloisite 30B, or MMT. It is clearly seen in Figure 6.12 that all

(a) (b)

200 nm 200 nm

(c)

200 nm

Figure 6.12. TEM images for: (a) (CS-B)/EVOH/Cloisite 15A nanocomposite;


(b) (CS-B)/EVOH/Cloisite 30B nanocomposite; and (c) (CS-B)/EVOH/MMT
nanocomposite. The nanocomposites were prepared by melt blending.

117
nanocomposites have large clay aggregates, indicating the presence of poor dispersion of

clay aggregates, which is consistent with the speculation made above in reference to

Figure 6.11. Figure 6.13 gives XRD patterns for the nanocomposites based on the

(CS-LF)/EVOH blend. As can be seen in Figure 6.13, the nanocomposites based on the

(CS-LF)/ EVOH blend and Cloisite 15A, Cloisite 30B, or MMT have d001 spacings of 3.5,

2.2, and 1.2 nm, respectively, only slightly higher than the d001 spacing of clay, which

suggests that the dispersion of clays in CS-LF would be poor. Figure 6.14 gives the TEM

images for the nanocomposites based on the (CS-LF)/EVOH blend. It is clearly seen that

all nanocomposites have large clay aggregates. The results of XRD patterns and TEM
Intensity (arb. unit)

3.5 nm

(c)
2.2 nm
(b)
1.2 nm
(a)

2 3 4 5 6 7 8 9 10
2θ (degree)
Figure 6.13. XRD patterns for: (a) (CS-LF)/EVOH/MMT nanocomposite; (b) (CS-LF)/
EVOH/Cloisite 30B nanocomposite; and (c) (CS-LF)/EVOH/Cloisite 15A
nanocomposite. The nanocomposites were prepared by melt blending.

118
(a) (b)

200 nm 200 nm

(c)

200 nm

Figure 6.14. TEM images for: (a) (CS-LF)/EVOH/Cloisite 15A nanocomposite;


(b) (CS-LF)/EVOH/Cloisite 30B nanocomposite; and (c) (CS-LF)/EVOH/MMT
nanocomposite. The nanocomposites were prepared by melt blending.

images given in Figures 6.11 to 6.14 indicate that the aggregates of both organoclays

(Cloisite 30B and Cloisite 15A) and natural clay (MMT) have a poor dispersion in the

cationically modified starch (CS-LF and CS-B).

Figure 6.15 gives XRD patterns for the nanocomposites based on the cationically

modified starch CS-B prepared by solution blending. It is worth noting that the

119
Intensity (arb. unit)
3.3 nm

(c)

(b)

2.2 nm
(a)

2 3 4 5 6 7 8 9 10
2θ (degree)

Figure 6.15. XRD patterns for: (a) (CS-B)/Cloisite 30B nanocomposite; (b) (CS-B)/
MMT nanocomposite; and (c) (CS-B)/Cloisite 15A nanocomposites. The
nanocomposites were prepared by solution blending.

nanocomposites containing Cloisite 15A or Cloisite 30B have the d001 spacings of 3.3 nm

or 2.2 nm, both slightly higher than the d001 spacing of Cloisite 15A and Cloisite 30B (see

Figure 6.1), which suggests that a poor dispersion of organoclay aggregates in the

nanocomposites would be expected. On the other hand, the nanocomposite containing

MMT has featureless XRD patterns shown in Figure 6.15, indicating that the dispersion

characteristic of layered silicates of MMT in the CS-B would be good. Figure 6.16 gives

TEM images for the nanocomposites based on CS-B, and Cloisite 15A, Cloisite 30B, or

MMT. It can be clearly seen in Figure 6.16 that there are large clay aggregates in the

nanocomposites containing Cloisite 15A or Cloisite 30B, indicating poor dispersion of

120
(a) (b)

200 nm 200 nm

(c)

200 nm

Figure 6.16. TEM images for: (a) (CS-B)/Cloisite 30B nanocomposite; (b) (CS-B)/MMT
nanocomposite; and (c) (CS-B)/Cloisite 15A nanocomposite. The nanocomposites were
prepared by solution blending.

clay aggregates in the CS-B, which is consistent with the XRD patterns shown in Figure

6.15. On the other hand, the TEM image for the nanocomposite containing MMT given

in Figure 6.16 shows a very high degree of dispersion of MMT aggregates in the CS-B

matrix.

Figure 6.17 gives XRD patterns, and Figure 6.18 gives TEM images for the

nanocomposites based on CS-LF which were prepared by solution mixing. It can be seen

121
Intensity (arb. unit)
3.4 nm

(c)
(b)

2.4 nm (a)

2 3 4 5 6 7 8 9 10
2θ (degree)
Figure 6.17. XRD patterns for: (a) (CS-LF)/Cloisite 30B nanocomposite; (b) (CS-LF)/
MMT nanocomposite; and (c) (CS-LF)/Cloisite 15A nanocomposite. The
nanocomposites were prepared by solution blending.

(a) (b)

200 nm 200 nm

(c)

200 nm

Figure 6.18. TEM images for: (a) (CS-LF)/Cloisite 30B nanocomposite; (b) (CS-LF)/
MMT nanocomposite; and (c) (CS-LF)/Cloisite 15A nanocomposite. The
nanocomposites were prepared by solution blending.

122
that the dispersion characteristics of clay aggregates in the nanocomposites based on the

CS-LF is similar to that of the nanocomposites based on the CS-B (see Figure 6.16),

namely, the aggregates of Cloisite 30B or Cloisite 15A have a poor dispersion in the

CS-LF matrix, whereas the aggregates of MMT have a good dispersion in the CS-LF

matrix.

In summary, we have observed that both organoclays, Cloisite 15A and Cloisite

30B, have poor dispersion characteristics in the nanocomposites based on cationically

modified starch, which were prepared by melt blending or solution blending. The poor

dispersion characteristics of the organoclay aggregates resulted from poor compatibility

between the layered silicates and the modified starch, which is attributed to insufficient

strength of attractive interactions between the surfactant residing at the surface of an

organoclay and the cationically modified starch. On the other hand, the experimental

results show poor dispersion characteristics of MMT aggregates in the nanocomposite

based on a cationically modified starch prepared by melt blending, whereas MMT

aggregates have a high degree of dispersion in the nanocomposite based on the same

cationically modified starch, which was prepared by solution blending. The above

observation points out that the preparation methods of nanocomposites have a significant

influence on the dispersion characteristics of MMT aggregates in a cationically modified

starch. The high degree of dispersion characteristics of MMT aggregates in the

nanocomposite based on a cationically modified starch, which was prepared by solution

blending, can be attributed to the presence of the ionic interaction between the positively

charged −N+(CH3)3 in the modified starch and the negatively charged surface of layered

silicates of MMT.

123
Figure 6.19 gives the FTIR spectra in C-N+ stretching mode of the

nanocomposites based on the cationically modified starch (CS-B) prepared by solution

blending. In the FTIR spectra of neat CS-B shown in Figure 6.19, the absorbance band at

957 cm−1 was designated to the asymmetric C-N+ stretching mode vas(C-N+), whereas the

absorbance bands locating at 911 cm−1 and 872 cm−1 were designated to the symmetric

C-N+ stretching modes vs(C-N+). As shown in Figure 6.19, the absorbance bands of

vs(C-N+) were shifted to higher wavenumbers of 923 cm−1 and 884 cm−1 from 911 cm−1

and 872 cm−1, respectively, in the nanocomposite having a high degree of dispersion of

MMT aggregates. On the other hand, there is no shift in the absorbance bands of vs(C-N+)

in nanocomposites containing organoclays (Cloisite 15A or Cloisite 30B).

vas(C-N+)
vs(C-N+)
Absorbance (arb. unit)

Starch O CH3
C C C N CH3
(d)
OH CH3
(c)

(b)

(a)

1000 950 900 850 800


−1
Wavenumber (cm )

Figure 6.19. FTIR spectra in the C-N+ stretching mode region of the nanocomposites
based on cationically modified starch prepared by solution blending: (a) neat CS-B; (b)
(CS-B)/MMT nanocomposite; (c) (CS-B)/Cloisite 15A nanocomposite; and (d) (CS-B)/
Cloisite 30B nanocomposite.

124
The number of absorbance bands of vs(C-N+) and their wavenumbers in infrared

spectroscopy or Raman spectroscopy were found to be very sensitive to the conformation

of O-C-C-N+ group.137-139 In the trans-conformation of the O-C-C-N+ group, the

absorbance bands of vs(C-N+) were reported to be at 925 cm−1 and 875 cm−1.139,140 In

gauche conformation, the absorbance bands of vs(C-N+) were observed at 905 cm−1 and

860 cm−1.139,140 Previous researchers137,141 reported that the gauche conformation of

O-C-C-N+ group in choline-containing structures is more stable than trans-conformation

in general. Thus, the gauche-conformation was reported to be dominant in choline

group.137,141-144 However, the trans-conformation becomes more stable than the gauche

conformation in some crystals or specific structures, which is attributable to specific

intermolecular interactions.137 In the investigation of thin silica film covered by a

1,2-dimyristoyl-sn-glycero-3-phosphocholine bilayer, infrared spectroscopy results

showed that O-C-C-N+ group preferably had the trans-conformation due to the ionic

interaction between the positively charged C-N+ and the negatively charged silicate

surface.140,145 Other research groups138 also reported on a conformation change of

O-C-C-N+ group arising from ionic interactions between the positively charged

O-C-C-N+ group and the negatively charged acid moiety in bisdesmosidic triterpenoid

glycosides, and infrared spectroscopy was used to confirm the interaction.

Thus, the shift in the absorbance bands of vs(C-N+) to higher wavenumbers in

nanocomposites containing MMT aggregates shown in Figure 6.19 can reasonably be

explained by the ionic interaction between the C-N+ in the modified starch and the

negatively charged surface of layered silicates in MMT aggregates. On the other hand, in

125
nanocomposites containing organoclays (Cloisite 15A or Cloisite 30B), the fact that there

is no shift in vs(C-N+) to higher wavenumbers in the FTIR spectra indicates that the

conformation of O-C-C-N+ has not changed, suggesting that no ionic interaction occurred

on the C-N+ group in the modified starch CS-B.

So far, XRD patterns and TEM images showed that MMT aggregates have a high

degree of dispersion characteristics in the nanocomposites based on the cationically

modified starch (CS-B), which was prepared by solution blending. We ascribed the high

degree of dispersion characteristics observed to the ionic interactions between the

positively charged −N+(CH3)3 and negatively charged surface of silicates in MMT,

which was confirmed by the FTIR spectra in the C-N+ stretching mode shown in Figure

6.19.

One might ask why the nanocomposite based on the same cationically modified

starch (CS-B) but prepared by melt blending, has poor dispersion characteristics of MMT

aggregates (refer to Figures 6.11 and 6.12). To answer the question, let us first show the

FTIR spectra in the C-N+ stretching mode region for the nanocomposites based on the

cationically modified starch, which were prepared by melt blending, as shown in Figure

6.20.

As described above in reference to Figure 6.19, the absorbance bands at 911 cm−1

and 872 cm−1 are ascribed to the symmetric stretching mode of C-N+(vs(C-N+)) in a

modified starch (CS-B). On the other hand, referring to Figure 6.20, the absorbance

bands of vs(C-N+) in the nanocomposites containing Cloisite 15A or Cloisite 30B have

not changed, indicating that no interaction occurred on the C-N+ group in a modified

126
vas(C-N+)
Absorbance (arb. unit)

vs(C-N+)
Starch O CH3
C C C N CH3
(d)
OH CH3
(c)
(b)

(a)

1000 950 900 850 800


−1
Wavenumber (cm )

Figure 6.20. FTIR spectrum in the C-N+ stretching mode region of nanocomposites
based on the cationically modified starch prepared by melt blending: (a) neat (CS-B)/
EVOH mixture; (b) (CS-B)/EVOH/Cloisite 15A nanocomposite; (c) (CS-B)/EVOH/
Cloisite 30B nanocomposite; and (d) (CS-B)/EVOH/MMT nanocomposite.

starch (CS-B). We further find from Figure 6.20 that the absorbance bands of vs(C-N+) in

the nanocomposite containing MMT have not changed, indicating that the proposed ionic

interaction between the positively charged −N+(CH3)3 in a modified starch (CS-B) and

the negatively charged surface of silicates in MMT aggregates have not occurred in the

nanocomposite prepared by melt blending. Thus, MMT aggregates have poor dispersion

characteristics in the nanocomposite prepared by melt blending, as has been described in

reference to Figures 6.11 to 6.14.

One might ask another question as to why the positively charged −N+(CH3)3 in a

cationically modified starch can form ionic interaction with the negatively charged

127
surface of silicates in MMT aggregates in the nanocomposite prepared by solution

blending, but fails to form ionic interaction with the same MMT in the nanocomposite

prepared by melt blending. We hasten to point out that water was used as the co-solvent

in the preparation of nanocomposites via solution blending. To answer the question

posed above, let us examine the chemical structure of MMT and the origins of the

negative charge at the surface of silicates in MMT.

Figure 6.21 gives schematic of the chemical structure of MMT.146 It should be

noted that the exchangeable cations in the interlayers of MMT are not quite accurately

shown in Figure 6.21. We will elaborate on this later. Referring to Figure 6.21, the

crystal lattice of MMT consists of layers made up of two tetrahedrally coordinated silicon

atoms fused to an edge-shared octahedral sheet of either aluminum or magnesium

hydroxide. The layer thickness is approximately 1 nm, and the lateral dimensions of

these layers vary from 30 nm to several microns.146 Normally, the net negative charge of

layers in MMT aggregates is generated either by tetrahedral substitution (silicon atoms in

Figure 6.21. Structure of MMT.146 (Reprinted with permission from Elsevier.)

128
tetrahedral sheet were substituted by aluminum atoms) or by octahedral substitution (in

octahedral sheet, Al3+ was replaced by Mg2+ or by Fe2+, or Mg2+ was replaced by

Li+).147 The negative surface charge of the layers in pristine MMT mainly resulted from

octahedral substitution.146,147 Figure 6.22 shows schematics of MMT layers from (a) the

side view and (b) the top view.

From the side view of the silicate layers in MMT aggregates shown in Figure

6.22(a), we can see that the tetrahedral layer consists of two consecutive layers of oxygen

atoms. In each layer of oxygen atoms, every six oxygen atoms form a cavity with a

diameter of 0.26 nm, shown in Figure 6.22(b).148 Normally, the electron orbital of 6

oxygen atoms in such a form as the cavity shown in Figure 6.22(b) will give rise to a

character of weak Lewis base,148 without considering the positive charge deficiency

resulted from octahedral substitution. With octahedral substitution, the resulted negative

charge will be distributed among oxygen atoms in cavities bonded to substituted

octahedral metal cations. Then the Lewis base behavior of the ditrigonal cavity is

reinforced and becomes sufficient to form complexes with those exchangeable cations.

Furthermore, ditrigonal cavities in two consecutive layers are preferable to superpose, as

shown in Figure 6.22(d), by rotation with a specific degree.147,148 Thus, the

exchangeable cations reside in the gallery of layered silicates in MMT by being housed

into such polyhedron structures. Figure 6.23 gives a schematic illustrating how

exchangeable cations reside at the surface of silicate layers in MMT.

Now we know that the negative charge resulted from octahedral substitution is

located at so-called ditrigonal cavities formed by oxygen atoms at the surface of silicate

129
a

b c d

Figure 6.22. Structure of MMT: (a) side view of layers in MMT; (b) top view of
tetrahedral layers of MMT; (c) the ditrigonal cavity formed among 6 oxygen atoms in
tetrahedral layers of MMT; and (d) coordination polyhedron formed by ditrigonal cavities
of two consecutive layers of oxygen atoms. In (b), (c), and (d), only oxygen atoms are
shown as black spots.148 (Reprinted with permission from Springer.)

130
a

cationically modified starch

Figure 6.23. Schematic of silicates layer in MMT with exchangeable cations: (a) top
view and (b) side view.

layers. Due to the electrostatic force, Na+ is absorbed at the surface of the silicate layer

and occupies the negatively charged cavities. Then the reason why there is no ionic

interaction between the positively charged −N+(CH3)3 in the modified starch (CS-B) and

the negatively charged surface of layered silicates in MMT in melt-blended

nanocomposites based on the (CS-B)/EVOH blend can be explained as follows: although

the surface of layered silicates is negatively charged, it is covered by a layer of positively

charged Na+. When the polymer molecules containing cationic moiety (cationically

modified starch in the present study) migrate into a gallery of layered silicates in MMT

aggregates, the presence of Na+ on the surface of silicates prevent the cationic moiety

131
from going close to the negatively charged ditrigonal cavities. Thus, the ionic interaction

could not be formed between the −N+(CH3)3 in the modified starch and negatively

charged ditrigonal cavities.

On the other hand, the ionic interaction between −N+(CH3)3 in the modified starch

(CS-B) and negatively charged ditrigonal cavities was found in nanocomposites prepared

by solution blending. It should be noted that water was used as a co-solvent in solution

blending. Because of its high hydrophilicity, the layered silicates of MMT change its

aggregation dramatically with the addition of water.147-149 When the dry MMT is

exposed to water, water molecules penetrate into the interlayer space of MMT. Because

of the high polarity of water, the negatively charged ditrigonal cavity at the surface of

silicates and the exchangeable cations will be hydrated.148,150-152 In this process, the

exchangeable cations move out of the ditrigonal cavities.147,148,152 The dissociation of an

ionic pair between the ditrigonal cavities and the exchangeable cations is schematically

shown in Figure 6.24.

Now we can explain why the ionic interaction between the positively charged

−N+(CH3)3 in the modified starch (CS-B) and the negatively charged surface of layered

silicates can be found in nanocomposites prepared by solution blending. The electrolytic

cationically modified starch

Figure 6.24. Schematic of electrolytic dissociation of ditrigonal cavity and the cations on
the surface of silicates in MMT.

132
dissociation of an ionic pair between the ditrigonal cavities and the exchangeable cations

resulted from water makes the negatively charged surfaces of silicates more accessible to

the positively charged −N+(CH3)3 in a modified starch. Then the ionic interaction can be

formed between the −N+(CH3)3 in the modified starch and negatively charged ditrigonal

cavities at the surface of silicates in MMT. Furthermore, some researchers148,149 reported

that the hydration of MMT in water increases the gallery distance of layered silicates.

The extent of hydration depends on the nature of the cations residing at the surface of

MMT aggregates. For Wyoming MMT saturated with Na+ residing at the surface of

layered silicates, the d001 spacing can be increased from 0.96 nm (without water) to 1.88

nm due to the increase of water partial pressure in the environment.148 The increase of

gallery distance of layered silicates in MMT resulted from water obviously helps achieve

a high degree of dispersion of MMT in a polymer matrix. However, it should be noted

that the strong attractive interaction between a polymeric matrix and the layered silicates

is the key factor in achieving a high degree of dispersion of MMT aggregate, not the

gallery distance of layered silicates.132,133

6.5 Crystallization and Crystalline Structures of EVOH in Nanocomposites

Figure 6.25 gives the DSC thermograms for the nanocomposites based on an

anionically modified starch AS-B or the (AS-B)/EVOH blend. It is worth noting that in

the nanocomposites containing organoclay, such as Cloisite 30B and Cloisite 15A, the

glass transition temperature has increased, from 85 oC ((AS-B)/EVOH blend) to

91 oC ((AS-B)/EVOH/Cloisite 30B nanocomposite) and 93 oC ((AS-B)/EVOH/

133
93 (a)
(b)
91
(c)
Endotherm

86
(d)
85 (e)
98 (f)
(g)
108

90

0 50 100 150 200


Temperature (oC)
Figure 6.25. DSC thermograms for: (a) (AS-B)/EVOH/Cloisite 15A nanocomposite; (b)
(AS-B)/EVOH/Cloisite 30B nanocomposite; (c) (AS-B)/EVOH/MMT nanocomposite; (d)
(AS-B)/EVOH blend; (e) (AS-B)/Cloisite 30B nanocomposite; (f) (AS-B)/Cloisite 15A
nanocomposite; (g) neat AS-B.

Cloisite 15A nanocomposite), and from 90 oC (neat AS-B) to 108 oC ((AS-B)/Cloisite

30B nanocomposite) and 98 oC ((AS-B)/Cloisite 15A nanocomposite). The increase in

the glass transition temperature was also observed in the nanocomposites based on neat

AS-LF and the (AS-LF)/EVOH blend, and Cloisite 30B or Cloisite 15A. The values of

glass transition temperatures determined from DSC thermograms are summarized in

Table 6.1. For the nanocomposites containing MMT, the glass transition temperature has

hardly changed. In previous sections, we have shown that the aggregates of Cloisite 15A

or Cloisite 30B have a high degree of dispersion (evidenced by XRD patterns and TEM

images) in the nanocomposites based on an anionically modified starch due to the ionic

134
Table 6.1. The changes in the thermal properties of nanocomposites*

Sample code ΔT g ΔTm ΔC


(oC) (oC) (100%)
Nanocomposites based on blends
AS-B/EVOH/Cloisite 30B 6 14 16.3
AS-B/EVOH/Cloisite 15A 8 8 7.1
AS-B/EVOH/MMT 1 1 1.2
AS-LF/EVOH/Cloisite 30B 7 16 18.7
AS-LF/EVOH/Cloisite 15A 10 7 5.6
AS-LF/EVOH/MMT 2 0 − 0.7
CS-B/EVOH/Cloisite 30B 1 2 0.7
CS-B/EVOH/Cloisite 15A 0 1 1.3
CS-B/EVOH/MMT 1 1 1.1
Nanocomposites based on neat starch
AS-B/Cloisite 30B 8 n/a n/a
AS-B/Cloisite 15A 18 n/a n/a
AS-B/MMT 0 n/a n/a
AS-LF/Cloisite 30B 12 n/a n/a
AS-LF/Cloisite 15A 9 n/a n/a
AS-LF/MMT 1 n/a n/a
CS-B/Cloisite 30B 2 n/a n/a
CS-B/Cloisite 15A 1 n/a n/a
CS-B/MMT 14 n/a n/a

*
The values listed in the table above are the difference between nanocomposites and
their blends or neat materials. The thermal properties of blend and neat material samples
were listed in Tables 5.1 and 5.2.

interaction between the negatively charged carboxylate group in a modified starch and

the positively charged N+ in the surfactant residing at the surface of organoclay. On the

other hand, MMT has quite poor dispersion characteristics in the matrix, because there is

no strong attractive interaction formed between MMT and the matrix.

In the nanocomposites based on the blends of an anionically modified starch and

EVOH, the reason for the observed increase in glass transition temperature lies in the

135
strong attractive interaction between the modified starch and silicates. However, the

possibility of forming attractive interaction between EVOH and the surfactant residing at

the surface of organoclay cannot be ruled out. The surfactant residing at the surface of

Cloisite 30B has hydroxyl groups which can form hydrogen bonding with EVOH.

Figures 6.26 to 6.28 give WAXD patterns of the nanocomposites based on

anionically modified starch (AS-B) and EVOH, and Cloisite 15A, Cloisite 30B, or MMT.

Referring to Figures 6.26 and 6.28, for nanocomposites containing Cloisite 15A or MMT,

the crystalline structure of EVOH in quenched samples takes orthorhombic lattice (the

single (110) reflection peak on WAXD patterns) and evolves into monoclinic lattice (the

(110) and (110) reflection peaks on WAXD patterns) as the crystallization temperature

was increased, which is similar to that of EVOH in the blends with a modified starch (see

Chapter V). On the other hand, referring to Figure 6.27 for the WAXD patterns for the

nanocomposite containing Cloisite 30B, we find that the crystalline structure of EVOH

takes orthorhombic lattice in quenched samples and does not evolve into monoclinic

lattice as the crystallization temperature is increased, which is different from that of

EVOH in the blends with a modified starch (see Chapter V). One should keep in mind

that both Cloisite 30B and Cloisite 15A have a very high degree of dispersion

characteristics in the nanocomposites based on AS-B and EVOH, whereas MMT has

poor dispersion characteristics in the nanocomposite, as described in Section 6.3.

For EVOH in the nanocomposites prepared in the present study, the interaction

between the hydroxyl groups in EVOH and the surfactant residing at the surface of

silicate sheets of Cloisite 30B could change the vinyl alcohol content on the molecules of

EVOH locally (in the vicinity of silicate sheets, but silicate sheets are exfoliated and

136
monoclinic lattice group

Intensity (arb. unit)


(1)

(2)

(3)

(4)
orthohombic lattice group

16 17 18 19 20 21 22 23 24 25
2θ (degree)
Figure 6.26. WAXD patterns for (AS-B)/EVOH/Cloisite 15A nanocomposite at different
annealing temperatures (oC): (1) 153; (2) 133; (3) 113; and (4) quench.
Intensity (arb. unit)

(1)
(2)
(3)
(4)

16 17 18 19 20 21 22 23 24 25
2θ (degree)
Figure 6.27. WAXD patterns for (AS-B)/EVOH/Cloisite 30B nanocomposite at different
annealing temperatures (oC): (1) 151; (2) 131; (3) 111; and (4) quench.

137
Intensity (arb. unit)

(1)
(2)
(3)
(4)

16 17 18 19 20 21 22 23 24 25
2θ (degree)

Figure 6.28. WAXD patterns for (AS-B)/EVOH/MMT nanocomposite at different


annealing temperatures (oC): (1) 146; (2) 126; (3) 106; and (4) quench.

highly dispersed all over the sample), and the crystalline structure predominantly takes

orthorhombic lattice groups. Furthermore, due to the rigid nature of silicate sheets

(compared with flexible polymer chains), the local change of vinyl alcohol content due to

the interaction between EVOH and silicate sheets does not vary with temperature nor the

hydrogen bonds between EVOH and the surfactant. That is the reason why a

transformation of crystalline structure of EVOH from orthorhombic lattice into

monoclinic lattice does not occur in the nanocomposite containing Cloisite 30B.

On the other hand, in the nanocomposites containing Cloisite 15A, its surfactant

does not have hydroxyl groups, and there is no attractive interaction between the silicate

138
sheets and EVOH. Thus, the crystalline structure of EVOH is quite the same as that of

neat EVOH.

Figures 6.29 to 6.31 give WAXD patterns for the nanocomposites based on the

cationically modified starch (CS-B) and EVOH, and Cloisite 15A, Cloisite 30B, or MMT.

One should keep in mind that the aggregates of Cloisite 15A, Cloisite 30B, and MMT

have a poor dispersion in the melt-blended nanocomposites based on the CS-B and

EVOH, as described in Section 6.4. Referring to Figures 6.29 and 6.31, the

nanocomposites containing Cloisite 15A or MMT have the similar crystalline structure of

EVOH to that of neat EVOH under different annealing temperatures, namely, the

crystalline structure of EVOH takes orthorhombic lattice in quenched sample and evolves
Intensity (arb. unit)

(1)

(2)
(3)
(4)

16 17 18 19 20 21 22 23 24 25
2θ (degree)

Figure 6.29. WAXD patterns for (CS-B)/EVOH/Cloisite 15A nanocomposite at different


annealing temperatures (oC): (1) 143; (2) 123; (3) 103; and (4) quench.

139
Intensity (arb. unit)
(1)

(2)

(3)

(4)

16 17 18 19 20 21 22 23 24 25
2θ (degree)
Figure 6.30. WAXD patterns for (CS-B)/EVOH/Cloisite 30B nanocomposite at different
annealing temperatures (oC): (1) 144; (2) 124; (3) 104; and (4) quench.
Intensity (arb. unit)

(1)
(2)

(3)
(4)

16 17 18 19 20 21 22 23 24 25
2θ (degree)

Figure 6.31. WAXD patterns for (CS-B)/EVOH/MMT nanocomposite at different


annealing temperatures (oC): (1) 144; (2) 124; (3) 104; and (4) quench.

140
into monoclinic lattice as the crystallization temperature is increased. Referring to Figure

6.30, the nanocomposite containing Cloisite 30B also has the similar crystalline structure

of EVOH to that of neat EVOH. Although there are hydroxyl groups in the surfactant

residing at the surface of Cloisite 30B aggregates, which might form hydrogen bonds

with EVOH, the poor dispersion characteristics of Cloisite 30B aggregates limit the

contact area between the silicate sheets and EVOH.

6.6 Mechanical Properties of Nanocomposites

Figures 6.32 to 6.35 give the tensile properties (tensile strength, Young’s modulus,

and elongation at break) of the nanocomposites based on a modified starch and EVOH,

and Cloisite 15A, Cloisite 30B, or MMT. In Chapter V, we have presented the tensile

properties of the blends based on a modified starch and EVOH (see Figures 5.19 to 5.22).

For the tensile properties of the nanocomposites based on an anionically modified

starch (AS-B or AS-LF) and EVOH, it is worth noting that the tensile strength and

Young’s modulus are enhanced in the nanocomposites containing organoclays (Cloisite

15A or Cloisite 30B), whereas the elongation at break is decreased in the same

nanocomposites. On the other hand, the tensile properties (tensile strength, Young’s

modulus, and elongation at break) of the nanocomposite containing MMT have not

changed much as compared with those of blends. The difference in the tensile properties

between the nanocomposites containing organoclay and the one containing MMT may be

explained as follows. The improved mechanical properties may be attributable to the

presence of the attractive interaction and large surface areas between the silicate sheets

and the matrix. In the nanocomposites containing an organoclay, the ionic interaction

141
140 7.0
130
(a) (b)
6.5

Young's modulus (GPa)


Tensile strength (MPa)

120
110 6.0
100
90 5.5

80
5.0
70
60 4.5
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction) EVOH (wt fraction)

60
(c)
50
Elongation at break (%)

40

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction)

Figure 6.32. Tensile properties of the nanocomposites based on AS-B and EVOH: (a)
tensile strength; (b) Young’s modulus; and (c) elongation at break. (●) (AS-B)/EVOH
blend; (□) (AS-B)/EVOH/Cloisite 30B nanocomposite; (△) (AS-B)/EVOH/Cloisite 15A
nanocomposite; and (▽) (AS-B)/EVOH/MMT nanocomposite.

142
140 7.5
(a) 7.0 (b)

Young's Modulus (GPa)


120
Tensile strength (MPa)

6.5

100 6.0
5.5
80 5.0
4.5
60
4.0
40 3.5
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction) EVOH (wt fraction)

60
(c)
50
Elongation at Break (%)

40

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction)

Figure 6.33. Tensile properties of the nanocomposites based on AS-LF and EVOH: (a)
tensile strength; (b) Young’s modulus; and (c) elongation at break. (●) (AS-LF)/EVOH
blend; (□) (AS-LF)/EVOH/Cloisite 30B nanocomposite; (△) (AS-LF)/EVOH/Cloisite
15A nanocomposite; and (▽) (AS-LF)/EVOH/MMT nanocomposite.

143
140 7.0
(a) (b)
120

Young's Modulus (GPa)


6.5
Tensile strength (MPa)

100 6.0

80 5.5

60 5.0

40 4.5
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction) EVOH (wt fraction)

60
(c)
Elongation at break (%)

50

40

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction)

Figure 6.34. Tensile properties of the nanocomposites based on CS-B and EVOH: (a)
tensile strength; (b) Young’s modulus; and (c) elongation at break. (●) (CS-B)/EVOH
blend; (□) (CS-B)/EVOH/Cloisite 30B nanocomposite; (△) (CS-B)/EVOH/Cloisite 15A
nanocomposite; and (▽) (CS-B)/EVOH/MMT nanocomposite.

144
140 7
(a) (b)
120

Young's Modulus (GPa)


Tensile strength (MPa)

100
5
80

4
60

40 3
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction) EVOH (wt fraction)

60
(c)
50
Elongation at break (%)

40

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
EVOH (wt fraction)

Figure 6.35. Tensile properties of the nanocomposites based on CS-LF and EVOH: (a)
tensile strength; (b) Young’s modulus; and (c) elongation at break. (●) (CS-LF)/EVOH
blend; (□) (CS-LF)/EVOH/Cloisite 30B nanocomposite; (△) (CS-LF)/EVOH/Cloisite
15A nanocomposite; and (▽) (CS-LF)/EVOH/MMT nanocomposite.

145
between the negatively charged carboxylate groups in a modified starch and the

positively charged N+ in the surfactant residing at the surface of an organoclay exfoliated

the layered silicates, yielding nanometer-sized platelets. These nanometer-sized platelets

brought large surface areas in contact with the matrix. The attractive interaction between

the silicate sheets and polymer molecules acts like a cross-linking point in the

nanocomposites. From the point of view of tensile properties in the nanocomposites

under the tensile tests, the stress is much more efficiently transferred from the polymer

matrix to the inorganic silicate sheets through the attractive interactions, resulting in

higher tensile strength and modulus. Referring to Figures 6.32 to 6.33, we find that the

nanocomposite containing Cloisite 30B has higher tensile strength and Young’s modulus

than the nanocomposite containing Cloisite 15A. On the other hand, for the

nanocomposite containing MMT, which has no functional groups to form attractive

interaction with the matrix, the tensile properties of the nanocomposite were not

improved as compared with those of the blend based on the same matrix.

On the other hand, referring to Figures 6.34 and 6.35, the tensile properties of the

nanocomposites based on a cationically modified starch (CS-B or CS-LF) and EVOH

were not improved by the addition of nanoclays (Cloisite 15A, Cloisite 30B, or MMT).

In the previous section, we have shown that the nanocomposites based on a cationically

modified starch and EVOH have poor dispersion characteristics of the aggregates of

Cloisite 15A, Cloisite 30B, or MMT.

146
6.7 Concluding Remarks

Nanocomposites based on a modified starch and EVOH were prepared in the

present study. The starch was modified with ionic groups (acetate sodium group or

trimethylammonium chloride group) and hydrophobic groups (hexanoyl group or benzoyl

group). The rationale behind the approach was to introduce ionic interaction between the

modified starch and nanoclay, so that the nanocomposites could have a high degree of

dispersion of nanoclay aggregates.

The dispersion characteristics of nanocomposites based on the anionically modified

starch (AS-B or AS-LF) or the cationically modified starch (CS-B or CS-LF) were

investigated by XRD and TEM. For the nanocomposites based on AS-B or AS-LF, the

experimental results have shown that organoclay aggregates (Cloisite 30B and Cloisite

15A) have a high degree of dispersion characteristics, whereas natural clay (MMT)

aggregates have very poor dispersion characteristics. We have ascribed the high degree

of dispersion of organoclay aggregates in nanocomposites based on the anionically

modified starch and EVOH to the ionic interaction formed between the negatively

charged carboxylate groups (−COO−) in the modified starch and positively charged N+ in

a surfactant (MT2EtOH or 2M2HT) residing at the surface of layered silicates of

organoclays. The FTIR spectra of nanocomposites confirmed the presence of ionic

interaction in the nanocomposites. On the other hand, there is no functional group

residing at the surface of MMT to form any specific attractive interaction with the matrix

in the nanocomposites, which we ascribed to the poor dispersion characteristics of MMT

aggregates.

147
For the nanocomposites based on the cationically modified starch (CS-B or CS-LF)

and EVOH, the experimental results have shown that both organoclays (Cloisite 30B or

Cloisite 15A) and natural clay (MMT) gave rise to poor dispersion characteristics in

nanocomposites. We have interpreted the results by presenting evidence that no

attractive interaction existed between the clay and the cationically modified starch.

However, we have learned that the nanocomposite based on the same cationically

modified starch, which was prepared by solution blending, has a very high degree of

dispersion of MMT aggregates. We ascribed the high degree of dispersion of MMT

aggregates in the nanocomposite to the ionic interaction between the negatively charged

surface of MMT and positively charged −N+(CH3)3 in the modified starch. The FTIR

spectra confirmed the presence of ionic interaction in the nanocomposite. Then we

explained why ionic interactions existed between the MMT and cationically modified

starch in the nanocomposites prepared by solution blending.

The nanocomposites prepared in the present study were characterized by DSC and

WAXD. The glass transition temperature was increased in those nanocomposites which

have a high degree of dispersion of organoclay aggregates. We ascribed the increased

glass transition temperature to the ionic interaction between the aggregates of organoclay

and matrix in the nanocomposites. The DSC thermograms showed that the melting point

and melting enthalpy of EVOH in the nanocomposites containing the aggregates of

Cloisite 30B were increased, suggesting that the dispersion characteristics of the

organoclay influenced the crystallization of EVOH in the nanocomposites. The WAXD

patterns have shown that in the nanocomposites containing exfoliated Cloisite 30B, the

crystalline structure of EVOH takes orthorhombic lattice in the quenched and annealed

148
samples at various annealing temperatures. In the nanocomposites based on EVOH and a

modified starch, and Cloisite 15A or MMT, the crystalline structure of EVOH was found

to be similar to that of neat EVOH. We ascribed the difference in crystallization and the

crystalline structure of EVOH between the nanocomposite containing Cloisite 30B and

the nanocomposites containing other clays (Cloisite 15A and MMT) to the presence of

hydrogen bonds formed between the hydroxyl group in EVOH and the hydroxyl group in

the surfactant (MT2EtOH) residing at the surface of silicate sheets of Cloisite 30B.

The tensile properties of the nanocomposites prepared in the present study were

characterized. We have found that an improvement in the tensile properties of the

nanocomposites based on a modified starch and EVOH can only be obtained if the clay

has a very high degree of dispersion in the nanocomposites, and strong attractive

interaction is present between the silicate sheets of clay and the matrix.

149
CHAPTER VII

CONCLUSIONS AND RECOMMENDATIONS

7.1 Conclusions

In this dissertation, we have synthesized four chemically modified starches

functionalized by ionic and hydrophobic groups: (1) a starch modified with anionic group

and benzoyl group (AS-B); (2) a starch modified with anionic group and hexanoyl group

(AS-LF); (3) a starch modified with cationic group and benzoyl group (CS-B); and (4) a

starch modified with cationic group and hexanoyl group (CS-LF). Subsequently, the

degree of substitution (DS) of the modified starches was determined by calculations from

the 1H-NMR spectra. The FTIR spectra have indicated that the strength of hydrogen

bonds in starch was weakened by chemical modification. The thermal stability of the

modified starches was also characterized. We have found that the thermal degradation

temperature was drastically decreased from over 300 oC to below 200 oC for the starches

modified only by ionic groups. However, when a starch was modified by both ionic and

hydrophobic groups, the thermal degradation temperature was increased, because the

number of hydroxyl groups in the starch was decreased. The glass transition temperature

of dried natural starch could not be determined because it is higher than the thermal

degradation temperature. We have found that the glass transition temperature of the

modified starch was decreased to below 100 oC, which would be useful for processing

starch as a thermoplastic material. The chemical modification destroyed the granular

150
semicrystalline structure and the hydrogen bonds in natural starch, consequently

increasing the mobility of starch molecules. No melting peak was found in the DSC

thermograms of the modified starches, suggesting that the modified starches are

amorphous.

Because of the chemical modification employed, we have changed starch from

hydrophilic to hydrophobic. The composting tests conducted showed that the chemically

modified starches have good biodegradability. However, the rate of weight loss of the

modified starches during compositing was slower than natural starch, suggesting that the

biodegradability of starch was lowered by chemical modification.

We also prepared the blends based on a modified starch and EVOH. Phase

diagrams of the modified starch/EVOH blends were obtained, and the blends exhibited

LCST. The DSC thermograms of the blends of a modified starch and EVOH showed a

single glass transition temperature, suggesting a good miscibility between the modified

starch and EVOH. We ascribed the miscibility of the blends to the presence of hydrogen

bonds, which were formed between the modified starch and EVOH, as confirmed by

FTIR spectroscopy. The DSC thermograms of the blends of a modified starch and

EVOH also showed that the melting point and melting enthalpy of EVOH were decreased,

suggesting that the crystallization of EVOH was interrupted. We ascribed the

interruption of crystallization to the presence of hydrogen bonds between the modified

starch and EVOH. The FTIR spectra showed that the hydrogen bonds formed between

the modified starch and EVOH was favored over the hydrogen bonds by self-association

within neat EVOH. In neat EVOH, the WAXD patterns have shown that the crystalline

structure of EVOH takes orthorhombic lattice in the quenched specimens and evolves

151
into monoclinic lattice when annealing temperature was increased. We have found that

the crystalline structure of EVOH in the blends of a modified starch and EVOH is similar

to that of neat EVOH, suggesting that the hydrogen bonds between the modified starch

and EVOH do not influence the crystalline structure of EVOH. Synergistic effects were

observed in the tensile properties of the blends of a modified starch and EVOH, owing to

a good miscibility between the modified starch and EVOH.

Nanocomposites were prepared based on a modified starch and EVOH. The

nanoclay employed in the preparation of nanocomposites include natural clay (MMT)

and two commercial organoclays (Cloisite 30B and Cloisite 15A). We have found that in

the nanocomposites based on the anionically modified starch (AS-B or AS-LF), the

aggregates of an organoclay have a very high degree of dispersion, whereas the

aggregates of MMT have poor dispersion characteristics. We ascribed the experimental

observations to the ionic interaction formed between the modified starch and the

organoclay. The FTIR spectra of the nanocomposites prepared have confirmed the

presence of ionic interaction between the negatively charged carboxylate groups in a

modified starch and positively charged N+ in a surfactant residing at the surface of

organoclay.

We have found poor dispersion characteristics of both organoclays (Cloisite 30B

and Cloisite 15A) and MMT in the nanocomposites based on the cationically modified

starch, when the nanocomposites were prepared by melt blending. In the nanocomposites

prepared by solution blending, we have found that MMT aggregates have a very high

degree of dispersion in the cationically modified starch. The FTIR spectra showed that

152
ionic interaction existed between the positively charged −N+(CH3)3 in the modified starch

and negatively charged surface of MMT aggregates.

To investigate how the dispersion characteristics of nanoclay aggregates might

influence the morphology and crystalline structure of nanocomposites, we characterized

the nanocomposites prepared by DSC and WAXD. The DSC thermograms have shown

that the glass transition temperature was increased in the nanocomposites having a high

degree of dispersion of organoclay aggregates. The WAXD patterns have shown that the

crystalline structure of EVOH in the nanocomposites containing exfoliated Cloisite 30B

takes orthorhombic lattice in both the quenched and annealed samples. On the other hand,

the crystalline structure of EVOH in the nanocomposites containing Cloisite 15A and

MMT is similar to that of neat EVOH.

The tensile properties of the nanocomposites were also measured. We have found

that an improvement in the tensile properties of the nanocomposites based on a modified

starch and EVOH could only be obtained when the nanoclay aggregates have a very high

degree of dispersion in the nanocomposites, which was attributable to the presence of

strong attractive interaction between the silicate sheets of clay and the matrix.

7.2 Recommendations

The present study has shown that the rate of biodegradation of a chemically

modified starch was lowered as compared to that of natural starch. The weight loss

during composting showed two different patterns between the cationically modified

starch and the anionically modified starch, suggesting that the functional groups might

have changed the biodegradation mechanism of starch. It is recommended that an

153
investigation be undertaken to find out the mechanism(s) of biodegradation of chemically

modified starch during composting.

We have found that the modified starch synthesized in the present study is miscible

with EVOH. It is recommended that an investigation be undertaken to find out what

degree of substitution of modified starch and the vinyl alcohol content in EVOH might

influence miscibility between the modified starch and EVOH.

The biodegradable starch-based material is a potential candidate for the replacement

of petroleum-based synthetic packaging materials widely used in the current market. We

recommended that the processability of modified starch be investigated.

154
REFERENCES

(1) Nikuni, Z. "Studies on Starch Granules." Starch/Staerke 1978, 30, 105.

(2) Sinha Ray, S.; Okamoto, M. "Polymer/layered silicate nanocomposites: a review from
preparation to processing." Progress in Polymer Science 2003, 28, 1539.

(3) Theng, B. K., The Chemistry of Clay-Organic Reactions. John & Wiley: New York,
1974.

(4) Grim, R. E., Clay Minerology. Second Edition ed.; McGraw Hill: New York, 1968.

(5) Wilhelm, H.-M.; Sierakowski, M.-R.; Souza, G. P.; Wypych, F. "The influence of
layered compounds on the properties of starch/layered compound composites." Polymer
International 2003, 52, 1035.

(6) Park, H.-M.; Li, X.; Jin, C.-Z.; Park, C.-Y.; Cho, W.-J.; Ha, C.-S. "Preparation and
Properties of Biodegradable Thermoplastic Starch/Clay Hybrids." Macromolecular
Materials and Engineering 2002, 287, 553.

(7) Yang, Z.; Han, C. D. "Rheology of Miscible Polymer Blends with Hydrogen
Bonding." Macromolecules 2008, 41, 2104.

(8) Hizukuri, S.; Takeda, Y.; Maruta, N.; Juliano, B. O. "Molecular structures of rice
starch." Carbohydrate Research 1989, 189, 227.

(9) Takeda, Y.; Hizukuri, S.; Takeda, C.; Suzuki, A. "Structures of branched molecules of
amyloses of various origins, and molar fractions of branched and unbranched molecules "
Carbohydrate Research 1987, 165, 139.

(10) Hizukuri, S.; Takeda, Y.; Yasuda, M.; Suzuki, A. "Multi-branched nature of amylose
and the action of debranching enzymes " Carbohydrate Research 1981, 94, 205.
155
(11) Imberty, A.; Chanzy, H.; Pérez, S.; Bulèon, A.; Tran, V. "The double-helical nature
of the crystalline part of A-starch." Journal of Molecular Biology 1988, 201, 365.

(12) Imberty, A.; Buléon, A.; Tran, V.; Péerez, S. "Recent Advances in Knowledge of
Starch Structure." Starch/Staerke 1991, 43, 375.

(13) Buléon, A.; Colonna, P.; Planchot, V.; Ball, S. "Starch granules: structure and
biosynthesis." International Journal of Biological Macromolecules 1998, 23, 85.

(14) Colonna, P.; Buleon, A.; Lemaguer, M.; Mercier, C. "Pisum sativum and vicia faba
carbohydrates: Part IV — Granular structure of wrinkled pea starch " Carbohydrate
Polymers 1982, 2, 43.

(15) Zobel, H. F. "Molecules to Granules: A Comprehensive Starch Review."


Starch/Staerke 1988, 40, 44.

(16) Gernat, C.; Dipl.-Krist; Radosta, S.; Anger, H.; Damaschun, G. "Crystalline Parts of
Three Different Conformations Detected in Native and Enzymatically Degraded
Starches." Starch/Staerke 1993, 45, 309.

(17) French, D. "Chemical and physical properties of starch." Journal of Animal Science
1973, 37, 1048.

(18) Kainuma, K.; French, D. "Naegeli amylodextrin and its relationship to starch granule
structure. II. Role of water in crystallization of B-starch." Biopolymers 1972, 11, 2241.

(19) Mizukami, H.; Takeda, Y.; Hizukuri, S. "The structure of the hot-water soluble
components in the starch granules of new Japanese rice cultivars." Carbohydrate
Polymers 1999, 38, 329.

(20) Bizot, H.; Le Bail, P.; Leroux, B.; Davy, J.; Roger, P.; Buleon, A. "Calorimetric
evaluation of the glass transition in hydrated, linear and branched polyanhydroglucose
compounds." Carbohydrate Polymers 1997, 32, 33.

(21) Van Soest, J. J. G.; Knooren, N. "Influence of glycerol and water content on the
structure and properties of extruded starch plastic sheets during aging." Journal of
Applied Polymer Science 1997, 64, 1411.

156
(22) Lourdin, D.; Ring, S. G.; Colonna, P. "Study of plasticizer-oligomer and plasticizer-
polymer interactions by dielectric analysis: maltose-glycerol and amylose-glycerol-water
systems." Carbohydrate Research 1998, 306, 551.

(23) Lourdin, D.; Bizot, H.; Colonna, P. ""Antiplasticization" in starch-glycerol films?"


Journal of Applied Polymer Science 1997, 63, 1047.

(24) Hulleman, S. H. D.; Helbert, W.; Chanzy, H. "Single crystals of V amylose


complexed with glycerol." International Journal of Biological Macromolecules 1996, 18,
115.

(25) Gaudin, S.; Lourdin, D.; Forssell, P. M.; Colonna, P. "Antiplasticisation and oxygen
permeability of starch–sorbitol films." Carbohydrate Polymers 2000, 43, 33.

(26) Miles, M. J.; Morris, V. J.; Ring, S. G. "Gelation of amylose." Carbohydrate


Research 1985, 135, 257.

(27) Putaux, J.-L.; Buléon, A.; Chanzy, H. "Network Formation in Dilute Amylose and
Amylopectin Studied by TEM." Macromolecules 2000, 33, 6416.

(28) Gidley, M. J. "Molecular mechanisms underlying amylose aggregation and


gelation." Macromolecules 1989, 22, 351.

(29) Gidley, M. J.; Bulpin, P. V. "Aggregation of amylose in aqueous systems: the effect
of chain length on phase behavior and aggregation kinetics." Macromolecules 1989, 22,
341.

(30) Shi, Y.-C.; Seib, P. A. "The structure of four waxy starches related to gelatinization
and retrogradation " Carbohydrate Research 1992, 227, 131.

(31) Ring, S. G.; Colonna, P.; I'Anson, K. J.; Kalichevsky, M. T.; Miles, M. J.; Morris, V.
J.; Orford, P. D. "The gelation and crystallisation of amylopectin " Carbohydrate
Research 1987, 162, 277.

(32) Kalichevsky, M. T.; Orford, P. D.; Ring, S. G. "The retrogradation and gelation of
amylopectins from various botanical sources " Carbohydrate Research 1990, 198, 49.

157
(33) Shi, Y.-C.; Seib, P. A. "Fine structure of maize starches from four wx-containing
genotypes of the W64A inbred line in relation to gelatinization and retrogradation "
Carbohydrate Polymers 1995, 26, 141.

(34) Würsch, P.; Gumy, D. "Inhibition of amylopectin retrogradation by partial beta-


amylolysis " Carbohydrate Research 1994, 256, 129.

(35) Gotlieb, K. F., Starch derivatization : fascinating and unique industrial


opportunities. Wageningen Academic Publishers, The Netherlands: 2005.

(36) Cui, S. W., Food carbohydrates : chemistry, physical properties, and applications
Taylor & Francis/CRC Press: Boca Raton, FL, USA, 2005.

(37) Tomasik, P.; Wang, Y. J.; Jane, J. L. "Complexes of Starch with Dioic Acids."
Starch/Staerke 1995, 47, 91.

(38) Tomasik, P.; Zaranyika, M. F. "Nonconventional methods of modification of


starch." Advances in Carbohydrate Chemistry and Biochemistry 1995, 51, 243.

(39) Jacobs, H.; Eerlingen, R. C.; Rouseu, N.; Colonna, P.; Delcour, J. A. "Acid
hydrolysis of native and annealed wheat, potato and pea starches—DSC melting features
and chain length distributions of lintnerised starches." Carbohydrate Research 1998, 308,
359.

(40) Hoover, R. "Acid-treated Starches." Food Review International 2000, 16, 369.

(41) Jyothi, A. N.; Sasikiran, K.; Sajeev, M. S.; Revamma, R.; Moorthy, S. N.
"Gelatinisation Properties of Cassava Starch in the Presence of Salts, Acids and
Oxidising Agents." Starch/Staerke 2005, 57, 547.

(42) Mussulman, W. C.; Wagoner, J. A. "Electron microscopy of unmodified and acid-


modified cornstarches." Cereal Chemistry 1968, 45, 162.

(43) Komiya, T.; Nara, S. "Changes in Crystallinity and Gelatinization Phenomena of


Potato Starch by Acid Treatment." Starch/Staerke 1986, 38, 9.

158
(44) Whistler, R. L.; BeMiller, J. N.; Paschall, E. F., Starch : chemistry and technology
Academic Press: Orlando, USA, 1984.

(45) Kainuma, K.; French, D. "Nägeli amylodextrin and its relationship to starch granule
structure. I. Preparation and properties of amylodextrins from various starch types."
Biopolymers 1971, 10, 1673.

(46) Jane, J.-l.; Wong, K.-s.; McPherson, A. E. "Branch-structure difference in starches


of A- and B-type X-ray patterns revealed by their Naegeli dextrins." Carbohydrate
Research 1997, 300, 219.

(47) Muhr, A. H.; Blanshard, J. M. V.; Bates, D. R. "The Effect of Lintnerisation on


Wheat and Potato Starch Granules." Carbohydrate Polymers 1984, 4, 399.

(48) Hoover, R.; Swamidas, G.; Vasanthan, T. "Studies on the physicochemical


properties of native, defatted, and heat-moisture treated pigeon pea (Cajanus cajan L)
starch " Carbohydrate Research 1993, 246, 185.

(49) Morrison, W. R.; Tester, R. F.; Gidley, M. J.; Karkalas, J. "Resistance to acid
hydrolysis of lipid-complexed amylose and lipid-free amylose in lintnerised waxy and
non-waxy barley starches " Carbohydrate Research 1993, 245, 289.

(50) Lauro, M.; Ring, S. G.; Bull, V. J.; Poutanen, K. "Gelation of Waxy Barley Starch
Hydrolysates." Journal of Cereal Science 1997, 26, 347.

(51) Watanabe, T.; French, D. "Structural features of naegeli amylodextrin as indicated


by enzymic degradation." Carbohydrate Research 1980, 84, 115.

(52) Matsunaga, N.; Seib, P. A. "Extraction of Wheat Starch with Aqueous Sodium
Hydroxide." Cereal Chemistry 1997, 74, 851.

(53) Kyungsoo, W.; Seib, P. A. "Cross-linking of wheat starch and hydroxypropylated


wheat starch in alkaline slurry with sodium trimetaphosphate." Carbohydrate Polymers
1997, 33, 263.

(54) Krochta, J. M.; Tillin, S. J.; Hudson, J. S. "Degradation of polysaccharides in


alkaline solution to organic acids: Product characterization and identification " Journal of
Applied Polymer Science 1987, 33, 1483.
159
(55) Krochta, J. M. "Thermochemical conversion of polysaccharides in concentrated
alkali to glycolic acid." Applied Biochemistry and Biotechnology 1988, 17, 23.

(56) Whistler, R. L.; Chang, P. K.; Richards, G. N. "Alkaline Degradation of Periodate-


oxidized Starch." Journal of the American Chemical Society 1959, 81, 3133.

(57) Hollingsworth, R. I. "The chemical degradation of starch: Old reactions and new
frontiers." Biotechnology Annual Review 1996, 2, 281.

(58) Wurzburg, O. B., Modified Starches: Properties and Uses. CRC Press, Inc: Boca
Raton, Florida, 1986.

(59) Caesar, G. V. "Starch nitrate." Advances in Carbohydrate Chemistry 1958, 13, 331.

(60) Mustafa, A.; Dawoud, A. F.; El-Shorbani, S. "Further Studies on Stabilization of


Polymeric Carbohydrate Nitrates of Starch, Amylose, Amylopectin and Glycogen."
Starch/Staerke 1968, 20, 55.

(61) Israelashvili, S. "Mechanism of the nitration of starch." Nature 1950, 165, 686.

(62) Mustafa, A.; Dawoud, A. F.; El-Shorbani, S. "Stability of the nitrates of starch,
amylose, and amylopectin." Starch/Staerke 1967, 19, 212.

(63) Tomasik, P.; Schilling, C. H. "Chemical Modification of Starch." Advances in


Carbohydrate Chemistry and Biochemistry 2004, 59, 175.

(64) Mustafa, A.; Dawoud, A. F.; Marawan, A. "Stabilization of polymeric carbohydrate


nitrates. Stability of dextrin nitrates." Starch/Staerke 1967, 19, 358.

(65) Grote, C.; Heinze, T. "Starch Derivatives of High Degree of Functionalization 11:
Studies on Alternative Acylation of Starch with Long-chain Fatty Acids Homogeneously
in N,N-dimethyl acetamide/LiCl." Cellulose 2005, 12, 435.

(66) Sagar, A. D.; Merrill, E. W. "Properties of Fatty-Acid Esters of Starch." Journal of


Applied Polymer Science 1995, 58, 1647.

160
(67) Thieboud, S.; Aburto, J.; Alric, I.; Borredon, E.; Bikiaris, D.; Prinos, J.; Panayiotou,
C. "Properties of fatty-acid esters of starch and their blends with LDPE." Journal of
Applied Polymer Science 1997, 65, 705.

(68) Aburto, J.; Alric, I.; Borredon, E. "Preparation of long-chain esters of starch using
fatty acid chlorides in the absence of solvent." Starch/Staerke 1999, 51, 132.

(69) Aburto, J.; Thieboud, S.; Alric, I.; Borredon, E.; Bikiaris, D.; Prinos, J.; Panayiotou,
C. "Properties of octanoated starch and its blends with polyethylene." Carbohydrate
Polymers 1997, 34, 101.

(70) Aburto, J.; Alric, I.; Thieboud, S.; Borredon, E.; Prinos, J.; Panayiotou, C.
"Synthesis, characterization, and biodegradability of fatty-acid esters of amylose and
starch." Journal of Applied Polymer Science 1999, 74, 1440.

(71) Aburto, J.; Hamaili, H. "Free-solvent synthesis and properties of higher fatty esters
of starch – Part 2." Starch/Staerke 1999, 51, 302.

(72) Bikiaris, D.; Aburto, J.; Alric, I.; Borredon, E.; Botev, M.; Betchev, C.; Panayiotou,
C. "Mechanical properties and biodegradability of LDPE blends with fatty-acid esters of
amylose and starch." Journal of Applied Polymer Science 1999, 71, 1089.

(73) Morita, H. "Characterization of starch and related polysaccharides by differential


thermal analysis." Analytical Chemistry 1956, 28, 64.

(74) Singh, V.; Tiwari, A. "Microwave-accelerated methylation of starch." Carbohydrate


Research 2008, 343, 151.

(75) van der Burgt, Y. E. M.; Bergsma, J.; Bleeker, I. P.; Mijland, P. J. H. C.; van der
Kerk-van Hoof, A.; Kamerling, J. P.; Vliegenthart, J. F. G. "Distribution of methyl
substituents in amylose and amylopectin from methylated potato starches." Carbohydrate
Research 2000, 325, 183.

(76) Yang, B. Y.; Montgomery, R. "Acylation of Starch using Trifluoroacetic Anhydride


Promoter." Starch/Staerke 2006, 58, 520.

161
(77) Steeneken, P. A. M.; Tas, A. C.; Woortman, A. J. J.; Sanders, P.; Mijland, P. J. H. C.;
de Weijs, L. G. R. "Substitution patterns in methylated potato starch as revealed from the
structure and composition of fragments in enzymatic digests." Carbohydrate Research
2008, 343, 2411.

(78) Trimnell, D.; Stout, E. I.; Doane, W. M.; Russell, C. R. "Preparation of starch 2-
hydroxy-3-mercaptopropyl ethers and their use in graft polymerizations." Journal of
Applied Polymer Science 1978, 22, 3579.

(79) Petzold, K.; Klemm, D.; Stein, A.; Günther, W. "Synthesis and NMR
characterization of regiocontrolled starch alkyl ethers " Designed Monomers & Polymers
2002, 5, 415.

(80) Steeneken, P. A. M. "Topochemical effects in the methylation of starch."


Carbohydrate Research 1991, 209, 239.

(81) Cho, K. Y.; Lim, S. T. "Preparation and Properties of Benzyl Corn Starches."
Starch/Staerke 1998, 50, 250.

(82) Mischnick-Lubbecke, P.; Koenig, W. A. "Determination of the substitution pattern


of modified polysaccharides. Part I. Benzyl starches." Carbohydrate Research 1989, 185,
113.

(83) Cao, X.; Wang, Y.; Zhang, L. "Effects of Ethyl and Benzyl Groups on the
Miscibility and Properties of Castor Oil-Based Polyurethane/Starch Derivative Semi-
Interpenetrating Polymer Networks." Macromolecular Bioscience 2005, 5, 863.

(84) Cao, X.; Zhang, L. "Effects of Molecular Weight on the Miscibility and Properties
of Polyurethane/Benzyl Starch Semi-Interpenetrating Polymer Networks."
Biomacromolecules 2005, 6, 671.

(85) Cao, X.; Zhang, L. "Miscibility and properties of polyurethane/benzyl starch semi-
interpenetrating polymer networks." Journal of Polymer Science: Part B: Polymer
Physics 2005, 43, 603.

(86) van Warners, A.; Lammers, G. "Kinetics of the diffusion and chemical reaction of
ethylene oxide in starch granules in a gas solid system." Starch/Staerke 1990, 42, 427.

162
(87) El-Hinnawy, S. I.; Fahmy, A.; Mohamed, H.; El-Shirbeeny, A. E. "Preparation and
evaluation of hydroxyethyl starch." Starch/Staerke 1982, 34, 65.

(88) Besheer, A.; Hause, G.; Kressler, J.; Mäder, K. "Hydrophobically modified
hydroxyethyl starch: synthesis, characterization, and aqueous self-assembly into nano-
sized polymeric micelles and vesicles." Biomacromolecules 2007, 8, 359.

(89) Xu, A.; Seib, P. A. "Determination of the Level and Position of Substitution in
Hydroxypropylated Starch by High-Resolution 1H-NMR Spectroscopy of Alpha-limit
Dextrins." Journal of Cereal Science 1997, 25, 17.

(90) Mohd, B. M. N.; Wootton, A., M. . "In vitro Digestibility of Hydroxypropyl Maize,
Waxy Maize and High Amylose Maize Starches." Starch/Staerke 1984, 36, 273.

(91) Doane, W. M. "USDA Research on Starch-Based Biodegradable Plastics." Starch -


Stärke 1992, 44, 293.

(92) Liu, Z. Q.; Yi, X. S.; Feng, Y. "Effects of glycerin and glycerol monostearate on
performance of thermoplastic starch." Journal of Materials Science 2001, 36, 1809.

(93) Ollett, A. L.; Parker, R.; Smith, A. C. "Deformation and fracture behaviour of wheat
starch plasticized with glucose and water." Journal of Materials Science 1991, 26, 1351.

(94) van Soest, J. J. G.; Benes, K.; de Wit, D.; Vliegenthart, J. F. G. "The influence of
starch molecular mass on the properties of extruded thermoplastic starch." Polymer 1996,
37, 3543.

(95) Yu, L.; Christov, V.; Gray.G; Dutt, U. "Effect of additives on gelatinization,
rheological properties, and biodegradability of thermoplastic starch." Macromolecular
Symposia 1999, 144, 371.

(96) Stepto, R. F. T. "Thermoplastic Starch and Drug Delivery Capsules." Polymer


International 1997, 43, 155.

(97) Stepto, R. F. T. "Thermoplastic Starch." Macromolecular Symposia 2000, 152, 73.

163
(98) van Soest, J. J. G.; de Wit, D.; Vliegenthart, J. F. G. "Mechanical properties of
thermoplastic waxy maize starch." Journal of Applied Polymer Science 1996, 61, 1927.

(99) Forssell, P.; Mikkila, J.; Suortti, T. "Plasticization of barley starch with glycerol and
water." Journal of Macromolecular Science, Pure and Applied Chemistry 1995, 33, 703.

(100) Dean, K.; Yu, L.; Wu, D. "Preparation and characterization of melt-extruded
thermoplastic starch/clay nanocomposites." Composites Science and Technology 2007,
67, 413.

(101) Kalambur, S. B.; Rizvi, S. S. "Starch-based nanocomposites by reactive extrusion


processing." Polymer International 2004, 53, 1413.

(102) McGlashan, S. A.; Halley, P. J. "Preparation and characterisation of biodegradable


starch-based nanocomposite materials." Polymer International 2003, 52, 1767.

(103) Park, H.-m.; Lee, W.-k.; Park, C.-Y.; Cho, W.-J.; Ha, C.-S. "Environmentally
friendly polymer hybrids: Part I Mechanical, thermal, and barrier properties of
thermoplastic starch/clay nanocomposites." Journal of Materials Science 2003, 38, 909.

(104) Chen, B.; Evans, J. R. G. "Thermoplastic starch-clay nanocomposites and their


characteristics." Carbohydrate Polymers 2005, 61, 455.

(105) Zhang, Q.; Yu, Z.; Xie, X.; Naito, K.; Kagawa, Y. "Preparation and crystalline
morphology of biodegradable starch/clay nanocomposites." Polymer 2007, 48, 7193.

(106) Chaudhary, D. S. "Understanding Amylose Crystallinity in Starch-Clay


Nanocomposites." Journal of Polymer Science: Part B: Polymer Physics 2008, 46, 979.

(107) Chivrac, F.; Pollet, E.; Schmutz, M.; Avérous, L. "New Approach to Elaborate
Exfoliated Starch-Based Nanobiocomposites." Biomacromolecules 2008, 9, 896.

(108) Huang, M.; Yu, J.; Ma, X.; Jin, P. "High performance biodegradable thermoplastic
starch-EMMT nanoplastics." Polymer 2005, 46, 3157.

(109) Kampeerapappun, P.; Aht-ong, D.; Pentrakoon, D.; Srikulkit, K. "Preparation of


cassava starch/montmorillonite composite film." Carbohydrate Polymers 2007, 67, 155.

164
(110) Qiao, X.; Jiang, W.; Sun, K. "Reinforced Thermoplastic Acetylated Starch with
Layered Silicates." Starch/Staerke 2005, 57, 581.

(111) Falk, H.; Stanek, M. "Two-Dimensional 1H and 13C NMR Spectrscopy and the
Strucural Aspects of Amylose and Amylopectin." Monatshefte für Chemie 1997, 128,
777.

(112) López-Rubio, A.; Lagaron, J. M.; Giménez, E.; Cava, D.; Hernandez-Muñoz, P.;
Yamamoto, T.; Gavara, R. "Morphological Alterations Induced by Temperature and
Humidity in Ethylene-Vinyl Alcohol Copolymers." Macromolecules 2003, 36, 9467.

(113) Takahashi, M.; Tashiro, K.; Amiya, S. "Crystal Structure of Ethylene-Vinyl


Alcohol Copolymers." Macromolecules 1999, 32, 5860.

(114) Faisant, J. B.; Ait-Kadi, A.; Bousmina, M.; Deschenes, L. "Morphology,


thermomechanical and barrier properties of polypropylene-ethylene vinyl alcohol
blends." Polymer 1998, 39, 533.

(115) Bershitan, V.; Egorov, V., Differential scanning calorimetry of polymers. Ellis
Horwood: Chichester, UK, 1994.

(116) Zhang, S. H.; Jin, X.; Painter, P. C.; Runt, J. "Composition-dependent dynamics in
miscible polymer blends: influence of intermolecular hydrogen bonding." Polymer 2004,
45, 3933.

(117) Nakamae, K.; Kameyama, M.; Matsumoto, T. "Elastic Moduli of the Crystalline
Regions in the Direction Perpendicular to the Chain Axis of Ethylene-Vinyl Alcohol
Copolymers." Polymer Engineering and Science 1979, 19, 572.

(118) Cerrada, M. L.; Perez, E.; Perena, J. M.; Benavente, R. "Wide-Angle X-ray
Diffraction Study of the Phase Behavior of Vinyl Alcohol-Ethylene Copolymers."
Macromolecules 1998, 31, 2559.

(119) Kojima, Y.; Usuki, A.; Kawasumi, M.; Fukushima, Y.; Okada, A.; Kurauchi, T.;
Kamigaito, O. "Mechanical Properties of Nylon 6-Clay Hybrid." Journal of Material
Research 1993, 8, 1185.

165
(120) Usuki, A.; Kawasumi, M.; Kojima, Y.; Okada, A.; Kurauchi, T.; Kamigaito, O.
"Swelling Behavior of Montmorillonite Cation-exchanged for ω-Amino Acids by ε-
Caprolactam." Journal of Material Research 1993, 8, 1174.

(121) Yano, K.; Usuki, A.; Okada, A.; Kurauchi, T.; Kurauchi, T.; Kamigaito, O.
"Synthesis and Properties of Polyimide-clay Hybrid." Journal of Polymer Science:
Polymer Chemistry 1993, 31, 2493.

(122) Maiti, P.; Okamoto, M. "Crystallization Controlled by Silicate Surfaces in Nylon 6-


Clay Nanocomposites." Macromolecular Materials Engineering 2003, 288, 440.

(123) Hasegawa, N.; Okamoto, H.; Kato, M.; Usuki, A.; Sato, N. "Nylon 6/Na
Montmorillonite Nanocomposites Prepared by Compounding Nylon 6 with
Na+Montorillonite Slurry." Polymer 2003, 44, 2933.

(124) Kojima, Y.; Usuki, A.; Kawasumi, M.; Okada, A.; Kurauchi, T.; Kamigaito, O.
"Synthesis of Nylon 6-clay Hybrid by Montmorillonite Intercalated with ε-Caprolactam."
Journal of Polymer Science: Polymer Chemistry 1993, 31, 983.

(125) Kojima, Y.; Usuki, A.; Kawasumi, M.; Okada, A.; Kurauchi, T.; Kamigaito, O.
"One-pot Synthesis of Nylon 6-clay Hybrid." Journal of Polymer Science: Polymer
Chemistry 1993, 31, 1755.

(126) Raym, S.; Yamada, K.; Okamoto, N.; Ogami, A.; Ueda, K. "New
Polylactide/Layered Silicate Nanocomposites. 3. High-Performance Biodegradable
Materials." Chemistry of Materials 2003, 15, 1456.

(127) Xu, R.; Mania, E.; Snyder, A. J.; Runt, J. "New Biomedical Poly(urethane urea)-
Layered Silicate Nanocomposites." Macromolecules 2001, 34, 337.

(128) Choi, S.; Lee, K. M.; Han, C. D. "Effects of Triblock Copolymer Architecture and
the Degree of Functionalization on the Organoclay Dispersion and Rheology of
Nanocomposites." Macromolecules 2004, 37, 7649.

(129) Huang, W.; Han, C. D. "Ruthenium(II) Complex-Induced Dispersion of


Montmorillonite in a Segmented Main-Chain Liquid-Crystalline Polymer Having Side-
Chain Terpyridine Group." Macromolecules 2006, 39, 8207.

166
(130) Huang, W.; Han, C. D. "Dispersion characteristics and rheology of organoclay
nanocomposites based on a segmented main-chain liquid-crystalline polymer having
side-chain azopyridine with flexible spacer." Polymer 2006, 47, 4400.

(131) Huang, W.; Han, C. D. "Dispersion Characteristics and Rheology of Organoclay


Nanocomposites Based on a Segmented Main-Chain Liquid-Crystalline Polymer Having
Pendent Pyridyl Group." Macromolecules 2006, 39, 257.

(132) Lee, K. M.; Han, C. D. "Rheology of organoclay nanocomposites: Effects of


polymer matrix/organoclay compatibility and the gallery distance of organoclay. ."
Macromolecules 2003, 36, 7165.

(133) Zha, W.; Choi, S.; Lee, K. M.; Han, C. D. "Dispersion Characteristics of
Organoclay in Nanocomposites Based on End-Functionalized Homopolymer and Block
Copolymer." Macromolecules 2005, 38, 8418.

(134) Zha, W.; Han, C. D.; Han, S. H.; Lee, D. H.; Kim, J. K.; Guo, M.; Rinaldi, P. L.
"Ion-dipole interactions in the dispersion of organoclay nanocomposites based on
polystyrene-block-poly(2-vinylpyridine) copolymer." Polymer 2009, 50, 2411.

(135) Kutsumizu, S.; Hara, H.; Tachino, H.; Shimabayashi, K.; Yano, S. "Infrared
Spectroscopic Study of the Binary Blends of Sodium and Zinc Salt Ionomers Produced
from Poly(ethylene-co-methacrylic acid)." Macromolecules 1999, 32, 6340.

(136) Tachino, H.; Hara, H.; Hirasawa, E.; Kutsumizu, S.; Yano, S. "Structure and
Properties of Ethylene Ionomers Neutralized with Binary Metal Cations."
Macromolecules 1994, 27, 372.

(137) Akutsu, H. "Direct determination by raman scattering of the conformation of the


choline group in phospholipid bilayers." Biochemistry 1981, 20, 7359.

(138) Demanaa, P. H.; Davies, N. M.; Hook, S.; Rades, T. "Analysis of Quil A–
phospholipid mixtures using drift spectroscopy." International Journal of Pharmaceutics
2007, 342, 49.

(139) Frengeli, U. P. "A new crystalline phase of L-α dipalmitoryl phosphatidylcholine


monohydrate." Biophysics Journal 1981, 34, 173.

167
(140) Zawisza, I.; Wittstock, G.; Boukherroub, R.; Szunerits, S. "Polarization Modulation
Infrared Reflection Absorption Spectroscopy Investigations of Thin Silica Films
Deposited on Gold. 2. Structural Analysis of a 1,2 Dimyristoyl-sn-glycero-3-
phosphocholine Bilayer." Langmuir 2008, 24, 3922.

(141) Hauser, H.; Guyer, W.; Pascher, I.; Skrabal, P.; Sundell, S. "Polar group
conformation of phosphatidylcholine. Effect of solvent and aggregation." Biochemistry
1980, 19, 366.

(142) Akutsu, H.; Kyogoku, Y. "Conformational difference in the polar groups of


phosphatidylcholine and phosphatidylethanolamine in aqueous phase." Chemistry and
Physics of Lipids 1977, 18, 285.

(143) Seelig, J.; Gally, G. U.; Wohlgemuth, R. "Orientation and flexibility of the choline
head group in phosphatidylcholine bilayers." Biochimica et Biophysica Acta 1977, 467,
109.

(144) Yeagle, P. L.; Hutton, W. C.; Huang, C.; Martin, R. B. "Phospholipid head-group
conformations; intermolecular interactions and cholesterol effects." Biochemistry 1977,
16, 4344.

(145) Zawisza, I.; Wittstock, G.; Boukherroub, R.; Szunerits, S. "PM IRRAS
Investigation of Thin Silica Films Deposited on Gold. Part 1. Theory and Proof of
Concept." Langmuir 2007, 23, 9303.

(146) Ray, S. S.; Okamoto, M. "Polymer/layered silicate nanocomposites: a review from


preparation to processing." Progress in Polymer Science 2003, 28, 1539.

(147) Yariv, S.; Cross, H., Organo-clay Complexes and Interactions. 1st ed.; CRC Press:
New York, 2001.

(148) Meunier, A., Clays. Springer Berlin Heidelberg: 2005.

(149) Fu, M. H.; Zhang, Z. Z.; Low, P. F. "Changes in the Properties of a


Montmorillonite-Water System During the Adsorption and Desorption of Water:
Hysteresis." Clay and Clay Minerals 1990, 38, 485.

168
(150) Mooney, R. W.; Keenan, A. G.; Wood, L. A. "Adsorption of Water Vapor by
Montmorillonite. II. Effect of Exchangeable Ions and Lattice Swelling as Measured by X-
Ray Diffraction." Journal of the American Chemical Society 1952, 74, 1371.

(151) Sposito, G.; Prost, R. "Structure of water adsorbed on smectites." Chemical


Reviews 1982, 82, 553.

(152) Sposito, G.; Prost, R.; Gaultier, J. P. "Infrared spectroscopic study of adsorbed
water on reduced-charge sodium/lithium-montmorillonites. ." Clay and Clay Minerals
1983, 91, 9.

169
APPENDICES

170
APPENDIX A
1
H-NMR SPECTRA OF CHEMICALLY MODIFIED STARCH

AS-B

171
AS-LF

CS-B

172
CS-LF

173
APPENDIX B

FTIR SPECTRA OF CHEMICALLY MODIFIED STARCH

AS-B

174
AS-LF

CS-B

175
CS-LF

176

You might also like