198
Effect of Laser Induced Crystallinity Modification on Degradation and Drug Release of Biodegradable Polymer Shan-Ting Hsu Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Graduate School of Arts and Sciences COLUMBIA UNIVERSITY 2013

Effect of Laser Induced Crystallinity Modification on

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Effect of Laser Induced Crystallinity Modification on Degradation and

Drug Release of Biodegradable Polymer

Shan-Ting Hsu

Submitted in partial fulfillment of the requirements for the degree of

Doctor of Philosophy in the Graduate School of Arts and Sciences

COLUMBIA UNIVERSITY

2013

© 2013 Shan-Ting Hsu

All rights reserved

ABSTRACT

Effect of Laser Induced Crystallinity Modification on Degradation and Drug Release of

Biodegradable Polymer

Shan-Ting Hsu

Biodegradable polymers such as poly(L-lactic acid) (PLLA) are promising in drug delivery

applications. In biodegradable polymer-based drug delivery systems, drugs are released at a

rate determined by polymer degradation. Polymer degradation exhibits an undesirable

induction period, during which a limited amount of embedded drug is released. As a

semi-crystalline polymer, PLLA degradation and drug release are affected by its crystallinity.

Control over PLLA crystallinity tailors drug release over time, and provides a potential solution

to shorten the induction period of polymer degradation and drug release.

The work presented in this thesis first investigates the crystalline morphology and crystallinity

developed during the PLLA film formation processes. Solvent casting, spin coating, and

subsequent annealing are conducted. The resulting morphology, crystallinity, molecular order,

conformation, and intermolecular interaction are examined using optical microscopy, wide-angle

X-ray diffraction, and Fourier transform infrared spectroscopy. Solvent casting produces

category 1 spherulites, and annealing the spin coated films leads to category 2 spherulites.

Crystal structure of the two kinds of films shows distinct features. The results advance the

understanding of PLLA crystal structures, which is essential for its medical applications.

Control over crystallinity allows for modification of PLLA degradation and drug release over

time. Laser irradiation is used in this study to induce surface melting and resolidification. The

high cooling rate of the laser treatment coupled with the slow polymer crystallization kinetics

leads to reduced surface crystallinity after the laser irradiation. Effects of laser irradiation on

the surface morphology, crystallinity, and chemical modifications are investigated via optical

microscopy, wide-angle X-ray diffraction, and X-ray photoelectron spectroscopy. The effect of

laser crystallinity modifications and drug loading concentration on PLLA biodegradation and

drug release is investigated. Degradation is characterized through molecular weight by gel

permeation chromatography. Drug release is measured by spectrophotometry. A finite

element model is developed to numerically examine the spatial and temporal temperature

profiles, as well as chemical modifications in the PLLA matrix after the laser treatment. Effect

of laser crystallinity modification on biodegradation and drug release is also numerically

investigated based on PLLA hydrolysis and diffusion mechanisms.

It has been demonstrated that laser irradiation reduces PLLA crystallinity. A working window

of laser fluence levels exists within which crystallinity is decreased while no appreciable

chemical modification is observed. The working window is enlarged for the higher crystalline

polymer as a result of the cage effect. The degradation and drug release tests show that the

addition of drug accelerates polymer degradation and drug release rate, because the porous

structure after drug release favors water penetration and hydrolysis based degradation. With a

low drug concentration, the slow polymer degradation kinetics results in an induction period of

drug release. The induction period is shortened by the laser treatment. It is demonstrated that

laser treated PLLA, with lower surface crystallinity, has a higher initial degradation rate while

the subsequent degradation is not modified. Because of the accelerated initial degradation, the

induction period of drug release is therefore shortened, while the drug release rate is kept

unmodified, which is desired in drug delivery applications.

i

Table of Contents

Chapter 1: Introduction ................................................................................................................... 1

1.1 Polymer Structure and Requirements for Crystallization ....................................... 1

1.2 Polymer Crystallization and Amorphization ........................................................... 3

1.2.1 Polymer Crystallization .................................................................................. 4

1.2.1.1 Polymer Crystalline Morphology ........................................................... 4

1.2.1.2 Polymer Crystallization Kinetics ............................................................ 5

1.2.2 Polymer Amorphization .................................................................................. 7

1.2.2.1 Polymer Amorphization Process ................................................................. 7

1.2.2.2 Polymer Amorphization Kinetics ............................................................ 8

1.3 Polymer Controlled Drug Release System .............................................................. 9

1.3.1 Mechanisms of Drug Release from Polymer Matrix .................................... 10

1.3.1.1 Diffusion Controlled Mechanism ......................................................... 10

1.3.1.2 Swelling Controlled Mechanism ........................................................... 11

1.3.1.3 Degradation Controlled Mechanism ..................................................... 11

1.3.2 Degradation of Biodegradable Polymer in Physiological Environments ..... 13

1.4 Factors Determining Polymer Degradation and Drug Release ............................. 14

1.4.1 Polymer Crystallinity .................................................................................... 14

1.4.2 Addition of Drug and Small Molecules ........................................................ 15

1.4.3 Irradiation ...................................................................................................... 17

1.4.3.1 γ-Irradiation ........................................................................................... 17

1.4.3.2 Laser Surface Irradiation ....................................................................... 18

1.4.4 Degradation and Drug Release Environment ............................................... 18

ii

1.5 Experimental Considerations ................................................................................ 20

1.5.1 Sample Preparation ....................................................................................... 20

1.5.1.1 Solvent-Involved Sample Formation Process ....................................... 20

1.5.1.2 Solvent-Free Sample Formation Process .............................................. 21

1.5.2 Laser System ................................................................................................. 23

1.5.3 Material Characterization .............................................................................. 25

1.5.3.1 Wide-Angle X-ray Diffraction .............................................................. 26

1.5.3.2 Gel Permeation Chromatography ......................................................... 27

1.5.3.3 Differential Scanning Calorimetry ........................................................ 30

1.6 Research Objectives and Organization of the Proposal ........................................ 31

Chapter 2: Effect of Film Formation Method and Annealing on Morphology and Crystal

Structure of Poly(L-Lactic Acid) Films ........................................................................................ 34

2.1 Introduction ........................................................................................................... 34

2.2 Background ........................................................................................................... 36

2.2.1 Nucleation and Crystallization ...................................................................... 37

2.2.2 Mobility of Polymer Molecules in Solution and Melt .................................. 37

2.2.3 Structure of Spherulites................................................................................. 39

2.3 Materials and Methods .......................................................................................... 40

2.3.1 Sample Preparation and Annealing ............................................................... 40

2.3.2 Characterization ............................................................................................ 40

2.4 Results and Discussion .......................................................................................... 41

2.4.1 Morphology................................................................................................... 41

2.4.2 Molecular Order ............................................................................................ 45

iii

2.4.3 Conformation of Molecules and Intermolecular Interaction ........................ 54

2.5 Conclusions ........................................................................................................... 60

Chapter 3: Effect of Excimer Laser Irradiation on Crystallinity and Chemical Bonding of

Biodegradable Polymer ................................................................................................................. 62

3.1 Introduction ........................................................................................................... 62

3.2 Background ........................................................................................................... 64

3.2.1 Melting and Crystallization........................................................................... 64

3.2.2 Interaction between Photons and Polymers .................................................. 65

3.2.3 Dissociation Quantum Yield ......................................................................... 66

3.3 Numerical Model ................................................................................................... 67

3.4 Materials and Methods .......................................................................................... 71

3.5 Results and Discussion .......................................................................................... 71

3.5.1 Effects of Laser Irradiation on Morphology ................................................. 71

3.5.2 Decrease of Crystallinity by Laser Irradiation .............................................. 73

3.5.3 Chemical Modifications ................................................................................ 78

3.5.3.1 Modifications of Chemical Bonds ........................................................ 78

3.5.3.2 Effect of Radical Mobility on the Amount of Chemical Modifications 83

3.6 Conclusions ........................................................................................................... 88

Chapter 4: Effect of Laser Induced Crystallinity Modification on Biodegradation Profile of

Poly(L-Lactic Acid) ....................................................................................................................... 90

4.1 Introduction ........................................................................................................... 90

4.2 Background ........................................................................................................... 92

4.2.1 Laser Melting of Polymer ............................................................................. 92

iv

4.2.2 Biodegradation of Polyester .......................................................................... 92

4.3 Numerical Model ................................................................................................... 95

4.4 Materials and Methods .......................................................................................... 98

4.5 Results and Discussion .......................................................................................... 99

4.5.1 Effect of Laser Irradiation on Crystallinity and Chemical Modifications .... 99

4.5.2 Effect of Laser Irradiation on Degradation ................................................. 101

4.5.2.1 Modification of Morphology .............................................................. 101

4.5.2.2 Modification of Crystallinity .............................................................. 104

4.5.2.3 Modification of Molecular Weight and Mass ..................................... 106

4.5.3 Concentration of Species during Degradation ............................................ 113

4.6 Conclusions ......................................................................................................... 115

Chapter 5: Effect of Drug Loading and Laser Surface Melting on Polymer Biodegradation and

Drug Release Profile ................................................................................................................... 116

5.1 Introduction ......................................................................................................... 116

5.2 Background ......................................................................................................... 119

5.2.1 Effect of Additives on Polymer Mobility .................................................... 119

5.2.2 Biodegradation of Polyester ........................................................................ 119

5.2.3 Drug Release from Biodegradable Polymers .............................................. 120

5.3 Numerical Model ................................................................................................. 121

5.4 Materials and Methods ........................................................................................ 125

5.5 Results and Discussion ........................................................................................ 127

5.5.1 Effect of Drug Loading on Chain Mobility and Polymer Crystallinity ...... 127

5.5.2 Effect of Drug Concentration on Polymer Degradation and Drug Release 132

v

5.5.2.1 Polymer Degradation during Drug Release ........................................ 132

5.5.2.2 Drug Release ....................................................................................... 136

5.5.2.3 Polymer Crystallization during Drug Release .................................... 140

5.5.3 Laser Modification of Polymer Degradation and Drug Release Profile ..... 142

5.5.3.1 Laser Modification of Polymer Crystallinity ...................................... 142

5.5.3.2 Polymer Degradation during Drug Release ........................................ 144

5.5.3.3 Drug Release ....................................................................................... 146

5.5.3.4 Polymer Crystallization during Drug Release .................................... 147

5.6 Conclusions ......................................................................................................... 148

Chapter 6: Conclusions ............................................................................................................... 150

6.1 Effect of Sample Formation Process on Morphology and Crystal Structure of

Biodegradable Polymer ....................................................................................... 150

6.2 Laser Crystallinity Modification of Biodegradable Polymer .............................. 151

6.3 Modification of Polymer Biodegradation by Laser Melting ............................... 152

6.4 Effect of Drug Loading on Degradation and Drug Release of Biodegradable

Polymer ............................................................................................................... 153

6.5 Shortening of Induction Period of Drug Release by Laser Melting .................... 154

6.6 Future Work ......................................................................................................... 155

References ................................................................................................................................... 157

Appendix ..................................................................................................................................... 171

A.1 Calculation of Polymer Crystallinity from Wide Angle X-Ray Diffraction Method ... 171

A.2 Finite Element Models ................................................................................................. 172

A.2.1 Model of Laser Induced Thermal and Chemical Modifications ................. 173

vi

A.2.2 Model of Polymer Biodegradation Process ................................................ 174

A.2.3 Model of Drug Release Process .................................................................. 175

A.3 Purification Process of Drug Loaded PLLA Matrix for GPC measurements ..... 176

A.4 Publications Under Candidature .......................................................................... 177

vii

List of Figures Chapter 1

Figure 1.1: Effects of defects in (a) a crystal composed of atoms or small molecules and (b) a

polymer crystal. The heterogeneous atom or small molecules can be accommodated in

the crystal, because of the relaxation of the surrounding atoms or molecules. The

existence of defects in polymer chains cannot be dissipated. The effect of chain defect

thus exists along the entire chain length (Schultz, 2001). ..................................................... 2

Figure 1.2: Sketch of the morphology of two categories of spherulites. (a) Category 1

spherulite, which features the central multidirectional growth. (b) Category 2 spherulite,

which features sheaf-like unidirectional growth and double circle appearance (Norton &

Keller, 1985). ......................................................................................................................... 5

Figure 1.3: (a) Melting of the lamellae at temperatures higher than the melting temperature.

Decrease of crystallinity occurs within ns. (b) Snapshots of gradually melting lamella at

different times at 370 K. The lamella growth front retreats with time. The vertical

yellow line is for the ease of identifying withdrawal of the growth font (Yamamoto, 2010).

............................................................................................................................................... 9

Figure 1.4: Mold used in this study for the thermal compression processes. .............................. 23

Figure 1.5: The excimer laser system used in the study. The system is composed of the

homogenizer and mask to generate a top hat pulse with a desired spot size. Laser pulse

irradiation is controlled by the computer. ............................................................................ 24

Figure 1.6: Sketch of the working principle of laser beam homogenization. .............................. 25

Figure 1.7: WAXD profile from semi-crystalline polyethylene. The dashed line shows the

estimated position of the amorphous scattering intensity (Campbell et al., 2000). ............. 27

Figure 1.8: GPC separation of polymer molecules with two different sizes. (a) the mixture of

viii

polymer molecules upon injection; (b) the mixture enters the porous column because of the

applied pressure gradient through the column length; (c) the small molecules are trapped

by the pores and flow slowly, but larger molecules are not, from which the molecules are

separated; (d) complete elution of large molecules. ............................................................ 29

Figure 1.9: DSC thermograms of poly (lactic acid) blended with plasticizer at different

concentrations (a) 0 wt %, (b) 5 wt %, (c) 10 wt %, (d) 15 wt %, (e) 20 wt %, (f) 25 wt %,

and (g) 30 wt %. The scanning rate is 40°C/min (Yeh et al., 2009). ................................ 31

Chapter 2

Figure 2.1: (a) The morphology of solvent cast film annealed at 140°C for 3 hours. The

morphology is similar to that of the as cast film. (b) The morphology of spin coated film

annealed at 140°C for 3 hours. The eye structure is observed. ........................................ 43

Figure 2.2: The morphology of spin coated film annealed at 140°C for 24 hours. Eye structure

remains observable while the boundary of spherulites can be more clearly observed. The

spherical symmetry does not extend to the center, and the sheaf-like features are preserved.

............................................................................................................................................. 45

Figure 2.3: (a) Sketch of the lamellar structure in solvent cast films. (b) Sketch of the lamellar

structure in spin coated films. .............................................................................................. 46

Figure 2.4: (a) WAXD profiles for solvent cast films annealed at different temperatures for 3

hours. Dotted line under the as cast film is the fitted curve for the amorphous region. (b)

WAXD profiles for spin coated films annealed at different temperatures. The (010) peak

start to emerge after 140°C annealing for 24 hours. Dotted line under the 80°C annealed

film is the fitted curve for the amorphous region. ............................................................... 49

ix

Figure 2.5: Intensity ratio I(110)/(200)/I(203) of the solvent cast and spin coated films annealed for 3

hours. The error bar represents the standard deviation. .................................................... 50

Figure 2.6: Location of the (110)/(200) peak for the solvent cast and spin coated films annealed

for 3 hours. The error bar represents the standard deviation. ........................................... 50

Figure 2.7: Degree of crystallinity of the solvent cast and spin coated films annealed for 3 hours.

Annealing temperature of 25°C means the as cast film. The error bar represents the

standard deviation. ............................................................................................................... 53

Figure 2.8: Degree of crystallinity and peak location of the spin coated films annealed at 140 °C

for 1, 3, 8, and 24 hours. ...................................................................................................... 54

Figure 2.9: (a) FTIR spectra of the solvent cast films before annealing and after annealed in

different conditions. (b) FTIR spectra of the spin coated films before annealing and after

annealed in different conditions. ......................................................................................... 56

Figure 2.10: (a) Curve fitting for the gt and tt conformers in the carbonyl C=O stretching region

of the as cast solvent cast film. (b) Curve fitting for the gt and tt conformers in the

carbonyl C=O stretching region of the solvent cast film annealed at 140°C for 3 hours. ... 57

Figure 2.11: Carbonyl C=O stretching region of the spin coated films annealed at different

temperatures for 3 hours. ..................................................................................................... 57

Figure 2.12: Second derivative spectra of the C=O stretching region for solvent cast and spin

coated films before annealing and annealed at different temperatures for 3 hours. ............ 58

Figure 2.13: Peak width of the resolved spectra for the gt and tt conformers of solvent cast films

before annealing and after annealed in different conditions. The error bar represents the

standard deviation of the five measurements. ..................................................................... 59

x

Chapter 3

Figure 3.1: Elementary events following the polymer absorption of photons with energies higher

than the polymeric bond energy........................................................................................... 66

Figure 3.2: Morphologies of the film with high radical mobility treated with fluence of (a) 2.50

J/cm2 and (b) 2.60 J/cm2, and film with low radical mobility treated with fluence of (c)

2.60 J/cm2 and (d) 2.70 J/cm2. ............................................................................................. 72

Figure 3.3: WAXD profiles of crystalline PLLA films with high radical mobility before and after

laser treatment. Note that the peaks decrease at higher fluences (2.40 and 2.60 J/cm2).

The same trend is observed for the films with low radical mobility. .................................. 74

Figure 3.4: Crystallinity of PLLA films derived from the WAXD results as a function of laser

fluence. ................................................................................................................................ 75

Figure 3.5: Simulation result of the typical temperature distribution in the film. This

simulation depicts the film with high radical mobility 1 μs after the onset of a 2.40 J/cm2

laser pulse. ........................................................................................................................... 75

Figure 3.6: Simulation results of the melting depth in the film and the normalized number

density of the CH3CHCOO units connecting at their ester group with other CH3CHCOO

units. .................................................................................................................................... 76

Figure 3.7: Simulation result of the typical film temperature profiles as a function of time after

laser irradiation. The simulation predicts the temperature profile on the surface of the

film with low radical mobility treated a single pulse with fluence of 1.80 and 2.80 J/cm2. 77

Figure 3.8: Typical C1s XPS spectra and the resolved peaks of the film before and after laser

treatment. The figures show the results of the film with high radical mobility (a) before

laser treatment, and treated with fluences of (b) 2.50 J/cm2 and (c) 2.60 J/cm2. Note that

xi

the peak area ratio does not agree with the theoretical value, 1:1:1; a possible reason is

given in Sec. 3.5.3.1. ........................................................................................................... 79

Figure 3.9: Area percentage of the three resolved peaks in the C1s XPS spectra of the films with

(a) high radical mobility and (b) low radical mobility. ....................................................... 80

Figure 3.10: Ratio of the O1s peak area to the C1s peak area derived from the XPS spectra for the

films with high and low radical mobility. ............................................................................ 81

Figure 3.11: Simulation results of the typical normalized number density of the CH3CHCOO

units connecting at their ester group with other CH3CHCOO units on the surface. Results

of the film with high radical mobility are shown. ............................................................... 84

Figure 3.12: Simulation results of the typical absolute value of the reaction rate of Eq. (9) for the

amorphous domain after the onset of laser pulse. Results of the film with high radical

mobility are shown. ............................................................................................................. 85

Figure 3.13: Simulation results of the typical absolute value of the reaction rate of Eq. (9) for the

crystalline domain after the onset of laser pulse. Results of the film with high radical

mobility are shown. ............................................................................................................. 87

Chapter 4

Figure 4.1: Schematic representation of (a) semicrystalline structure and (b) crystalline residue

of polymer after hydrolysis, which preferentially occurs in amorphous region and on

crystal fold surface (Tsuji & Ikarashi, 2004)....................................................................... 94

Figure 4.2: Non-degraded PLLA sample (a) before and (b) after laser treatment with a fluence of

3 J/cm2. Laser irradiated spots show less transparency due to increased surface roughness.

The laser-treated sample degraded for 14 days is given in (c) and its high crystallinity

xii

reduces the transparency. ................................................................................................... 100

Figure 4.3: PLLA crystallinity and the ratio of O1s to C1s as a function of laser fluence. The

error bar represents the standard deviation of 3 data points. ............................................. 100

Figure 4.4: Surface morphology of the laser treated sample (a) before and after degradation for

(b) 3, (c) 8, and (d) 14 days under the stereomicroscope. The squares in (a) are laser spots.

The edges of laser spots become less defined with degradation period, suggesting the

erosion of laser melted layer. Degradation in the non-melted bulk volume occurs in the

later stage. .......................................................................................................................... 102

Figure 4.5: Cross section of the laser-treated sample degraded for 14 days. The bulk remains

solid, suggesting that autocatalysis is not dominant. ......................................................... 102

Figure 4.6: Simulated spatial distribution of monomer concentration in the laser treated sample

degraded for (a) 0.5 days and (b) 6 days. .......................................................................... 103

Figure 4.7: WAXD profiles of the (a) non-laser treated and (b) laser treated samples degraded

for regular periods. Intensity of crystalline peaks increases with degradation period,

suggesting a higher crystallinity. Profiles are shifted in y direction for viewing clarity. 105

Figure 4.8: Crystallinity as a function of degradation period determined from WAXD.

Crystallinity increases with degradation period. A significant increase occurs on day 0.5

for both types of samples. Crystallinity also increases on day 8 and day 5 for the

non-laser treated and laser treated samples, respectively. The error bar represents the

standard deviation of 3 data points. ................................................................................... 105

Figure 4.9: GPC profiles of the (a) non-laser treated and (b) laser treated samples after regular

degradation periods. For (a), the distribution becomes wider and shifts left as the

degradation period increases to day 5, representing the random chain scission in the

xiii

amorphous region. After day 8, a distinct new peak is developed due to selective chain

scission of the fold surface of crystals. For (b), the MW distribution extends to the left

before day 3, signifying the random chain scission of the laser melt layer. At day 5, two

distinct peaks are developed due to the selective chain scission of the partially melted

crystal fold surfaces. Profiles are shifted in y direction for viewing clarity. .................. 107

Figure 4.10: Mw, Mn, and PDI of the (a) non-laser treated and (b) laser treated samples after

regular degradation periods. Mw and Mn decrease at a higher rate for the laser treated

samples, as a result of fast degradation in the laser melted layer. The non-homogeneous

degradation of the melted layer and bulk increases PDI. The error bar represents the

standard deviation of 3 data points. ................................................................................... 108

Figure 4.11: Experimental results of sample mass with and without laser treatments. Mass

decrease is observed after day 8 for non-laser treated sample and after day 3 for laser

treated sample. The error bar represents the standard deviation of 3 data points. ......... 109

Figure 4.12: Simulated MW of the non-laser treated and laser treated samples. ....................... 111

Figure 4.13: Simulated mass change of the non-laser treated and laser treated samples. ......... 112

Figure 4.14: Molar concentration of the simulated species as a function of degradation period.

(a) In the center of the non-laser treated sample. (b) In the laser melted layer of the laser

treated sample .................................................................................................................... 114

Chapter 5

Figure 5.1: Non-laser treated PLLA matrices loaded with 5 % drug (a) before degradation and

drug release, and degraded and released for (b) 35, (c) 49, (d) 70 days. ........................... 127

Figure 5.2: WAXD profiles of PLLA with different drug loading concentrations. Intensity of

xiv

crystalline peaks increases with drug concentration, suggesting a higher crystallinity.

Profiles are shifted in y direction for viewing clarity. ....................................................... 128

Figure 5.3: DSC thermograms of pure PLLA and PLLA loaded with 1 %, 5 %, 10 %, and 20 %

drug (a) heating from 50 to 200 °C and (b) around the glass transition temperature. The

heating rate is 5 °C/min. .................................................................................................... 130

Figure 5.4: Glass transition temperature and melting temperature of drug loaded PLLA matrices

as a function of drug concentration. .................................................................................. 131

Figure 5.5: Crystallinity as a function of drug loading concentration obtained from WAXD and

DSC. Crystallinity increases with drug concentration based on both measurements. The

error bar represents the standard deviation of 3 data points. ............................................. 131

Figure 5.6: GPC profiles of the non-laser treated PLLA matrices loaded with (a) 0 % and (b) 1

% drug as a function of degradation and drug release period. .......................................... 133

Figure 5.7: Number average and weight average molecular weights of (a) pure PLLA and PLLA

loaded with (b) 1 % (c) 5 % (d) 10 % (e) 20 % drug with polymer degradation and drug

release period. For pure PLLA and PLLA loaded with 1 %, 5 %, and 10 % drug, effect of

laser melting on degradation is also studied. ..................................................................... 136

Figure 5.8: (a) Absorbance spectra of the drug/PBS solutions with different drug concentrations.

(b) Linear relationship between absorbance at 552 nm and drug concentration in PBS.

Rhodamine B is used as the model drug. .......................................................................... 137

Figure 5.9: Drug release profiles of PLLA with different drug loading concentrations. ........... 138

Figure 5.10: Simulated spatial distribution of drug concentration in non-laser treated PLLA

loaded with 5 % drug after release for (a) 30 and (b) 70 days. Drug concentration is

represented as a percentage of initial value. ...................................................................... 139

xv

Figure 5.11: Crystallinity of non-laser treated PLLA loaded with 1 %, 5 %, 10 %, and 20 % drug

as a function of polymer degradation and drug release period. ......................................... 141

Figure 5.12: WAXD profiles of PLLA before and after laser treatment. Intensity of crystalline

peaks decreases after laser treatment, suggesting a reduced crystallinity. Profiles are

shifted in y direction for viewing clarity. .......................................................................... 143

Figure 5.13: Crystallinity of 1 %, 5 %, and 10 % drug loaded PLLA before and after laser

treatment. Laser treatment reduces crystallinity. Crystallinity decreases at a smaller

extent for higher drug concentrated PLLA. ....................................................................... 143

Figure 5.14: GPC profiles of the laser treated PLLA matrices loaded with (a) 0 % and (b) 1 %

drug as a function of degradation and drug release period. ............................................... 145

Figure 5.15: Modification of drug release profiles through laser melting. The induction period

of drug release is shortened due to laser treatment. ........................................................... 147

Figure 5.16: Crystallinity of laser treated PLLA loaded with 1 %, 5 %, and 10 % drug as a

function of polymer degradation and drug release period. ................................................ 148

xvi

Acknowledgements

The completion of this thesis cannot be possible without the aid of a number of individuals and

groups. First of all, I would like to express my deepest gratitude towards my academic advisor,

Prof. Y. Lawrence Yao. His academic guidance, expert advice, attention to detail, and hard

work have set an example I hope to achieve some day. I would also like to acknowledge my

thesis committee members, Prof. Jeffrey W. Kysar, Prof. James S. Im, Prof. Christopher J.

Durning, and Prof. Qiao Lin, for their time and precious comments. Financial supports from

Columbia University and National Science Foundation are also gratefully acknowledged.

Great gratitude goes to my colleagues in the Manufacturing Research Lab, Gen Satoh, Panjawat

Kongsuwan, Hongliang Wang, Huade Tan, and Grant Brandal, for their friendship and generous

help. I would also like to thank Dr. Young Suk Park at Brookhaven National Lab for his help

with the DSC and a portion of GPC measurements conducted in this thesis. Technical and

practical guidance during experiments from Dr. Chien-Yang Chiu, Dr. Po-Hua Lee, Dr.

Benjamin Dach, Dr. Xuegong Lei, Dr. Yongjun Li, and Dr. Xia Li is gratefully appreciated.

Many thanks are owed to my parents who have been standing by my side and supporting me in

various ways. Finally, I would like to thank my girlfriend to whom I am deeply grateful, for her

understanding and support during my Ph.D. studies.

1

Chapter 1: Introduction

1.1 Polymer Structure and Requirements for Crystallization

A polymer is a compound with high molecular weight. The structure of a polymer is composed

of chains of small repeat units known as monomers. Configuration and conformation are the

two terms commonly used to describe polymer structures. Configuration refers to the

organization of atoms or groups along the polymer chain and is determined during the

polymerization process. Configuration can only be changed by breaking and reforming existing

chemical bonds. Conformation refers to the arrangement of atoms or groups, and can be altered

through rotating the atoms or groups around a single bond. Different configurations result from

different placements of the monomers, including the head-to-head, tail-to-tail, and head-to-tail

placements, as well as stereoregular arrangements. Based on different monomer placements, a

polymer chain can possess isotactic configuration, syndiotactic configuration, or atactic

configuration. In isotactic configuration, all the side groups lie on the same side of the plane of

the chain. If the side groups lie alternately above and below the plane, the chain has a

syndiotactic configuration. If the side groups have a random sequence of positions, the

configuration is atactic. Usually the thermodynamically and spatially preferred configuration is

the syndiotactic configuration. Conformation includes the trans, gauche, and cis arrangements

of consecutive C-C single bonds and helical arrangements of crystalline polymers. The trans

conformation is staggered and the rotation angle of R-C-C-R is 180°. The trans conformation is

the most energetically stable. The gauche form is the next most stable conformation. The cis

conformation corresponds to the R-C-C-R rotation angle 0°, and is the least energetically stable.

2

(a) Crystal composed of atoms or small molecules (left) and a defect and the region it affects

(right).

(b) Polymer crystal (left) and a defect and the region it affects (right).

Figure 1.1: Effects of defects in (a) a crystal composed of atoms or small molecules and (b) a polymer crystal. The heterogeneous atom or small molecules can be accommodated in the crystal, because of the relaxation of the surrounding atoms or molecules. The existence of

defects in polymer chains cannot be dissipated. The effect of chain defect thus exists along the entire chain length (Schultz, 2001).

Structural regularity determines the crystallizability of a polymer. Crystallizable polymers

typically possess chemically and geometrically regular molecules. Irregularity of chain, or

chain defect, limits crystallizability of polymer. Chain defects can be the branches from the

backbone, chemically different monomers, steric irregularity (e.g., isotactic, syndiotactic, and

atactic configurations), and different monomer attachment directions (e.g., head-to-head,

tail-to-tail, and head-to-tail placements). As compared with small molecules, the tolerance for

3

chain defect in polymer crystals is relatively low. This is because of the connectivity of

monomers along the chain, as illustrated in Fig. 1.1 for an example of a chain consisting of a

chemically different monomer (Schultz, 2001). In crystals composed of atoms of small

molecules, the defect induced by heterogeneous atoms or molecules can be dissipated within a

short distance by small displacement of the near atoms or molecules. However, the effect of

defects on polymer chains is significant and exists along the entire chain. Therefore, chain

regularity is stringent for polymer crystallization. It is expected that chains in the crystalline

phase are almost free from defects, and that the packing of defect-free portion of the chains leads

to polymer crystallization.

1.2 Polymer Crystallization and Amorphization

Polymers with high degree of chain regularity can crystallize, and the crystallization conditions

determine its crystallinity and the crystalline morphology. Polymer crystallization is composed

of three stages: nucleation, development of crystals, and impingement of crystal regions.

Polymer crystallization occurs under temperatures above the glass transition temperature and

below the melting temperature. Polymer amorphization is a process in which crystalline chains

detach from the crystals, and the crystals diminish. The amorphization process can be achieved

by melting above the melting temperature, and occurs within ns. Polymer amorphization

begins with reduction of lamella thickness along chain direction, followed by the reduction of

lamella lateral dimensions. The melting temperature is higher than the glass transition

temperature for crystallizable polymers. For PLLA the glass transition temperature is around

60 °C and the melting temperature is around 180 °C. Due to the different behaviors of polymer

crystallization and amorphization, they are individually considered as below.

4

1.2.1 Polymer Crystallization

Polymer crystallization is a process in which packing of chains occurs to reduce their energy

state. Crystalline polymers are linear and have regular molecular structures in terms of

chemical composition and tacticity. Existence of irregularities in polymer chains, such as chain

branching, change of conformation and configuration, limits the extent of crystallization. The

morphology of polymer crystals and crystallization kinetics are essential to considering polymer

crystallization.

1.2.1.1 Polymer Crystalline Morphology

Polymer crystallization involves nucleation and packing of polymer molecules onto each other,

such that crystal structures are developed. Nucleation accounts for an induction period during

which crystal structures do not develop. According to the nucleation growth theory

(Wunderlich, 1976), the induction period is a time period to obtain steady nuclei. Once a stable

nucleus is formed, it provides a surface on which the chain folds back on itself and crystallizes in

a position adjacent to the previous crystalline chains, forming crystalline regions in an unstrained

melt or dilute solution (Hoffman et al., 1976). The resulting polymer crystal structure is the

lamellae, which are thin and flat, and of about 10 nm thick and several μm or mm in lateral

dimensions. Lamellae form a more prominent crystal structure in polymer known as the

spherulite, which is a spherical aggregate ranging from μm to mm in size.

5

Figure 1.2: Sketch of the morphology of two categories of spherulites. (a) Category 1 spherulite, which features the central multidirectional growth. (b) Category 2 spherulite, which features sheaf-like unidirectional growth and double circle appearance (Norton & Keller, 1985).

Based on the structure, spherulites have been classified into two categories (Norton & Keller,

1985). In Category 1 spherulites, a central nucleating entity initiates lamella growth in all

directions, as shown in Fig. 1.2(a). The lamellae grow uncorrelated and remain so during the

spherulite development process. The spherical symmetry extends to the central region. The

development of Category 2 spherulites initiates from the unidirectional growth of one single

lamella. The spherulite morphology is attained through continuous branching and fanning from

the growing lamellae. The resulted morphology is shown in Fig. 1.2(b). The spherical

symmetry does not extend to the central region, which preserves its sheaf-like features with a

characteristic double circle appearance. It is noted that the source of nucleation of the initial

crystal itself has no consequence for the final spherulite morphology.

1.2.1.2 Polymer Crystallization Kinetics

Analyses of polymer crystallization kinetics (Schultz, 2001) follow the theory proposed by

Kolmogorov (Kolmogorov, 1937), Johnson and Mehl (Johnson & Mehl 1939), and Avrami

(Avrami, 1939; Avrami, 1939; Avrami, 1941). It is assumed that spherulites nucleate randomly

in time and space, at a constant rate n spherulites per second. Consider a spherulite nucleateing

at time τ and has grown at a steady radial velocity v from time τ. By time t, the radius of this

6

spherulite has become v(t-τ) and its volume is

Ω1(t)=4π

3v t-τ 3 (1)

Eq. (1) gives the volume of one spherulite nucleated at time τ. In the entire system, existing

spherulites continues to grow, and new spherulites start to nucleate in other places at any time.

It is reasonable to assume that once the volume is occupied by a spherulite, any other spherulites

cannot occupy the same volume. If multiple spherulites can share the volume, an expression

for the total transformed volume at time t is given as

Ω'(t)=4π

3n(τ) v t-τ 3dτ

0

t

π

3nv3t4 (2)

The increment of transformed volume between time t and time t+dt is

dΩ'(t)=4π

3nv3t3dt (3)

which accounts for the increase in transformed volume without considering the fact that the

material is consumed as the transformation proceeds. The volume available for nucleation or

for expansion of a growing spherulite is the complement, Ω0-Ω(t), of the volume already

transformed, Ω(t), where Ω0 is the total volume of the system and Ω(t) is the volume

transformed in the real system. The fractional volume transformed ϕ(t) is expressed as

ϕ(t)=Ω(t)

Ω0 (4)

The increase in actual transformed volume dΩ(t) is related to dΩ'(t) by the following relation

dΩ t = 1-ϕ(t) dΩ'(t) (5)

which gives

7

ϕ(t)=1-exp(-π

3nvv

3t4) (6)

where nv is the number of nuclei formed per second per unit volume. This expression

explicitly gives the time-dependence of the development of the overall degree of crystallization.

1.2.2 Polymer Amorphization

Polymer amorphization is a process in which crystalline polymer molecules detached from the

crystals, which leads to lamellae thinning and size reduction of crystals. Amorphization can be

achieved by melting, and polymer amorphization kinetics is relatively high as compared with

crystallization kinetics.

1.2.2.1 Polymer Amorphization Process

Polymer crystal amorphization occurs through a melting process, in which crystalline chain

segments consecutively detach from the crystal surface and diffuse into the melt. The melting

procedure is composed of two stages as theoretically considered (Belyayev, 1988) and

numerically verified (Yamamoto, 2010). The first stage of crystal melting occurs at the fold

surface, leading to a decrease of longitudinal crystal dimensions. This fold-surface melting is

very quick and reversible, giving increasingly thinner lamellae at higher temperatures. Large

molecular mobility in the crystal is essential prerequisite for this specific mode of melting. The

second melting stage becomes apparent right after the end of rapid surface melting. This stage

accounts for the chain detachment from the crystal surface, and thus the lateral dimension of

crystal decreases. The second stage of melting is slower and irreversible.

8

1.2.2.2 Polymer Amorphization Kinetics

Melting of polymer crystals occur within a temperature range, from Tm to Tm+∆Tm. The

crystal fraction melting in a temperature increment dTm within Tm and Tm+∆Tm can be

expressed as ϕ(t,Tm)dTm. The total crystallinity is then expressed as (Toda et al., 1998)

Φ(t)= ϕ(t,Tm)dTm

Tm+∆Tm

Tm

(7)

The change in ϕ(t,Tm) during melting is given by a rate equation

dt(t,Tm)=-Rm(ΔT)ϕ(t,Tm) (8)

where Rm is the melting rate coefficient and assumed as a function of superheating, ΔT=T-Tm,

expressed as

Rm=0, for ΔT<0

Rm(ΔT), for ΔT>0 (9)

Molecular dynamics simulations have suggested that crystallinity decreases within ns, and Rm is

of the order of 109 s-1 (Yamamoto, 2010). The simulation results of polymer melting process

are given in Fig. 1.3. When the lamellae are subjected to temperatures above its melting

temperature, the melting of lamella fold surfaces rapidly takes place in the first 1 ns. The fold

surface melting is then followed by an irreversible melting process, which shortens the lamellae

along their initial growth directions. Fig. 1.3(a) shows the crystallinity after the polymer is

subjected to different temperatures higher than the melting point. Within the initial 1 ns, the

crystallinity rapidly decreases due to fold surface melting. The subsequent slower melting

represents the melting of lamellae, which leads to withdraws of lamella growth front, as shown

in Fig. 1.3(b). It can be noted that crystalline lamella melting and crystallinity reduction occur

on the order of ns. Due to the relatively slow crystallization kinetics, rapid cooling after

9

melting prevents polymer molecules from recrystallization.

(a) (b)

Figure 1.3: (a) Melting of the lamellae at temperatures higher than the melting temperature. Decrease of crystallinity occurs within ns. (b) Snapshots of gradually melting lamella at

different times at 370 K. The lamella growth front retreats with time. The vertical yellow line is for the ease of identifying withdrawal of the growth font (Yamamoto, 2010).

1.3 Polymer Controlled Drug Release System

Biodegradable polymers have been used for pharmaceutical applications, especially for the

controlled drug delivery. In drug delivery applications, drug molecules are embedded in a

polymer matrix. The polymer matrix can protect the drug from biological degradation before it

is released. Depending on the properties of the polymer matrix, drug can be released in a

controlled manner. The development of drug loaded polymeric devices starts with

non-biodegradable polymers, in which drug is released through diffusion from the polymer

matrix. Subsequently, biodegradable polymers have been used, in which swelling and erosion

10

can take place.

1.3.1 Mechanisms of Drug Release from Polymer Matrix

Based on the physical or chemical characteristics of polymer, drug release mechanism from a

polymer matrix can be categorized in accordance to three main mechanisms: diffusion-controlled

mechanism, swelling-controlled mechanism, polymer degradation-controlled mechanism (Arifin

et al., 2006; Leong & Langer, 1987). In all three mechanisms, diffusion is always involved.

For a non-biodegradable polymer matrix, drug release is due to the concentration gradient by

either diffusion or matrix swelling. Matrix swelling enhances drug diffusion. For a

biodegradable polymer matrix, release is normally controlled by the hydrolytic cleavage of

polymer chains that lead to matrix erosion. Matrix degradation and erosion accelerate diffusion

of drug molecules embedded in the matrix.

1.3.1.1 Diffusion Controlled Mechanism

For the diffusion controlled system, drug diffuses from a non-degradable polymer system, and

the drug release profile is obtained by solving Fick’s second law of diffusion subject to

appropriate boundary conditions. For one-dimensional drug release from a microsphere, the

second Fick’s law of diffusion is given by:

∂C

∂t=

1

r2

∂rDr2 ∂C

∂r (10)

where D and C are the diffusion coefficient and drug concentration in the polymer matrix. The

amount of total drug released from a diffusion-controlled system into a release medium acting

essentially as a perfect sink is determined by the Higuchi’s model, given as (Higuchi, 1961)

Q= Dt(2A-Cs)Cs (11)

11

where Q is the amount of released drug after time t per unit exposed area, A is the total amount

of drug present in the matrix per unit volume, and Cs is the solubility of the drug in the polymer

matrix.

1.3.1.2 Swelling Controlled Mechanism

A swelling polymer can provide more control over the release of drug, especially when its

diffusivity in polymer is very low. For this purpose, a swellable polymer based system is

commonly made using a hydrophilic polymer so that water molecules can easily penetrate into

the polymer matrix. The imbibing water decreases the polymer concentration and thus the level

of polymer disentanglement. The polymer matrix disentanglement also leads to matrix swelling

that increases the free volume between polymer chains. The increased free volume favors drug

diffusion and thus drug release. At the interface of the polymer matrix and release medium,

polymer concentration is low due to swelling, such that polymer chains may diffuse into the

release medium. Therefore, the drug release is not only controlled by drug diffusion inside the

matrix, but also by the polymer chain disentanglement and diffusion processes.

1.3.1.3 Degradation Controlled Mechanism

A great deal of attention and research effort is being concentrated on biodegradable polymers.

These materials degrade within the human body, which avoids the need to remove a drug

delivery system after the drug release period. Drug is released as a result of polymer

degradation and erosion. Biodegradable polymers are versatile materials for a variety of

biomedical applications, especially for drug delivery systems, since their chemistry and surfaces

can be tailored to stabilize macromolecular agents and enhance the tissue site-targeting. More

importantly, the erosion kinetics can be tailored by careful selection of polymer and a variety of

12

techniques of encapsulation to control the drug release profile. In the simplest manner, the

erosion kinetics can be altered by modifying copolymer composition or the degree of

crystallinity as crystalline and amorphous polymers erode at different rates. In the thesis,

degradation refers to the polymer chain/bond cleavage/scission reaction (chemical process),

whereas the erosion designates the loss of polymer material in either monomers or oligomers

(chemical and physical process). Here, the erosion may consist of several chemical and

physical steps, including degradation. Since erosion is a more general term to capture the

overall mechanism of the bioerodible system, this section will mainly utilize the term erosion.

But the term of degradation will still be used when specific degradation processes, e.g. polymer

backbone cleavage and autocatalytic process, are involved in the model.

Polymer degradation leads to a porous matrix. The pores are then filled with liquid,

accelerating drug diffusion. Drug concentration within a matrix during polymer degradation is

as a function of space (r,z) and time (t), and is calculated from Fick’s second law (Saltzman &

Langer, 1989). The effective diffusion coefficient is as a function of porosity

Deff=Dε(r,z,t) (12)

where D is drug diffusivity via pores, and ε(r,z,t) is the matrix porosity from 0 to 1, given as

(Rothstein et al., 2009)

ε(r,z,t)=1‐1

2erf

Mw0-Mw r,z,t

√2σ2+1 (13)

The variance (σ2 ) depends on the crystallinity of the polymer matrix and corresponding

distribution of degradation rates. Mw0 is the initial molecular weight, and Mw(r,z,t) is the

molecular weight at location (r,z) and time t.

13

1.3.2 Degradation of Biodegradable Polymer in Physiological Environments

PLA degradation in physiological environments occurs via hydrolysis, in which water penetrates

into the polymer matrix through its free volume, attacking the ester bonds and causing chain

scission. During hydrolysis, water molecules attack the ester bonds of PLA via the following

hydrolysis reaction.

C C O C C O

H HO O

CH3 CH3

2OH C C

H O

CH3

OH HO C C O

H O

CH3

Hydrolysis causes chain scission, and produces shorter chains with carboxylic (-COOH) groups

and alcohol (-OH). Water molecules readily penetrate into the amorphous region, but hardly

into the crystalline region, because the polymer chains are highly packed and densely ordered in

crystals (Chu, 1981). Degradation of ester bonds occurs faster in the amorphous phase because

of its more permeable structure. Hydrolysis in the crystalline phase begins at the fold surfaces

and progresses inwards as controlled by chain scission (Tsuji & Ikada, 1998). Hydrolysis in the

crystalline phase occurs less preferentially than in the amorphous phase. A sample composed

of both crystalline and amorphous phases will more rapidly undergo hydrolysis in the amorphous

phase than the crystalline phase.

PLLA hydrolysis is accelerated by autocatalysis (Siparsky et al., 1998). The rate of hydrolysis

increases as the concentration of reaction products increases. Hydrolysis of polyester produces

shorter chains with acid and alcohol end groups. Acid end groups dissociate, leading to an

acidic environment, which accelerates hydrolysis. Therefore, the diffusion of shorter chains out

of the polymer plays a key role in controlling the overall degradation rate. Another factor that

(14)

14

complicates biodegradation is the increase of crystallinity during degradation (Zong et al., 1999;

Renouf-Glausera et al., 2005). Preferential degradation in the amorphous region leaves the

crystalline phase behind, leading to a mostly crystalline material. Chain scission as a result of

degradation also increases chain mobility which facilitates crystallization in the amorphous

phase.

1.4 Factors Determining Polymer Degradation and Drug Release

Polymer degradation and drug release are complicated processes, which are determined by a

variety of factors. These factors include polymer composition, polymer crystallinity, polymer

molecular weight, addition of drug, polymer matrix size, drug release environment, and effect of

radiation (Alexis, 2005). In physiological environments, PLLA degrades through hydrolysis, in

which water penetrates into polymer matrix through its free volume, and break the ester bonds.

In the current study, polymer crystallinity, drug addition, irradiation, and degradation and drug

release environment determine polymer degradation and drug release. These factors are

discussed below.

1.4.1 Polymer Crystallinity

It has been proposed that degradation proceeds through two main stages: in the amorphous

region, and then in the crystalline regions (Chu & Campbell, 1982). When the material is

immersed in an aqueous solution, water molecules diffuse through the polymer, and

accommodate themselves in the amorphous regions, rather than in the crystalline regions. The

chains in the amorphous regions are less packed, and thus water penetrates easily. On the other

hand, chains in the crystalline regions are aligned with a densely ordered crystalline structure,

such that little or no water can penetrate. Accordingly, hydrolytic degradation starts in the

15

amorphous regions, and chain folds in these regions degrade into fragments. As degradation

proceeds, the size of the fragments reduces and the fragments do not disentangle with other

polymer chains. At this stage, the fragments can diffuse into the release medium. This

diffusion removes these amorphous fragments, leading to a loss of materials. The regions

originally occupied by these segments become vacant and allows for further diffusion of species

with small molecular weights.

Polymer crystallinity is increased by annealing above its glass transition temperature. High

temperature annealing, however, may reduce polymer molecular weight due to thermal

decomposition, which increases the degradation rate. It has been reported (Nakamura et al.,

1989) that, after three-month in vivo implantation, the decrease in molecular weight of the

amorphous PLLA was 30-40 %, while that of the crystalline PLLA was 70-80 %. In addition,

the amorphous PLLA retained over 90 % of its initial bending strength, whereas the crystalline

PLLA lost more than 50 % of its initial value over the same degradation period. The reason

why the crystalline PLLA showed a higher degradation rate is attributed to the high

crystallization temperature of 200 °C. Thermal degradation occurs to PLLA when the

processing temperature is higher than its decomposition temperature around 200°C. After

thermal degradation, chain molecular weights reduce such that the number of end groups

increases. The end groups do not align with the crystalline chains, and offer additional free

volume. Water molecules can thus penetrate into such structure more easily, leading to faster

hydrolysis.

1.4.2 Addition of Drug and Small Molecules

While there is free volume around molecules, the long chains entangle with each other and

16

reduce their mobility. Plasticizers and drug molecules, generally small molecules with

molecular weights of several hundred g/mol, are able to permeate into the free volume around

the polymer chains and increase chain mobility. First, the existence of small molecules reduces

the effects of secondary bonding forces between polymer chains, which tend to keep the chains

to attract each other. Secondly, it increases intermolecular distance which results in an

increased free volume. Finally, as a result of the decreased secondary van der Waals bonding

forces and increased free volume, the small molecule acts as a lubricant which enhances chain

mobility. The polymer molecules are thus more easily able to move and slip past each other in

response to an applied load.

Through rendering additional chain mobility, the addition of a plasticizer lowers the glass

transition temperature of the material. It has been shown that the glass transition temperature of

PLA can be depressed by up to 45°C due to the addition of citrate esters as the plasticizer

(Labrecque et al., 1997). The increased chain mobility also affects polymer crystallization

kinetics. According to the DSC thermograms, crystallinity of annealed PLA increases with the

increasing contents of the plasticizer (Ljungberg & Wesslen, 2002). This indicates that the

existence of the plasticizer facilitates the crystallization process of PLA. While allowing

additional chain mobility, plasticizers increase free volume around chains. The additional free

volume favors water penetration into the polymer matrix, and thus hydrolytic degradation. The

extent of polymer weight loss during hydrolysis has been shown to increase with an increase in

plasticizer concentration (Labrecque et al., 1997). The addition of plasticizer may thus affect

polymer degradation in a combined way in terms of increased crystallinity and chain mobility.

17

1.4.3 Irradiation

Polymer degradation and drug release can be influenced by irradiation, including γ-irradiation

and laser irradiation. γ-irradiation is commonly used for sterilization of the drug loaded

polymer matrices for clinical applications. With the extremely high photon energy, γ-irradiation

can induce extensive bond breaking, causing undesired molecular weight reduction and chemical

structure modifications. As a heating source, on the other hand, laser irradiation has been

carried out to modify polymer crystallinity. Through crystallinity modification, polymer

degradation and thus drug release profiles over time can be tailored to meet individual needs in

clinical applications.

1.4.3.1 γ-Irradiation

γ-irradiation has been used to sterilize drug loaded polymer matrices for clinical applications,

while the γ-ray is also known to induce chemical modifications of polymer chains, such as

scission and crosslinking. The chemical modifications undesirably alter the biodegradable and

biocompatible properties of the polymer. From the GPC analysis, the γ-irradiated PLGA also

shows decreased molecular weights (Rothen-Weinhold et al., 1997). Such changes influence

the degradation rate and drug release profiles over time of drug loaded biodegradable polymers.

It has been reported that the release of 5-fluorouracil from PLGA microparticles depends on the

applied γ-irradiation dose. The exposure to the γ irradiation causes random polymer chain

scission with a decreased molecular weight. Species with smaller molecular weights do not

aligned with crystalline chains and allow for increased free volume, which leads to significant

effect on drug release. It has been observed that the initial burst of drug release is enhanced by

γ-irradiation, an undesired phenomenon in drug delivery applications (Faisant et al., 2002).

18

1.4.3.2 Laser Surface Irradiation

Polymer crystallinity plays an important role in determining its properties. Laser irradiation,

with advantage of high temporal and spatial resolutions, has been serving as a heat source to melt

polymer and reduce its crystallinity. Irradiated by the Nd:YAG laser with the photon energy

lower than PLLA bond energies, PLLA crystallinity has been reduced as quantified by the wide

angle X-ray diffraction (WAXD) measurements (Bhatla & Yao, 2009). However, the

non-uniform laser spatial intensity profile makes it difficult to achieve a uniform surface

treatment. Because of the incoherent light generated by the excimer laser, the beam can be

homogenized without interference, and the homogenized beam is spatially uniform and favorable

for surface treatment. Using the excimer laser, Lazare and Benet treat poly(ethylene

terephthalate) (PET) films, and demonstrate the presence of a thin amorphous surface layer

(Lazare & Benet, 1993). Lazare and Benet report that photochemical reactions are possible in

addition to photothermal effects, since excimer lasers operate at UV wavelengths with photon

energies higher than certain bond energies in polymers. In another attempt to decrease PET

film crystallinity using the excimer laser (Dunn & Ouderkirk, 1990), the possible chemical

changes are studied via X-ray photoelectron spectroscopy (XPS). It is found that low fluences

cause very little chemical changes. These chemical changes include the dissociation of ester

bonds and reduce of the number of oxygen atoms due to the emission of small molecules

(Watenabe et al., 1993); these modifications could alter the original properties of the polymers.

1.4.4 Degradation and Drug Release Environment

PLLA degradation and its drug release are a strong function of its degrading and release

environments. Two main factors are the temperature and pH. In a physiological environment,

19

the complete degradation of PLLA has been reported to take several years (Tsuji et al., 2000).

During the initial research and development processes, it is desired to proceed with a reduced

degradation and drug released period. To accelerate the degradation of biodegradable

polyesters, conducting degradation tests at elevated temperatures (Tsuji & Tsuruno, 2010; Weir et

al., 2004) and at acidic or alkaline pH (Belbella et al., 1996; Lam et al., 2008) are commonly

attempted. Under the alkaline pH, although the degradation mechanism is still based on

hydrolysis, molecular weight and mass loss results differ due to different degradation pathways

followed (Tsuji & Ikada, 1998; Lam et al., 2008). For accelerated conditions surface

degradation dominates, while bulk degradation pathway dominates for simulated physiological

conditions. The dominant surface degradation in the accelerated conditions can be attributed to

the fact that acidic or basic medium favors degradation while the diffusion rate remains the same,

so that it hydrolyzes the surface region before it diffuses into the bulk. Also, bi-modal

molecular distribution as a result of heterogeneous degradation of crystalline and amorphous

domains is not detected.

On the other hand, it has been shown that increased temperature appears to be suitable for

accelerating PLLA degradation relative to its physiological degradation. The evaluations of

molecular weight and film mass have similar patterns in different degradation temperatures, and

degradation at elevated temperature proceeds via the same mechanism to that at 37°C in vitro

and in vivo The degradation rate of PLLA crystals is given as a function of temperature, T

(Tsuji & Tsuruno, 2010)

kdegradg

mol·h=1.20×1012exp

-9.04×103

T (15)

The results obtained in this study can thus be potentially applied to predict the PLLA degradation

20

behavior in the physiological environments at lower temperatures.

1.5 Experimental Considerations

In the study, experiments are composed of three major parts: sample preparation, laser treatments,

and characterization. PLLA samples are prepared in both solvent-involved and solvent free

processes. Laser treatment is conducted on a KrF excimer laser operating at a wavelength of

248 nm and pulse duration of 30 ns. Appropriate characterization methods are used to

determine sample properties. The three parts are detailed as follows.

1.5.1 Sample Preparation

Depending on whether solvent is used, sample preparation methods can be divided into

solvent-involved process and solvent-free process. The solvent-involved process includes spin

coating and solvent casting. The solvent-free process includes thermal compression and

extrusion. The solvent-involved processes have been commonly used for fundamental studies,

while products in drug delivery industry are generally fabricated through the solvent-free

processes, as discussed below.

1.5.1.1 Solvent-Involved Sample Formation Process

In fundamental studies concerning PLLA crystallization, PLLA samples are generally fabricated

through solvent-involved processes such as solvent casting or spin coating. Solvent casting is a

simple film formation technique in which polymer is dissolved in an appropriate solvent, and the

quiescent polymer solution is deposited on a substrate, while evaporation of solvent leaves the

solid polymer film. This process generally takes hours. During the casting process

crystallization could occur in solution if the evaporation rate is low enough for polymer

21

molecules to diffuse to the crystal growth front and overcome the energy barrier of deposition, so

that crystals can grow. Spin coating generates polymer films with thinner and more uniform

thickness. It involves rapid rotation of polymer solution, with the centrifugal force pushing it to

flow radially outward, decreasing its thickness. The solvent evaporates simultaneously. This

process typically finishes within several seconds or minutes. Due to rapid evaporation, it is less

likely that the material can crystallize.

The solvent-involved processes are affected by multiple variables including solution

concentration, solvent vaporization rate, the solvent used, and subsequent melting and

recrystallization. Fundamental understanding of crystallization and crystalline morphology has

been established through the investigation of the effects of these variables. Because of the wide

variety of process parameters allowed in the solvent-involved process, both solving casting and

spin coating methods are used in this thesis to investigate crystalline morphology developed in

different environments.

1.5.1.2 Solvent-Free Sample Formation Process

Solvent-involved methods, although commonly used in fundamental studies, have intrinsic

drawbacks which prevent their usage in biomedical applications, including non-uniform crystal

structure along film thickness (Zilberman et al., 2001), non-uniform dissolution of polymer

blends in solvent (McDonald et al., 2010), and issues related to residual solvent (Koegler et al.,

2002, which is toxic to cell (Mikos & Temeoff, 2000). Thus, solvent-free processes, including

thermal compression and extrusion, have been commonly used in biomedical applications such

as drug delivery. Solvent-free processes also allow for fabrication of matrices with complicated

shape for bio-applications (Jing et al., 2006), while the film thickness generated by

22

solvent-involved processes is limited because different solvent vaporization rate along solution

depth leads to non-uniform film structure along thickness. In addition, PLA-based polymer are

also usually blended with other polymers such as PCL to adjust the degradation and drug release

rate (Tanaka et al., 2006), while solubility of each component is different, which results in

incompatible blends and affects drug release rate (Liau & Chang, 1999).

Because of the above-mentioned constraints and potential issues associated with solvent

involvement in biomedical applications, the solvent-free sample formation process is used to

produce samples for biodegradation and drug release tests. The solvent-free method is realized

by thermal molding. The mold used in this study is composed of three parts, as shown in Fig.

1.4. The mold has six cavities, each with a diameter of 10 mm. Six pins are also shown in Fig.

1.4. Appropriate weight loading is applied onto the pin during the molding process to reduce

the porosity of the prepared samples.

The degradation patterns among the solvent-involved and solvent-free sample formation

processes are similar. It has been shown that PLLA films prepared in the two methods both

degrade via the same hydrolysis route: initial cleavage of ester bonds and reduction of molecular

weights, followed by the mass loss as a result of diffusion of degradation products (Tsuji &

Tsuruno, 2010; Weir et al., 2004). Thus, the results obtained based on the samples prepared

with the solvent-free methods can be applicable to those prepared with the solvent-involved

methods.

23

Figure 1.4: Mold used in this study for the thermal compression processes.

1.5.2 Laser System

Laser treatment, with the advantages of high spatial and temporal resolutions, is conducted as a

heating source to modify polymer crystallinity in this study. A laser system generating ns

pulses is desired, because short pulse duration allows for fast cooling of polymer, and prevents

its recrystallization. The applicability of laser treatment on polymer crystallinity modifications

has been demonstrated by lasers with a wide range of wavelengths, from UV to IR. Lasers

operating at IR generate photons with energy lower than polymer binding energies, and

photochemical effects can be neglected (Bhatla & Yao, 2009). UV irradiation, with photon

energies higher than polymer binding energies has also been conducted, while the effects on

polymer chemical structures are not well understood. Lazare and Benet have suggested that

polymer amorphization process induced by UV laser irradiation involves both photothermal and

photochemical processes (Lazare & Benet, 1993). On the other hand, Dunn and Ouderkirk

proposed that no significant chemical changes occur after UV laser amorphization (Dunn &

Ouderkirk, 1990). The understanding of UV laser irradiation effects on polymer chemical

structures is diverged and needs further investigation, so as to broaden the applicability of laser

treatment on polymer crystallinity modifications.

Accordingly, laser treatments of PLLA in this study are performed on the KrF excimer laser

2 cm

24

system, which is operated at a wavelength of 248 nm, and produces laser pulses with pulse

duration of 30 ns. The system is shown in Fig. 1.5. The maximum repetition rate is 50 Hz,

and the maximum average power is 30 W. The system is composed of a homogenizer and a

mask. The homogenizer ensures a spatially stable laser beam output by dividing the laser beam

into multiple subsections, and then superimposing each other. The working principle of a

homogenizer is illustrated in Fig. 1.6. An array of lens divides the incident beam into a number

of subsections. Each subsection, after passing through the spherical lens, spreads over the focal

plane, on which the beam is homogenized. The homogenized beam possesses a top hat spatial

profile. The homogenized beam is then masked. The size of the mask is selected such that the

output beam size is 1 mm by 1 mm.

Figure 1.5: The excimer laser system used in the study. The system is composed of the homogenizer and mask to generate a top hat pulse with a desired spot size. Laser pulse

irradiation is controlled by the computer.

25

Spherical lens focal length=F

Homogenized planeArray of lens Spherical lens

Figure 1.6: Sketch of the working principle of laser beam homogenization.

Laser pulse is generated in the laser tube filled with excimer laser gas mixture, Kr and F2.

Before lasing, the Kr and F atoms are weakly bound, and exist in form of Kr and F2. Kr and F

atoms are bound only in their excited state. To generate laser pulses, a high voltage is supplied

to ionize the Kr and F atoms such that the Kr+ and F- ions are created. The Kr+ and F- ions are

combined to form the excited KrF molecule. The excited KrF molecule is not stable, and

transitions rapidly to the ground state. When the KrF molecule transitions to the unbound lower

energy state, an ultraviolet photon at a wavelength of 248 nm is emitted.

1.5.3 Material Characterization

In the study, PLLA samples before and after laser treatments, degradation and drug release tests

have been characterized. The wide-angle X-ray diffraction (WAXD) detects the crystalline

phases, and also allows for quantitative determination of crystallinity. The Fourier transform

infrared (FTIR) spectroscopy captures the interactions within polymer chains, which reveal the

information of polymer structures. The X-ray photoelectron spectroscopy (XPS) detects

surface chemical compositions before and after laser treatments. The gel permeation

chromatography (GPC) captures the molecular weight change after degradation. The

spectrophotometry detects the amount of drug release into the release medium. The differential

26

scanning calorimetry (DSC) measures the thermal properties of polymer, which indicates the

mobility of polymer chains. The working principle of GPC, as well as the applications of

WAXD and DSC to determine polymer crystallinity and chain mobility, respectively, is detailed

below.

1.5.3.1 Wide-Angle X-ray Diffraction

WAXD is an X-ray-diffraction technique generally used to investigate polymer crystal structures.

WAXD specifically refers to the analysis of crystalline peaks scattered to large angles, with

2θ>5° where 2θ is the angle between the directions of the scattered X-ray and the incident X-ray.

According to the Bragg’s law, the angle θ is determined as a function of the X-ray wavelength, λ,

the distance between two atomic planes, d, given as

nλ=2dsinθ (16)

While diffraction occurs in arbitrary directions, the equation demonstrates that only when the

angle θ satisfies the relationship can the diffracted rays be positively reinforced and observed.

For all values of θ that do not satisfy the equation, the reflected X-rays from multiple atomic

planes will be out of phase with one another and cancel each other, such that no diffraction is

observed (Alexander, 1969).

Diffraction from the crystalline phase gives well-defined peaks based on the Bragg’s law. On

the other hand, diffraction from an amorphous phase occurs in all directions, while a broad hump

can still be observed. This is because the existence of favored interactomic bond lengths allows

for weak positive reinforcement in particular directions. It can be assumed that the diffracted

intensity by an atom in a crystal is the same as that in an amorphous region. Based on the

assumption, crystallinity of a semi-crystalline polymer can be determined from the measured

27

WAXD profiles. An example is given by the semi-crystalline polyethylene, with its WAXD

profile given in Fig. 1.7 (Campbell et al., 2000). Prominent diffraction peaks are observed and

attributed to (110) and (200) diffractions from the orthorhombic polyethylene crystals. Because

the crystalline peaks are much more prominent than the amorphous hump, it is often possible to

separate them. The integrated intensity of the crystal peaks divided by the integrated intensity

of the amorphous hump can represent the ratio of the number of atoms in the crystalline phase to

the number of atoms in the amorphous phase. The crystallinity ϕ is therefore given as

ϕ=Ic

Ic+Ia (17)

where Ic is the total area attributed to crystalline peaks and Ia is the area of the amorphous

hump. Practically, the amorphous hump can be obtained by preparing a completely amorphous

sample on which the WAXD measurement can be conducted.

Figure 1.7: WAXD profile from semi-crystalline polyethylene. The dashed line shows the estimated position of the amorphous scattering intensity (Campbell et al., 2000).

1.5.3.2 Gel Permeation Chromatography

GPC is a type of size exclusion chromatography (SEC) which separates the material based on the

size. The purpose of using the GPC measurements is to provide the information of molecular

28

weight distribution of a given polymeric material. The obtained results typically depict the

response from one or several detectors on the ordinate versus the chromatographic elution time

or the logarithm of molecular weight on the abscissa.

In GPC, a porous column is used to separate the materials by their sizes. Smaller molecules

pass through the porous column at a slower rate, while larger molecules elute at a higher rate.

The separation of a mixture of two sizes of polymeric molecules is given in Fig. 1.8. During

the measurement, a high pressure gradient is applied along the column length direction, and a

liquid mobile phase is continuously passed through the column at a fixed flow rate, generally 1

mL/min. Upon injection of the polymeric sample, illustrated in Fig. 1.8(a), the material is

located on the head of the column. Because of the applied pressure gradient, the sample

polymeric molecules tend to pass through the porous column, Fig. 1.8(b). The smaller

polymeric molecules are trapped by the pores in the column, while the larger molecules cannot

be accommodated in the pores, as shown in Fig. 1.8(c). Therefore, the smaller molecules move

more slowly through the column, and the larger molecules move more rapidly. Eventually,

molecules with two different sizes are separated and pass through the column at different elution

time as shown in Fig. 1.8(d). The detectors are located at the end of the column, and generate

chromatographic bands for polymeric molecules with different sizes. The intensity of the

chromatographic bands is proportional to the material concentration.

(a)

29

(b)

(c)

(d)

Figure 1.8: GPC separation of polymer molecules with two different sizes. (a) the mixture of polymer molecules upon injection; (b) the mixture enters the porous column because of the

applied pressure gradient through the column length; (c) the small molecules are trapped by the pores and flow slowly, but larger molecules are not, from which the molecules are separated; (d)

complete elution of large molecules.

Practically, the GPC chromatographic bands are broad, which reflects the fact that not all the

polymer molecules have the same degree of polymerization, and thus the molecular weight. To

consider the distribution of the molecular weight of a given polymeric material, different average

values of molecular weight have been defined. The most commonly used statistical methods

include the calculation of the number average molecular weight (Mn) and the weight average

molecular weight (Mw). Mn is defined as

Mn=∑ NiMi

Ni=1

∑ NiNi=1

(18)

where Ni is the number of moles of species i, and Mi is the molecular weight of species i. Mn

30

gives the average over the number of the molecules. Mw is defined as

Mw=∑ NiMi

2Ni=1

∑ NiMiNi=1

(19)

Mw gives the average over the weight of each polymer molecules. The ratio of Mw to Mn is

known as the polydispersity index (PDI), which gives a quantitative estimate of the molecular

weight distribution for the polymer. A large PDI represents a broad distribution of molecular

weight. A polymer with a PDI of unity contains molecules with the same molecular weight.

1.5.3.3 Differential Scanning Calorimetry

DSC is a thermoanalytical technique, which allows for the determination of thermal properties of

a material. During the DSC measurements, the sample and the reference are heated

independently while both are maintained at the same temperature during heating. To maintain

the same temperature, different amount of heat needs to be supplied to the sample during the

measurement. The difference of the provided heat is recorded so as to obtain the enthalpy and

entropy changes during phase transformations of the sample. In addition, the information of the

glass transition temperature (Tg), crystallization temperature (Tc), and melting temperature (Tm)

of a given sample can be determined.

While directly providing with the thermal properties of polymeric samples, information obtained

from the DSC measurements also reveal the polymer chain mobility in a given sample. This is

because chains with higher mobility require lower energy to crystallize and melt, and thus their

crystallization and melting occur at a lower temperature. This is experimentally demonstrated

in Fig. 1.9 (Yeh et al., 2009), in which PLA is blended with triacetine (TAc), an effective

plasticizer for PLA, at different concentrations. After blending, the values of Tg and Tm reduce

31

with increasing TAc concentration. Tg reduces from 62.3°C to 29.2°C and Tm reduces from

158.8°C to 151.9°C as the TAc concentration increase from 0 to 30 wt %. The reduction of Tg

and Tm is attributed the existence of TAc, which as the effective plasticizer for PLA increases the

free volume within PLA molecules and thus their mobility.

Figure 1.9: DSC thermograms of poly (lactic acid) blended with plasticizer at different concentrations (a) 0 wt %, (b) 5 wt %, (c) 10 wt %, (d) 15 wt %, (e) 20 wt %, (f) 25 wt %, and (g)

30 wt %. The scanning rate is 40°C/min (Yeh et al., 2009).

1.6 Research Objectives and Organization of the Proposal

The purpose of this research is to study the effect of film formation methods on PLLA crystalline

morphology, to study the effect of laser treatments as a heating source on crystallinity

modification of PLLA, and to evaluate effect of laser crystallinity modification on drug release

profile from PLLA in the drug delivery applications. Specifically, the aims of this research are

as follows.

(1) To investigate the morphology and crystal structure in the PLA films formed through solvent

32

casting and spin coating followed by annealing.

(2) To reduce surface crystallinity of PLLA films through melting by excimer laser treatment,

while minimizing possible chemical modifications.

(3) To investigate the effects of laser surface treatment on PLLA degradation in terms of

molecular weight, sample mass, and crystallinity.

(4) To investigate the effects of polymer degradation and drug loading on drug release

mechanisms, as well as the effect of laser crystallinity modification on drug release profile.

In Chapter 2, the effects of film formation method and annealing on the crystals formed in PLLA

films are investigated. Two solvent involved film formation processes, solvent casting and spin

coating, are conducted to prepare the PLLA films. Subsequent annealing on the prepared films

is carried out. The resulting crystalline morphology, molecular order, conformation, and

intermolecular interaction are examined using optical microscopy, wide-angle X-ray diffraction,

and Fourier transform infrared spectroscopy. It is observed that solvent casting produces

category 1 spherulites while annealing the spin coated films leads to spherulites of category 2.

The crystal structure of the two kinds of films also shows distinct features. The results enable

better understanding of the crystallites in PLLA, which is essential for its drug delivery

applications.

In Chapter 3, the excimer laser is used to induce surface crystallinity modifications of PLLA.

The effects of excimer laser irradiation on the surface morphology, crystallinity, and chemical

modifications are investigated via optical microscopy, wide-angle X-ray diffraction, and X-ray

photoelectron spectroscopy. A model is developed to numerically examine the spatial and

33

temporal temperature profiles, as well as the amount of chemical modifications. It is found that

PLLA crystallinity decreases as a function of laser fluence, and that the amount of chemical

modifications is minimized by annealing before laser treatment and reducing laser fluences,

which decrease free radical mobility and thus the dissociation quantum yield. A working

window is demonstrated in which PLLA crystallinity decreases with no measurable chemical

modifications, and the working window can be enlarged by decreasing free radical mobility.

In Chapter 4, the effect of laser induced PLLA crystallinity reduction on its biodegradation is

investigated. The degradation is characterized in terms of molecular weight, sample mass, and

crystallinity during the degradation process. Samples having lower initial surface crystallinity

are shown to have higher rates of molecular weight reduction and earlier mass loss than

non-laser treated samples, as observed from gel permeation chromatography and mass change.

Wide-angle X-ray diffraction measurements show that crystallinity increases with degradation.

A numerical model is implemented from hydrolysis and diffusion mechanisms to investigate the

effect of laser irradiation on biodegradation. Controlled laser treatment of PLLA offers a

method for constant drug release through the reduction of surface crystallinity.

34

Chapter 2: Effect of Film Formation Method and Annealing on Morphology and Crystal

Structure of Poly(L-Lactic Acid) Films

2.1 Introduction

There is a significant interest in use of biodegradable polymers due to their biocompatibility and

biodegradability. Poly lactic acid (PLA) is a biodegradable polymer and can be obtained from

renewable sources such as corn starch. It has a wide range of applications in food packaging

and tissue engineering. It is also preferred in drug delivery because it can degrade into

bioabsorbable products in physiological environments. Its degradation mainly comes from the

cleavage of its ester groups, and is sensitive to chemical hydrolysis. In this application, drug

molecules are encapsulated in the polymer by dispersing or dissolving in the polymeric solution,

formed through melting the polymer or dissolving it in a solvent. During the process polymer

may crystallize, which influences its degradation (Nampoothiri et al., 2010). Tsuji and Ikada

(Tsuji & Ikada, 1998) proposed that the degradation of PLA begins in the amorphous region

between the lamellae, followed by the disorientation of the lamellae and disappearance of the

spherulitic structure. Namely, crystallites affect PLA degradation and thus play a significant

role when the material is used in the drug delivery system.

Efforts have been done to investigate the crystallization behavior and resulting crystal structures

of PLA films. Among them, solvent casting (Tsuji & Ikada, 1995; Bhatla & Yao, 2009) and

spin coating (Li et al., 2009) are the two commonly used film formation methods. Solvent

casting is a simple film formation technique in which quiescent polymer solution is deposited on

a substrate, while evaporation of solvent leaves the solid polymer film. This process generally

takes hours. During the casting process crystallization could occur in solution if the

35

evaporation rate is low enough for polymer molecules to diffuse to the crystal growth front and

overcome the energy barrier of deposition, so that crystals can grow. Category 1 spherulites, in

which lamellae grow in all directions and the spherical symmetry extends to the central region as

defined by Norton and Keller (Norton & Keller, 1985), can be commonly observed in the as cast

films (Tsuji & Ikada, 1995; Takeda, 2005). Spin coating generates polymer films with thinner

and more uniform thickness. It involves rapid rotation of polymer solution, with the centrifugal

force pushing it to flow radially outward, decreasing its thickness. The solvent evaporates

simultaneously. This process typically finishes within several seconds or minutes. Due to

rapid evaporation, it is less likely for crystallites to develop in the as coated films.

To increase crystallinity in as cast/coated PLA films, further annealing has been attempted. For

the as cast films fully covered by spherulites formed during casting, subsequent annealing

increases the degree of crystallinity, but does not alter the morphology (Tsuji & Ikada, 1995;

Bhatla & Yao, 2009). As coated films are generally amorphous. Annealing such films from

the glassy state can generates crystal structures different from those developed in as cast films.

One common feature is the nodular structure (Zhao et al., 2009; Su et al., 2009; Hsu et al., 1986);

the nodule is the nanometer-sized crystalline particle developed during annealing from the glassy

state (Ogawa et al., 1985). As annealing time and/or temperature increase, the nodules enlarge,

and may develop into larger needle-like crystals (Zhao et al., 2009). It is also observed that

nodules develop into lath-like lamellae through merging or lateral aggregation (Su et al., 2009;

Hsu et al., 1986). At larger scale, the sheaf-like structures can be observed (Bassett et al., 1988;

Gao et al., 2004), and category 2 spherulites, in which lamellae grow unidirectionally in the

central region and the spherical symmetry does not extend to the center (Norton & Keller, 1985),

are developed (Ivanov & Jonas, 1998). Crystal structures developed by annealing from the

36

glassy state thus show distinct features from those developed in solution.

In spite of the fact that different film formation methods and subsequent annealing lead to

different morphologies and crystal structures, their effects have not been fully explored. Pluta

and Galeski (Pluta & Galeski, 2002) conducted the wide-angle X-ray diffraction (WAXD)

measurement on PLA films crystallized from melt and glass, and peculiarity in the WAXD

profiles for the latter is noticed. It reveals further information of the crystal structure, which is

not clear. Li et al. observed various PLA film morphologies crystallized from glass, while the

crystal structure is not addressed (Li et al., 2009). For annealed solvent cast PLA films, crystal

structures developed in the two crystallization steps are expected to be different. The difference

has not drawn much attention. It is also worth considering the causes leading to different

morphologies in the films.

The objective of this study is to investigate the morphology and crystal structure in the PLA

films formed through solvent casting and spin coating followed by annealing. The morphology

is observed by optical microscopy, and crystal structure is studied through the WAXD and

Fourier transform infrared (FTIR) measurement. The former is utilized to study the molecular

order and the latter is to investigate the conformation of molecules and intermolecular

interaction.

2.2 Background

Poly (L-lactic acid) (PLLA) has a glass transition and melting temperatures of around 60°C and

175°C respectively, and can crystallize in α, β, and γ forms depending on crystallization

conditions. The α form exists in solvent cast and annealed samples, and the unit cell structure is

orthorhombic with two 103 polymeric helices (Kobayashi et al., 1995; Miyata & Masuko, 1997;

37

Sasaki & Asakura, 2003). The lattice parameters are a=10.66 , b=6.16 , and c=28.88

(Miyata & Masuko, 1997), but the values may be slightly different (Kobayashi et al., 1995;

Sasaki & Asakura, 2003; Davis et al., 1968). In the orthorhombic unit cell, PLLA chains are

folded in the <110> direction (Miyata & Masuko, 1997; Kikkawa et al., 2002).

2.2.1 Nucleation and Crystallization

Polymer crystallization in dilute solution is composed of induction period, primary

crystallization, followed by secondary crystallization. The induction period is the time to

obtain steady nuclei before crystallization can occur (Imai et al., 1992). In primary

crystallization, spherulites are developed and eventually impinge with each other.

Crystallization can continue after the impingement, known as the secondary crystallization. It

is contributed by the amorphous chains within the lamellae with slower crystallization rates due

to chain imperfection (Schultz, 2001). Crystallization from the melt follows a different route, in

which nanometer-sized nodules are developed first followed by coalescing with each other to

form larger crystals such as lamellae (Su et al., 2009). The concurrent orientation adjustment of

neighboring nodules is required for prefect coalescence. This crystallization route accounts for

annealing of the as cast/coated film and the resulted crystal structures.

2.2.2 Mobility of Polymer Molecules in Solution and Melt

Polymer molecules move via diffusion, and the diffusion coefficient D is proportional to

mobility μ via D=µkBT, with kB the Boltzmann constant and T temperature. The value D is

related to solution concentration, which is divided into dilute, semidilute, and concentrated.

Molecules in dilute solutions are separated, occupying a spherical region of radius Rg=b√N/6

where b and N are the length and number of submolecules, respectively (Doi & Edwards, 1988).

38

As the concentration increases, the polymer coils start to overlap at the concentration c* given

as (Cotton et al., 1976)

MNA

43 πRg

3≤ c*≤

MNA

Rg3 (1)

where M is molecular weight and NA is the Avogadro constant. In the concentrated solution,

such as the polymer bulk above its glass transition temperature, the concentration fluctuation is

negligible. The cross-over concentration from semidilute to concentrated c** is (Doi &

Edwards, 1988)

c**v

MsNA

b6 (2)

where v and Ms are the excluded volume and molecular weight of the segment, respectively.

For dilute solution, the Stokes-Einstein equation gives the diffusion coefficient

D=kBT

6πηRH (3)

where η and RH are the solvent viscosity and hydrodynamic radius, respectively. The latter is

related to the radius of gyration Rg by RH=0.537Rg (Akcasu & Han, 1979). If the

concentration is extremely high, the chains are entangled and their movements are described on

the basis of the reptation theory (de Gennes, 1971), and the diffusion coefficient is then given by

(Doi & Edwards, 1988)

D=kBT

Nζ (4)

where ζ is a Rouse model parameter.

39

2.2.3 Structure of Spherulites

Spherulites are composed of the crystalline, the amorphous phases, as well as the rigid

amorphous phase (RAP). Chains in RAP are immobile and remain vitrified even above the

glass transition temperature (Menczel & Wunderlich, 1981). In terms of architectures,

spherulites can be developed into two categories (Norton & Keller, 1985), depending on the

competition between the order in the crystal structure and the disorder of the crystallographic

orientation; the disorder is associated with the randomization of secondary nucleation at the

growth front (Granasy et al., 2005). The randomness disrupts the crystalline anisotropy. For

category 1 spherulites, disrupt occurs early, and multiple directional growth is achieved in the

beginning. The disorder disrupts the crystalline anisotropy later for category 2 spherulites,

leading to the threadlike structure at the center and sheaf outside.

The origin of disorder during the secondary nucleation is caused by three reasons (Granasy et al.,

2005). The first reason is the static spatial heterogeneity such as foreign particles, which

deflect the tip of growing lamellae and perturb their growth directions. The second reason is

due to the small rotational diffusion coefficient when compared with translational diffusion

coefficient, which suggests that the molecule has difficulty changing its orientation when

depositing on the growth front and thus causes disorder. The third reason is the

non-crystallographic branching. From direct AFM observation (Li et al., 2001), it may occur at

a short distance away from the mother lamellae. This gives evidence that branching is formed

by the loose loop or protruding cilia of molecules trapped in the mother lamellae. The third

reason dominates based on simulation results (Granasy et al., 2005).

40

2.3 Materials and Methods

2.3.1 Sample Preparation and Annealing

PLLA granule samples were provided by PURAC and used as received. The inherent viscosity

(η) in chloroform at 25°C is 1.6 dl/g, and the molecular weight is M=56,274 g/mol from the

Mark-Houwink equation, η=0.454×10-4M0.73 (Schindler & Harper, 1979). To prepare solvent

cast films, the granules were dissolved in methylene chloride (CH2Cl2) from Sigma Aldrich, with

concentration of 0.1 g PLLA/3 mL CH2Cl2. The solution is stirred with a magnetic stirrer for 4

hours, and 15 mL solution was cast in a covered Petri dish with a glass slide as a substrate. The

solution was left in the hood at 25°C for 24 hours. The film formed on the glass substrate was

used as the solvent cast films. Its thickness, measured by profilometry, is 20-25 μm. The

solution with doubled concentration was spin-coated on a glass substrate with a rotational

acceleration of 255 rev/min2 and rotational velocity of 1000 rpm. The film thickness is around

2.5 μm. The thermophysical property and crystallization behavior in films thinner than 1 μm

differs from that in thicker films (Frank et al., 1996). Since both solvent cast and spin coated

films are thicker than 1 μm, the effects of film thickness are insignificant. The films were

annealed at 80°C, 110°C, and 140°C for 1 and 3 hours; some spin coated films were annealed at

140°C for 8 and 24 hours. The annealed films were quenched at room temperature after

annealing.

2.3.2 Characterization

Morphology of PLLA films was observed with an optical microscope (Olympus BX60) in the

transmission mode. WAXD measurements were accomplished using the Inel X-ray

diffractometer. Film specimens were exposed to the monochromatic CuKα radiation with

41

wavelength λ=0.15418 nm at 40 kV and 30 mA. Care was taken to adjust the incident angle of

the X-ray to prevent penetration into the substrate. A Perkin Elmer Spectrum 400 FTIR

spectrometer with attenuated total reflectance (ATR) attachment using ZnSe crystal was used to

measure the IR spectrum of the films peeled from the glass substrate. The data were recorded

with a 4 cm-1 resolution.

2.4 Results and Discussion

2.4.1 Morphology

Spherulites already appear and impinge in solvent cast films before annealing because solvent

evaporates at a rate low enough for crystals to grow during the casting process. The spherulitic

structure resembles category 1, and remains unchanged after annealing. The morphology for

the annealed solvent cast film is given in Fig. 2.1(a). No crystallite exists in spin coated films

before annealing as will be confirmed by WAXD. This is due to the fact that during coating,

the solvent evaporates faster than the induction period. It is pointed out that for linear

poly(decamethylene terephthalate) with molecular weight approximately 40,000 crystallized in

solution at 32.1°C, the induction period is between 6 and 34 minutes for solution concentration

between 10 and 2.5% (Lanceley & Sharples, 1966). Because of the similarity of the molecular

weight and crystallization temperature, the induction time of the PLLA in this study is expected

to be on the order of minutes, which is much longer than the solvent evaporation time (less than

45 seconds, assuming all the solvent evaporates by the end of coating process). It is reasonable

to obtain an amorphous structure through spin coating.

After annealing the spin coated films, paired dots are developed with a bright threadlike fiber in

between, which are category 2 spherulites, as shown in Fig. 2.1(b). The difference between the

42

two categories results from the order of the secondary nucleation: if disorder disrupts the

orientation of crystal growth early, category 1 spherulites can be generated, and

non-crystallographic branching is a main reason causing the disorder (Granasy et al., 2005).

The different branching behaviors in the two films results from the different amount of supplying

chains at crystallization sites. The chains are carried by a compensation flow induced by the

volume shrinkage because of crystallization (Schultz, 2001), and chain mobility is essential. To

investigate the mobility, the concentration and diffusion coefficient are calculated. For the

solvent cast films, the spherulites are developed in solution, whose overlap concentration,

calculated from Eq. (1), is between 0.6 and 2.3 g/ml. (For calculation, the monomer is taken as

the submolecule, and its length is estimated by the size of the unit cell (Miyata & Masuko, 1997).

The number of submolecules is therefore the degree of polymerization, 56,274/72=782.) Since

the solution concentration, 0.033 g/ml, is much less than the overlap concentration, the solution

is dilute and Eq. (3) is used to determine the diffusion coefficient to be 3 10-6 cm2/s. Due to

extremely high concentration, the molecules in spin coated films experience entanglement during

crystallization. The diffusion coefficients for PLLA bulk above the glass transition temperature

is as a function of temperature and degree of polymerization (Zhang et al., 2007). For the

temperature around 400 K and the degree of polymerization of 782, the diffusion coefficient is

estimated on the order of 10-9 cm2/s. Their results agree with the relation between diffusion

coefficient and chain length stated by Eq. (4). Therefore, during the periods when the

spherulites are developed, chain mobility in the spin coated films is much less than that in the

solvent cast films. The scenario of chain movement in the crystallization front in both films can

thus be described as follows. For solvent cast films, the high chain mobility enables a flow of

surrounding material to the crystallization front to compensate the volume shrinkage, supplying

43

molecules to the branching growth surface, so that branching could easily happen. For spin

coated films, due to low mobility, the polymer chains connecting crystalline and non-crystalline

regions are entangled with adjacent molecules and hard to relax. The shrinkage is thus not

compensated by flow, but by local extension of molecules to the front (Schultz, 2001).

Therefore, the supplying material is less available, which is unfavorable to the growth of

branches. This explains the origin of different spherulitic structures observed in solvent cast

and spin coated films.

(a)

(b)

Figure 2.1: (a) The morphology of solvent cast film annealed at 140°C for 3 hours. The morphology is similar to that of the as cast film. (b) The morphology of spin coated film

annealed at 140°C for 3 hours. The eye structure is observed.

44

For the spin coated films annealed at 140 °C for longer time, the boundaries can be observed

more clearly, while the number of observable eye structures is reduced, as shown in Fig. 2.2.

The “disappeared” eyes are essentially enclosed by the lamellae growing in the direction outward

from the film, suggesting a later stage of the development of category 2 spherulites. Although

the eyes are enclosed, the spherical symmetry does not extend to the center, and the sheaf-like

features are preserved, which agrees with the description of category 2 spherulite proposed by

Norton and Keller (Norton & Keller, 1985) and differentiates from the spherulites formed in the

solvent cast films. The existence of the eye structures suggests that lamellae do not branch into

the region even in the later stage of development. The possibilities can be twofold. First, the

lamellae develop at the expense of the crystallizable material, and the non-crystallizable material,

such as atactic components or branched chains, does not participate in crystallization and are

accumulated around the growing lamellae. As the lamellae grow and splay, the

non-crystallizable material may be trapped in the eyes, which prevents lamellae from branching

into it. Another reason could be the low chain mobility within this region. Some

crystallizable chains may be highly entangled and their low mobility makes it hard to relax and

crystallize. As such chains accumulate in the eyes, it is more difficult for lamellae to grow into

the region.

The morphology of PLLA crystals developed in spin coated films after annealing is a function of

film thickness and annealing temperature. Spherulites are observed in the films thicker than 50

nm, annealed at temperatures between 100 and 145 °C, while single crystals and dendrites are

formed in a thinner film or at higher annealing temperatures (Maillard, & Prud’homme, 2008).

This agrees with the spherulitic morphology observed in our spin coated films, developed within

a 2.5-μm film thickness and annealed at temperatures up to 140 °C. The difference of

45

crystalline morphologies as a function of film thickness can be related to the local chain mobility

near the polymer/substrate interface, depending on the interaction between the polymer and

substrate. A reduced glass transition temperature has been observed for a PLLA film thinner

than 100 nm, suggesting a higher local chain mobility near the surface than in the bulk

(Narladkar et al., 2008). Similar results are proposed as stated in Sec. 2.3.1 (Frank et al., 1996).

The thickness effect is insignificant in the current study due to the much larger film thickness.

Figure 2.2: The morphology of spin coated film annealed at 140°C for 24 hours. Eye structure

remains observable while the boundary of spherulites can be more clearly observed. The spherical symmetry does not extend to the center, and the sheaf-like features are preserved.

2.4.2 Molecular Order

Chain mobility influences lamellar structures. The lamellae observed in the solvent cast films

form in dilute solution, in which chains have high mobility to rearrange into an ordered state and

to travel to lamella growth fronts to crystallize. This leads to more uniform lamellae and more

regular fold surfaces, as shown in Fig. 2.3(a). In spin coated films, the lamellae are developed

from the melt during the annealing process, and chains entangle with each other experiencing

less mobility. Crystallization in the melt thus does not follow the same route as in the dilute

solution, and proceeds in a pathway involving a transient mesomorphic structure (Strobl, 2000).

46

Crystal growth direction withgrowth planes

Chain folding direction <110 >

a

b

c

Fold surface ismore regular Chain folding

Protruding cilium

Lamella

To spherulitecenter

(a)

Not well-aligned nodules

Chain foldingdirection <110 >

a

bc

a

bc

To spherulitecenter

Well-aligned nodules

Fold surface isless regular

Nodule

(b)

Figure 2.3: (a) Sketch of the lamellar structure in solvent cast films. (b) Sketch of the lamellar structure in spin coated films.

47

In the beginning, chains close to the lamellae growth front form a mesomorphic structure while

they are crystallizing. The crystallizing chains in the mesomorphic structure are ordered while

still include conformational defects. The mesomorphic structure thickens with time and with

the distance from the lamella growth front. Chain rearrangements occur during the thickening

process, and higher ordered crystals are developed eventually. The resulting structure is

composed of nanometer-sized nodules in a planar assembly. The nodules coalesce to develop

lamellae (Su, et al., 2009), while fringed micellar crystallization can also occur simultaneously

with the folding structure, which roughens the fold surface (Wunderlich, 1976). The sketch of

the lamellar structures in the annealed spin coated film is given in Fig. 2.3(b).

The WAXD profiles of solvent cast and spin coated films are given in Figs. 2.4(a) and 2.4(b).

For solvent cast films, the profiles show three stronger peaks, (010), (110)/(200), and (203),

while only (110)/(200) and (203) peaks exist for the spin coated films. The (110)/(200) peak

remains prominent in all crystalline samples, including the non-annealed solvent cast film with

lower degree of crystallinity. (The degree of crystallinity will be discussed below; see Fig. 2.5.)

This suggests the crystal structure in the (110) and/or (200) direction keeps ordered throughout

the crystallization process. Since PLLA crystallization begins with chain folding in the <110>

direction, forming a crystal surface for the subsequent chains to attach (Miyata & Masuko, 1997),

the (110)/(200) peak may be mainly due to the order of chain folding in the <110> direction.

The (110)/(200) peak gradually increases with the annealing temperature for solvent cast films,

while for spin coated films the peak remains prominent even at lower annealing temperature.

This is because the degree of crystallinity of spin coated films is always higher than that of

solvent cast films under the same annealing conditions. It is also noted that the (203) peak

becomes less prominent and the (010) peak even disappears for the spin coated films. The

48

observed intensity is determined by the polarization factor , Lorentz factor , multiplicity

factor , absorption factor , and structure factor , or · · · · | | (Alexander,

1969). Since we consider the intensity difference of a single peak in the same material, the first

four factors are taken to be the same. The structure factor is thus of our interest, and is

given as

2 5

in which f is the atomic scattering factor accounting for the amplitude of the wave scattered by

an atom in a given direction, h, k, and l are the Miller indices of crystallographic planes, and x, y,

and z are the atomic positions in a unit cell (Alexander, 1969). This equation suggests that the

intensity is the summation of the wave amplitude scattered by the N atoms, and that the atoms

need to be located at certain positions for diffraction to occur. The number of atoms is

proportional to the area of the reflection plane. This area can be approximated by knowing the

sizes of the lateral dimensions of lamellae (on the order of μm), the lamellar thickness and the

dimensions of the nodules (on the order of nm), and the separation between atomic planes (on

the order of ). Since the degree of crystallinity is higher in spin coated films, the total area is

expected to be larger. This calculation therefore cannot explain the decrease of the (203) peak

and disappearance of the (010) peak in spin coated films. The cause is thus given by the atomic

positions, or chain packing, which affects the structure factor and thus the observed intensity.

Accordingly, in spin coated films, although chains fold orderly, generating strong (110) peak,

their packing is less orderly, diminishing the two peaks related to chain packing, (010) and (203).

If the spin coated films are annealed longer (8 and 24 hours), the (010) peak begins to emerge, as

given in Fig. 2.4(b) for 24-hour annealing. This shows that longer annealing time favors better

49

packing of the folded chains and more ordered reflection planes.

10 15 20 25 30 35 400

2000

4000

6000

8000

10000

12000

As castAnnealed at 80oCAnnealed at 110oC

(110)/(200)

(010)

(203)

Inte

nsity

2 theta (degree)

Annealed at 140oC

(a)

10 15 20 25 30 35 400

4000

8000

12000

16000

20000

24000

(010)Annealed at 140oC for 24 hours

As coatedAnnealed at 80oC for 3 hoursAnnealed at 110oC for 3 hours

Inte

nsity

2 theta (degree)

(110)/(200)

(203)

Annealed at 140oC for 3 hours

(b)

Figure 2.4: (a) WAXD profiles for solvent cast films annealed at different temperatures for 3 hours. Dotted line under the as cast film is the fitted curve for the amorphous region. (b)

WAXD profiles for spin coated films annealed at different temperatures. The (010) peak start to emerge after 140°C annealing for 24 hours. Dotted line under the 80°C annealed film is the

fitted curve for the amorphous region.

50

0 20 40 60 80 100 120 1400

2

4

6

8

10

12

Inte

nsity

rat

io

Annealing temperature (oC)

Spin coated film Solvent cast film

Figure 2.5: Intensity ratio I(110)/(200)/I(203) of the solvent cast and spin coated films annealed for 3

hours. The error bar represents the standard deviation.

0 20 40 60 80 100 120 14016.6

16.7

16.8

16.9

17.0

17.1

Pea

k lo

catio

n (2

the

ta)

Annealing temperature (oC)

Solvent cast film Spin coated film

2 theta=16.82o [17]

Figure 2.6: Location of the (110)/(200) peak for the solvent cast and spin coated films annealed

for 3 hours. The error bar represents the standard deviation.

The intensity ratio of the two strongest peaks, (110)/(200) and (203), is calculated based on the

area under the peaks. The results for 3-hour annealed films are given in Fig. 2.5, while the

results for 1-hour annealing are similar. The ratio for the spin coated films is always higher

because of the smaller (203) peaks as discussed above. This ratio does not change with

51

annealing temperature, which suggests that similar crystal structures are formed at the three

annealing temperatures. For solvent cast films, the ratio gradually increases with annealing

temperature. This higher ratio could come from the newly formed crystallites, since the degree

of crystallinity increases with annealing temperature. This increase is because the newly

formed crystallites do not develop in solution, but in an environment similar to the spin coated

films. From the above discussion, the chains may not pack well and have a smaller (203) peak.

Figure 2.6 shows the location of the (110)/(200) peak for 3-hour annealed films. Similar trend

is observed for 1-hour annealing. Similar trend is observed for 1-hour annealing. The peak

location for the spin coated films is smaller than that for solvent cast films. From the Bragg’s

law, spin coated films have larger d-spacings and thus less ordered and compact crystal structure,

which agrees with the discussion above. This can originate from the effect of folds (Alexander,

1968) irregular fold surfaces prevent ordered chain packing, increasing the basal area (a and b)

of the unit cell, while c remains unchanged. As discussed, the lamellar fold surface in the spin

coated films is rougher, which leads to larger a and b. Based on the plane spacing equation for

orthorhombic cells

1

d2 =h2

a2 +k2

b2 +l2

c2 (6)

where d is the d-spacing value, larger a and b result in larger d, and therefore smaller 2θ values.

Higher annealing temperature gives energy to the molecules to achieve equilibrium state, and

also enables them to change their conformations to favor chain packing. This explains the

smaller d-spacing in the 140°C-annealed spin coated film. We do not observe the change of 2θ

values in solvent cast films with annealing temperature. This is because the dominant

spherulitic structure is formed during the same casting process, and the effect of annealing is less

52

significant. The 2θ values differ in solvent cast films and spin coated films, and the calculated

value based on the lattice parameters is 16.82° (Miyata & Masuko, 1997). The values vary

slightly due to different processing and crystallization conditions as discussed, which

demonstrates that the lattice parameters of polymers can vary even for the same material.

The WAXD profiles in Fig. 2.4 are composed of amorphous diffraction and crystal diffraction.

The dotted line, obtained from amorphous PLLA, is the amorphous diffraction. The integrated

intensity of the amorphous diffraction, Ia, is the area under the amorphous diffraction curve,

while the integrated intensity of crystal diffraction, Ic, is the peak area. Crystallinity is the ratio

of Ic to Ia+Ic and can be calculated accordingly (Campbell et al., 2000) with results given in Fig.

2.7. The degree of crystallinity increases with annealing temperature, as expected. Similar

results are obtained for the 1-hour annealed films, except that the degree of crystallinity of the

1-hour annealed spin coated films is slightly less than that of the 3-hour annealed spin coated

films by 3%. The degree of crystallinity for 1-hour and 3-hour annealed solvent cast films are

the same, and 1 hour is found to be long enough to complete the secondary crystallization in

solvent cast films. It is also noticed that the degree of crystallinity for the solvent cast films is

always lower than the spin coated films, which suggests part of amorphous molecules in the

solvent cast films cannot crystallize. This could be related to the structure in spherulites.

Optical microscopy reveals that the non-annealed solvent cast film is already filled with

spherulites, but its degree of crystallinity is only 26%; the rest 74% is composed of the

amorphous chains. Some of them are inherently non-crystallizable due to branching or atactic

configuration, while some are crystallizable. Part of the crystallizable chains may be trapped by

the existing lamellar fold surfaces, forming the RAP. The fraction of RAP in crystalline PLLA

can be as large as around 20-30% (Wang et al., 2006; Sarasua et al., 2008), which inhibits the

53

development of crystal structure between the lamellae. For spin coated films, no RAP exists in

the beginning of the crystallite development, and chains can get crystallized before being trapped

by the formed lamellae, leading to a higher degree of crystallinity.

0 20 40 60 80 100 120 1400

10

20

30

40

50

60

70 Spin coated film Solvent cast film

De

gre

e o

f cry

stal

linity

(%

)

Annealing temperature (oC) Figure 2.7: Degree of crystallinity of the solvent cast and spin coated films annealed for 3 hours.

Annealing temperature of 25°C means the as cast film. The error bar represents the standard deviation.

Figures 2.2 and 2.4(b) give the crystal information of the spin coated film annealed for a longer

time (24 hours), and its degree of crystallinity and peak location are given as a function of

annealing time, as given in Fig. 2.8 which shows that both parameters increase with annealing

time. The increase of degree of crystallinity is expected, while the increase of 2θ value

suggests that the crystal structure is more compact and ordered. This figure also shows that the

increasing trends tend to level off, which suggests that the crystal structure in the 24-hour

annealed film is close to a stable structure and less likely to have further change for even longer

annealing time.

54

0 5 10 15 20 25

48

52

56

60

64

68

De

gre

e o

f cry

stal

linity

(%

)

Annealing time (hour)

Degree of crystallinity

16.7

16.8

16.9

17.0

Peak location

Pea

k lo

catio

n (2

thet

a, d

egre

e)

Figure 2.8: Degree of crystallinity and peak location of the spin coated films annealed at 140 °C

for 1, 3, 8, and 24 hours.

2.4.3 Conformation of Molecules and Intermolecular Interaction

The conformation of molecules and intermolecular interaction are studied through FTIR. The

spectra of the solvent cast and spin coated films are given in Figs. 2.9(a) and 2.9(b), respectively.

The carbonyl (C=O) stretching region, CH3 and CH bending region, and the skeletal (C-O-C)

stretching region show prominent peaks. It is also observed that there is no absorption band at

921 cm-1 for the amorphous film (the as coated film), while this peak emerges as the material

crystallizes. This peak is attributed to the coupling of the CH3 rocking vibration with the C-C

backbone stretching, and is the characteristic of PLLA α crystals (Kister et al., 1998).

Along the molecular chain, there are three skeletal bonds: C-O (ester), O-Cα, and Cα-C, in which

the ester bond can be assumed trans (Meaurio et al., 2006). The O-Cα bond has two minimal

energies when its bond rotation angle φ is -160° and -73°, while the Cα-C bond has two minimal

energies when its bond rotation angle ψ is 160° and -48°. The rotation angles φ=-160° and

ψ=160° are assigned as trans (t), while φ=-48° and ψ=-73° are assigned as gauche (g).

55

Accordingly, there are four conformers leading to four energy minima: tt, tg, gt, and gg, which

correspond to 21, 51, 103, and 41 helices (Kang et al., 2001). The gt conformer has the lowest

energy and constitutes the crystal structure. Because of the four conformers, the C=O

stretching region can be split into four peaks due to intramolecular and intermolecular

interactions, and the four peaks are located at 1776 cm-1, 1767 cm-1, 1759 cm-1, and 1749 cm-1,

corresponding to the gg, tg, gt, and tt conformers, respectively (Meaurio et al., 2006). The

spectra of this region of solvent cast and spin coated films are given in Figs. 2.10(a), 2.10(b), and

2.11. For the solvent cast films, annealing makes more prominent band splitting. (The peak

resolving will be discussed below.) However, band splitting cannot be visually observed for

spin coated films.

1800 1600 1400 1200 1000 8000.00

0.02

0.04

0.06

0.08

0.10

0.12CH3 and CH

bending region

As cast

80oC, 1h110oC, 1h

140oC, 1h

80oC, 3h

110oC, 3h

Abs

orba

nce

Wavenumber (cm-1)

140oC, 3h

C=Ostretchingregion

C-O-C stretching region

(a)

56

1800 1600 1400 1200 1000 8000.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35A

bsor

banc

e

Wavenumber (cm-1)

140oC, 3h

110oC, 3h

80oC, 3h

140oC, 1h

110oC, 1h80oC, 1h

As coated

(b)

Figure 2.9: (a) FTIR spectra of the solvent cast films before annealing and after annealed in different conditions. (b) FTIR spectra of the spin coated films before annealing and after

annealed in different conditions.

1780 1770 1760 1750 1740 1730 17200.000

0.005

0.010

0.015

0.020

0.025

Abs

orb

an

ce

Wavenumber (cm-1)

gt tt

(a)

57

1780 1770 1760 1750 1740 1730 17200.000

0.005

0.010

0.015

0.020

0.025

ttgt

Abs

orba

nce

Wavenumber (cm-1) (b)

Figure 2.10: (a) Curve fitting for the gt and tt conformers in the carbonyl C=O stretching region of the as cast solvent cast film. (b) Curve fitting for the gt and tt conformers in the carbonyl

C=O stretching region of the solvent cast film annealed at 140°C for 3 hours.

1780 1770 1760 1750 1740 17300.00

0.02

0.04

0.06

0.08

0.10

0.12

Abs

orba

nce

Wavenumber (cm-1)

Annealed at 140oC

Annealed at 110oC

Annealed at 80oC As coated

Figure 2.11: Carbonyl C=O stretching region of the spin coated films annealed at different

temperatures for 3 hours.

To investigate the components in the spectra, their second derivatives are illustrated in Fig. 2.12.

For solvent cast films, two prominent peaks can be seen at 1757 cm-1 and 1747 cm-1, while two

inconspicuous peaks are at 1777 cm-1 and 1766 cm-1. These four peaks represent the gt, tt, gg,

and tg conformers, respectively. The most prominent two peaks, gt and tt are resolved, as given

in Figs. 2.10(a) and 2.10(b). The intensity ratio of gt/tt increases as the film is annealed, which

58

confirms that the crystal structure is preferred by the gt conformer. The peak width (defined as

the full width at half maximum) of the resolved peaks is shown in Fig. 2.13. It is observed that,

as the films are annealed, the width of the gt peaks decreases, while that of the tt peaks almost

does not alter. The peak width in FTIR spectrum is related to the homogeneity in the structure:

more chemical environments lead to more complicated intermolecular interactions and broadens

the peaks, while fewer chemical environments narrow the peaks (Smith, 1999). Since the gt

peak is responsible for crystal structure, the narrowing gt peaks is because the crystal structure

becomes more ordered as the structure gets crystallized, which leads to a more uniform and

homogeneous chemical environment. The band width of the tt conformer does not change

significantly, because the tt conformers do not contribute to crystallization, and remain

amorphous after annealing.

1780 1775 1770 1765 1760 1755 1750 1745 1740

-0.0020

-0.0015

-0.0010

-0.0005

0.0000

0.0005

17421753

1746, tt1759, gt

1766, tg1777, gg

1747, tt

Spin coated filmsS

econ

d de

rivat

ive

for

spin

coa

ted

film

s

Sec

ond

deriv

ativ

e fo

r so

lven

t cas

t fil

ms

Wavenumber (cm-1)

As cast/as coated

Annealed at 80oC

Annealed at 110oC

Annealed at 140oC

Solvent cast films

1757, gt

17541742

-0.0010

-0.0005

0.0000

0.0005

0.0010

0.0015

Figure 2.12: Second derivative spectra of the C=O stretching region for solvent cast and spin

coated films before annealing and annealed at different temperatures for 3 hours.

It is observed that there are two shoulders appearing at 1754 cm-1 and 1742 cm-1 in the derivative

spectrum for the non-annealed solvent cast film (Fig. 2.12). The two shoulders cannot be

59

attributed to any of the four conformers, and represent two peaks splitting from the gt and tt

peaks. They both split toward the lower wavenumbers. Stress in the film is therefore a

possible cause, in that it increases the length of a covalent bond, which decreases the bond’s

atomic force constant, and thus decreases the infrared absorption frequency of bonds for their

stretching vibrations. Band shift has been observed to occur at stresses on the order of MPa

(Roylance & DeVries, 1971). For solvent cast films, the stress can be generated during the

casting process: as the solvent evaporates, the volume shrinks but the area is constrained by

adhesion to the substrate. This constraint creates an internal stress in the films, which is also on

the order of several MPa (Croll, 1979). This solidification process can thus account for the

stress which stretches the bonds and thus the observed shoulders in the spectra. After annealing,

the 1754 cm-1 shoulder disappears, and the one at 1742 cm-1 also becomes less prominent,

because annealing eliminates the internal stress.

0 20 40 60 80 100 120 14020.5

21.0

21.5

22.0

22.5

23.0

23.5

Pea

k w

idth

(F

WH

M, c

m-1)

Annealting temperature (oC)

gt conformer, annealed for 1h gt conformer, annealed for 3h tt conformer, annealed for 1h tt conformer, annealed for 3h

Figure 2.13: Peak width of the resolved spectra for the gt and tt conformers of solvent cast films

before annealing and after annealed in different conditions. The error bar represents the standard deviation of the five measurements.

60

For the spin coated films, the two prominent peaks are gt and tt, located at 1759 cm-1 and 1746

cm-1, respectively. There are two additional peaks shown at 1753 cm-1 and 1742 cm-1, which

are also located on the right to the gt and tt peaks, respectively; therefore, internal stress could be

the reason of the splitting. In addition to solidification, the centrifugal force during the spinning

process is another source of the internal stress (Vorotilov et al., 1995). However, this stress is

only on the order of only several Pa (Chang et al., 2003), much less than the stress required to

split the band and is less likely to be the reason. The stress induced by solidification is a more

reasonable cause. As the film is annealed, the peaks tend to become less prominent, suggesting

the decrease of stress. It can also be seen that, in the derivative spectrum for the as coated film,

the tt peak is much more prominent than the gt peak, which is consistent with the fact that this

film is amorphous, and may also explain why the band is located at a lower wavenumber in the

original spectrum, as seen in Fig. 2.11.

2.5 Conclusions

It has been shown that both film formation method and annealing have significant influences on

morphology and crystal structure of PLLA films. In solvent cast films, category 1 spherulites

are developed in solution during the casting process, and further annealing increases the degree

of crystallinity without affecting the morphology. Spin coating leads to amorphous structure,

and subsequent annealing generates category 2 spherulites. Molecular order in the two kinds of

films is also different as investigated via WAXD measurements. The crystal structure is more

compact in solvent cast films than in spin coated films, while the lamellae in solvent cast films

formed during casting may be unfavorable to further crystallization. The FTIR measurements

demonstrate the conformation of molecules and intermolecular interaction in the crystallites, and

also show possible stress in the films. Longer annealing time does not change the spherulitic

61

structure, which confirms that the crystal structure is mainly determined by film formation

method and the mobility of polymer chains at which annealing is conducted.

62

Chapter 3: Effect of Excimer Laser Irradiation on Crystallinity and Chemical Bonding of

Biodegradable Polymer

3.1 Introduction

Biodegradable polymers have been attracting wide attention due to their biocompatibility and

biodegradability. Being biodegradable, poly(lactic acid) (PLA) is used in food packaging and

tissue engineering. Its chemical structure also makes it favorable in drug delivery as it

hydrolyzes in the human body into lactic acid, which is then excreted showing no toxicity. In

drug delivery, drugs are embedded in a polymer matrix and released as it degrades. The

advantages of using biodegradable polymers in drug delivery system can be manifested by the

controlled drug release profiles against time (Amass et al., 1998). However, due to the

induction period of polymer degradation caused by the time for water molecules to penetrate into

the matrix, embedded drugs cannot release at the designed rate in the early stage.

Efforts have been made to modify the degradation rate of PLA. Its degradation rate has been

modified by adjusting temperature and pH value in the degradation environment (Zhao et al.,

2004). This attempt is limited since both parameters are invariants in the human body.

Blending PLA with another biodegradable polymer, poly(glycolic acid), to adjust the

degradation rate is also investigated (Miller et al., 1977), while the desired properties of PLA

may be altered in the copolymer. Tsuji (Tsuji, 2002) investigate the effects of tacticity, content

of L-lactide unit, and blending of poly(L-lactic acid) (PLLA) and poly(D-lactic acid) (PDLA) on

the degradation rate of PLA. Although the PDLLA film can have the highest degradation rate,

it requires specific chemical reaction environments to synthesize PDLLA with a desired ratio of

PLLA and PDLA.

63

A way to modify PLA degradation rate free from the above limitations is to change its

crystallinity (Nampoothiri et al., 2010). It is proposed that PLA degradation begins in the

amorphous regions between lamellae, followed by the disorientation of the lamellae and

disappearance of the spherulitic structure (Tsuji & Ikada, 1998). Less crystalline regions are

expected to degrade faster. Adjusting crystallinity is a relatively simple process, and can be

accomplished through thermal treatments such as annealing, melting and quenching (Tsuji &

Ikada, 1995). However, they affect the overall properties within the bulk and have limitation

since accelerating the initial degradation may only require decreasing surface crystallinity while

keeping the bulk intact. Recently, laser irradiation as a heating source to modify the surface

crystallinity has drawn increasing attention. It is also attractive due to flexibility, ease of use,

and spatial controllability.

Using the Nd:YAG laser with the photon energy lower than PLLA bond energies, Bhatla and

Yao (Bhatla & Yao, 2009) reduced the PLLA crystallinity, and quantified it by the wide angle

X-ray diffraction (WAXD) measurements. However, the non-uniform laser spatial intensity

profile makes it difficult to achieve a uniform surface treatment. Because of the incoherent

light generated by the excimer laser, the beam can be homogenized without interference, and the

homogenized beam is spatially uniform and favorable for surface treatment. Using the excimer

laser, Lazare and Benet (Lazare & Benet, 1993) treat poly(ethylene terephthalate) (PET) films,

and demonstrate the presence of a thin amorphous surface layer. Lazare and Benet report that

photochemical reactions are possible in addition to photothermal effects, since excimer lasers

operate at UV wavelengths with photon energies higher than certain bond energies in polymers.

In another attempt to decrease PET film crystallinity using the excimer laser (Dunn & Ouderkirk,

1990), the chemical changes are studied via X-ray photoelectron spectroscopy (XPS). It is

64

found that low fluences cause little chemical changes. These chemical changes include the

dissociation of ester bonds and reduce of the number of oxygen atoms due to the emission of

small molecules (Watenabe et al., 1993); these modifications could alter the original properties

of the polymers.

The primary objective of this work is to reduce crystallinity of PLLA films with excimer laser

surface treatment, while minimizing the chemical modifications. The optical microscopy,

WAXD, and XPS measurements are utilized to study the morphology, crystallinity, and chemical

modifications, respectively. A 2D model considering the thermal and chemical effects is

proposed to numerically examine the spatial and temporal temperature profiles and the amount

of chemical changes. The effects of radical mobility in the PLLA films and laser fluence have

also been considered.

3.2 Background

3.2.1 Melting and Crystallization

Polymer melting is an amorphization process in which polymer chains detach from crystals, and

the crystal dimension decreases longitudinally and transversely (Belyayev, 1988). As a rapid

process, chain detachment and crystallinity decrease can occur within nanoseconds (Yamamoto,

2010); melting polymers by a nanosecond laser is thus feasible. The melting temperature

of PLLA is between 170°C to 180°C. Due to the difference in the fold surface energy and

lateral surface energy of crystals, thicker crystals are more thermodynamically stabile and thus

have higher (Cormia et al., 1962; Hoffman & Week, 1962):

12∆

1

65

where is the equilibrium melting temperature, ∆ is the fusion heat, and is the crystal

thickness. Given sufficient chain mobility, lamellar crystals thicken to reduce surface energy.

Annealing enhances chain mobility, and can increase and by 5-10°C for PLLA (Tsuji &

Ikada, 1995; Mitomo, 1988). Temperatures higher than 200°C can cause thermal degradation

over minutes.

Crystallization begins with the nucleation process. Nucleation of PLLA occurs over minutes.

PLLA crystal growth is given by the non-isothermal crystallization kinetics

1 2

where is the crystallinity, is crystallization time, is the reaction rate constant,

and is the Avrami exponent. For crystal development is also on the order of minutes. A

cooling rate higher than 50 °C/min prevents molten PLLA from recrystallization (Di Lorenzo,

2005). Since laser treatment can induce an extremely high cooling rate, recrystallization during

the cooling process of laser treatment does not occur (Bhatla & Yao, 2009).

3.2.2 Interaction between Photons and Polymers

Polymer absorption of photons with energies exceeding the bond energy can break the bonds.

A series of events is given in Fig. 3.1 (Lazare & Granier, 1989). The photons excite electrons

from the ground state Sg to the exited state Se, Event 1. Part of the energy is relaxed, Event 2;

part of it breaks the bond, forming free radicals, Event 3. Some radicals generate decomposed

products via Event 4, while some recombine, Event 5. Radical reactions release heat. In

situations where oxidation does not occur, UV irradiation on PLA breaks the CO bonds in its

backbone, separate the CH3CHCOO units, and generates free radicals R1 and R2 following Eq.

66

(3) (Janorkar et al., 2007)

C C O C C O

H HO O

CH3 CH3

C C O

H O

CH3

C C O

H O

CH3 (R2)(R1)

This reaction occurs between two CH3CHCOO units. As the reaction continues, the number of

scissions increases, which decreases the molecular weight and number of connected

CH3CHCOO units. R1 and R2 can generate smaller molecules such as CO and CO2 via further

reactions (Inagaki et al., 2002). The reactions last for microseconds after the nanosecond UV

laser pulse (Schneider et al., 1998). After laser treatment, bubbles can be observed on the

polymer surface (Inagaki et al., 2002; Rebollar et al., 2006; Efthimiopoulos et al., 2008).

Free radicals Decomposedproducts

12

Sg

Se

3

4

5

Figure 3.1: Elementary events following the polymer absorption of photons with energies higher

than the polymeric bond energy.

3.2.3 Dissociation Quantum Yield

Competition of Events 4 and 5 in Fig. 3.1 determines the dissociation quantum yield , which is

the number of dissociations occurring per photon absorbed, defined by

⁄ 1 4

where is the sample weight, is the initial molecular weight, is the molecular weight

3

67

at time , is the number of absorbed photons per unit time. The number of scissions per

chain, ⁄ 1, increases with radical mobility based on the cage effect. The quantum

yield in the mobile state is generally larger by an order under UV irradiation (Dan & Guillet,

1973), and is proportional to the dissociation reaction rate constant through (Aleksandrov et

al., 1988)

5

where is the photon absorption cross section with photon frequency .

3.3 Numerical Model

To numerically investigate the spatial and temporal distribution of temperature and the amount

of chemical modifications, a 2D symmetric thermal and chemical model is developed. Since

PLLA is semitransparent at UV, laser irradiation energy is absorbed by the bulk, and the bulk

temperature is governed by the heat equation

· , 6

where is density, is specific heat, is conductivity, is laser irradiation depth,

, is the laser power density, and is the release heat of radical reactions and will be

determined in Eq. (12). For an incompressible material, change in during the phase

transition can be approximated by ∆ ∆ ⁄ , where ∆ is the heat of fusion. The

specific heat is calculated as

∆ , ∆ 7

where is the specific heat in the solid state and is a step function with value 0 for

∆ and 1 for ∆ . The specific heat of polymeric materials is typically

68

around 1000 J/(kg·K) (Boudenne et al., 2011); thus is assumed 1000 J/(kg·K). ∆ , 10°C,

is half the transition temperature span. Laser power density, , , is a function of space and

time

, · · exp · 4 22

8

where is the absorptivity, 0.07, is the absorption coefficient, 103 cm-1 (Ohki et al., 2007),

is the pulse width, 25 ns, and is the peak power density obtained by ⁄ with

the laser fluence. Convection and radiation are considered on the boundaries while insulation is

specified along the symmetry line. The initial temperature is set 20°C. Change of crystal size

due to annealing is not of interest, while its effect on is considered by assuming different

’s for the non-annealed films (175°C) and annealed films (180°C). From the backgrounds of

crystallization and cooling rate in Sec. 3.2.1, it is assumed that no crystallization occurs during

cooling; the assumption will be validated by our simulation results in Sec. 3.5.2. Crystallinity

decreases as the material melts; therefore, the crystallinity change is related to the melting depth

in the film.

The chemical modifications occur via Eq. (3) if R1 and R2 do not recombine. The amount of

modifications depends on the number of scissions, Eq. (4). As scission occurs, the connections

of the CH3CHCOO units at their ester group with another unit are broken, and the number of

such connected CH3CHCOO units decreases. Therefore, the number of the CH3CHCOO units

with connection at their ester group to another unit in a unit volume, or its number density, is

modeled. For incompressible materials, the number density is governed by the mass

conservation equation

69

· · 9

where is the number density with an initial value of , is the velocity vector, is the

diffusion coefficient, and is the reaction rate. Velocity and diffusion gradients are assumed

to be negligible during laser treatment, and thus ⁄ . The reaction rate is given by the

crystallinity ,

· 1 · 10

where and are the reaction rates of the crystalline and amorphous domains, respectively.

The photochemical reaction rate is proportional to the number density (Bityurin et al., 2003).

Thus, · and · where and are the reaction rate constants of the

crystalline and amorphous domains, respectively. Temperature dependence of reaction rate is

expressed by the Arrhenius expressions for the reaction rate constant (Astrath et al., 2009); the

proportionality between the reaction rate constant and quantum yield (Aleksandrov et al., 1988)

is also considered.

· · 11

where or , and is the pre-exponential factor of the Arrhenius expression, assumed

107 1/s, is the activation energy, assumed 25 kJ/mol, and is the universal gas constant.

and are the step functions accounting for the quantum yields in solid and mobile states.

To consider the one-order-of-magnitude difference of the quantum yields in the mobile and solid

states (Dan & Guillet, 1973), is assumed 1 if and 0.1 if and is

assumed 1 if and 0.1 if . Temperature is imported from the thermal analysis.

It is assumed that the time period for the photochemical process is 20 and that no

flux across the boundaries. The release heat of radical reactions is considered as

70

2 12

where is the number density of the free radicals to be recombined or react, which is 2 if

all molecules in the ground state can be excited to generate free radicals (Event 1 in Fig. 3.1).

Also in Eq. (12), is the reaction heat by each event, around 600 J/mol (Currie et al., 1969),

and is the step function whose value is 1 if and 0 if to consider the

time period of the photochemical process. The reaction heat is then output to the thermal

analysis. Part of the crystalline chain becomes amorphous during melting, while its motion

could be constrained by surrounding crystals. Thus, at the same , the initial crystalline

domain is assumed to have the same mobility and reaction rate before and after melting. Since

the amount of chemical modifications on the polymer chains induced via Eq. (3) is of interest

and the generated free radicals only react with other radicals instead of the polymer chains

(Inagaki et al., 2002), the subsequent chemical reactions and products are not considered in the

simulation.

The coupled partial differential equations 6 and 9 with boundary and initial conditions can

be solved through the finite element method. The software COMSOL Multiphysics 4.1 is used.

A 2D symmetric model is developed and composed of thermal and chemical analyses. Its

domain is 0.7 mm × 20 µm with laser beam size 0.5 mm. The thermal analysis is realized by

the Heat Transfer in Solids Interface of the Heat Transfer Module. The chemical analysis is

accomplished by the Transport of Diluted Species Interface of the Chemical Reaction

Engineering Module, which is able to compute the number density of a single species under

varying reaction rates. The COMSOL smoothed Heaviside function, flc2hs, is used as the step

functions.

71

3.4 Materials and Methods

PLLA granules were provided by PURAC and used as received. The inherent viscosity of

PLLA in chloroform at 25°C is 1.6 dl/g, and the molecular weight is thus estimated as 5.6×104

g/mol (Schindler & Harper, 1979). Films were prepared by the solvent casting method

described in reference (Hsu & Yao, 2011). The film thickness is 20-25 μm. Dissociation

quantum yield changes with radical mobility, which can be reduced by higher crystallinity.

Annealing increases crystallinity, and can obtain films with lower radical mobility when radicals

are formed due to laser irradiation. Non-annealed films are termed films with high radical

mobility. Annealing is conducted at 140°C for 1 hour. A KrF excimer laser with wavelength

248 nm and pulse width 25 ns is used. The homogenized laser beam has spatially uniform

intensity favorable for a uniform surface treatment. The film is radiated by a single pulse with

fluence 1.80 J/cm2 to 2.80 J/cm2 in argon atmosphere with the flow rate of 15 standard cubic feet

per hour to prevent oxidation. Morphology of PLLA films is observed via optical microscopy.

WAXD measurement is accomplished using the Inel X-ray diffractometer. Monochromatic

CuKα radiation with wavelength 0.15418 nm at 40 kV and 30 mA is used. X-ray photoelectron

spectroscopy (PHI 5500 ESCA) is used to analyze chemical constituents of pre- and post-treated

films. The O1s and C1s spectra are captured, and the take-off angle is 45°.

3.5 Results and Discussion

3.5.1 Effects of Laser Irradiation on Morphology

The morphologies of the laser treated films are given in Figs. 3.2(a) to 3.2(d). High fluence

irradiation generates appreciable bubbles as a result of laser energy penetration into the bulk due

to the low absorption coefficient (103 cm-1) of PLLA at UV wavelength. The same phenomenon

72

has been reported in poly(methyl methacrylate) (PMMA) under the excimer laser irradiation

(Rebollar et al., 2006; Efthimiopoulos et al., 2008). Its absorption coefficient is on the same

order as PLLA, and the fluences of several J/cm2 generate bubbles in both materials.

(a) (b)

(c) (d) Figure 3.2: Morphologies of the film with high radical mobility treated with fluence of (a) 2.50

J/cm2 and (b) 2.60 J/cm2, and film with low radical mobility treated with fluence of (c) 2.60 J/cm2 and (d) 2.70 J/cm2.

The bubble formation in laser treated polymers has been explained by different theories.

Srinivasan (Srinivasan, 1993) suggests that the bubbles are the gas trapped by the melted and

re-solidified material, while others believe that the bubbles are the gaseous products from the

decomposition of original materials (Rebollar et al., 2006). The origin of bubble formation can

73

be revealed by comparing the morphology shown in Fig. 3.2 and the XPS results shown in Figs.

3.8 to 3.10. (The XPS measurements and chemical modifications will be discussed in more

detail in Sec. 3.4.3.) Figure 3.2 shows that appreciable bubbles are generated in the film with

high radical mobility if the fluence achieves 2.60 J/cm2 (Fig. 3.2(b)). However, in the film with

low radical mobility, the appreciable bubbles form only when the fluence reaches 2.70 J/cm2 (Fig.

3.2(d)). The results agree with the XPS measurements, which show that appreciable chemical

modifications occur at 2.60 J/cm2 and 2.65 J/cm2 for the film with high and low radical mobility,

respectively. Thus, bubble formation is believed to be a chemical process involving the

decomposition of the original material. The components in the bubbles include small molecules

such as CO and CO2 (Inagaki et al., 2002).

Laser fluences used in this study are on the order of J/cm2, while the fluences treating PET are

only several mJ/cm2 (Lazare & Benet, 1993; Dunn & Ouderkirk, 1990). This difference comes

from different absorption coefficients of PET (105 cm-1) and PLLA at UV wavelength (Ohki et al.,

2007), which determines the laser penetration depth. The penetration depth, defined as the

depth at which the laser intensity in the bulk falls to 1/e of the original intensity, is the reciprocal

of the absorption coefficient based on the Beer-Lambert Law. The penetration depth is on the

order of 0.1 μm for PET and 10 μm for PLLA. Therefore, PET concentrates laser energy on the

surface, reducing fluences required to induce surface modification, while PLLA allows larger

energy penetration and higher fluences are needed to modify the material.

3.5.2 Decrease of Crystallinity by Laser Irradiation

Crystallinity is investigated via WAXD, and the results for the films with high radical mobility

are given in Fig. 3.3. The peaks become less prominent for the films treated with higher

74

fluences, suggesting a lower crystallinity. Similar results are obtained for the films with low

radical mobility, while more prominent peaks are observed due to its higher crystallinity caused

by annealing. The crystallinity is calculated based on reference (Alexander, 1969) with results

given in Fig. 3.4. Figure 3.4 quantitatively shows that annealing results in a higher crystallinity;

radical mobility may thus be reduced. It is also shown in Fig. 3.4 that crystallinity decreases at

a certain fluence, which suggests a thermal process involving polymer melting.

10 15 20 25 30 35

Treated with F = 2.40 J/cm2

Treated with F = 2.60 J/cm2

Treated with F = 2.00 J/cm2

Treated with F = 2.20 J/cm2

(010)

Inte

nsity

2 theta (degree)

(100)/(200)

(203)

Not laser treated

Figure 3.3: WAXD profiles of crystalline PLLA films with high radical mobility before and after

laser treatment. Note that the peaks decrease at higher fluences (2.40 and 2.60 J/cm2). The same trend is observed for the films with low radical mobility.

From Fig. 3.4, the crystallinities of the two kinds of films both decrease at 2.30 J/cm2, suggesting

that the increase of due to annealing does not affect the onset of crystallinity decrease.

This is because the temperature increment caused by different fluences is larger than the

difference caused by annealing. However, the difference causes different amount of

crystallinity change. For the film with high radical mobility (the non-annealed film),

crystallinity decreases from 31 % to 14 % after treated with a 2.70 J/cm2 pulse, which is 55 %

decrease. For the film with low radical mobility (the annealed film), the crystallinity decreases

75

from 48 % to 34 % under the same condition, which accounts for 29 % decrease. Namely, less

amount of material is melted in the annealed film. Since WAXD measures the crystallinity in

the bulk due to the high penetration depth of X-ray, a possible scenario is that the lower in

the non-annealed film facilitates melting, and thus the amount of crystallinity decrease is larger.

0.00 0.20 1.80 2.00 2.20 2.40 2.60 2.80

10

20

30

40

50

Fluence (J/cm2)

Deg

ree

of C

ryst

allin

ity (

%)

Film with low radical mobility Film with high radical mobility

Figure 3.4: Crystallinity of PLLA films derived from the WAXD results as a function of laser fluence.

Figure 3.5: Simulation result of the typical temperature distribution in the film. This

simulation depicts the film with high radical mobility 1 μs after the onset of a 2.40 J/cm2 laser pulse.

76

To numerically investigate the temperature distribution and melting depth in the film, a 2D

thermal model is developed. A typical temperature distribution is given in Fig. 3.5. The

highest temperature occurs on the surface due to the highest local laser energy density, and the

temperature decreases with depth. Since the excimer pulse intensity is spatially uniform, the

temperature distribution is not a function of the distance to the beam center.

The melting depth is also determined from the simulation results, as given in Fig. 3.6. For both

kinds of films, melting begins once the fluence reaches a certain level (2.10 J/cm2 in the

simulation), and melting depth increases with fluence thereafter. The simulation agrees with

the WAXD measurements in that small difference in of both films does not affect the onset

of melting. Simulation results confirm that crystallinity decreases from the surface, and that

different melting depths are obtained with different ’s. The non-annealed film has larger

melting depth because of its lower , which in turn allows crystallinity to decrease at a larger

extent, as observed in the WAXD. Therefore, the amount of crystallinity decrease is a bulk

phenomenon and can be explained by the melting depth in the film.

1.80 2.00 2.20 2.40 2.60 2.804

3

2

1

0

Mel

ting

dept

h (u

m)

Fluence (J/cm2)

Melting depth (low radical moblity) Melting depth (high radical moblity)

1.80 2.00 2.20 2.40 2.60 2.80

65

70

75

80

85

90

95

100

Nor

mal

ized

num

ber

dens

ity (

%)

Number denstiy (low radical moblity) Number denstiy (high radical moblity)

Figure 3.6: Simulation results of the melting depth in the film and the normalized number

density of the CH3CHCOO units connecting at their ester group with other CH3CHCOO units.

77

To determine the cooling rate after the laser pulse, the temperature time history is obtained from

the model. A typical result is given in Fig. 3.7, which describes the temperature time history on

the surface of the film with low radical mobility treated with 1.80 and 2.80 J/cm2. It is

observed that the temperature drops below at a cooling rate of 105 °C/min. At temperatures

lower than , the structure becomes solid and no crystallization can occur. Since a cooling

rate higher than 50 °C/min prevents PLLA recrystallization (Di Lorenzo, 2005), recrystallization

does not occur in our case, as agrees with reference (Bhatla & Yao, 2009) and our assumption.

The temperature can exceed the thermal degradation temperature (~200°C) due to laser

irradiation. However, thermal degradation requires minutes to occur (Wachsen et al., 1997) and

therefore is negligible during this rapid cooling process. To emphasize the cooling rate induced

by laser irradiation, forced convection due to argon flow as experimentally conducted is not

considered in simulation. The forced convection leads to a higher cooling rate as compared to

the simulation results.

0.00 0.02 0.04 0.06 0.08 0.100

50

100

150

200

250

Glass transition

temperature=60oCTem

pera

ture

(o C

)

Time (sec)

Fluence=2.80 J/cm2

Fluence=1.80 J/cm2

Melting temperature=180oC

Figure 3.7: Simulation result of the typical film temperature profiles as a function of time after

laser irradiation. The simulation predicts the temperature profile on the surface of the film with low radical mobility treated a single pulse with fluence of 1.80 and 2.80 J/cm2.

78

3.5.3 Chemical Modifications

3.5.3.1 Modifications of Chemical Bonds

XPS measurements are performed to investigate the chemical modifications on the film surface.

O1s and C1s spectra are recorded to determine the amount of O and C elements and possible bond

breakages. C1s spectra of the film with high radical mobility before and after laser treatment are

given in Figs. 3.8(a) to 3.8(c). The profiles are composed of three peaks, which come from the

three carbon groups in the structure: the C-H groups, the C-O groups in the backbone, and the

O-C=O groups in the ester bonds. They have the binding energies of 285, 287, and 289 eV, and

contribute to Peak 1, Peak 2, and Peak 3, respectively (Beamson & Briggs, 1992). The resolved

peaks are also given in Fig. 3.8. It is observed that the profiles remain similar until a high

fluence is utilized.

Areas under the resolved peaks in Fig. 3.8 are calculated. The calculation results are shown in

Figs. 3.9(a) and 3.9(b) for both kinds of films as a function of laser fluence. Peak areas change

only when the laser fluence reaches certain levels, which are 2.60 J/cm2 and 2.65 J/cm2 for the

film with high and low radical mobility, respectively. The excimer laser used in this study

generates photons with 248-nm wavelength, whose energy is 5.0 eV. Chemical bond energies

in PLLA are 3.6 eV for the C-C bonds, 3.8 eV for the C-O bonds, 7.5 eV for the C=O bonds, and

4.3 eV for the C-H bonds. Since the photochemical reactions occur whenever the energy of a

single photon exceeds the dissociation energy, chemical bond breakage always occurs. The

threshold behavior shown in the XPS results thus suggests that low-fluence laser treatment

induces an insignificant amount of chemical modifications not reflected in the measurements

(Dunn & Ouderkirk, 1990).

79

292 290 288 286 284 282

Peak 3 Peak 2

Cou

nts

/ se

c

Binding energy (eV)

Original data Resultant fitted curve

Peak 1

(a)

292 290 288 286 284 282

Cou

nts

/ se

c

Binding energy (eV)

Original data Resultant fitted curve

Peak 2Peak 3

Peak 1

(b)

292 290 288 286 284 282

Cou

nts

/ se

c

Binding energy (eV)

Original data Resultant fitted curve

Peak 2

Peak 3

Peak 1

(c)

Figure 3.8: Typical C1s XPS spectra and the resolved peaks of the film before and after laser treatment. The figures show the results of the film with high radical mobility (a) before laser treatment, and treated with fluences of (b) 2.50 J/cm2 and (c) 2.60 J/cm2. Note that the peak area ratio does not agree with the theoretical value, 1:1:1; a possible reason is given in Sec.

3.5.3.1.

C C O

H O

CH31

2 3

C C O

H O

CH31

2 3

C C O

H O

CH31

2 3

80

0.00 0.20 2.40 2.60 2.8010

20

30

40

50

60

70

Pea

k ar

ea (

%)

Fluence (J/cm2)

Peak 1 Peak 2 Peak 3

(a)

0.00 0.20 2.40 2.60 2.8010

20

30

40

50

60

70 Peak 1 Peak 2 Peak 3

Pea

k ar

ea (

%)

Fluence (J/cm2)

(b) Figure 3.9: Area percentage of the three resolved peaks in the C1s XPS spectra of the films with

(a) high radical mobility and (b) low radical mobility.

XPS results also reveal possible chemical reactions. Decrease of Peaks 2 and 3 in Fig. 3.9

suggests, respectively, the loss and/or breakage of the C-O groups in the backbone and the

O-C=O groups in the ester bonds. The O/C ratio decreases shown in Fig. 3.10 suggests the loss

81

of oxygen atoms. Since bubbles are composed of small gaseous products from chemical

decomposition as discussed in Sec. 3.5.1, a portion of the oxygen atoms is believed to emit from

the film in form of CO and CO2. A possible series of chemical reactions following the

generation of the free radicals R1 and R2 through Eq. (3) is given in Appendix (Inagaki et al.,

2002). The generation of R1 and R2 via Eq. (3) requires the breakage of the C-O bonds in the

backbone, which reduces the amount of the carbon atoms contributing to Peak 2. Due to the

instability of R1 and R2, two R1’s may react to produce R3 and CO2 following Eq. (A.1). The

evolution of CO2 decreases the amount of oxygen atoms as well as the carbon group accounting

for Peak 3. Two R3’s may react to generate R4 and eliminate small molecules CHO-CH3

through Eq. (A.2). R4 may decompose into R3 and CO through Eq. (A.3), forming a chain

reaction. Two R2’s may react to produce R5 and C2H4 through Eq. (A.4). R5 can further

decompose into R2 and CO2 through Eq. (A.5); this further decreases the amount of oxygen and

the area under Peak 2, and also initiates a chain reaction between R2 and R5.

O/C

Film with high radical mobility

0.00 0.20 2.40 2.60 2.80

0.3

0.4

0.5

0.6

Film with low radical mobility

Fluence (J/cm2)

Figure 3.10: Ratio of the O1s peak area to the C1s peak area derived from the XPS spectra for the films with high and low radical mobility.

82

It has been observed that modifications of crystallinity and chemical structures demonstrate

non-linearity with laser energy. To reduce crystallinity, polymer melting temperature has to be

achieved, and thus a certain level of laser energy is required. Chemical modification depends

on polymer chain state and temperature. At temperatures higher than the glass transition

temperature and melting temperature, amorphous chains and crystalline chains becomes mobile,

respectively. The mobile chains can be chemically modified with larger quantum yields based

on the cage effect as discussed in Sec. 3.5.3.2. Chemical reaction rate also exponentially

depends on temperature. Both factors contribute to the non-linearity of chemical modification

with respect to laser energy.

It is noticed that the area under Peak 1 is larger than those under Peaks 2 and 3, and the peak area

ratio does not agree with the theoretical value, 1:1:1. The theoretical value of the peak area

can be determined through the equation

13

where is the instrument constant, is the atomic concentration, is the photoionization

cross section for an element, is the inelastic mean-free path length for photoelectrons, is

the area of the sample, and is the analyzer transmission function. Sample area , being a

constant, is not expected to cause the deviation from the theoretical value. and depend

on chemical elements and do not cause the deviation since the element of interest is carbon.

Both and are device dependent constants and are not the cause either. Atomic

concentration can be the reason. Larger area of Peak 1 suggests a higher concentration of

the side group CH3 on the surface than the other two carbon groups in the backbone. Since

XPS is sensitive to a monolayer of atoms on the surface, the side group is believed to be more

abundant on the surface. This abundance could be due to the fact that the crystalline chains are

83

well-aligned in the lamellae, and their side groups uniformly protrude outward from the chains.

Such side groups may form the outmost atom layer on the lateral surface of the lamellae, and

thus the surface of spherulites and the film. Therefore, in the ordered structure the side group

may have higher concentration on the film surface. The explanation is supported by a separated

XPS measurement on an amorphous PLLA film prepared by spin coating with procedures

detailed in reference (Hsu & Yao, 2011). Results of both area ratio and atomic ratio O/C (not

shown) agree with the theoretical values, since the distribution of the three carbon groups is more

uniform in a less ordered structure.

3.5.3.2 Effect of Radical Mobility on the Amount of Chemical Modifications

The amount of chemical modifications can be revealed by the XPS results given in Figs. 3.9 and

10, which show that appreciable chemical modifications occur at a lower fluence (2.60 J/cm2) for

the film with high radical mobility and higher fluence (2.65 J/cm2) for the film with low radical

mobility. Peak areas and the O/C ratios also change more for the film with high radical

mobility. To numerically investigate the amount of chemical modifications, the number density

of the CH3CHCOO units connecting at their ester group with another unit is simulated and

represented in terms of normalized values based on the initial number density in Eq. (9).

The normalized number density is thus defined as ⁄ 100% and represents the percentage

of the initial value . The normalized number density of the film with high radical mobility is

given in Fig. 3.11, which shows that the number density decreases with time and fluence. The

same trends are obtained for the film with low radical mobility. The simulation is computed for

the film surface and thus comparable with the XPS results. The normalized number density is

also plotted as a function of fluence as given in Fig. 3.6, which shows that the film with high

radical mobility experiences more chemical modifications under the same fluence, as agrees with

84

the XPS results.

0 5 10 15 2075

80

85

90

95

100

Nor

mal

ized

num

ber

den

sity

(%

)

Time (us)

F=1.80 J/cm2

F=2.00 J/cm2

F=2.20 J/cm2

F=2.40 J/cm2

F=2.60 J/cm2

F=2.80 J/cm2

Figure 3.11: Simulation results of the typical normalized number density of the CH3CHCOO

units connecting at their ester group with other CH3CHCOO units on the surface. Results of the film with high radical mobility are shown.

The reason that the film with less radical mobility has less chemical modifications is because

radicals with low mobility less easily react with other free radicals to generate decomposed

products. This is known as the cage effect, which states that the free radicals generated by the

dissociation of molecules are unable to move apart due to the existence of surrounding molecules,

with the result that the dissociation products may recombine and return to the initial state. To

react with each other it is required that the free radicals meet in a single cage, which is the

volume in which the probability of reaction is significantly higher. The cage radius is affected

by radical mobility, which is higher in a high temperature or the molten state. The cage radius

can reach several tens of angstroms at a temperature close to the melting temperature (Butiagin,

1972). Due to a large cage radius, the radicals have high possibility to react with other radicals

within the cage. When the radical mobility is hindered, such as in the crystalline domain, the

effective cage radius decreases to the dimension of the elementary crystal unit cell, and the

number of active radicals in the cage is much less. Free radicals can also diffuse from cage to

cage to react with others. However, the diffusion is slow in a polymer matrix with a diffusion

85

coefficient on the order of 10-15 to 10-18 cm2/s. This slow process leads to a high probability of

free radical pair recombination in the cage, which is dominant in the crystalline domain. This

accounts for the phenomenon that higher fluence is required to cause chemical modifications in

the film with high radial mobility and its larger amount of chemical modifications.

Figure 3.6 shows that chemical modifications always occur even at small fluences, and no

chemical modification threshold exists. This differs from the XPS results while agrees with the

photochemical decomposition theory. The threshold observed in the XPS measurements, Figs.

3.8-3.10, is due to insignificant amount of chemical modifications. Appreciable amounts of

chemical modifications are generated by higher fluences. Chemical changes obtained from

experiments tend to level off at high fluences, while in the simulation the trend becomes steeper.

Small molecules generated by laser irradiation are ejected from the surface and are not detected

by the XPS measurements. However, the molecular ejection is not considered in the model,

which causes the steeper trend of chemical change in the simulation results.

0 5 10 15 200.0

5.0x105

1.0x106

1.5x106

2.0x106

2.5x106

3.0x106

Rea

ctio

n ra

te o

f E

q. (

9) (

mol

/m3 /s

)

Time (us)

F=2.80 J/cm2

F=2.60 J/cm2

F=2.40 J/cm2

F=2.20 J/cm2

F=2.00 J/cm2

F=1.80 J/cm2

Figure 3.12: Simulation results of the typical absolute value of the reaction rate of Eq. (9) for the amorphous domain after the onset of laser pulse. Results of the film with high radical mobility

are shown.

86

The reaction rate of Eq. (3), , defined in Eq. (9) as ⁄ , is determined. Figure 3.12 gives

its absolute value for the amorphous domain on the surface of the film with high radical mobility;

similar trends are obtained for the film with low radical mobility. The amorphous domain

becomes mobile as the temperature reaches , which is 60°C for PLLA. During the period

when the photodegradation can last, the temperature remains higher than . Therefore, the

amorphous domain is in mobile state and has a high quantum yield according to the cage effect.

The dissociation rate is maintained on the order of 106 mol/(m3s). The rate changes with

temperature and thus increases with fluences and decreases exponentially with time.

Figure 3.13 gives the absolute value of the reaction rate for the original crystalline domain on

surface of the film with high radical mobility. The crystal structure becomes mobile as the

temperature increases up to , which is assumed 175°C for the film with high radical mobility

and 180°C for the film with low radical mobility due to annealing. The temperature exceeds

because of high laser fluences and subsequently drops below it. When the temperature is

higher than , the crystals are melted and the dissociation rate is significantly increased due to

the high free radical mobility. The rate is maintained around 106 mol/(m3s) as in the

amorphous structure, Fig. 3.12. As the temperature drops below , radical mobility is limited

by the remaining crystal structure, and is assumed to be that of the crystal structure. The

dissociation rate drops to the order of 105 mol/(m3s). However, for smaller fluences, 1.80, 1.90,

and 2.00 J/cm2 for both films, the dissociation rate is kept on the order of 105 mol/(m3s). At

these fluence levels, no surface melting occurs under these fluences, and mobility is therefore

low. The results also agree with the simulated melting depth, Fig. 3.11, which demonstrates

that 2.10 J/cm2 is the lowest fluence causing surface melting for both films. Similar trends are

obtained for the film with low radical mobility, while the dissociation rate maintains at higher

87

value for a shorter time. A fluence of 2.80 J/cm2 maintains a high dissociation rate for the film

with high radical mobility for 12 μs, while it only maintains for 9 μs for another type of film.

The film with low radical mobility, due to annealing, has a higher , and therefore temperature

drops below the sooner.

0 5 10 15 200.0

5.0x105

1.0x106

1.5x106

2.0x106

2.5x106

3.0x106

Rea

ctio

n ra

te o

f E

q. (

9) (

mol

/m3 /s

)

Time (us)

F=2.80 J/cm2

F=2.60 J/cm2

F=2.40 J/cm2

F=2.20 J/cm2

F=2.10 J/cm2

F=2.00 J/cm2

F=1.80 J/cm2

0 1 2 30.0

5.0x104

1.0x105

1.5x105

2.0x105

2.5x105

3.0x105

Rea

ctio

n ra

te o

f E

q. (

9) (

mol

/m3 /s

)

Time (us)

Figure 3.13: Simulation results of the typical absolute value of the reaction rate of Eq. (9) for the crystalline domain after the onset of laser pulse. Results of the film with high radical mobility

are shown.

(a)

(a)

88

3.6 Conclusions

PLLA crystallinity changes and chemical modifications under excimer laser irradiation are

experimentally and numerically investigated. Crystallinity decrease due to laser melting is

demonstrated, and the effects of radical mobility on chemical modifications are evaluated. The

amount of chemical modifications is determined by the free radical mobility, which is decreased

by increasing the crystallinity through annealing before laser treatment. Both the WAXD and

simulation results show that the decrease of crystallinity initiates at the same fluence for both

films with high and low radical mobility. According to the XPS measurements and simulation,

the film with low radical mobility has a smaller amount of chemical modifications, and larger

fluences are required to cause appreciable chemical modifications. Films with smaller radical

mobility have a larger working window, in which the surface melting occurs while the chemical

modifications are limited. A larger working window increases the applicability of surface

treatment by the excimer laser.

Appendix

A possible series of chemical reactions following generating the free radicals R1 and R2 by UV

laser irradiation through Eq. (3) is given as follows (Inagaki et al., 2002).

C C O

H O

CH3

C C O

H O

CH3

C C O

H O

CH3

C

H

CH3

C

H

CH3

CO2

(R1) (R1) (R3)

C C O

H O

CH3

C

H

CH3

C

H

CH3

C C O

H O

CH3

C

H

CH3

C

H

CH3

O

(R3) (R3)

C C

H O

CH3

C C

H O

CH3

O C

H

CH

O

3

2

(R4)

A. 1

A. 2

89

C C

H O

CH3

C C

H O

CH3

O

(R4)

C C

H O

CH3

C

H

CH3

O

(R3)

CO

C C O

H O

CH3 (R2)

C C O

H O

CH3 (R2)

C C O

H O

CH3 (R5)

C O

O

C H2 4

C C O

H O

CH3 (R5)

C O

O

C C O

H O

CH3 (R2)

CO2

A. 3

A. 4

A. 5

90

Chapter 4: Effect of Laser Induced Crystallinity Modification on Biodegradation Profile of

Poly(L-Lactic Acid)

4.1 Introduction

Poly(L-lactic acid) (PLLA) is attractive in drug delivery, food packaging, and tissue engineering

applications because of its biocompatible and biodegradable properties. PLLA is especially of

interest in drug delivery applications because it hydrolyzes in the human body into lactic acid, a

product which is excreted by the body with no toxicity. In drug delivery applications, drugs are

embedded in a polymer matrix and released as it degrades. Biodegradable polymers offer a

means to control the drug delivery in time (Amass et al., 2008). PLLA degradation exhibits an

induction period, a duration of time required for water molecules to penetrate into the matrix

before degradation can occur. During this time, embedded drugs are not released at a linear

rate.

PLLA degradation in a physiological environment occurs via hydrolysis, in which water

penetrates into the polymer matrix, attacking the ester bonds and causing chain scission. Water

molecules readily penetrate into the amorphous region, but hardly into the crystalline region,

because the polymer chains are highly packed and densely ordered in crystals (Chu, 1981).

Degradation of ester bonds occurs faster in the amorphous phase because of its more permeable

structure. Hydrolysis in the crystalline phase begins at the fold surfaces and progresses inwards

as controlled by chain scission (Tsuji & Ikada, 1998). Hydrolysis in the crystalline phase

occurs less preferentially than in the amorphous phase. A sample composed of both crystalline

and amorphous phases will more rapidly undergo hydrolysis in the amorphous phase than the

crystalline phase.

91

PLLA hydrolysis is accelerated by autocatalysis (Siparsky et al., 1998). The rate of hydrolysis

increases as the concentration of reaction products increases. Hydrolysis of polyester produces

shorter chains with acid and alcohol end groups. Acid end groups dissociate, leading to an

acidic environment, which accelerates hydrolysis. Therefore, the diffusion of shorter chains out

of the polymer plays a key role in controlling the overall degradation rate. Another factor that

complicates biodegradation is the increase of crystallinity during degradation (Zong et al., 1999;

Renouf-Glausera et al., 2005). Preferential degradation in the amorphous region leaves the

crystalline phase behind, leading to a mostly crystalline material. Chain scission as a result of

degradation also increases chain mobility which facilitates crystallization in the amorphous

phase.

Because changing the surface crystallinity of PLLA has the potential to tailor the initial

degradation rate, laser heat treatment to reduce surface crystallinity is of interest. PLLA

crystallinity reduction by laser irradiation with infra-red (IR) and green wavelengths has been

quantified through wide-angle X-ray diffraction (WAXD) measurements (Bhatla & Yao, 2009).

The same effects have been observed by excimer laser processes operating at ultra violet (UV)

wavelengths with photon energies higher than bond energies (Dunn & Ouderkirk, 1990; Hsu et

al., 2012). It has been shown that surface crystallinity can be reduced with no measurable

chemical modifications due to the low radical mobility in the crystal structure (Hsu et al., 2012).

The effects of laser treatment on polymer degradation over time have yet to be studied.

The objective of this work is to investigate the effects of laser surface treatment on PLLA

degradation, characterized by the change in mass, molecular weight (MW), and crystallinity.

Since PLLA degradation behaves similarly at elevated temperatures (Weir et al., 2004),

degradation is conducted at elevated temperatures to shorten the total test time. MW and

92

crystallinity are determined through gel permeation chromatography (GPC) and WAXD,

respectively. A numerical model is developed to capture the degradation process.

4.2 Background

4.2.1 Laser Melting of Polymer

Polymer melting is an amorphization process in which crystalline polymer chains detach from

crystals. Crystalline polymer chains are held by weak van der Waals forces and hydrogen

bonds, which are broken during melting to form an amorphous structure. Melting initiates at

the crystal fold surface, followed by the progressive unfolding of chains towards the center of the

crystal. Polymer melting occurs within nanoseconds (Yamamoto, 2010), and thus nanosecond

laser treatment is used to induce polymer melting. Crystallinity is reduced after laser melting as

a result of slow crystallization kinetics compared to the rapid melting and cooling during laser

processing (Bhatla & Yao, 2009; Dunn & Ouderkirk, 1990; Hsu et al., 2012).

Polymer absorption of photons with energies exceeding the bond energy risk braking chemical

bonds. In PLA, this reaction occurs between two CH3CHCOO units. If no oxidation occurs

during processing, the separated CH3CHCOO units may recombine. According to the cage

effect, high PLLA crystallinity has been shown to reduce the amount of bond breaking and lead

to non-measurable chemical modifications under UV laser treatment (Hsu et al., 2012).

4.2.2 Biodegradation of Polyester

PLA is a biodegradable polyester which breaks down in the human body through hydrolysis.

Water molecules attack the ester bonds via the following reaction

93

C C O C C O

H HO O

CH3 CH3

2OH C C

H O

CH3

OH HO C C O

H O

CH3

Hydrolysis causes chain scission, and produces shorter chains with carboxylic (-COOH) groups

and alcohol (-OH). The rate of hydrolysis is proportional to the number of ester bonds present

in each monomer constituting the polymer. The number of reactive bonds in the polymer

decreases as hydrolysis goes on. Hydrolysis depends on the molar concentrations of the

monomer still present in the polymer chain, Ce, and water, Cw (Pitt & Gu, 1987), at the rate of

dCe

dt=-k1CeCw 2

where the constant k1 depends on temperature and does not vary with reaction. Carboxylic

end groups generated from this reaction have a high degree of dissociation and can act as a

catalyst to accelerate the hydrolysis. Hydrolysis of polyesters may become autocatalytic if

carboxylic end groups remain in the bulk (Li et al., 1990). During autocatalyzed hydrolysis, the

reaction rate depends on the concentration of the carboxylic end groups, CCOOH, as well. The

rate of autocatalyzed hydrolysis is given by (Lyu et al., 2007)

dCe

dt=-k2CeCw CCOOH

n (3)

where k2 is the rate constant for the autocatalysis reaction, and n accounts for the dissociation of

the carboxylic groups and can be used as an empirical parameter to reflect the reactions (Wang et

al., 2008).

Amorphous chains hydrolyze in three stages before generating monomers (Stephens et al., 2008).

In stage 1, chain scission commences in the intact amorphous chains, resulting in the breaking of

(1)

94

amorphous tie chains. In stage 2, chain scission occurs on already broken amorphous chains,

creating unattached oligomers. In stage 3, hydrolytic reactions occur on short amorphous

segments protruding from the crystals and the oligomers produced in stage 2. Hydrolysis in this

stage generates highly mobile monomers which can diffuse out of the polymer matrix.

(a)

(b)

Figure 4.1: Schematic representation of (a) semicrystalline structure and (b) crystalline residue of polymer after hydrolysis, which preferentially occurs in amorphous region and on crystal fold

surface (Tsuji & Ikarashi, 2004).

PLLA can be semicrystalline, and the amorphous region experiences a higher hydrolysis rate

than the crystalline region (Chu, 1981; Tsuji & Tsuruno, 2010). Hydrolysis of the amorphous

phase occurs randomly, but occurs preferentially in the crystal phase at the fold surfaces. A

schematic representation of semicrystalline polymer chains before and after extensive hydrolytic

degradation is given in Fig. 4.1 (Tsuji & Ikarashi, 2004). Random hydrolysis generates chains

Chain with free end Folded chain Loosely packed structure

Tie chainAmorphous

region Amorphous

region Crystalline region Crystalline region Amorphous region

Selective chain scission on the fold surface and

in the amorphous region

95

with a wide length distribution; preferential hydrolysis on fold surfaces can lead to the chains

representing the integral folds the crystalline residues (Tsuji & Tsuruno, 2010; Fischer et al.,

1973). Mobile segments reorganize themselves from a disordered to an ordered state due to

intermolecular hydrogen bonds and van der Waals forces, leading to crystallization during

hydrolysis (Chu, 1981).

4.3 Numerical Model

A 2D symmetric model is developed to investigate the effect of laser treatments on

biodegradation profiles. Laser energy absorbed by semi-transparent bulk PLLA generates heat

governed by the heat equation

ρCp T∂T

∂t= · k T +q z,t (4)

where ρ is mass density, Cp T is specific heat as a function of temperature T, k is thermal

conductivity, q(z,t) is the laser power density as a function of depth from the laser irradiated

surface z and time t, expressed as

q z,t =Q0e-αz+β

ttp

-22

(5)

where Q0 is peak power density, α is absorption coefficient, tp is pulse width, and β is -4ln2.

Parameters used in Eqs. (4) and (5) and the change in Cp T during the phase transition are

further detailed in ref. (Hsu et al., 2012). Polymer crystals melt within a temperature range,

from Tm to Tm+∆Tm. The crystal fraction melting in a temperature increment dTm within Tm

and Tm+∆Tm can be expressed as ϕ(t,Tm)dTm . The total crystallinity is then

Φ(t)= ϕ(t,Tm)dTmTm+∆Tm

Tm. The change in ϕ(t,Tm) during melting is given by (Toda et al., 1998)

96

dt(t,Tm)=-Rm(ΔT)ϕ(t,Tm) (6)

The melting rate coefficient Rm is a function of superheating, ΔT=T-Tm , expressed as

Rm= Rm(ΔT) if ΔT>0 and Rm=0 if ΔT<0. Molecular dynamics simulations have suggested

that crystallinity decreases within ns, and Rm is of the order of 109 s-1 (Yamamoto, 2010).

Spatially resolved crystallinity after laser processing is imported into the degradation model as

the initial condition.

Biodegradation is captured using a phenomenological model (Wang et al., 2008).

Semicrystalline PLLA is modeled to be consistent of 7 species during degradation: non-degraded

amorphous chains, degraded amorphous chains in stages 1, 2, and 3 (Stephens et al., 2008),

crystalline chains, monomers, and water molecules. Non-degraded and degraded amorphous

chains hydrolyze via the scission of the ester bonds contained in the species, but have zero

diffuse-ability due to restricted mobility. Monomers have high mobility. Amorphous polymer

chains crystallize during degradation. Crystalline polymer chains cannot diffuse and are

assumed to hydrolyze ten times slower than amorphous chains (Li et al., 2009), generating chain

scission on the crystal fold surface. Water molecules are assumed abundant in time and space,

and the size distribution of polymer chains is neglected. Acid end groups of the monomers,

generated during chain scission, accelerate the hydrolysis through autocatalysis. Assuming

water concentration is constant temporally and spatially, the molar concentration of the

non-degraded amorphous chain is expressed as

dC0

dt=-γ0C0-ε0C0Cm

n -κ0dCc

dt (7)

where C0 and Cc are the molar concentrations of monomers in the non-degraded amorphous

chains and crystalline chains, respectively. The dissociation of the acid end groups n is

97

assumed to be unity. The first, second, and third terms on the right of Eq. (7) account for rates

of non-autocatalysis, autocatalysis, and crystallization from amorphous chains, respectively.

Values of γ0, ε0, and κ0 are the corresponding phenomenological rate constants. Hydrolysis

of non-degraded amorphous chains, Eq. (7), generates degraded amorphous chains in stage 1,

which is then hydrolyzed into stage 2 and contributes to crystallinity. Similarly, hydrolysis of

chains in stage 1 leads to stage 2, and hydrolysis of chains in stage 1 leads to stage 2. Part of

the hydrolyzed chains also crystallize during hydrolysis. The molar concentration of monomers

in each stage is expressed as

dCi

dt=(γi-1+εi-1Cm

n )Ci-1-(γi+εiCmn )Ci-κi

dCc

dt (8)

where i=1, 2, and 3, representing degradation stages 1, 2, and 3. Ci is the molar concentration

of the monomers in stage i. γi, εi, and κi are the phenomenological rate constants accounting

for non-autocatalysis, autocatalysis, and crystallization due to hydrolysis in stage i. During

degradation, the crystallization of linear polyesters are modeled using the Avrami equation (Zong

et al., 2009) ϕ=1-exp(-kctm) where ϕ is monomer fraction in the crystalline domain, kc is the

Avrami constant and m is the Avrami exponent (Avrami, 1941). Hydrolysis of the stage 3

species generates monomers which do not connect to other monomers and have high mobility to

diffuse out of the polymer matrix. Assuming Fick’s second law for monomer diffusion, which

predicts the change of monomer concentration with space and time, the molar concentration Cm

of the monomers with high mobility is modeled by

dCm

dt= γ3+ε3Cm

n C3+ ·(D Cm) (9)

where D is the monomer diffusivity. Appropriate units are used for the phenomenological rate

constants so that each term in Eqs. (7)-(9) is in mole per volume per time. The values of the

98

rate constants are selected to capture experimental results (Wang et al., 2008).

The coupled partial differential Eqs. (4) and (6) are solved to determine crystallinity change due

to laser treatments. Solutions of these equations are used as initial conditions to solve the

coupled Eqs. (7)-(9) to capture degradation profiles. The two sets of coupled equations are

solved through the finite element method in COMSOL Multiphysics 4.1. A 2D axial symmetric

domain is used, in which the 1 mm by 5 mm film is immersed in a 10 mm by 10 mm aqueous

medium. Laser treated area covers both sides of the film. The film is initially composed of

non-degraded crystalline and amorphous chains with crystallinity determined experimentally.

Monomers generated during degradation can diffuse out of the film into the surrounding aqueous

medium. The time domain in the simulation is 14 days.

4.4 Materials and Methods

PLLA granules were provided by PURAC and used as received. PLLA samples were prepared

through thermal compression of PLLA granules under 5.7×104 Pa at 180°C for 4 hours, and

cooled down in air. Crystals develop during the cooling process. The obtained sample is

around 1 mm thick. Laser treatment was conducted on both sides of the sample by a KrF

excimer laser with 248 nm wavelength 248 nm and 25 ns pulse width 25 ns. The homogenized

excimer laser beam has a spatially uniform intensity favorable for a uniform surface treatment.

The sample is radiated by a single pulse in argon atmosphere with a flow rate of 15 standard

cubic feet per hour to prevent oxidation. To determine crystallinity, WAXD measurements are

made using an Inel X-ray diffractometer to determine crystallinity. Monochromatic CuKα

radiation with wavelength λ=0.15418 nm at 40 kV and 30 mA is used. The chemical

compositions of the PLLA samples are measured using X-ray photoelectron spectroscopy (XPS,

99

PHI 5500 ESCA). From XPS measurements, the O1s and C1s spectra are captured, and the

take-off angle is 45°.

During degradation tests, each sample was placed in a vial and fully immersed in a 10 mL

phosphate buffered saline (PBS) with a pH of 7.4 purchased from Life Technologies. Vials

were placed in a water bath at 70°C, and the PBS was changed every two days. After

degradation, samples were dried in vacuum for two days. The sample mass was recorded

before and after vacuum drying. The weight-average MW (Mw) and number-average MW (Mn)

were determined in tetrahydrofuran (THF) from gel permeation chromatography (GPC) at room

temperature. The GPC was calibrated with polystyrene standards. Crystalline PLLA was first

dissolved in methylene chloride and rapidly dried with a rotary evaporator to obtain the

amorphous form with higher solubility in THF. PLLA/THF solutions were prepared with 1

mg/1 mL concentration for GPC measurements. Crystallinity is determined by the WAXD, and

morphology is observed using stereomicroscopy and optical profilometry.

4.5 Results and Discussion

4.5.1 Effect of Laser Irradiation on Crystallinity and Chemical Modifications

Images of samples before and after laser treatment as well as after 14 days of degradation are

presented in Fig. 4.2. The non-laser treated sample is semi-transparent because of the existence

of crystalline phase. Laser treatment generates opaque spots, a result of strong light scattering

from the roughened surface as confirmed by optical profilometry. Roughness is induced by

laser surface melting and resolidification. WAXD profiles are given in Fig. 4.7. The 16.7°

WAXD peak becomes less prominent for the samples treated with higher laser fluences, due to

crystallinity reduction. Crystallinity is calculated based on WAXD profiles (Alexander, 1969),

100

as given in Fig. 4.3, which shows that the crystallinity decreases at 2.4 J/cm2 as a result of laser

melting.

Figure 4.2: Non-degraded PLLA sample (a) before and (b) after laser treatment with a fluence of

3 J/cm2. Laser irradiated spots show less transparency due to increased surface roughness. The laser-treated sample degraded for 14 days is given in (c) and its high crystallinity reduces the

transparency.

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

0.3

0.4

0.5

0.6

Cry

stal

linity

(%

)

O/C

Fluence (J/cm2)

O/C

20

40

60

80

100

Crystallinity

Figure 4.3: PLLA crystallinity and the ratio of O1s to C1s as a function of laser fluence. The

error bar represents the standard deviation of 3 data points.

To investigate possible chemical modifications under laser irradiation, XPS measurements are

performed. Chemical reactions caused by laser irradiation with photon energies higher than the

binding energies produce small molecules such as CO and CO2, and reduce the total ratio of

oxygen to carbon (Inagaki et al., 2002). The ratio of O1s and C1s, shown in Fig. 4.3, remains

(a) (b) (c)

101

constant below 3.0 J/cm2, and drops when fluence reaches 3.2 J/cm2, indicative of chemical

modifications due to laser treatment. PLLA samples remain chemically unmodified below 3.0

J/cm2 due to the cage effect, which states that free radicals generated by the dissociation of

molecules cannot move apart because of the confinement of surrounding molecules (Hsu et al.,

2012). This results in the recombination of the dissociation products, which then return to the

initial state. A fluence of 3.0 J/cm2 induces the maximum amorphization with non-measurable

chemical modifications, and is used as the energy level to treat samples in this study.

4.5.2 Effect of Laser Irradiation on Degradation

4.5.2.1 Modification of Morphology

Laser treated samples after having been degraded for 14 days is shown in Fig. 4.2. A uniformly

opaque morphology is observed as a result of higher crystallinity developed during degradation.

Detailed morphology investigations are given by stereomicroscopic images of laser treated

samples as a function of degradation period in Fig. 4.4. A laser treated sample before

degradation is given in Fig. 4.4(a). The 1 mm by 1 mm square spot is generated by laser

irradiation. As degradation progresses, spot edges become less defined, Fig. 4.4(b), suggesting

the degradation and erosion (mass loss) of laser melted layer. After 8 days, Fig. 4.4(c), spot

edges become more indistinct than samples degraded for shorter periods. During prolonged

testing, the laser melted layer is eroded and degradation starts to occur in the non-melted volume

below. Eventually, the morphology of laser treated sample, Fig. 4.4(d), is similar to that of the

non-laser treated sample. It is observed that light scattering is not uniform over the surface.

Surface roughness is caused by the selective degradation on the amorphous phase, leaving the

crystalline phase with strong local light scattering.

102

Figure 4.4: Surface morphology of the laser treated sample (a) before and after degradation for

(b) 3, (c) 8, and (d) 14 days under the stereomicroscope. The squares in (a) are laser spots. The edges of laser spots become less defined with degradation period, suggesting the erosion of

laser melted layer. Degradation in the non-melted bulk volume occurs in the later stage.

Figure 4.5: Cross section of the laser-treated sample degraded for 14 days. The bulk remains

solid, suggesting that autocatalysis is not dominant.

The cross section of a laser treated sample after being degraded for 14 days is given in Fig. 4.5.

It is observed that the bulk remains solid, which suggests that autocatalytic degradation is not

(c)

(a)

(d)

(b)

103

dominant as a result of easy diffusion of acidic monomers out of the sample. Spatial

degradation profiles of monomer molar concentration are given in Fig. 4.6. After 0.5 days,

monomer concentration on the surface is higher than that in the bulk, suggesting accelerated

degradation in the 80-μm thick laser melted layer. In the later stage of degradation, Fig. 4.6(b),

monomer concentration in the non-melted bulk starts to increase. The simulation results

capture that laser melted layer experiences a higher degradation rate.

(a)

(b)

Figure 4.6: Simulated spatial distribution of monomer concentration in the laser treated sample degraded for (a) 0.5 days and (b) 6 days.

104

4.5.2.2 Modification of Crystallinity

To investigate the changes in crystallinity during degradation, WAXD measurements have been

conducted on samples after regular degradation intervals, as depicted in Fig. 4.7. Laser

treatments cause surface melting and a reduced crystalline peak, as discussed in Sec. 4.5.1. For

both types of samples, crystalline peaks become more prominent with time, suggesting the

occurrence of crystallization during degradation. Sample crystallinity, as calculated from

WAXD measurements (Alexander, 1969), is given as a function of time in Fig. 4.8, which shows

that crystallinity increases with degradation period, in agreement with Fig. 4.7. For both types

of samples a significant increase of crystallinity is observed at day 0.5, when the sample MW

begins to decrease, as is discussed in Sec. 4.5.2.3. The decrease of MW favors chain

reorganization and crystallization, and increases crystallinity. Lower crystallinity and greater

amounts of crystallizable material in non-degraded samples also favors crystallization (Avrami,

1939). Crystallization slows between day 0.5 and day 5 for the non-laser treated sample,

because part of the amorphous chains is less crystallizable, including the rigid amorphous phase

(RAP) (Menczel & Wunderlich, 1981) and entangled chains. The RAP and entangled chains

hinder chain movements and make it difficult for amorphous chains to relax and crystallize.

Crystallinity also significantly increases at day 8 for the non-laser treated sample, and day 3 for

the laser treated sample. Crystallinity increase at this stage is mainly caused by the loss of

amorphous chains, Fig. 4.11, as discussed in Sec. 4.5.2.3.

105

10 15 20 25 30

0

20

40

60

80

100

120

140

Inte

nsity

(cp

s)

2 theta (deg)

Degraded for 8 days Degraded for 0.5 days Before degradation

(110)/(200)

(203)

(a)

10 15 20 25 30

0

20

40

60

80

100

120

140

(203)

(110)/(200)

Inte

nsi

ty (

cps)

2 theta (deg)

Degraded for 3 days Degraded for 0.5 days Before degradation

(b)

Figure 4.7: WAXD profiles of the (a) non-laser treated and (b) laser treated samples degraded for regular periods. Intensity of crystalline peaks increases with degradation period, suggesting

a higher crystallinity. Profiles are shifted in y direction for viewing clarity.

0 2 4 6 8 10 12 14

40

50

60

70

80

90

Experimental results of non-laser treated film Experimental results of laser treated film

Cry

sta

llin

ity (

%)

Degradation period (day)

Figure 4.8: Crystallinity as a function of degradation period determined from WAXD. Crystallinity increases with degradation period. A significant increase occurs on day 0.5 for

both types of samples. Crystallinity also increases on day 8 and day 5 for the non-laser treated and laser treated samples, respectively. The error bar represents the standard deviation of 3 data

points.

106

4.5.2.3 Modification of Molecular Weight and Mass

Degradation is characterized by sample MW and sample mass. MW distribution profiles as a

function of degradation period are given in Fig. 4.9, which plots the mass fraction of chains (w)

per increment of MW in a logarithmic scale (logM) versus MW. The number average MW (Mn)

and weight average MW (Mw) are determined from MW distributions. The polydispersity

index (PDI), defined as Mw/Mn, is calculated as a measure of distribution of MW. Mn, Mw, and

PDI for the non-laser treated and laser treated samples are given in Fig. 4.10. Sample mass as a

function of degradation period is given in Fig. 4.11. Simulated sample MW and sample mass

for both types of samples are given in Figs. 4.12 and 4.13.

10000 1000000

2

4

6

8

10

Before degradation

2-day degradation

Molecular weight (g/mol)

5-day degradation

8-day degradation

11-day degradation

Inte

nsi

ty (

dw

/dlo

gM

)

14-day degradation

3-day degradation

(a)

107

10000 1000000

2

4

6

8

10

Inte

nsi

ty (

dw

/dlo

gM

)

Molecular weight (g/mol)

Before degradation

2-day degradation

3-day degradation

5-day degradation

8-day degradation

11-day degradation

14-day degradation

(b)

Figure 4.9: GPC profiles of the (a) non-laser treated and (b) laser treated samples after regular degradation periods. For (a), the distribution becomes wider and shifts left as the degradation

period increases to day 5, representing the random chain scission in the amorphous region. After day 8, a distinct new peak is developed due to selective chain scission of the fold surface of

crystals. For (b), the MW distribution extends to the left before day 3, signifying the random chain scission of the laser melt layer. At day 5, two distinct peaks are developed due to the selective chain scission of the partially melted crystal fold surfaces. Profiles are shifted in y

direction for viewing clarity.

For non-laser treated samples, the GPC profile before degradation is centered at around 120000

g/mol and the distribution is narrow, as shown in Fig. 4.9(a). As the degradation period

increases, days 2, 3 and 5, the distribution gradually becomes wide and shifts left. A small

hump is developed at around 60000 g/mol at day 5. Sample MW decreases while PDI increases

as a result of random chain scission, as given in Fig. 4.10(a). Random chain scission before day

5 occurs in the amorphous region, causing chains to degrade into segments with various MWs

(Tsuji & Tsuruno, 2010). There is no obvious mass decrease before day 5 for non-laser treated

sample, as given in Fig. 4.11, because the degraded segments are not small enough to diffuse out

108

of the sample. The degraded segments have higher mobility, and reorganize themselves from a

disordered to an ordered state as a result of intermolecular hydrogen bonds and van der Waals

forces during degradation, which increases the crystallinity for the first 5 days, as shown in Fig.

4.8.

0 2 4 6 8 10 12 140

20

40

60

80

100

Pol

ydis

pers

ity in

dex

Degradation period (day)

Mo

lecu

lar

we

igh

t (%

)

Mw Mn

1.0

1.5

2.0

2.5

3.0

Polydispersity index

(a)

0 2 4 6 8 10 12 140

20

40

60

80

100

Mo

lecu

lar

we

igh

t (%

)

Degradation period (day)

Mw Mn

1.0

1.5

2.0

2.5

3.0

Pol

ydis

pers

ity in

dex

Polydispersity index

(b)

Figure 4.10: Mw, Mn, and PDI of the (a) non-laser treated and (b) laser treated samples after regular degradation periods. Mw and Mn decrease at a higher rate for the laser treated samples, as a result of fast degradation in the laser melted layer. The non-homogeneous degradation of the melted layer and bulk increases PDI. The error bar represents the standard deviation of 3

data points.

109

At day 8, the GPC profile of the non-laser treated samples shifts left significantly, Fig. 4.9(a).

A distinct peak, corresponding to a smaller MW of 10300 g/mol, is observed. This shift is due

to the selective scission of the crystal fold surface. Preferential scission generates chains with

MW representing the integral folds the crystalline residues (Tsuji & Tsuruno, 2010; Fischer et al.,

1973). The left peak is mainly composed of the MW of crystalline residues, and the right peak

mainly comes from the amorphous region experiencing random chain scission. As degradation

approaches 11 and 14 days, the right peak starts to diminish due to the continuous scission of

amorphous chains, and the overall distribution is mainly composed of a peak centered at 10300

g/mol. The PDI thus decreases, as shown in Fig. 4.10(a).

0 2 4 6 8 10 12 14

70

80

90

100

Film

ma

ss (

%)

Degradation period (day)

Non-laser treated, experimental results Laser treated, experimental results

Figure 4.11: Experimental results of sample mass with and without laser treatments. Mass decrease is observed after day 8 for non-laser treated sample and after day 3 for laser treated

sample. The error bar represents the standard deviation of 3 data points.

The lowest MW, around 3000 g/mol, is detected after day 8 in Fig. 4.9(a), while mass decrease

also occurs at day 8, Fig. 4.11. This finding suggests that chains with MW smaller than 3000

g/mol are diffused out of the sample. Figure 4.8 shows that crystallinity increases significantly

at day 8 for the non-laser treated sample. Thus, the materials diffusing out of the sample are

110

mainly amorphous. The reduction of amorphous material increases the crystallinity in the

sample and leads to stronger local light scattering shown in Fig. 4.4(d).

GPC profiles of the laser treated sample are given in Fig. 4.9(b). Before degradation, the GPC

profile of the laser treated sample has a peak at 120000 g/mol similar to the non-laser treated

sample, which suggests that laser treatment does not modify PLLA MW. When compared with

the non-laser treated sample, MW decreases at a high rate at the early stage for the laser treated

sample. On days 2 and 3, significant changes of GPC profiles are observed in Fig. 4.9(b). At

day 2, the peak at 120000 g/mol decreases, while the distribution extends to the smaller MW

region. The peak diminishes at day 3, and a hump at smaller MW region is developed.

During this period, a random chain scission of the laser melted layer occurs.

Non-homogeneous degradation at the laser melted layer and the bulk widens the MW

distribution and thus increases the PDI, Fig. 4.10(b). The high degradation rate of the laser

treated sample is caused by water molecules penetrating into the less ordered structure caused by

laser melting. Faster water penetration into the bulk is also responsible for the diminished

original peak at day 3.

At day 5, the hump is split into two peaks for the laser treated sample, as shown in Fig. 4.9(b).

The distinct distribution of two peaks is suggestive of selective chain scission on crystal fold

surfaces. Same phenomenon is observed in the non-laser treated sample at day 8. The earlier

occurrence of this selective scission for the laser treated sample may be due to the less ordered

structure of the partially melted crystals. For partially melted crystals, the crystal thickness is

diminished, which increases the molecular loop length on the fold surface (Belyayev, 1988).

This longer loop length enlarges the free volume, the volume not occupied by polymer molecules

in a polymer matrix. A larger free volume allows for easy water penetration, and facilitates

111

hydrolysis on the fold surface. Toward the late stage of the degradation, the right peak

diminishes and the left peak increases, as shown in Fig. 4.9(b). This left peak is composed of

the chains with MW representing the integral folds the crystalline residues, as well as the

smallest MW distribution of chains which can remain in the matrix. The narrowing MW

distribution at the late degradation stage decreases the PDI, as given in Fig. 4.10(b) after day 5.

Laser surface melting accelerates MW reduction, which in turn shortens the induction periods of

mass loss. Mass loss occurs at day 3 for the laser treated samples, compared to day 8 for the

non-laser treated sample, Fig. 4.11. The decrease of sample mass is accompanied with the

increase of crystallinity, Fig. 4.8, a phenomenon observed for both types of samples. The laser

melted layer is composed of amorphous chains, part of which degrades and diffuses out at a

higher rate, while the remainder crystallizes. At day 3, the degraded products in the laser

melted layer have small enough MW such that diffusion out of the matrix occurs, which leads to

the mass loss as shown in Fig. 4.11. After the erosion of the laser melted layer, the sample MW

and mass decrease at similar rates as the non-laser treated sample, as observed in Figs. 4.10 and

4.11.

0 2 4 6 8 10 12 140

20

40

60

80

100

Mol

ecul

ar w

eigh

t (%

)

Degradation period (day)

Non-laser treated, simulation results Laser treated, simulation results

Figure 4.12: Simulated MW of the non-laser treated and laser treated samples.

112

Simulation results of MW change as a function of degradation period are given in Fig. 4.12.

Numerical results of the laser treated sample experience a larger MW decrease, in agreement

with experiments. Experimentally, the MW of the non-laser treated sample does not decrease

as significantly as predicted in simulation before day 1. This difference is because it takes time

for water molecules to diffuse into the bulk in experiments. Water diffusion time is not

considered in simulation, which causes the overestimation of the MW decrease.

The simulated mass changes with degradation, Fig. 4.13, capture the phenomenon that

appreciable mass decrease occurs after an induction period. During the induction period in

which mass loss does not occur, the original amorphous polymer chains degrade into fragments

not small enough for diffusion out of the sample. Fragments experience further degradation

into monomers, which can diffuse and lead to mass loss. Experimentally, the mass change is

more significant when compared with simulation. This is because the erosion of crystalline

material also occurs on the surface during degradation (Lam et al., 2008), which is not

considered in simulation, leading to a conservative prediction.

0 2 4 6 8 10 12 14

70

80

90

100

Film

ma

ss (

%)

Degradation period (day)

Non-laser treated, simulation results Laser treated, simulation results

Figure 4.13: Simulated mass change of the non-laser treated and laser treated samples.

113

4.5.3 Concentration of Species during Degradation

The molar concentration of species considered in the simulation is plotted with the degradation

period, given in Fig. 4.14(a). Amorphous chains start to degrade in the early stage of the

degradation period, generating the species in stage 1. Stage 1 is sequentially degraded into

stages 2 and 3. Chains in stage 3, upon further hydrolysis, are decomposed into monomers,

which diffuse out of the sample. Only a small amount of monomers remain on the sample

surface. Degradation also enhances chain mobility, which leads to crystallization. The

simulated concentration of crystalline chains increases as observed in experiments.

Simulated crystallinity increases via two stages, as agreed with the experiments. The first stage

is predicted by the Avrami crystallization theory (Avrami, 1939). Experimentally, crystallinity

increases to a significant extent before day 0.5, not captured in the simulation. The discrepancy

results from the fact that according to the Avrami theory, crystallization begins with an induction

period due to the formation of nuclei. Pre-existing nuclei in the experiment samples favor

crystal growth without experiencing the induction period. The second stage of crystallinity

increase is due to the loss of amorphous material, which increases crystallinity as experimentally

observed and numerically simulated. Hydrolysis of crystals occurs on the fold surfaces.

Concentration of non-degraded crystalline chains decreases, which leads to the increased

concentration of the hydrolyzed crystals and the reduced MW.

Simulation results of the species concentration in the laser melted layer are shown in Fig. 4.14(b).

Laser irradiation melts the crystal structure, generating an amorphous layer near the surface.

Amorphous chains are degraded to stages 1, 2, 3, and monomers in sequence. Because the

species are near the surface, the major part of monomers readily diffuses out of the sample and

114

thus the concentration is lower as compared to that in the sample bulk. Although most

amorphous chains degrade, crystallization takes place in the laser melted layer, mainly at the

early stage.

0 2 4 6 8 10 12 140

2000

4000

6000

8000

10000

Co

nce

ntr

atio

n (

mo

l/m3 )

Degradation period (day)

Crystalline chain Crystalline chain, scission on fold surface Non-degraded crystalline chain Non-degraded amorphous chain Amorphous chain in degradation stage 1 Amorphous chain in degradation stage 2 Amorphous chain in degradation stage 3 Monomer

(a)

0 2 4 6 8 10 12 140

2000

4000

6000

8000

10000

12000

14000

Co

nce

ntr

atio

n (

mo

l/m3 )

Degradation period (day)

Non-degraded amorphous chain Amorphous chain in degradation stage 1 Amorphous chain in degradation stage 2 Amorphous chain in degradation stage 3 Monomer Crystalline chain

(b)

Figure 4.14: Molar concentration of the simulated species as a function of degradation period. (a) In the center of the non-laser treated sample. (b) In the laser melted layer of the laser treated

sample.

115

4.6 Conclusions

The effect of laser irradiation on PLLA biodegradation has been studied experimentally and

numerically. Excimer laser irradiation has been shown to melt the PLLA surface. The melted

layer showed lower crystallinity and no observable chemical changes according to WAXD and

XPS measurements. Optical micrographs taken during degradation testing show that

degradation initiates on the sample surface, and the melted layer allows for the accelerated initial

degradation as is also captured through numerical simulations. Laser treated samples

experience a faster initial MW reduction, which leads to a shorter induction period of mass loss

as observed from GPC. Heterogeneous degradation of PLLA samples caused by

inhomogeneous crystallinity is observed in both laser treated and non-laser treated samples from

bimodal distribution of MW from GPC measurements. Laser treated samples exhibited higher

crystallinity after degradation as a result of preferred degradation and higher chain mobility in

the amorphous phase, as confirmed from the WAXD measurements. Changes in MW, sample

mass, and species evolution during degradation are captured in simulation results of both laser

treated and non-laser treated samples. Laser modification to the crystallinity of PLLA has been

shown to reduce the degradation induction period, a desired property in drug delivery

applications.

116

Chapter 5: Effect of Drug Loading and Laser Surface Melting on Polymer Biodegradation

and Drug Release Profile

5.1 Introduction

Poly(lactic acid) (PLA) has been attracting wide attention due to its biocompatible and

biodegradable properties. PLA has been of interest in drug delivery applications because it

hydrolyzes in the human body into lactic acid, a product which is then excreted by the body with

no toxicity. In drug delivery applications, drug molecules are embedded in a polymer matrix

and released into the human body, with the release profile determined by polymer matrix

degradation. The advantages of using biodegradable polymers in a drug delivery system are

demonstrated by the controlled drug release profiles over time. With the controlled release, the

concentration of the released drug in the human body is stably maintained between the minimum

effective level and toxic level (Amass et al., 1998). The prolonged drug effective period also

reduces drug taking frequency and improves the life quality of patients.

Drug release profile is determined by the degradation of polymer matrix. In a physiological

environment, PLA degradation occurs via hydrolysis, in which water penetrates into the polymer

matrix, breaking the ester bonds and causing chain scission. Crystallinity affects PLA

hydrolysis, because water molecules readily penetrate into the amorphous region, while are

hardly accommodated in the crystalline region with highly packed and densely ordered structures

(Chu, 1981). A higher water concentration increases hydrolytic degradation rate. Therefore,

degradation rate of ester bonds is higher in the amorphous region, and disorientation of the

lamellae and disappearance of the crystal structure occur later (Tsuji & Ikada, 1998). The slow

degradation kinetics of crystalline PLA leads to an induction period of drug release, during

117

which the limited amount of drug is released. The induction period is undesirable because it

delays drug release and effectiveness. Since PLA degradation is a function of crystallinity, as

an attempt to shortening the induction period, laser irradiation has been conducted on the PLA

surface to reduce its crystallinity (Hsu et al., 2012a). It has been shown that with a reduced

surface crystallinity, PLA degrades at a higher initial rate and experiences a shorter time period

before occurrence of mass loss (Hsu et al., 2012b). The results demonstrate the potentiality to

shorten the induction period of drug release through laser crystallinity modification.

In addition to polymer crystallinity, drug release is also influenced by drug loading. Drugs are

typically small molecules. Blending of small molecules into polymer affects polymer

crystallinity and thermal properties, a process known as plasticization. Plasticization increases

chain mobility because the existence of small molecules renders extra free volume between

chains (Flory, 1940). From differential scanning calorimetry (DSC), the plasticized PLA has

been reported to have a reduced glass transition temperature (Tg) and melting temperature (Tm)

(Ljungberg & Wesslen, 2002; Xiao et al., 2009). The reduced Tg and Tm are a result of higher

chain mobility (Yeh et al., 2009). Chain mobility is a factor which affects PLA degradation and

therefore drug release. Plasticization also increases PLA crystallinity as compared to the

non-plasticized PLA, because mobile chains are easy to reorganize themselves to an ordered and

energetically stable state (Xiao et al., 2009; Yeh et al., 2009). PLA with higher crystallinity

exhibits a slower degradation, which in turn slows down drug release. The combined effects of

modified chain mobility and crystallinity by drug loading on drug release are not clear and

require further investigation.

Drug release can also be resulted from drug dissolution and diffusion, which leads to an initial

high release rate, known as the initial burst. One mechanism of initial burst is the rapid

118

dissolution of a large amount of drug molecules on the surface of polymer matrix (Lao et al.,

2009). The burst phenomenon is morphologically illustrated on the surface of drug loaded

polymer matrix before and after the immersion in release medium. Drug particles are observed

on the polymer surface before immersion. The particles are quickly dissolved into the release

medium after immersion, leaving behind a vacant space with no polymer molecules. The

released drug from the surface layer accounts for the initial burst. The vacant space can also

connect with each other, forming channels which allow for fast drug diffusion into the release

medium (Spenlehauer, 1988). At a higher drug loading concentration, channels formed by drug

release are more closely connected which facilitates drug diffusion. Drug release during the

initial burst has less correlation with polymer degradation. Water molecules will diffuse into

the vacant space. Since water concentration determines the rate of hydrolytic degradation,

water diffusion into the polymer matrix is expected to affect polymer degradation.

Accordingly, drug release from a polymer system is complicated by multiple factors, including

polymer crystallinity, polymer mobility, and the vacant space left behind after drug release.

Investigation of the combined effects is prerequisite to better control the drug release profiles.

By considering these combined effects, the objective of this work is to investigate the effects of

drug loading on polymer degradation and drug release, as well as the modification of drug

release profiles through laser surface melting. Polymer degradation is characterized by the

molecular weight of polymer matrix measured from the gel permeation chromatography (GPC).

Effect of drug loading on polymer matrix is characterized by its crystallinity and thermal

properties from the wide-angle X-ray diffraction (WAXD) and differential scanning calorimetry

(DSC). The amount of drug release in a release medium is determined by spectrophotometry.

A numerical model is developed to capture polymer degradation and drug release processes.

119

5.2 Background

5.2.1 Effect of Additives on Polymer Mobility

The addition of small molecules, such as drugs, into polymer matrices can plasticize the material.

Theories explaining polymer plasticization include the lubricity theory (Kirkpatrick, 1940) and

the gel theory (Aiken et al., 1947), while a more precise and widely accepted explanation is

provided by the free volume theory (Flory, 1940). The free volume can be divided into two

fractions: the free oscillation volume, which accounts for molecule oscillations and increases

slightly as the temperature rises below glass transition temperature (Tg), and the free

torsion-oscillation volume, which increases greatly with the temperature above Tg, as the

molecules have enough energy to move, bend, or rotate (Ueberreiter & Kanig, 1952). By

providing additional free volume, the introduction of plasticizer molecules into the polymer

matrix reduces viscosity and thus Tg. As the free volume theory predicts, the introduction of

plasticizers leads to more flexibility and easy molecule movement. When a small quantity of

plasticizer is added, the free volume is increased. Therefore, a new redistribution of the

configurations is allowed, which increases the number and size of polymer crystallites as well as

the overall crystallinity (Marcilla & Beltran, 2004).

5.2.2 Biodegradation of Polyester

PLA is a biodegradable polyester which hydrolyzes in the human body, leading to chain scission

at ester bonds. The hydrolysis reaction is given as follows.

120

C C O C C O

H HO O

CH3 CH3

2OH C C

H O

CH3

OH HO C C O

H O

CH3

Hydrolysis causes chain scission, and produces shorter chains with carboxylic (-COOH) groups

and alcohol (-OH). The reaction is associated with an enthalpy of -0.80 kJ/g (Wadso &

Karlsson, 2013), leading to a more thermodynamically stable state. The thermodynamic

stability provides the driving force of PLA hydrolysis. Hydrolysis rate is proportional to the

molar concentrations of water molecules and ester bonds (Pitt & Gu, 1987), given as

dCe

dt=-k1CeCw 2

where k1 is a rate constant. Carboxylic end groups generated from this reaction have a high

degree of dissociation and can act as a catalyst to accelerate the hydrolysis. Hydrolysis of

polyesters may become autocatalytic if carboxylic end groups remain in the bulk. During

autocatalyzed hydrolysis, the reaction rate depends on the concentration of the carboxylic end

groups, CCOOH, as well. The rate of autocatalyzed hydrolysis is given by (Lyu et al., 2007)

dCe

dt=-k2CeCw CCOOH

n (3)

where k2 is the rate constant for the autocatalysis reaction, and n accounts for the dissociation of

the carboxylic groups.

5.2.3 Drug Release from Biodegradable Polymers

Drug can be released from a polymer delivery system as controlled by both drug diffusion and

polymer degradation. Diffusion occurs when drug passes through the polymer matrix into the

surrounding release medium. A purely diffusion controlled drug release from polymer is first

(1)

121

quantitatively considered in the Higuchi’s model (Higuchi, 1961). For the past decades, the

importance of polymer degradation on drug release has drawn increasing attention, and models

accounting for polymer degradation have been proposed to explain physical phenomena

observed experimentally. Based on model prediction and experimental data, drug release

profile is composed of up to four phases: initial burst, induction period, degradation controlled

release, and terminal release phase (Rothstein et al., 2008). Upon immersion into the release

medium, the polymer matrix begins to be hydrated by the surrounding liquid environment,

leading to the initial burst, phase 1. After phase 1, drug release is then controlled by polymer

degradation. Before polymer chains degrade into chain segments with molecular weights small

enough, drugs are held in the polymer matrix and cannot release, which accounts for the

induction period as demonstrated in phase 2. Once chain segments with small enough

molecular weights are formed, drugs can be released, as observed in phase 3. By the end of the

release period, release occurs at a reduced rate, phase 4. The reduced rate in phase 4 is a result

of the longer diffusion path length of the drug located in the bulk center, which is released in the

late stage.

5.3 Numerical Model

Simulation is developed to investigate the effect of drug loading and laser treatment on polymer

degradation and drug release profiles. Laser energy absorbed by semi-transparent bulk PLLA

generates heat and is governed by the heat equation

ρCp T∂T

∂t= · k T +q z,t

∂Hm

∂t (4)

where ρ is mass density, Cp T is specific heat as a function of temperature T, and k is

thermal conductivity. During laser heating, polymer experiences glass transition and melting.

122

Polymer glass transition is a second order transition, in which specific heat changes while no

latent heat is involved. Polymer melting is a first order transition, in which specific heat

changes and melting enthalpy is involved (Painter & Coleman, 1997). q(z,t) is the laser power

density as a function of depth z from the matrix surface and time t, expressed as

q z,t =Q0eαz+β

ttp

22

(5)

where Q0 is peak power density, α is absorption coefficient, tp is pulse width, and β is a

constant 4ln2.

The results of laser melting depth serve as the initial condition of the polymer degradation and

drug release model. The process of polymer degradation and drug release is captured with a

phenomenological model (Wang et al., 2008; Hsu et al., 2012b). During the polymer

degradation and drug release process, the PLLA matrix is modeled to be composed of 9 species:

non-degraded amorphous chains, degraded amorphous chains in stages 1, 2, and 3, crystalline

chains, degraded crystalline chains, monomers, water molecules, and drug molecules. Details

of the species of polymeric material generated during degradation process are given in Chap. 4

Degradation of amorphous chains and crystalline chains are considered separately. To consider

amorphous chain degradation, the concentration of the non-degraded amorphous chain is

expressed as

dC0

dt= 0C0Cw ε0C0CwCm

n κ0dCc

dt (6)

where C0 , Cc , and Cw are the molar concentrations of monomers in the non-degraded

amorphous chains, non-degraded crystalline chains, and water molecules, respectively. The

dissociation of the acid end groups n is assumed to be unity. Values of 0, ε0, and κ0 are the

123

corresponding phenomenological rate constants, accounting for non-autocatalysis, autocatalysis,

and crystallization due to hydrolysis of non-degraded amorphous chains. Degradation of

amorphous chains experiences three stages before monomers are generated (Stephens et al.,

2008), and thus the molar concentration of monomers in each degradation stage is expressed as

dCi

dt=(γi-1+εi-1Cm

n )Ci-1Cw (γi+εiCmn )CiCw κi

dCc

dt (7)

where i=1, 2, and 3, representing degradation stages 1, 2, and 3. Ci is the molar concentration

of the monomers in stage i. γi, ε , and κi are the phenomenological rate constants accounting

for non-autocatalysis, autocatalysis, and crystallization due to hydrolysis in stage i. Hydrolysis

of the stage 3 species generates monomers which do not connect to other monomers and have

high mobility to diffuse out of the polymer matrix. Assuming Fick’s second law for monomer

diffusion, which predicts the change of monomer concentration with space and time, the molar

concentration Cm of the monomers with high mobility is modeled by

dCm

dt= γ3+ε3Cm

n C3Cw+ ·(Dm,eff Cm) (8)

where Dm,eff is the effective monomer diffusivity as a function of matrix porosity induced by

degradation. The pores are defined as regions in a polymer matrix with a molecular weight low

enough to allow the embedded drug to release. Dm,eff is then expressed as

Dm,eff=Dmε(r,z,t) (9)

where Dm is the monomer diffusivity via pores, and ε(r,z,t) is the matrix porosity from 0 to 1.

Assuming the molecular weights which form the pores follow a normal distribution, ε(r,z,t) is

given as (Rothstein et al., 2009)

ε(r,z,t)=1-1

2erf

Mw(r,z,t)-Mwp

√2σ2+1 (10)

124

where σ2 accounts for the variation of degradation. Mw(r,z,t) is the molecular weight at

location (r,z) and time t. Mwp is the average molecular weight at which pores start to form,

allowing the diffusion of small molecules such as monomers and drug molecules.

Degradation of crystalline chains is assumed to be a one-step process, in which chain scission

only occurs on the fold surfaces of lamellae, generating the crystalline region composed of the

integral folds the crystalline chains. Crystal degradation is thus given as

dCc

dt= κ0+κi

dCc

dtγcCcCw εcCcCwCm

n (11)

where κ0+κi , γc , and εc are the phenomenological rate constants accounting for

crystallization of amorphous chains during their degradation, non-autocatalysis, and

autocatalysis of crystalline chains, respectively.

Drug concentration within a matrix during polymer degradation is as a function of space (r,z) and

time t, and is calculated from Fick’s second law as (Saltzman & Langer, 1989)

∂Cd(r,z,t)

∂t= (Dd,eff Cd) (12)

where Cd is the concentration of drug molecules, Dd,eff is the effective diffusivity accounting

for the porosity during polymer degradation and drug release period.

Drug molecules located in the layer below polymer matrix surface are subjected to initial burst

because of the contact with the release medium. Due to the connection of vacant space left

behind by the released drug, the layer accounting for initial burst is thicker for higher drug

loading concentration. It is assumed in the simulation that in the layer from the matrix surface

S to a depth of S‐dib, drug is subject to initial burst and the matrix porosity ε is unity. In the

bulk from matrix center to S‐dib, drug release is a function of porosity generated by polymer

125

degradation, such that

Dd,eff=Ddε(r,z,t) (13)

where Dd is drug diffusivity via pores, and ε(r,z,t) is the matrix porosity from 0 to 1 as

expressed in Eq. (10). Due to fast water diffusion into PLLA matrix as compared to the slow

PLLA hydrolysis kinetics (Lyu & Untereker, 2009), water concentration, Cw, is assumed to be

saturated over the degradation and drug release period, and loss of monomer and drug is

replaced by water molecules in simulation.

Eqs. (4) and (5) are solved to determine the laser melting depth. Solutions are used as the initial

conditions to solve the coupled Eqs. (6) to (11) to capture degradation profiles. Release of

embedded drug is simulated with Eqs. (12) and (13). Appropriate units are used for the

phenomenological rate constants so that concentration change is expressed in mole per volume

per time. The values of the rate constants are selected to capture experimental results. The

equations are solved through the finite element method in COMSOL Multiphysics 4.1. A 2D

axisymmetric model is used for the matrix with a diameter of 10 mm and thickness of 1 mm

experimentally tested. In the spatial domain, the 1 mm by 5 mm film is immersed in a 10 mm

by 10 mm aqueous medium. Laser treated area covers both sides of the film domain. The

film domain is initially composed of drug molecules, non-degraded crystalline and amorphous

chains with crystallinity determined experimentally. Monomers generated during degradation

can diffuse out of the film into the surrounding aqueous medium. The time domain in the

simulation corresponds to the experiment time span.

5.4 Materials and Methods

PLLA granules were provided by PURAC and used as received. Rhodamine B (RB) from

126

Sigma Aldrich is used as the model drug. To assure homogeneous drug loading, RB was loaded

into PLLA through solvent casting. 2 g PLLA granules were dissolved in 45 mL

dichloromethane from Sigma Aldrich by sonication for 1.5 hours. During sonication, 20, 100,

200, and 400 mg RB powder was dissolved in the PLLA solution to prepare 1 %, 5 %, 10 %, and

20 % drug loading concentrations, respectively. The drug loaded solution was cast in a covered

Petri dish, and left in a hood at 25°C for 72 hours for dichloromethane to vaporize. A solid

PLLA film with homogeneous drug loading is left behind. 100 mg of the film was thermally

compressed under 5.7×104 Pa at 185°C for 1 hour, and cooled down in air. The cooling process

lasts for around 2 hours before room temperature is reached, allowing for polymer crystallization.

The obtained PLLA matrix has 1 mm thickness and 10 mm diameter. Sample crystallinity and

thermal properties were determined by WAXD and DSC measurements. The WAXD system is

equipped with monochromatic CuKα radiation with wavelength λ=0.15418 nm at 40 kV and 30

mA. For DSC measurement, fragment of matrix (around 5 mg) was heated from 50 to 200 °C

at a rate of 5°C/min under a nitrogen gas flow. To study the effect of laser treatment on

shortening the drug release induction period, the 1 %, 5 %, and 10 % drug loaded matrices were

treated on both sides by a KrF excimer laser with a 248 nm wavelength, 25 ns pulse width, and

3.0 J/cm2 fluence (Hsu et al., 2012b).

Drug release tests were conducted such that each sample was placed in a vial and fully immersed

in a 10 mL phosphate buffered saline (PBS) with a pH of 7.4. The drug release period lasts for

up to 84 days. Vials were placed in a water bath at 37°C to simulate body temperature, and the

PBS was changed every 7 days. The amount of released drug in the PBS was determined by

spectrophotometry, in which the absorbance of RB at 552 nm was recorded and is a function of

RB concentration. After drug release, samples were rinsed with distill water and dried in

127

vacuum for two days. The PLLA matrices loaded with 5 % drug before and after different

periods of drug release are shown in Fig. 5.1. The matrix color becomes lighter with drug

release period, demonstrating that a larger amount of drug is release. To characterize polymer

degradation during the drug release period, the weight average molecular weight (Mw) and

number average molecular weight (Mn) were determined in chloroform from gel permeation

chromatography (GPC) at 30°C. The GPC was calibrated with polystyrene standards. The

drug loaded samples were purified to remove the RB before GPC measurements. For

purification, the drug loaded PLLA were dissolved in chloroform, and methanol was added to the

solution to precipitate PLLA. The mixture was separated by centrifugation to collect the

precipitated PLLA for GPC measurements. The PLLA/chloroform solution with a

concentration of 1.5 mg/mL was prepared for GPC measurements. The refractive index and

differential pressure detectors were used.

Figure 5.1: Non-laser treated PLLA matrices loaded with 5 % drug (a) before degradation and drug release, and degraded and released for (b) 35, (c) 49, (d) 70 days.

5.5 Results and Discussion

5.5.1 Effect of Drug Loading on Chain Mobility and Polymer Crystallinity

The addition of drug molecules in polymer chains may change the free volume between chains,

which affects chain mobility and crystallization. As stated in Sec. 5.4, polymer crystallization

(a) (b) (c) (d)

128

occurs during the cooling process after thermal molding. The cooling process lasts for around 2

hours. Crystallinity of the resulted polymer matrices is characterized by WAXD, with profiles

given in Fig. 5.2. The 16.7° WAXD crystalline peak becomes more prominent for the samples

loaded with higher concentrated drug. Crystallinity, calculated based on ref. (Alexander, 1969),

is given in Fig. 5.5, in which the crystallinity calculated from DSC results is also shown and will

be discussed later. Figure 5.5 shows that crystallinity increases with the increasing drug

concentration. The increased crystallinity is a result of high chain mobility allowed by the

addition of drug molecules. With a small molecular weight of 479.02 g/mol relatively to

polymer chains, drug molecules serve as plasticizers and render free volume within chains. As

the free volume theory predicts, the introduction of plasticizers facilitates chain movements by

increasing free volume, which tends to increase the number and size of crystallites since a new

redistribution of the configurations is allowed (Marcilla & Beltran, 2004). Therefore, during

the cooling process of thermal molding, PLLA loaded with a higher concentrated drug has higher

mobility and is ended up with higher crystallinity.

10 15 20 25 300

20

40

60

80

100(110)/(200)

Inte

nsity

(C

PS

)

2 (deg)

PLLA loaded with 20 % drug PLLA loaded with 10 % drug PLLA loaded with 5 % drug PLLA loaded with 1 % drug Pure PLLA

(203)

Figure 5.2: WAXD profiles of PLLA with different drug loading concentrations. Intensity of crystalline peaks increases with drug concentration, suggesting a higher crystallinity. Profiles

are shifted in y direction for viewing clarity.

129

The crystals developed in PLLA matrices during the cooling process of thermal molding may

also influence chain mobility. To investigate the combined effect of drug loading and polymer

crystallinity, the thermal properties of the molded matrices with different drug loading

concentrations are characterized by the DSC. The DSC thermograms of the PLLA matrices are

given in Fig. 5.3(a). The thermograms demonstrate that the PLLA matrices, during DSC

scanning, experience the glass transition at around 60-65 °C, cold crystallization at around

95-105°C, and melting at around 160-190 °C. The occurrence of cold crystallization is because

at a temperature higher than Tg, polymer chains gain extra mobility when compared to their

initial status below Tg, which allows chain crystallization to achieve a more energetically stable

state. A weak exotherm is demonstrated at around 160 °C, which is slightly lower than the

onset temperature of melting. The exotherm is generated because a part of polymer chains start

to melt at that temperature, which increases chain mobility and induces the crystallization. The

melting peak then begins right after the exotherm. The initial crystal and the crystal developed

during the two crystallization processes of DSC scanning are melted, generating the melting

peak.

60 80 100 120 140 160 180 2000.0

0.5

1.0

1.5

2.0

2.5

Hea

t flo

w (

mW

/mg)

Temperature (oC)

Pure PLLA

PLLA loaded with 5 % drug PLLA loaded with 10 % drug PLLA loaded with 20 % drug

Exo.

PLLA loaded with 1 % drug

Endo.

(a)

130

56 58 60 62 64 66 68 70

0.3

0.4

0.5

0.6

0.7

Hea

t F

low

(m

W/m

g)

Temperature (oC)

Pure PLLA PLLA loaded with 1 % drug PLLA loaded with 5 % drug

Endo.

Exo.

PLLA loaded with 10 % drug PLLA loaded with 20 % drug

(b)

Figure 5.3: DSC thermograms of pure PLLA and PLLA loaded with 1 %, 5 %, 10 %, and 20 % drug (a) heating from 50 to 200 °C and (b) around the glass transition temperature. The heating

rate is 5 °C/min.

The thermograms around Tg is given in Fig. 5.3(b). It is demonstrated that both Tg and Tm shift

with drug concentrations. Tg and Tm are plotted with drug concentration in Fig. 5.4. When the

drug concentration is low (below 5 %), Tg and Tm reduce with an increasing drug concentration.

The reduction of Tg and Tm suggests increased chain mobility, because of the plasticization effect

induced by drug small molecules. However, at high drug concentration (above 5 %), an

opposite trend is observed. Tg and Tm increase with an increasing drug concentration,

suggesting reduced chain mobility. The reduced chain mobility is a result of high crystallinity

in the samples with high drug concentration. High crystallinity limits chain movement and thus

increases Tg and Tm.

In addition to the WAXD, crystallinity of drug loaded samples is also derived from the DSC.

Based on the DSC results, the crystallinity ϕ of the PLLA matrices is evaluated according to the

following equation.

131

ϕ (∆Hm+∆Hc)/∆Hm0 (14)

where ∆Hm is the melting enthalpy represented by the melting peak shown in Fig. 5.3, ∆Hc is

the total crystallization enthalpy represented by the two crystallization exotherms in Fig. 5.3, and

∆Hm0 is the melting enthalpy of PLLA crystals with infinite crystal thickness, 93 J/g (Fischer et

al., 1973). The calculated crystallinity is given in Fig. 5.5. Crystallinity increases with drug

concentration, which agrees with the crystallinity derived from the WAXD results.

0 5 10 15 2060

61

62

63

64

65

Mel

ting

tem

per

atu

re (

o C)

Gla

ss t

ran

sitio

n te

mpe

ratu

re (

o C)

Drug concentration (%)

Glass transition temperature

174

175

176

177

178

179

Melting temperature

Figure 5.4: Glass transition temperature and melting temperature of drug loaded PLLA matrices

as a function of drug concentration.

0 5 10 15 2025

30

35

40

45

50

Cry

sta

llin

ity (

%)

Drug concentration (%)

Crystallinity derived from WAXD Crystallinity derived from DSC

Figure 5.5: Crystallinity as a function of drug loading concentration obtained from WAXD and

DSC. Crystallinity increases with drug concentration based on both measurements. The error bar represents the standard deviation of 3 data points.

132

5.5.2 Effect of Drug Concentration on Polymer Degradation and Drug Release

Drug loading affects polymer degradation and can affect resulted drug release profile over time.

The effect of drug loading concentration on polymer degradation and drug release profiles are

investigated. Crystallinity of PLLA matrix during the degradation and drug release period is

also recorded and discussed.

5.5.2.1 Polymer Degradation during Drug Release

The effect of drug loading concentration on polymer degradation is characterized by the

molecular weight through the GPC. GPC profiles of non-laser treated samples are given in Fig.

5.6 for pure PLLA and PLLA loaded with 1 % drug. The profiles give the information of

molecular weight distributions. Minor change of GPC profile is observed for the non-drug

loaded PLLA, suggesting insignificant degradation over time. On the other hand, the GPC

profiles of drug loaded PLLA shift left, demonstrating a decrease of molecular weight as a result

of degradation. Bimodal distributions can be observed at the late stage of degradation for the

drug loaded matrices. The occurrence of the bimodal distribution is because of the preferential

chain scission on the lamella fold surfaces, which generates integral folds of crystalline chains

and therefore a peak located at a lower molecular weight. The bimodal distribution is also

observed for non-drug loaded PLLA at a late stage under an accelerated degradation condition

previously reported in literature (Hsu et al., 2012b). For 5 %, 10 %, and 20 % drug loaded

PLLA, the GPC profiles also shift left with time and the bimodal distributions are observed.

The profiles shift left at a higher rate with a higher drug concentration, suggesting a faster

degradation.

133

0 1 2 3 4 5 6 70

10

20

30

40

Inte

nsity

(m

V)

Log Molecular Weight

Before degradation

Non-laser treated PLLA, no drug loading

Degraded for 84 days

(a)

0 1 2 3 4 5 6 70

10

20

30

40

Inte

nsi

ty (

mV

)

Log Molecular Weight

Before degradationNon-laser treated PLLA, loaded with 1 % drug

Degraded for 7 days Degraded for 21 days Degraded for 35 days Degraded for 56 days Degraded for 70 days

(b)

Figure 5.6: GPC profiles of the non-laser treated PLLA matrices loaded with (a) 0 % and (b) 1 % drug as a function of degradation and drug release period.

The Mw and Mn determined from GPC, as well as the Mn obtained from simulation, are given in

Fig. 5.7. The molecular weights decrease insignificantly for the non-laser treated PLLA with

no drug loading, which agrees with Fig. 5.6(a) in which similar GPC profiles are observed before

and after degradation. When the matrix is loaded with drug, the degradation rate is accelerated.

The accelerated degradation can be a result of the enlarged free volume by drug small molecules,

which favors water diffusion into the matrix and hydrolysis. It is also observed that 10 % and

134

20 % drug loaded PLLA significantly degrades in the first three days, which corresponds to the

initial burst period, as discussed in Sec. 5.5.2.2. This suggests that, in addition to enlarged free

volume by drug, polymer degradation is also accelerated by the initial burst occurring on the first

several days. As a result of the initial burst, the drug on the polymer matrix surface is released.

The space initially occupied by drug molecules becomes vacant. The vacant space favors water

diffusion into the matrix and accelerates hydrolysis in the first three days, Fig. 5.7(d) and (e).

The accelerated hydrolysis generates a structure with higher porosity, and favors subsequent

water diffusion and hydrolysis. The degradation profiles are captured by the model, in which

the molecular weight experiences a rapid initial drop followed by decrease at a slow rate. In the

late stage of degradation, the simulation curve of molecular weight tends to level off, while the

experimental data demonstrate a more significant decrease. A possible reason is that, due to

significant amount of drug release in the late stage, polymer matrix starts to disintegrate, which

is experimentally observed as cleavage and holes on the degraded samples. The effect of such

macro-scale morphology changes are not considered numerically, while further favor hydrolysis

and drug release.

0 10 20 30 40 50 60 70 80 900

20000

40000

60000

80000

100000

120000

140000

160000

Mol

ecul

ar w

eig

ht (

g/m

ol)

Polymer degradation and drug release period (day)

Mw, non-laser treated, experiment Mn, non-laser treated, experiment

Mw, laser treated, experiment Mn, laser treated, experiment

Mn, non-laser treated, simulation

Mn, laser treated, simulation

(a)

135

0 10 20 30 40 50 60 70 80 900

20000

40000

60000

80000

100000

120000

140000

160000

Polymer degradation and drug release period (day)

Mol

ecul

ar w

eigh

t (g

/mol

)

Mw, non-laser treated, experiment Mn, non-laser treated, experiment

Mw, laser treated, experiment Mn, laser treated, experiment

Mn, non-laser treated, simulation

Mn, laser treated, simulation

(b)

0 10 20 30 40 50 60 70 80 900

20000

40000

60000

80000

100000

120000

140000

160000

Mol

ecul

ar w

eigh

t (g

/mol

)

Polymer degradation and drug release period (day)

Mw, non-laser treated, experiment Mn, non-laser treated, experiment

Mw, laser treated, experiment Mn, laser treated, experiment

Mn, non-laser treated, simulation

Mn, laser treated, simulation

(c)

0 10 20 30 40 50 60 70 80 900

20000

40000

60000

80000

100000

120000

140000

160000

Mol

ecul

ar w

eigh

t (g

/mol

)

Polymer degradation and drug release period (day)

Mw, non-laser treated, experiment Mn, non-laser treated, experiment

Mw, laser treated, experiment Mn, laser treated, experiment

Mn, non-laser treated, simulation

Mn, non-laser treated, simulation

(d)

136

0 10 20 30 40 50 60 70 80 900

20000

40000

60000

80000

100000

120000

140000

160000

Mol

ecul

ar w

eigh

t (g

/mol

)

Polymer degradation and drug release period (day)

Mw, non-laser treated, experiment Mn, non-laser treated, experiment Mn, non-laser treated, simulation

(e)

Figure 5.7: Number average and weight average molecular weights of (a) pure PLLA and PLLA loaded with (b) 1 % (c) 5 % (d) 10 % (e) 20 % drug with polymer degradation and drug release

period. For pure PLLA and PLLA loaded with 1 %, 5 %, and 10 % drug, effect of laser melting on degradation is also studied.

It is observed that polymer degradation accelerates with drug concentration, which is not a

function of chain mobility as revealed in Fig. 5.4. Chain mobility therefore plays a less

important role in determining polymer degradation under the current test conditions. Instead,

due to the enlarged free volume and the vacant space after drug release, drug release is a strong

function of drug loading concentration.

5.5.2.2 Drug Release

Drug release from PLLA matrix into the release medium is measured by spectrophotometry.

The model drug, RB, has a characteristic absorbance peak at 552 nm. The absorbance peak

height is a function of RB concentration. To determine the relationship between the peak height

and concentration, the RB/PBS solution with different concentrations are prepared and measured

by spectrophotometry. The absorbance profiles as a function of solution concentration is given

in Fig. 5.8(a). A linear relationship between absorbance at 552 nm and RB concentration in

137

PBS is described by A=0.058cRB-0.017, where A is the RB absorbance at 552 nm and cRB is

the RB concentration in μmol/L.

460 480 500 520 540 560 580 6000.0

0.5

1.0

1.5

2.0

2.5

Abs

orba

nce

Wavelength (nm)

47.8 mol/L 20.9 mol/L 10.4 mol/L 5.2 mol/L 2.6 mol/L 1.3 mol/L

(a)

0 5 10 15 20 25 30 35 40 450.0

0.5

1.0

1.5

2.0

2.5

Abs

orba

nce

Drug concentration (mol/L)

y=0.058x-0.017

R2=0.9998

Figure 5.8: (a) Absorbance spectra of the drug/PBS solutions with different drug concentrations.

(b) Linear relationship between absorbance at 552 nm and drug concentration in PBS. Rhodamine B is used as the model drug.

The drug release profiles are given in Fig. 5.9, with the y-axis the amount of released drug as a

percentage of the initial drug loaded in the matrix. Simulation results are also provided. A

higher release rate is observed in matrix loaded with a higher concentrated drug, because a

138

higher drug concentration accelerates water penetration and polymer degradation, as discussed in

Sec. 5.5.2.1. Different drug release profiles are observed in PLLA loaded with low (1 %),

medium (5 % and 10 %), and high (20 %) drug concentrations. For low drug concentration,

drug release begins with an induction period during which drug release is controlled by polymer

degradation. Drug is not released before polymer molecules degrade into small enough chain

segments.

0 10 20 30 40 50 60 70 800

20

40

60

80

100

Am

ount

of

rele

ased

dru

g (%

)

Polymer degradation and drug release period (day)

20 % drug, experiment 10 % drug, experiment 5 % drug, experiment 1 % drug, experiment

20 % drug, simulation 10 % drug, simulation 5 % drug, simulation 1 % drug, simulation

Figure 5.9: Drug release profiles of PLLA with different drug loading concentrations.

For medium drug concentration, the release profiles begin with an initial burst, followed by an

induction period. The initial burst is a result of immediate release of drug molecules located on

the matrix surface and exposed to the release medium directly. Drug release during initial burst

is independent from polymer degradation. Once the initial burst period comes to the end, drug

release is controlled by polymer degradation, demonstrating the induction period. The initial

burst is more significant and the induction period is shorter as the drug concentration increases.

At 20 % drug concentration, no induction period is observed. With a higher drug concentration,

the space occupied by the drug molecules has higher opportunity to connect with each other,

139

forming channels which allow for fast drug release into the medium and water diffusion into the

matrix. Therefore, the degradation is accelerated, leading to a short induction period. The

higher degradation rate causes a higher drug release rate as shown in Fig. 5.9.

(a)

(b)

Figure 5.10: Simulated spatial distribution of drug concentration in non-laser treated PLLA loaded with 5 % drug after release for (a) 30 and (b) 70 days. Drug concentration is represented

as a percentage of initial value.

The model captures physical phenomena obtained experimentally, as initial burst, induction

period, and subsequent release. Deviation between simulation and experiment, however, is

observed in the late stage of release period. In simulation, the drug release rate slows down at

the end. This stems from the different diffusion time for drug located near the surface and in

the center. As given in Fig. 5.10, the spatial drug distribution is simulated for the PLLA with 5

140

% drug right after the end of induction period (30 days) and in the end of drug release (70 days).

Drug release starts on the matrix surface. Because of the shorter path length to the surface,

drug diffusion requires shorter time. The drug molecules located in the matrix center, on the

other hand, are released in the late stage. Due to the longer path length of diffusion to the

surface, they are released at a slower rate. Experimentally, however, drug release in the late

stage can be accelerated by matrix disintegration as discussed in Sec. 5.5.2.1. The cleavage and

holes generated during this period enlarge the matrix contact area with release medium and

accelerate drug release, which accounts for the deviation between simulation and experiment in

the late stage.

5.5.2.3 Polymer Crystallization during Drug Release

Crystallization during drug release period is monitored using the WAXD. Crystallinity of

non-laser treated PLLA matrices with different drug concentrations is given in Fig. 5.11.

Crystallinity increases with polymer degradation and drug release period. Crystallization is a

result of polymer chain degradation, because degraded chains have smaller molecular weights

and thus a higher mobility. With enough mobility, the degraded chains reorganize into the

crystalline state, which is energetically stable. For pure PLLA, crystallinity does not increase

over the test period, because there is no significant degradation as shown in Figs. 5.6(a) and

5.7(a). Therefore, chain mobility in the pure PLLA is not high enough for crystallization. It is

noticed that for the drug loaded matrices, crystallinity increases with a pattern similar to the drug

release profiles as shown in Fig. 5.9. Crystallinity increases after an induction period, within

which crystallinity increase is limited. The induction period is shorter for PLLA loaded with

higher concentrated drug. The similarities between drug release profiles, Fig. 5.9, and

crystallinity changes, Fig. 5.11, demonstrate that both drug release and crystallization are a result

141

of polymer degradation. Polymer matrix with a reduced molecular weight has larger free

volume and higher chain mobility, which favor drug release and chain crystallization.

0 10 20 30 40 50 60 70 80

20

30

40

50

60

70

Cry

stal

linity

(%

)

Polymer degradation and drug release period (day)

20 % drug in non-laser treated PLLA 10 % drug in non-laser treated PLLA 5 % drug in non-laser treated PLLA 1 % drug in non-laser treated PLLA 0 % drug in non-laser treated PLLA

Figure 5.11: Crystallinity of non-laser treated PLLA loaded with 1 %, 5 %, 10 %, and 20 % drug

as a function of polymer degradation and drug release period.

Water can hardly penetrate into polymer crystalline region, and thus higher crystallinity can

reduce polymer degradation and retard drug release. However, based on Figs. 5.9 and 5.11,

polymer degradation and drug release do not slow down even the crystallinity increases. This is

because drug release and polymer degradation left behind a porous structure. The pores

accelerate water diffusion and hydrolytic degradation, even if the crystallinity increases. The

effect of higher crystallinity to slow down degradation and drug release is therefore canceled out

and does not dominate during the process.

It is also noticed that the crystallinity of drug loaded PLLA drops in the first three days of drug

release period. The drop may be due to the significant release of drug molecules as a result of

initial burst. A portion of drug molecules was initially trapped in the crystalline phase and held

142

the crystalline chains through the intermolecular force. The loss of drug molecules reduces the

intermolecular force between crystalline chains, and leads to less ordered chain packing.

5.5.3 Laser Modification of Polymer Degradation and Drug Release Profile

It has been shown in the previous section that at low drug concentrations (1 %, 5 % and 10 %),

drug release profile demonstrates an induction period during which the amount of drug release is

limited. The induction period is undesirable in drug delivery applications. During the

induction period, drug release is prohibited by the slow polymer degradation kinetics. Polymer

degradation is a function of crystallinity. Laser treatment is conducted to reduce polymer

surface crystallinity. The effect of laser treatment on polymer degradation and drug release is

investigated. Crystallinity of the laser treated PLLA matrix during degradation and drug release

period is also monitored and discussed.

5.5.3.1 Laser Modification of Polymer Crystallinity

Modification of polymer crystallinity by laser treatment is studied by WAXD with results shown

in Fig. 5.12. The crystalline peak is located at 16.7°. The height of crystalline peak reduces

for the laser treated samples, suggesting a reduced crystallinity due to laser treatment. The

calculated crystallinity based on ref. (Alexander, 1969) is given in Fig. 5.13 as a function of drug

concentration. As agreed with Fig. 5.12, crystallinity decreases after laser treatment. The

reduced crystallinity is a result of fast melting kinetics as compared to slow crystallization

kinetics of polymer. It is noticed that crystallinity decreases at a smaller extent for matrices

with higher drug concentrations. This can be due to the reason that the drug molecules have an

absorption peak near the laser wavelength, 248 nm. Part of laser energy is absorbed by the drug

molecules, and therefore less energy can penetrate into the matrix and melt the polymer.

143

10 15 20 25 300

20

40

60

80

100

(203)

(110)/(200)

Inte

nsity

(C

PS

)

2 (deg)

PLLA with 10 % drug, non-laser treated PLLA with 10 % drug, laser treated PLLA with 5 % drug, non-laser treated PLLA with 5 % drug, laser treated PLLA with 1 % drug, non-laser treated PLLA with 1 % drug, laser treated Pure PLLA, non-laser treated Pure PLLA, laser treated

Figure 5.12: WAXD profiles of PLLA before and after laser treatment. Intensity of crystalline peaks decreases after laser treatment, suggesting a reduced crystallinity. Profiles are shifted in

y direction for viewing clarity.

0 2 4 6 8 105

10

15

20

25

30

35

40

45

Cry

stal

linity

abs

olu

te d

ecre

ase

afte

r la

ser

tre

atm

ent

(%)

Cry

stal

linity

pe

rcen

tage

de

crea

seaf

ter

lase

r tr

eatm

ent

(%)

Cry

stal

linity

(%

)

Drug concentration (%)

before laser treatment after laser treatment

10

20

30

40

50

60

70 crystallinity percentage decrease

4

6

8

10

12

14

16 crystallinity absolute decrease

Figure 5.13: Crystallinity of 1 %, 5 %, and 10 % drug loaded PLLA before and after laser

treatment. Laser treatment reduces crystallinity. Crystallinity decreases at a smaller extent for higher drug concentrated PLLA.

Before investigating the effect of laser treatment on polymer degradation and drug release,

chemical structure of laser treated matrices needs consideration, because chemical modifications

144

may alter the biodegradable and biocompatible properties of PLLA and the effectiveness of

drugs. The effect of laser irradiation on potential chemical modifications of polymer matrices

has been considered previously (Hsu et al., 2012a). The laser energy level used in the current

study, 3.0 J/cm2, induces non-measurable chemical modification on PLLA matrices. The

chemical effect of laser treatment on drug molecules is studied through the absorption by

spectrophotometry, since the absorbance spectrum is the characteristic of a specific chemical

structure (Wilhelm & Stephan, 2007). Drug loaded PLLA films with a thickness similar to

laser melting depth (around 100 μm) are formed through the same thermal molding process as

stated in Sec. 5.4. The samples, before and after laser treatment, were analyzed by

spectrophotometry. No modification on absorbance peak height is observed for the laser treated

samples, which suggests that under the current conditions laser treatment does not cause

chemical modification on the drug.

5.5.3.2 Polymer Degradation during Drug Release

Degradation of laser treated matrices is characterized through molecular weight measurement by

GPC. The GPC profiles of laser treated PLLA are given in Fig. 5.14 for the pure PLLA and 1

% drug loaded PLLA. In both cases, GPC profiles shift left, suggesting the occurrence of

degradation. The comparison between the GPC profiles of the non-laser treated pure PLLA,

Fig. 5.6(a), and the laser treated pure PLLA, Fig. 5.14(a), demonstrates the effect of laser

crystallinity modification on polymer degradation. A reduced crystallinity by laser melting

accelerates polymer degradation. For the laser treated 1 % drug loaded PLLA, Fig. 5.14(b), the

GPC profiles shift left within a shorter period as compared to the non-laser treated 1 % drug

loaded PLLA, Fig. 5.6(a), suggesting an accelerated degradation. The accelerated shift of GPC

profiles is also observed for laser treated PLLA loaded with 5 % and 10 % drug.

145

0 1 2 3 4 5 6 70

10

20

30

40

Laser treated PLLA, no drug loading

Inte

nsity

(m

V)

Log Molecular Weight

Before degradation Degraded for 14 days Degraded for 84 days

(a)

0 1 2 3 4 5 6 70

10

20

30

40 Laser treated PLLA, loaded with 1 % drug

Inte

nsity

(m

V)

Log Molecular Weight

Before degradation Degraded for 7 days Degraded for 21 days Degraded for 28 days Degraded for 56 days Degraded for 63 days Degraded for 70 days

(b)

Figure 5.14: GPC profiles of the laser treated PLLA matrices loaded with (a) 0 % and (b) 1 % drug as a function of degradation and drug release period.

Based on the GPC measurements, the Mw and Mn are calculated with the results given in Fig. 5.7,

in which simulated Mn is also given. An accelerated degradation induced by laser melting is

demonstrated for all laser treated PLLA matrices with and without drug loading. As clearly

shown in Fig. 5.7, laser treatment accelerates the initial degradation of polymer matrices,

because laser melted material, with a lower crystallinity, degrades at a higher rate. Once the

degradation of laser melted material comes to the end, the matrix can be seen as composed of a

structure similar to the non-laser treated matrix. At this later stage, therefore, the laser treated

146

matrix degrades at a rate similar to the non-laser treated matrix.

5.5.3.3 Drug Release

Drug release from laser treated PLLA is investigated with results given in Fig. 5.15, in which the

non-laser treated PLLA and simulation results are also provided for comparison. The induction

period is defined as a period after which the drug release rate increases. It is demonstrated that

the induction periods of laser treated matrices have been shortened. For 1 % drug loaded PLLA,

the induction period is shortened by around two weeks, from day 63 to day 49. For 5 % drug

loaded PLLA, the induction period is shortened by around one week, from day 28 to day 21.

For 10 % drug loaded PLLA, the induction period is shortened by around one week, from day 21

to day 14, while the difference of drug release on day 21 is insignificant. The shortened

induction period is because of the accelerated initial degradation induced by laser melting. The

time for the molecular weight to reduce to the critical value for drug to be released thus reduces,

which in turn shortens the induction period of drug release. The induction period shortens at a

smaller extent for matrix loaded with a higher drug concentration (10 %), which is because drug

molecules absorbs laser energy and reduces the portion of laser energy to melt polymer, as

demonstrated and discussed in Fig. 5.13. The shortening of induction period is also captured by

the model, while the deviation from the experiment results is observed. The possible reason is

discussed in Sec. 5.5.2.2 and applicable to the laser treated cases.

It has been noticed that both drug loading and laser surface melting accelerates polymer

degradation. However, the accelerated degradation by drug loading and laser surface melting

shows different profiles. Drug loading accelerates overall degradation of the bulk, and gives a

higher degradation rate until the end of degradation. Effect of laser melting, on the other hand,

147

is limited within a layer below matrix surface, and keeps the bulk intact. Laser melting

therefore only accelerates the initial degradation, while the degradation in the later stage remains

unchanged. Therefore, drug loading and laser treatment lead to distinct drug release profiles.

Drug loading accelerates overall degradation, and thus results in a shorter induction period of

drug release and a higher drug release rate. Laser surface melting accelerates initial

degradation and thus shortens the induction period of drug release. However, laser melting

keeps the subsequent drug release similar to the non-laser treated sample, which is desired since

drug release rate needs to be maintained within a specific range for the drug to be non-toxic and

effective.

0 10 20 30 40 50 60 70 800

10

20

30

40

50

60

70

80

Am

oun

t of

rel

ease

d dr

ug (

%)

Polymer degradation and drug release period (day)

10 % non-laser, experiment 5 %, non-laser, experiment 1 %, non-laser, experiment 10 %, laser, experiment 5 %, laser, experiment 1 %, laser, experiment 10 %, laser, simulation

5 %, laser, simulation 1 %, laser, simulation

Figure 5.15: Modification of drug release profiles through laser melting. The induction period

of drug release is shortened due to laser treatment.

5.5.3.4 Polymer Crystallization during Drug Release

Crystallinity of laser treated PLLA over the degradation and drug release period is monitored

using the WAXD, with results given in Fig. 5.16. The crystallinity increases with a similar

148

pattern of the drug release profiles as shown in Fig. 5.15. The similarity between crystallinity

change and drug release profiles is also observed for the non-laser treated PLLA, as given in Figs.

5.9 and 5.11. The similarity is again attributed to the fact that drug release and crystallinity

increase are both determined by polymer degradation, as discussed in Sec. 5.5.2.3. It is also

noticed that, for the laser treated PLLA, the induction periods of crystallinity change are shorter

than those of the non-laser treated PLLA. In addition, the crystallinity of the laser treated pure

PLLA increases at the late stage of drug release, as oppose to a constant crystallinity of the

non-laser treated pure PLLA over the testing period. Both phenomena are a result of

accelerated polymer degradation by laser melting.

0 10 20 30 40 50 60 70 80

20

30

40

50

60

70

Cry

stal

linity

(%

)

Polymer degradation and drug release period (day)

10 % drug in laser-treated PLLA 5 % drug in laser-treated PLLA 1 % drug in laser-treated PLLA 0 % drug in laser-treated PLLA

Figure 5.16: Crystallinity of laser treated PLLA loaded with 1 %, 5 %, and 10 % drug as a

function of polymer degradation and drug release period.

5.6 Conclusions

The effects of drug loading concentration and laser surface melting on PLLA biodegradation and

drug release have been investigated. It has been shown that PLLA biodegradation is a strong

149

function of drug loading, with a higher drug concentration leading to faster degradation. The

accelerated biodegradation caused by drug loading is attributed to the porous structure in the

polymer matrix after drug release. The porous structure favors water diffusion into the matrix

and accelerates hydrolytic degradation. The accelerated biodegradation reduces the induction

period of drug release and increases the drug release rate. Drug loading also influences chain

crystallinity and mobility, while in the current study both factors do not dominantly determine

PLLA biodegradation and drug release.

Laser melting reduces surface crystallinity of PLLA matrix, which accelerates polymer

degradation in the early stage and shortens the induction period of drug release. Laser

treatment only melts a layer below matrix surface and keeps the bulk intact. Therefore, after

laser melted material degrades, polymer degradation proceeds at a similar rate as the non-laser

treated samples. Similar polymer degradation rate results in similar drug release rate after laser

treatment. Accordingly, laser crystallinity modification has been shown to reduce the induction

period of drug release, while keep the drug release rate unmodified, which is a desired property

in drug delivery applications.

150

Chapter 6: Conclusions

In this study, the knowledge of morphology and crystal structure of poly(L-Lactic Acid) (PLLA)

as a biodegradable polymer has been advanced. Laser treatment has been conducted to reduce

surface crystallinity of biodegradable polymer. The reduced crystallinity has modified polymer

degradation profile, and successfully shortened the induction period of drug release. Major

findings and contributions of this study are summarized as follows. Suggestions for future

work are also listed.

6.1 Effect of Sample Formation Process on Morphology and Crystal Structure of

Biodegradable Polymer

It has been systematically demonstrated that PLLA morphology and crystal structure depend on

sample formation processes. Sample formation processes determine polymer chain mobility

and crystallization time, which are the important factors affecting the development of polymer

crystals. Solvent casting and spin coating, as well as subsequent annealing, have been

conducted for the detailed investigation of the process effects and resulted crystal structures.

Different categories of spherulites have been observed on solvent cast and spin coated samples.

Development of different spherulites stems from different chain mobility during the spherulite

development process. In solvent casting, solvent is present and chain mobility is high. For the

annealed spin coated films, spherulites are formed in viscous state with low chain mobility.

The higher mobility in solvent casting also allows for a better chain packing as revealed by the

wide angle X-ray diffraction (WAXD) observations. Annealing increases crystallinity of both

solvent cast and spin coated samples. After annealing under the same conditions, the

crystallinity for the solvent cast sample is lower than that in the spin coated sample, showing the

151

possible existence of the rigid amorphous phase in the solvent cast sample.

The detailed investigation advances the knowledge of PLLA crystallization and the effect of

sample formation processes. In addition, PLLA crystallinity strongly determines its

biodegradation. The advanced knowledge of the process effect, therefore, contributes to the

application of PLLA as a biodegradable material in various medical devices such as stent,

scaffold, and drug delivery system.

6.2 Laser Crystallinity Modification of Biodegradable Polymer

It has been numerically and experimentally demonstrated that under laser irradiation with photon

energy higher than polymer bond energies, there exists a working window of laser energy levels

which reduce polymer crystallinity and incur insignificant chemical modifications. The

working window size has also been shown to be a function of polymer crystallinity. Higher

crystallinity leads to lower chain mobility. Mobility determines melting point and quantum

yield of chemical decomposition.

Annealing has been conducted to increase polymer crystallinity for subsequent laser treatment.

After laser melting, crystallinity decreases for both annealed and non-annealed samples based on

the WAXD observations. Because of higher chain mobility, the non-annealed samples have a

lower melting temperature, and its melting depth is slightly larger. In the current study,

however, the melting points of both types of samples are not significantly different to affect the

laser energy level to reduce the crystallinity. Both annealed and non-annealed samples require

the same laser energy level to reduce their crystallinity.

Under laser irradiation with the photon energy higher than polymer bond energies, molecule

152

dissociation will occur, generating free radicals. Based on the cage effect, the free radicals are

difficult to move apart as confined by surrounding molecules with lower mobility such as

crystals. As a result, a large portion of free radicals recombine and return to the initial state,

incurring inappreciable chemical changes with a small quantum yield, which is desire to keep the

biocompatibility and biodegradability of PLLA. The cage effect is also revealed by the fact that

the annealed sample, with lower mobility, requires higher laser energy to induce chemical

modifications. Therefore, the working window is larger for samples with higher initial

crystallinity.

Since biodegradation of PLLA is a strong function of its crystallinity, laser treatment can be

utilized to modify PLLA biodegradation. PLLA biodegradation also determines drug release

profiles in the drug delivery application. Therefore, laser crystallinity modification further

provides a potential method to tailor the drug release profiles.

6.3 Modification of Polymer Biodegradation by Laser Melting

Effect of laser crystallinity modification on polymer biodegradation has been investigated for the

first time. To characterize the degradation process, polymer molecular weight, morphology,

and sample mass have been monitored. It has been shown that laser treated samples experience

a faster initial reduction of molecular weight, which suggests the preferential degradation of laser

melted material due to the reduced crystallinity.

To study the degradation of laser melted layer, careful morphology observation has been

conducted. As degradation progresses, it is observed that the surface morphology of laser

treated sample becomes more similar to that of the non-laser treated sample. Namely, the laser

treated surface regions are progressively eroded. This shows that degradation preferentially

153

occurs in the laser melted layer, which allows for the accelerated initial degradation. In

addition, the degraded sample possesses a solid cross section, which suggests that autocatalytic

degradation is not dominant, as a result of easy diffusion of acidic monomers out of the sample.

The accelerated initial degradation facilitates the diffusion of the degraded products with smaller

molecular weights, which in turn shortens the time period before the occurrence of sample mass

loss. This demonstrates the potentiality of laser crystallinity modification to shorten the

induction period of drug release, which is desired in drug delivery applications.

6.4 Effect of Drug Loading on Degradation and Drug Release of Biodegradable Polymer

The addition of drug into polymer matrix modifies matrix properties in many ways, including

changing chain mobility, crystallinity, and porosity during drug release. The combined effects

determine polymer degradation and drug release profiles. In this study, detailed investigation of

these combined effects has been carried out as a function of drug loading concentration. Drug

concentration influences polymer crystallinity and chain mobility as characterized by WAXD and

differential scanning calorimetry. Crystallinity increases with drug concentration due to the

plasticization effect, which enhances chain mobility and favors crystallization. The developed

crystals, however, reduce chain mobility. Therefore, with increasing drug concentration,

polymer chain mobility increases before decreasing at an even higher concentration.

Distinct PLLA biodegradation and drug release profiles have been observed with different drug

concentrations. At low concentration, drug release begins with an induction period during

which drug release is hindered due to the slow PLLA degradation kinetics. Higher drug

concentration results in faster degradation and thus drug release. The accelerated degradation is

attributed to the porous structure in the polymer matrix after drug release rather to the polymer

154

mobility. The effect of porosity also cancels out the effect of crystallinity; the latter tends to

slow down polymer degradation and drug release, which is not observed in the current study.

Therefore, with higher porosity, higher drug loaded PLLA leads to more porous structure during

drug release. The porous structure favors water diffusion and accelerates hydrolysis, which in

turn reduces the induction period of drug release and increases the release rate.

6.5 Shortening of Induction Period of Drug Release by Laser Melting

The slow degradation kinetics of PLLA results in an induction period of drug release, during

which a limited amount of drug is released. The induction period is undesirable because it

delays drug release and effectiveness. In this study, laser crystallinity modification has been

conducted to tailor the drug release profile and induction period of drug release for the first time.

It has been shown that laser melting accelerates polymer degradation in the early stage and

shortens the induction period of drug release, due to the early degradation of laser melted

material. Laser treatment keeps the bulk intact. Therefore, polymer degradation rate remains

similar after laser melted material degrades. Accordingly, laser treatment reduces the induction

period of drug release, while keeps the drug release rate originally designed for a drug delivery

system.

Both drug loading and laser surface melting accelerates polymer degradation and drug release.

However, it has been demonstrated the effects of drug loading and laser melting on the

accelerated degradation and drug release profiles are distinct. Drug loading accelerates the

overall degradation of the matrix until the end of the drug release period. Effect of laser

melting, on the other hand, is limited within a layer below matrix surface and therefore only

accelerates the initial degradation, while the degradation in the late stage remains similar.

155

Therefore, drug loading and laser treatment lead to distinct drug release profiles. Drug loading

accelerates overall degradation, and results in a shorter induction period of drug release and a

higher drug release rate. Laser surface melting accelerates initial degradation, and thus shortens

the induction period of drug release while keeps the originally designed release rate, which is

desired in drug delivery applications.

6.6 Future Work

Effects of laser crystallinity modification of biodegradable polymer have been investigated.

Laser treatment reduces polymer crystallinity. The reduced crystallinity accelerates polymer

degradation and therefore shortens the induction period of drug release. The results open the

possibility to precisely tailor the drug release profiles from biodegradable polymers through laser

treatments. To achieve the objective, possible required research is given below.

This study has demonstrated that one layer of laser melted material successfully modifies

polymer biodegradation and shortens the induction period of drug release. Future investigation

could involve the development of polymer matrix having multiple laser melted layers with

different crystallinities. With the multi-layer structure, a more graduate crystallinity gradient

along matrix thickness can be realized, which potentially allows for tuning drug release profiles

in a more accurate manner. To develop the multi-layer structure, multiple laser irradiations,

with different operating parameters, on the polymer matrix surface may be required. Effects of

laser parameters such as repetition rate, pulse number, wavelength, and pulse energy on

crystallinity modification thus need further investigation, in order to achieve better control over

crystallinity along matrix thickness. Other irradiation sources such as flash lamps, with the

advantages of tunable wavelengths, will also be considered for further investigation.

156

Spatially resolved measurements of crystallinity along the specimen thickness direction are

required, especially for the multi-layer structure mentioned in the previous paragraph. Micro

X-ray diffraction (μXRD) method may be a potential characterization method. μXRD allows

for a spatial resolution of the order of micron, and can be investigated to measure the cross

sections of laser treated polymeric materials. Crystallinity as a function of thickness can be

obtained. To carry out the μXRD measurements and to interpret the results, the issue associated

with the X-ray beam size with respect to the polymer crystal structure size will be addressed to

ensure correct and accurate characterization.

In this study, degradation and drug release tests are conducted on matrices with a fixed geometry.

Geometry of matrices determines the contact area with the body fluid, as well as the diffusion

path lengths of degraded products and drug molecules. Effect of matrix geometry requires

further investigation. During drug release period, complex processes such as hydrolysis,

crystallization, and diffusion occur. To investigate the processes in detail, a numerical model

with better predictability is desired. Predictable model is also favorable to further investigate

laser processing to modify drug release, because long term experiments of drug release, on the

order of weeks or months, slow down the investigation and development processes.

Undesired initial burst has been observed on high drug concentrated matrices in this study. To

reduce initial burst, matrices can be covered by a layer with lower drug concentration. Laser

treatment can be applied on this covering layer to shorten the induction period. Through

adjusting drug spatial concentration within the matrix, and applying suitable laser treatment,

problems with initial burst and induction period can be simultaneously resolved.

157

References

Chapter 1

Alexander, L. E., 1969, X-Ray Diffraction Methods in Polymer Science, John Wiley & Sons, New York, Chap. 1.

Alexis, F., 2005, “Factors Affecting the Degradation and Drug-Release Mechanism of Poly(lactic acid) and Poly[(lactic acid)-co-(glycolic acid)],” Polym. Int., 54, pp. 36-46.

Arifin, D. Y., Lee, L. Y., and Wang, C.-H., 2006, “Mathematical Modeling and Simulation of Drug Release from Microspheres: Implications to Drug Delivery Systems,” Adv. Drug Delivery Rev., 58, pp. 1274-1325.

Avrami, M., 1939, “Kinetics of Phase Change. I. General Theory,” J. Chem. Phys., 7, pp. 1103-1112.

Avrami, M., 1940, “Kinetics of Phase Change. I. Transformation-Time Relations for Random Distribution of Nuclei,” J. Chem. Phys., 8, pp. 212-224.

Avrami, M., 1941, “Granulation, Phase Change, and Microstructure: Kinetics of Phase Change. III,” J. Chem. Phys., 9, pp. 177-184.

Belbella, A., Vauthier, C., Fessi, H., Devissaguet, J.-P., Puisieux, F., 1996, “In Vitro Degradation of Nanospheres from Poly(D,L-Lactides) of Different Molecular Weights and Polydispersities,” Int. J. Pharm., 129, pp. 95-102.

Belyayev, O. F., 1988, “Mechanism of Melting of Oriented Polymers,” Polym. Sci. U. S. S. R., 30, pp. 2545-2552.

Bhatla, A. and Yao, Y.L., 2009, “Effect of Laser Surface Modification on the Crystallinity of Poly(L-lactic acid),” J. Manuf. Sci. Eng., 131, 051004.

Campbell, D., Pethrick, R. A., and White, J. R., 2000, Polymer Characterization, Stanley Thornes Ltd., Chap. 8.

Chu, C. C. and Campbell, N. D., 1982, “Scanning Electron Microscopic Study of the Hydrolytic Degradation of Poly(glyco1ic acid) Suture,” J. Biomed. Mater. Res., 16, pp. 417-430.

Chu, C. C., 1981, “Hydrolytic Degradation of Polyglycolic Acid: Tensile Strength and Crystallinity Study,” J. Appl. Polym. Sci., 26, pp. 1727-1734.

Dunn, D. S. and Ouderkirk, A. J., 1990, “Chemical and Physical Properties of Laser-Modified Polymers,” Macromol., 23, pp. 770-774.

Faisant, N., Siepmann, J., Oury, P., Laffineur, V., Bruna, E., Haffner, J., Benoit, J.P., 2002, “The Effect of Gamma-Irradiation on Drug Release from Bioerodible Microparticles: a Quantitative Treatment,” Int. J. Pharm., 242, pp. 281-284.

158

Higuchi, T., 1961, “Rate of Release of Medicaments from Ointment Bases Containing Drugs in Suspension,” J. Pharm. Sci., 50, pp. 874-875.

Hoffman, J. D., Davis, G. T., and Lauritzen, J. I., 1976, Treatise on Solid State Chemistry, Hannay, N. B., ed., Plenum Press, New York, Vol. 3, Chap. 7.

Jing, D, Wu, L, and Ding, J., 2006, “Solvent-Assisted Room-Temperature Compression Molding Approach to Fabricated Porous Scaffolds for Tissue Engineering,” Macromol. Biosci., 6, pp. 747-757.

Koegler, W.S., Patrick, C., Cima, M.J., and Griffith, L.G., 2002, “Carbon Dioxide Extraction of Residual Chloroform from Biodegradable Polymers,” J. Biomed. Mater. Res., 63, pp. 567-576.

Labrecque, L. V., Kumar, R. A., Dave, V., Gross, R. A., McCarthy, S. P., 1997, “Citrate Esters as Plasticizers for Poly (lactic acid),” J Appl. Polym. Sci., 66, pp. 1507-1513.

Lam, C.X.F., Savalani, M.M., Teoh, S.-H., and Hutmacher, D.W., 2008, “Dynamics of in Vitro Polymer Degradation of Polycaprolactone-Based Scaffolds: Accelerated versus Simulated Physiological Conditions,” Biomed. Mater. 3, 034108.

Lazare, S. and Benet, P., 1993, “Surface Amorphization of Mylar Films with the Excimer Laser Radiation Above and Below Ablation Threshold: Ellipsometric Measurements,” J. Appl. Phys., 74, pp. 4953-4957.

Leong, K. W. and Langer, R., 1987, “Polymeric Controlled Drug Delivery,” Adv. Drug Delivery Rev., 1, pp. 199-233

Liau, W. B. and Chang, C. F., 1999, “Casting Solvent Effect on Crystallization Behavior and Morphology of Poly(Ethylene Oxide)/Poly(Methyl Methacrylate),” J. Appl. Polym. Scie., 76, pp. 1627-1636.

Ljungberg, N and Wesslen, B, 2002, “Preparation and Properties of Plasticized Poly(lactic acid) Films,” J Appl. Polym. Sci., 86, pp. 1227-1234.

McDonald, P.F., Lyons, J.G., Geever, L.M., Higginbotham, C.L., 2010, “In Vitro Degradation and Drug Release from Polymer Blends Based on Poly(DL-Lactide), Poly(L-Lactide-Glycolide) and Poly(ε-Caprolactone),” J. Mater. Sci., 45, pp. 1284-1292.

Mikos, A. G. and Temeoff, J. S., 2000, “Formation of Highly Porous Biodegradable Scaffolds for Tissue Engineering,” EJB Electronic Journal of Biotechnology, 3, fulltext-5.

N. Kolmogorov, “On the Statistical Theory of Crystallization of Metals [in Russian],” Izv. Akad. Nauk SSSR, Ser. Mat., No. 3, 1937, pp. 355-359.

Nakamura, T., Hitomi, S., Watanabe, S., Shimizu, Y., Jamshidi, K., Hyon, S.-H., and Y. Ikada, 1989, “Bioabsorption of Polylactides with Different Molecular Properties,” J. Biomed. Mater. Res., 23, pp. 1115-1130.

159

Norton, D. R., and Keller, A., 1985, “The Spherulitic and Lamellar Morphology of Melt-Crystallized Isotactic Polypropylene,” Polymer, 26, pp. 704-716.

Renouf-Glausera, A. C., Roseb, J., Farrarb, D. F., and Cameron, R. E., 2005, “The Effect of Crystallinity on the Deformation Mechanism and Bulk Mechanical Properties of PLLA,” Biomater., 26, pp. 5771-5782.

Rothen-Weinhold, A., Besseghir, K., and Gurny, R, 1997, “Analysis of the Influence of Polymer Characteristics and Core Loading on the in Vivo Release of a Somatostatin Analogue,” Eur. J. Pharm. Sci., 5, pp. 303-313.

Rothstein, S. N., Federspiel, W. J., and Little, S. R., 2009, “A Unified Mathematical Model for the Prediction of Controlled Release from Surface and Bulk Eroding Polymer Matrices,” Biomater., 30, pp. 1657-1664.

Saltzman, W. M. and Langer, R., 1989, “Transport Rates of Proteins in Porous Materials with Known Microgeometry,” Biophys. J., 55, pp. 163-171.

Schultz, J. M., 2001, Polymer Crystallization: The Development of Crystalline Order in Thermoplastic Polymers, Oxford University Press, New York, Chap 9.

Siparsky, G. L., Voorhees, K. J., and Miao, F., 1998, “Hydrolysis of Polylactic Acid (PLA) and Polycaprolactone (PCL) in Aqueous Acetonitrile Solutions: Autocatalysis,” J. Environmental Polym. Degrad., 6, pp. 31-41.

Tanaka, T., Tsuchiya, T., Takahashi, H., Taniguchi, M., and Lloyd, D. R., 2006, “Microfiltration Membrane of Polymer Blend of Poly(L-Lactic Acid) and Poly(ε-Caprolactone),” Desalination, 193, pp. 367-374.

Toda, A., Tomita, C., Hikosaka, M., and Saruyama, Y., 1998, “Melting of Polymer Crystals Observed by Temperature Modulated D.S.C. and Its Kinetic Modeling,” Polym., 39, pp. 5093-5104.

Tsuji, H. and Tsuruno, T., 2010, “Accelerated Hydrolytic Degradation of Poly(L-lactide)/Poly(D-lactide) Stereocomplex up to Late Stage,” Polym. Degrad. Stab., 95, pp. 477-484.

Tsuji, H., and Ikada, Y., 1998, “Properties and Morphology of Poly(L-Lactide). ii. Hydrolysis in Alkaline Solution,” J. Polym. Sci. A: Polym. Chem., 36, pp. 59-66.

Tsuji, H., Mizuno, A., and Ikada, Y., 2000, “Properties and Morphology of Poly(L-lactide). III. Effects of Initial Crystallinity on Long-Term in Vitro Hydrolysis of High Molecular Weight Poly(L-lactide) Film in Phosphate-Buffered Solution,” J. Appl. Polym. Sci., 77, pp. 1452-1464.

W. A. Johnson and R. F. Mehl, “Reaction Kinetics in Processes of Nucleation and Growth,” Transactions of the American Institute of Mining and Metallurgical Engineers, Vol. 135, 1939, pp. 416-442.

160

Watenabe, H., Takata, T., Tsuge, M., 1993, “Polymer Surface Modification due to Excimer Laser Radiation—Chemical and Physical Changes in the Surface Structure of Poly(ethylene terephthalate),” Polym. Int., 31, pp. 247-254.

Weir, N.A., Buchanan, F.J., Orr, J.F., Farrar, D.F., and Dickson, G.R., 2004, “Degradation of Poly-L-Lactide. Part 2: Increased Temperature Accelerated Degradation,” Proc. Inst. Mech. Eng., H: J. Eng. Med, 218, pp. 321-330.

Wunderlich, B., 1976, Macromolecular Physics, Vol. 2, Academic Press, New York.

Yamamoto, T., 2010, “Molecular Dynamics of Reversible and Irreversible Melting in Chain-Folded Crystals of Short Polyethylene-Like Polymer,” Macromol., 43, pp. 9384-9393.

Yeh, J.-T., Huang, C.-Y., Chai, W.-L., Chen, K.-N., 2009, “Plasticized Properties of Poly (lactic acid) and Triacetine Blends,” J. Appl. Polym. Sci., 112, 2757-2763.

Zilberman, M., Schwade, N. D., Meidell, R. S., and Eberhart, R. C., 2001, “Structured Drug-Loaded Bioresorbable Films for Support Structures,” J. Biomater. Sci. Polymer Edn., 12, pp. 875-892.

Zong, X. H., Wang, Z. G., Hsiao, B. S., Chu, B., Zhou, J. J., and Jamiolkowski, D. D., 1999, “Structure and Morphology Changes in Absorbable Poly(glycolide) and Poly(glycolide- co-lactide) during in Vitro Degradation,” Macromol., 32, pp. 8107-8114.

Chapter 2

Akcasu, A. Z., and Han, C. C., 1979, “Molecular Weight and Temperature Dependence of Polymer Dimensions in Solution,” Macromolecules, 12(2), pp. 276-280.

Alexander, L. E., 1969, X-Ray Diffraction Methods in Polymer Science, John Wiley & Sons, New York, Chap. 1.

Bassett, D. C., Olley, R. H., and Al Raheil, I. A. M., 1988, “On Crystallization Phenomena in PEEK,” Polymer, 29, pp. 1745-1754.

Bhatla, A., and Yao, Y. L., 2009, “Effect of Laser Surface Modification on the Crystallinity of Poly(L-Lactic Acid),” J. Manufacturing Science and Engineering, 131, pp. 051004-1-051004-11.

Campbell, D., Pethrick, R. A., and White, J. R., 2000, Polymer Characterization, 2nd ed., Stanley Thornes, Cheltenham.

Chang, C.-C., Wei, K.-H., and Chen, W.-C., 2003, “Spin-Coating of Polyimide-Silica Hybrid Optical Thin Films,” J. Electrochem. Soc., 150(8), pp. F147-F150.

Cotton, J. P., Nierlich, M., Boue, F., Daoud, M., Farnoux, B., Jannink, G., Duplessix, R., and Picot, C., 1976, “Experimental Determination of the Temperature-Concentration Diagram of Flexible Polymer Solutions by Neutron Scattering,” J. Chem. Phys., 65(3), pp. 1101-1108.

161

Croll, S. G., 1979, “The Origin of Residual Internal Stress in Solvent-Cast Thermoplastic Coatings,” J. Appl. Polym. Sci., 23, pp. 847-858.

Davis, G. T., Eby, R. K., and Martin, G. M., 1968, “Variations of the Unit-Cell Dimensions of Polyethylene: Effect of Crystallization Conditions, Annealing, and Deformation,” J. Appl. Phys., 39(11), pp. 4973-4981.

de Gennes, P. G., 1971, “Reptation of a Polymer Chain in the Presence of Fixed Obstacles,” J. Chem. Phys., 55(2), pp. 572-579.

Doi, M., and Edwards, S. F., 1988, The Theory of Polymer Dynamics, Oxford University Press, New York, Chaps. 2 and 5.

Frank, C. W., Rao, V., Despotopoulou, M. M., Pease, R. F. W., Hinsberg, W. D., Miller, R. D., and Rabolt, J. F., 1996, “Structure in Thin and Ultrathin Spin-Cast Polymer Films,” Science, 273, pp. 912-915.

Gao, X., Hou, W., Zhou, J., Li, L., and Zhao, L., 2004, “Relationships Between the Crystal Structures and the Multiple Melting Behaviors of Poly(Ethylene 2,6-Naphthalate),” Macrom. Mater. and Eng. 289, pp. 174-180.

Granasy, L., Pusztai, T., Tegze, G., Warren, J. A., and Douglas, J. F., 2005, “Growth and Form of Spherulites,” Phys. Rev. E, 72, pp. 011605-1-011605-15.

Hsu, C. C., Geil, P. H., Miyaji, H., and Asai, K., 1986, “Structure and Properties of Polypropylene Crystallized from the Glassy State,” J. Polym. Sci. B, 24, pp. 2379-2401.

Imai, M., Mori, K., Mizukami, T., Kaji, K., and Kanaya, T., 1992, “Structural Formation of Poly (Ethylene Terephthalate) during the Induction Period of Crystallization: 1. Ordered Structure Appearing before Crystal Nucleation,” Polymer, 33(21), pp. 4451-4456.

Ivanov, D. A., and Jonas, A. M., 1998, “Isothermal Growth and Reorganization upon Heating of a Single Poly(Aryl-Ether-Ether-Ketone) (PEEK) Spherulite, as Imaged by Atomic Force Microscopy,” Macromolecules, 31, pp. 4546-4550.

Kang, S., Hsu, S. L., Stidham, H. D., Smith, P. B., Leugers, M. A., and Yang, X., 2001, “A Spectroscopic Analysis of Poly(Lactic Acid) Structure,” Macromolecules, 34, pp. 4542-4548.

Kikkawa, Y., Abe, H., Iwata, T., Inoue, Y., and Doi, Y., 2002, “Crystallization, Stability, and Enzymatic Degradation of Poly(L-Lactide) Thin Film,” Biomacromolecules, 3(2), pp. 350-356.

Kister, G., Cassanas, G., and Vert, M., 1998, “Effects of Morphology, Conformation and Configuration on the IR and Raman Spectra of Various Poly(Lactic Acid)s,” Polymer, 39(2), pp. 267-273.

Kobayashi, J., Asahi, T., Ichiki, M., Oikawa, A., Suzuki, H., Watanabe, T., Fukada, E., and Shikinami, Y., 1995, “Structural and Optical Properties of Poly Lactic Acids,” J. Appl. Phys. 77(7), pp. 2957-2973.

162

Lanceley, H. A., and Sharples, A., 1966, “A Kinetic Study of Polymer Crystallization from Solution,” Die Makromolekulare Chemie, 94, pp. 30-41.

Li, H., Nie, W., Deng, C., Chen, X., and Ji, X., 2009, “Crystalline Morphology of Poly(L-Lactic Acid) Thin Films,” Eur. Polym. J., 45, pp. 123-130.

Li, L., Chan, C.-M., Yeung, K. L., Li, J.-X., Ng, K.-M., and Lei, Y., 2001, “Direct Observation of Growth of Lamellae and Spherulites of a Semicrystalline Polymer by AFM,” Macromolecules, 34, pp. 316-325.

Maillard, D., and Prud’homme, R. E., 2008, “The Crystallization of Ultrathin Films of Polylactides-Morphologies and Transitions,” Can. J. Chem., 86, pp. 556-563.

Meaurio, E., Zuza, E., Lopez-Rodrıguez, N., and Sarasua, J. R., 2006, “Conformational Behavior of Poly(L-Lactide) Studied by Infrared Spectroscopy,” J. Phys. Chem., 110, pp. 5790-5800.

Menczel, J., and Wunderlich, B., 1981, “Heat Capacity Hysteresis of Semicrystalline Macromolecular Glasses,” J. Polym. Sci.: Polym. Lett. Ed., 19, pp. 261-264.

Miyata, T., and Masuko, T., 1997, “Morphology of Poly(L-Lactide) Solution-Grown Crystals,” Polymer, 38(16), pp. 4003-4009.

Nampoothiri, K. M., Nair, N. R., and John, R. P., 2010, “An Overview of the Recent Developments in Polylactide (PLA) Research,” Bioresource Technology, 101, pp. 8493-8501.

Narladkar, A., Balnois, E., Vignaud, G., and Grohens, Y., 2008, “Difference in Glass Transition Behavior Between Semi Crystalline and Amorphous Poly(lactic acid) Thin Films,” Macromol. Symp., 273, pp. 146-152.

Norton, D. R., and Keller, A., 1985, “The Spherulitic and Lamellar Morphology of Melt-Crystallized Isotactic Polypropylene,” Polymer, 26, pp. 704-716.

Ogawa, T., Miyaji, H., and Asai, K., 1985, “Nodule Structure of Polypropylene,” J. Phys. Soc. Japan, 54(10), pp. 3668-3670.

Pluta, M., and Galeski, A., 2002, “Crystalline and Supermolecular Structure of Polylactide in Relation to the Crystallization Method,” J. Appl. Polym. Sci., 86, pp. 1386-1395.

Roylance, D. K., and DeVries, K. L., 1971, “Determination of Atomic Stress Distribution in Oriented Polypropylene by Infrared Spectroscopy,” Polym. Lett., 9, pp. 443-447.

Sarasua, J.-R., Zuza, E., Imaz, N., and Meaurio, E., 2008, “Crystallinity and Crystalline Confinement of the Amorphous Phase in Polylactides,” Macromol. Symp., 272, pp. 81-86.

Sasaki, S., and Asakura, T., 2003, “Helix Distortion and Crystal Structure of the α-Form of Poly(L-Lactide),” Macromolecule, 36, pp. 8385-8390.

Schindler, A., and Harper, D., 1979, “Polylactide. II. Viscosity-Molecular Weight Relationships

163

and Unperturbed Chain Dimensions,” J. Polym. Sci.: Polym. Chem. Ed., 17, pp. 2593-2599.

Schultz, J. M., 2001, Polymer Crystallization: The Development of Crystalline Order in Thermoplastic Polymers, Oxford University Press, New York, Chaps. 4 and 10.

Smith, B., 1999, Infrared Spectral Interpretation: A Systematic Approach, CRC Press, Boca Raton, FL, Chap. 1.

Strobl, G., 2000, “From the Melt via Mesomorphic and Granular Crystalline Layers to Lamellar Crystallites: A Major Route Followed in Polymer Crystallization?” Eur. Phys. J. E 3, pp. 165-183.

Su, C. H., Jeng, U., Chen, S. H., Lin, S. J., Wu, W. R., Chuang, W.-T., Tsai, J. C., and Su, A. C., 2009, “Nanograin Evolution in Cold Crystallization of Syndiotactic Polystyrene as Illustrated via In-Situ Small/Wide Angle X-Ray Scattering and Differential Scanning Calorimetry,” Macromolecules, 42, pp. 6656-6664.

Takeda, H., Nakashima, C., and Nasu, N., 2005, “Studies on Structure of Poly(ε-L-Lysine) Spherulite Grown from Solution,” Sen’I Gakkaishi, 61, pp. 61-66.

Tsuji, H., and Ikada, Y., 1995, “Properties and Morphology of poly(L-Lactide). 1. Annealing Condition Effects on Properties and Morphologies of Poly(L-Lactide),” Polymer, 36(14), pp. 2709-2716.

Tsuji, H., and Ikada, Y., 1998, “Properties and Morphology of Poly(L-Lactide). II. Hydrolysis in Alkaline Solution,” J. Polym. Sci. A: Polym. Chem., 36, pp. 59-66.

Vorotilov, K., Petrovsky, V., and Vasiljev, V., 1995, “Spin Coating Process of Sol-Gel Silicate Films Deposition: Effect of Spin Speed and Processing Temperature,” J. Sol-Gel Sci. and Tech., 5, pp. 173-183.

Wang, Y., Funari, S. S., and Mano, J. F., 2006, “Influence of Semicrystalline Morphology on the Glass Transition of Poly(L-Lactic Acid),” Macromol. Chem. Phys., 207, pp. 1262-1271.

Wunderlich, B., 1976, Macromolecular Physics, Academic Press, New York, Vol. 2, Chap. 5.

Zhang, J., Liang, Y., Yan, J., and Lou, J., 2007, “Study of the Molecular Weight Dependence of Glass Transition Temperature for Amorphous Poly(L-Lactide) by Molecular Dynamics Simulation,” Polymer, 48, pp. 4900-4905.

Zhao, J., Qiu, J., Niu, Y., and Wang, Z., 2009, “Evolutions of Morphology and Crystalline Ordering Upon Annealing of Quenched Isotactic Polypropylene,” J. Polym. Sci. B, 47, pp. 1703-1712.

Chapter 3

Aleksandrov, A. P., Kitai, M. S., and Varbanskaya, R. A., 1988, “Photochemical Autocatalytic Propagation Reaction of Polyene Sequences in Photodegradation of Polyvinyl Chloride,” Polym.

164

Sci. U. S. S. R., 30, pp. 1687-1696.

Alexander, L. E., 1969, X-ray Diffraction Methods in Polymer Science, New York: John Wiley & Sons.

Amass, W., Amass, A., Tighe, B., 1998, “A Review of Biodegradable Polymers: Uses, Current Developments in the Synthesis and Characterization of Biodegradable Polyesters, Blends of Biodegradable Polymers and Recent Advances in Biodegradation Studies,” Polym. Int., 47, pp. 89-144.

Astrath, N. G. C., Astrath, F. B. G., Shen, J., Zhou, J., Michaelian, K. H., Fairbridge, C., Malacarne, L. C., Pedreira, P. R. B., Santoro, P. A., and Baesso, M. L., 2009, “Arrhenius Behavior of Hydrocarbon Fuel Photochemical Reaction Rates by Thermal Lens Spectroscopy,” Appl. Phys. Lett., 95, 191902.

Beamson, G. and Briggs, D., 1992, High Resolution XPS of Organic Polymers, Chichester: John Wiley & Sons.

Belyayev, O. F., 1988, “Mechanism of Melting of Oriented Polymers,” Polym. Sci. U. S. S. R., 30, pp. 2545-2552.

Bhatla, A. and Yao, Y. L., 2009, “Effect of Laser Surface Modification on the Crystallinity of Poly(L-Lactic Acid),” J. Manuf. Sci. Eng., 131, 051004.

Bityurin, N., Luk’yanchuk, B. S., Hong, M. H., and Chong, T. C., 2003, “Models for Laser Ablation of Polymers,” Chem. Rev., 103, pp. 519-552.

Boudenne, A., Ibos, L., Candau, Y., and Thomas, S., editors, 2011, Handbook of Multiphase Polymer Systems, West Sussex, John Wiley & Sons, Chap. 10.

Butiagin, P. J., 1972, “The Decay of Free Radicals in Polymer Media,” Proc. Int. Conf. Chem. Transform. Polym., pp. 57-76.

Cormia, R. L., Price, F. P., and Turnbull, D., 1962, “Kinetics of Crystal Nucleation in Polyethylene,” J. Chem. Phys., 37, pp. 1333-1340.

Currie, J. A., Bohm, G. G. A., and Dole, M., 1969, “Calorimetric Detection on Free Radicals in Irradiated Polyethylene,” Polym. Lett., 7, pp. 535-539.

Dan, E. and Guillet, J. E., 1973, “Photochemistry of Ketone Polymers. X. Chain Scission Reactions in the Solid State,” Macromol., 6, pp. 230-235.

Di Lorenzo, M. L., 2005, “Crystallization Behavior of Poly(L-Lactic Acid),” Eur. Polym. J., 41, pp. 569-575.

Dunn, D. S. and Ouderkirk, A. J., 1990, “Chemical and Physical Properties of Laser-Modified Polymers,” Macromol., 23, pp. 770-774.

165

Efthimiopoulos, T., Kiagias, H., Christoulakis, S., and Merlemis, N., 2008, “Bubble Creation and Collapse during Excimer Laser Ablation of Weak Absorbing Polymer,” Appl. Surf. Sci., 254, pp. 5626-5630.

Hoffman, J. D. and Weeks, J. J., 1962, “Melting Process and the Equilibrium Melting Temperature of Polychlorotrifluoroethylene,” J. Res. Nat. Bureau of Standards-A. Phys. Chem., 66A, pp. 13-28.

Hsu, S.-T. and Yao, Y. L., 2011, “Effect of Film Formation Method and Annealing on Crystallinity of Poly (L-Lactic Acid) Films,” Proc. Int. Manuf. Sci. Eng. Conf., MSEC2011-50205.

Inagaki, N., Narushima, K., Tsutsui, Y., and Ohyama, Y., 2002, “Surface Modification and Degradation of Poly(Lactic Acid) Films by Ar-Plasma,” J. Adhes. Sci. Technol., 16, pp. 1041-1054.

Janorkar, A. V., Metters, A. T., and Hirt, D. E., 2007, “Degradation of Poly(L-Lactide) Films under Ultraviolet-Induced Photografting and Sterilization Conditions,” J. Appl. Polym. Sci., 106, pp. 1042-1047.

Lazare, S. and Benet, P., 1993, “Surface Amorphization of Mylar Films with the Excimer Laser Radiation Above and Below Ablation Threshold: Ellipsometric Measurements,” J. Appl. Phys., 74, pp. 4953-4957.

Lazare, S. and Granier, V., 1989, “Ultraviolet Laser Photoablation of Polymers: A Review and Recent Results,” Laser Chem., 10, pp. 25-40.

Miller, R. A., Brady, J. M., and Cutright, D. E., 1977, “Degradation Rates of Oral Resorbable Implants (Polylactates and Polyglycolates): Rate Modification with Changes in PLA/PGA Copolymer Ratios,” J. Biomed. Mater. Res., 11, pp. 711-719.

Mitomo, H., 1988, “Correspondence of Lamellar Thickness to Melting Point of Nylon-6,6 Single Crystal,” Polym., 29, pp. 1635-1642.

Nampoothiri, K. M., Nair, N. R., and John, R. P., 2010, “An Overview of the Recent Developments in Polylactide (PLA) Research,” Bioresour. Technol., 101, pp. 8493-8501.

Ohki, Y., Hirai, N., Fuse, N., Tanaka, T., Kohtoh, M., and Okabe, S., 2007, “Search for Adequate Biodegradable Polymer as an Eco-Friendly Electrical Insulating Material,” Proc. Int. Symp. EcoTopia Sci., pp. 488-493.

Rebollar, E., Bounos, G., Oujja, M., Georgiou, S., and Castillejo, M., 2006, “Effect of Molecular Weight on the Morphological Modifications Induced by UV Laser Ablation of Doped Polymers,” J. Phys. Chem. B, 110, pp. 16452-16458.

Schindler, A. and Harper, D., 1979, “Polylactide. II. Viscosity-Molecular Weight Relationships and Unperturbed Chain Dimensions,” J. Polym. Sci.: Polym. Chem. Ed., 17, pp. 2593-2599.

166

Schneider, S., Richter, F., and Brem, B., 1998, The Photodegradation of Poly(2,6-dimethyl-1,4-phenylene Oxide)—a Flash Photolysis Study of Polymer and Model Compounds,” Polym. Degrad. Stab., 61, pp. 543-464.

Srinivasan, R., 1993, “Pulsed Ultraviolet Laser Interactions with Organic Polymers Dependence of Mechanism Upon Laser Power,” Polym. Degrad. Stab., 43, pp. 101-107.

Tsuji, H. and Ikada, Y., 1995, “Properties and Morphology of Poly(L-Lactide). 1. Annealing Condition Effects on Properties and Morphologies of Poly(L-Lactide),” Polym., 36, pp. 2709-2716.

Tsuji, H. and Ikada, Y., 1998, “Properties and Morphology of Poly(L-Lactide). II. Hydrolysis in Alkaline Solution,” J. Polym. Sci., Part A: Polym. Chem., 36, pp. 59-66.

Tsuji, H., 2002, “Autocatalytic Hydrolysis of Amorphous-Made Polylactides: Effects of L-Lactide Content, Tacticity, and Enantiomeric Polymer Blending,” Polym., 43, pp. 1789-1796.

Wachsen, O., Platkowski, K., and Reichert, K. H., 1997, “Thermal Degradation of Poly-L-Lactide—Studies on Kinetics, Modelling and Melt Stabilization,” Polym. Degrad. Stab., 57, pp.87-94.

Watenabe, H., Takata, T., and Tsuge, M., 1993, “Polymer Surface Modification due to Excimer Laser Radiation—Chemical and Physical Changes in the Surface Structure of Poly(Ethylene Terephthalate),” Polym. Int., 31, pp. 247-254.

Yamamoto, T., 2010, “Molecular Dynamics of Reversible and Irreversible Melting in Chain-Folded Crystals of Short Polyethylene-Like Polymer,” Macromol., 43, pp. 9384-9393.

Zhao, Y., Fu, J., Ng, D. K. P., and Wu, C., 2004, “Formation and Degradation of Poly(D,L-Lactide) Nanoparticles and Their Potential Application as Controllable Releasing Devices,” Macromol. Biosci., 4, pp. 901-906.

Chapter 4

Alexander, L. E., 1969, X-Ray Diffraction Methods in Polymer Science, John Wiley & Sons, New York, Chap. 1.

Amass, W., Amass, A., and Tighe, B., 2008, “A Review of Biodegradable Polymers: Uses, Current Developments in the Synthesis and Characterization of Biodegradable Polyesters, Blends of Biodegradable Polymers and Recent Advances in Biodegradation Studies.” Polym. Int., 47, pp. 89-144.

Avrami, M., 1939, “Kinetics of Phase Change. I: General Theory,” J. Chem. Phys., 7, pp. 1103-1112.

Avrami, M., 1941, “Granulation, Phase Change, and Microstructure: Kinetics of Phase Change. III,” J. Chem. Phys., 9, pp. 177-184.

167

Belyayev, O. F., 1988, “Mechanism of Melting of Oriented Polymers,” Polym. Sci. U. S. S. R., 30, pp. 2545-2552.

Bhatla, A., and Yao, Y. L., 2009, “Effect of Laser Surface Modification on the Crystallinity of Poly(L-Lactic Acid),” J. Manuf. Sci. Eng., 131, 051004.

Chu, C. C., 1981, “Hydrolytic Degradation of Polyglycolic Acid: Tensile Strength and Crystallinity Study,” J. Appl. Polym. Sci., 26, pp. 1727-1734.

Dunn, D. S., and Ouderkirk, A. J., 1990, “Chemical and Physical Properties of Laser-Modified Polymers,” Macromol., 23, pp. 770-774.

Fischer, E. W., Sterzel, H. J., and Wegner, G., 1973, “Investigation of the Structure of Solution Grown Crystals of Lactide Copolymers by Means of Chemical Reactions,” Kol-ZuZ Polymere, 251, pp. 980-990.

Hsu, S.-T., Tan, H., and Yao, Y. L., 2012, “Effect of Excimer Laser Irradiation on Crystallinity and Chemical Bonding of Biodegradable Polymer,” Polym. Degrad. Stab., 97, pp. 88-97.

Inagaki, N., Narushima, K., Tsutsui, Y., and Ohyama, Y., 2002 “Surface Modification and Degradation of Poly(Lactic Acid) Films by Ar-Plasma,” J. Adhes. Sci. Technol., 16, pp. 1041-1054.

Lam, C. X. F., Savalani, M. M., Teoh, S. H., and Hutmacher, D. W., 2008, “Dynamics of in Vitro Polymer Degradation of Polycaprolactone-Based Scaffolds: Accelerated versus Simulated Physiological Conditions,” Biomed. Mater., 3, 034108.

Li, H., Nie, W., Deng, C., Chen, X., and Ji, X., 2009, “Crystalline Morphology of Poly(L-Lactic Acid) Thin Films,” Eur. Polym. J., 45, pp. 123-130.

Li, S. M., Garreau, H., and Vert, M., 1990, “Structure-Property Relationships in the Case of the Degradation of Massive Poly(α-Hydroxy Acids) in Aqueous Media: Part 2 Degradation of Lactide-Glycolide Copolymers: PLA37.5GA25 and PLA75GA25,” J. Mater. Sci.: Mater. Med., 1, pp. 131-139.

Lyu, S., Schley, J., Loy, B., Lind, D., Hobot, C., Sparer, R., and Untereker, D., 2007, “Kinetics and Time-Temperature Equivalence of Polymer Degradation,” Biomacromol., 8, pp. 2301-2310.

Menczel, J., Wunderlich, B., 1981, “Heat Capacity Hysteresis of Semicrystalline Macromolecular Glasses,” J. Polym. Sci.: Polym. Lett. Ed., 19, pp. 261-264.

Pitt, C. G., and Gu, Z., 1987, “Modification of the Rates of Chain Cleavage of Poly(ε-Caprolactone) and Related Polyesters in the Solid State,” J. Controlled. Release, 4, pp. 283-292.

Renouf-Glausera, A. C., Roseb, J., Farrarb, D. F., and Cameron, R. E., 2005, “The Effect of Crystallinity on the Deformation Mechanism and Bulk Mechanical Properties of PLLA,” Biomater., 26, pp. 5771-5782.

168

Siparsky, G. L., Voorhees, K. J., and Miao, F., 1998, “Hydrolysis of Polylactic Acid (PLA) and Polycaprolactone (PCL) in Aqueous Acetonitrile Solutions: Autocatalysis,” J. Environmental Polym. Degrad., 6, pp. 31-41.

Stephens, C. H., Whitmore, P. M., Morris, H. R., and Bier, M. E., 2008, “Hydrolysis of the Amorphous Cellulose in Cotton-Based Paper,” Biomacromol., 9, pp. 1093-1099.

Toda, A., Tomita, C., Hikosaka, M., and Saruyama, Y., 1998, “Melting of Polymer Crystals Observed by Temperature Modulated D.S.C. and Its Kinetic Modeling,” Polym., 39, pp. 5093-5104.

Tsuji, H., and Ikada, Y., 1998, “Properties and Morphology of Poly(L-Lactide). ii. Hydrolysis in Alkaline Solution,” J. Polym. Sci. A: Polym. Chem., 36, pp. 59-66.

Tsuji, H., and Ikarashi, K., 2004, “In Vitro Hydrolysis of Poly(L-Lactide) Crystalline Residues as Extended-Chain Crystallites: II. Effects of Hydrolysis Temperature,” Biomacromol., 5, pp. 1021-1028.

Tsuji, H., and Tsuruno, T., 2010, “Accelerated Hydrolytic Degradation of Poly(L-Lactide)/Poly(D-Lactide) Stereocomplex up to Late Stage,” Polym. Degrad. Stab., 95, pp. 477-484.

Wang, Y., Pan, J., Han, X., Sinka, C., and Ding, L., 2008, “A Phenomenological Model for the Degradation of Biodegradable Polymers,” Biomater., 29, pp. 3393-3401.

Weir, N. A., Buchanan, F. J., Orr, J. F., Farrar, D. F., and Dickson, G. R., 2004, “Degradation of Poly-L-Lactide. Part 2: Increased Temperature Accelerated Degradation,” Proc. Instn. Mech. Engrs. Part. H: J. Eng. in Med., 218, pp. 321-330.

Yamamoto, T., 2010, “Molecular Dynamics of Reversible and Irreversible Melting in Chain-Folded Crystals of Short Polyethylene-Like Polymer,” Macromol., 43, pp. 9384-9393.

Zong, X. H., Wang, Z. G., Hsiao, B. S., Chu, B., Zhou, J. J., and Jamiolkowski, D. D., 1999, “Structure and Morphology Changes in Absorbable Poly(glycolide) and Poly(glycolide- co-lactide) during in Vitro Degradation,” Macromol., 32, pp. 8107-8114.

Chapter 5

Amass, W., Amass, A., and Tighe, B., 2008, “A Review of Biodegradable Polymers: Uses, Current Developments in the Synthesis and Characterization of Biodegradable Polyesters, Blends of Biodegradable Polymers and Recent Advances in Biodegradation Studies,” Polym. Int., 47, pp. 89-144.

Aiken, W., Alfrey, T., Janssen, A., and Mark, H., 1947, “Creep Behavior of Plasticized Vinylite VYNW,” J. Polym. Sci., 2, pp. 178-198.

Alexander, L. E., 1969, X-Ray Diffraction Methods in Polymer Science, John Wiley & Sons, New York, Chap. 1.

169

Chu, C. C., 1981, “Hydrolytic Degradation of Polyglycolic Acid: Tensile Strength and Crystallinity Study,” J. Appl. Polym. Sci., 26, pp. 1727-1734.

Fischer, E. W., Sterzel, H. J., and Wegner, G., 1973, “Investigation of the Structure of Solution Grown Crystals of Lactide Copolymers by Means of Chemical Reactions,” Kolloid-Z. u. Z. Polym., 251, pp. 980-990.

Flory, P. J., 1940, “Viscosities of Linear Polyesters. An Exact Relationship between Viscosity and Chain Length,” J. Am. Chem. Soc., 62, pp. 1057-1070.

Higuchi, T., 1961, “Rate of Release of Medicaments from Ointment Bases Containing Drugs in Suspensions,” J. Pharm. Sci., 50, pp. 874-875.

Hsu, S.-T., Tan, H., and Yao, Y. L., 2012a, “Effect of Excimer Laser Irradiation on Crystallinity and Chemical Bonding of Biodegradable Polymer,” Polym. Degrad. Stab., 97, pp. 88-97.

Hsu, S.-T., Tan, H., and Yao, Y. L., 2012b, “Effect of Laser Induced Crystallinity Modification on Biodegradation Profile of Poly(L-Lactic Acid),” J. Manuf. Sci. Eng., submitted.

Kirkpatrick, A., 1940, “Some Relations between Molecular Structure and Plasticizing Effect,” J. Appl. Phys., 11, pp. 255-261.

Lao, L. L., Venkatraman, S. S., and Peppas, N. A., 2009, “A Novel Model and Experimental Analysis of Hydrophilic and Hydrophobic Agent Release from Biodegradable Polymers,” J. Biomed. Mater. Res. A., 90, pp. 1054-1065.

Ljungberg, N. and Wesslen, B., 2002, “The Effects of Plasticizers on the Dynamic Mechanical and Thermal Properties of Poly(Lactic Acid),” J. Appl. Polym. Sci., 86, pp. 1227-1234.

Lyu, S. and Untereker, D., 2009, “Degradability of Polymers for Implantable Biomedical Devices,” Int. J. Mol. Sci., 10, pp. 4033-4065.

Lyu, S., Schley, J., Loy, B., Lind, D., Hobot, C., Sparer, R., and Untereker, D., 2007, “Kinetics and Time-Temperature Equivalence of Polymer Degradation,” Biomacromol., 8, pp. 2301-2310.

Marcilla, A. and Beltran, M., 2004, “Mechanisms of Plasticizers Action,” in: Wypych, G., ed., Handbook of Plasticizers, ChemTec Publishing, Toronto, Chap. 5.

Painter, P. C. & Coleman, M. M., Fundamentals of Polymer Science: An Introductory Text, Technomic Publishing, Lancaster, Chap. 8.

Pitt, C. G. and Gu, Z., 1987, “Modification of the Rates of Chain Cleavage of Poly(ε-Caprolactone) and Related Polyesters in the Solid State,” J. Controlled Release, 4, pp. 283-292.

Rothstein, S. N. and Federspiel, W. J., Little, S. R., 2009, “A Unified Mathematical Model for the Prediction of Controlled Release from Surface and Bulk Eroding Polymer Matrices,” Biomater., 30, pp. 1657-1664.

170

Rothstein, S. N., Federspiel, W. J., and Little, S. R., 2008, “A Simple Model Framework for the Prediction of Controlled Release from Bulk Eroding Polymer Matrices,” J. Mater. Chem., 18, pp. 1873-1880.

Saltzman, W. M. and Langer, R., 1989, “Transport Rates of Proteins in Porous Materials with Known Microgeometry,” Biophys. J., 55, pp. 163-171.

Spenlehauer, G., Vert, M., Benoit, J.-P., Chabot, F., and Veillard, M., 1988, “Biodegradable Cisplatin Microspheres Prepared by the Solvent Evaporation Method: Morphology and Release Characteristics,” J. Controlled Release, 7, pp. 217-229.

Stephens, C. H., Whitmore, P. M., Morris, H. R., and Bier, M. E., 2008, “Hydrolysis of the Amorphous Cellulose in Cotton-Based Paper,” Biomacromol., 9, pp. 1093-1099.

Tsuji, H. and Ikada, Y., 1998, “Properties and Morphology of Poly(L-Lactide). ii. Hydrolysis in Alkaline Solution,” J. Polym. Sci. A: Polym. Chem., 36, pp. 59-66.

Ueberreiter, K. and Kanig, G., 1952, “Self-Plasticization of Polymers,” J. Colloid Sci., 7, pp. 569-583.

Wadso, L. and Karlsson, O. J., 2013, “Alkaline Hydrolysis of Polymers with Ester Groups Studied by Isothermal Calorimetry,” Polym. Deg. Stab., 98, pp. 73-78.

Wang, Y., Pan, J., Han, X., Sinka, C., and Ding, L., 2008, “A Phenomenological Model for the Degradation of Biodegradable Polymers,” Biomater., 29, pp. 3393-3401.

Wilhelm, P. and Stephan, D., 2007, “Photodegradation of Rhodamine B in Aqueous Solution via SiO2@TiO2 Nano-Spheres,” J. Photochem. Photobio. A: Chem., 185, pp. 19-25.

Xiao, H., Lu, W., and Yeh, J., 2009, “Effect of Plasticizer on the Crystallization Behavior of Poly(Lactic Acid),” J. Appl. Polym. Sci., 113, pp. 112-121.

Yeh, J., Huang, C., Chai, W., and Chen, K., 2009, “Plasticized Properties of Poly (Lactic Acid) and Triacetine Blends,” J. Appl. Polym. Sci., 112, pp. 2757-2763.

171

Appendix

This appendix details data processing, numerical models, sample preparation process for gel

permeation chromatography (GPC), and the publications associated with the work achieved in

this thesis. The appendix is divided into four parts. The first part explains the calculation of

polymer crystallinity from the wide angle x-ray diffraction (WAXD) method. The second part

describes the numerical models developed in this thesis. The third part gives the purification

process of drug loaded samples for the GPC measurements. Publications based on the work

done in this thesis are included in the fourth part.

A.1 Calculation of Polymer Crystallinity from Wide Angle X-Ray Diffraction Method

Under the assumption that the x-ray diffraction intensity from the atom in the amorphous region

is the same as that in the crystalline region, polymer crystallinity is determined based on the

WAXD profiles. The crystallinity is the ratio of the crystalline peak area to the area under the

overall profile. To calculate the area, the background of WAXD profiles needs to be removed

first. A MATLAB program, crystallinity.m, has been written to remove the background. In

the program, the increment and range of WAXD measurement is defined in terms of 2θ values as

x_sec=0.029;

x_min=0.295;

x_max=100.751;

The experiment result, in the .txt format, is imported into the program through

y=load('WAXD result.txt');

The data points to determine the background are defined at both ends of the WAXD profiles.

172

For the left end of the WAXD profile, the data points are

x_linefit_left1=100;

x_linefit_right1=200;

which means the 100th to 200th data point from the left end of the WAXD profile are used to

determine the background. Similarly, for the right end of the WAXD profile, the data points are

x_linefit_left2=3000;

x_linefit_right2=3165;

The background is fitted through the polyfit function. The profile with background subtracted

is plotted through

fid=fopen('WAXD result_output.txt','wt');

for i=1:sec

fprintf(fid,'%d\n',round(y_cut(1,i)));

end

The processed profile is then saved as WAXD result_output.txt. To determine the crystallinity,

the processed profile is then fitted with the WAXD profiles obtained from the amorphous sample.

The difference between the total area under the processed profile and the area under the

amorphous profile represents the contribution from the crystalline phase. The area contributed

from the crystalline phase divided by the total area is the calculated crystallinity.

A.2 Finite Element Models

In this thesis, three models have been developed based on the finite element method through

COMSOL Multiphysics. The first model, developed in Chap. 3, considers the laser induced

thermal and chemical modifications. The second model, developed in Chap. 4, considers the

173

polymer biodegradation process. The third model, developed in Chap. 5, considers the drug

release process. The three models are described as follows.

A.2.1 Model of Laser Induced Thermal and Chemical Modifications

This 2D symmetric model simulated the coupled thermal and chemical modifications induced by

a single excimer laser pulse. Material properties are defined in Parameters under Global

Definitions. Variables are written in Definitions. Heat capacity is given as 1000[J/(kg*K)] +

(dH/Tm) * flc2hs((T-Tm)[1/K],dT[1/K]) where flc2hs is a smoothed Heaviside function and dT

specifies the interval within which the heat capacity changes due to melting. The rate factor, k,

for chemical reaction is given as A*exp(-E/(8.314[J/(mol*K)]*T))*if(T>Tm,1,0.1) to consider its

dependence with temperature based on the Arrhenius relationship, as well as the polymer chain

state. If the temperature is higher than the melting temperature, the crystalline chains become

mobile and have high rate factor than below the melting temperature. Through if(T>Tm,1,0.1)

it is assumed that the polymer chain below the melting temperature has a rate factor one order of

magnitude smaller than the chain above the melting temperature.

The thermal effect induced by laser irradiation is simulated in the Heat Transfer (ht) model, and

the chemical modification is simulated in the Transport of Diluted Species (chds) model. The

two models are coupled such that the temperature calculated in the Heat Transfer (ht) model

determines the chemical reaction rate in the Transport of Diluted Species (chds) model, and

chemical reaction heat is imported into the Heat Transfer (ht) model. In Heat Transfer (ht),

laser energy is defined in Heat Source 1 as a function of depth, exp((ab)*y[1/m]) and a function

of time, exp(-4*log(2)*((t-2*tp)/tp)^2). The top hat profile of laser energy in space is described

through if(x>bw,0,1), where bw is the beam width. The calculated temperature is imported into

174

the Transport of Diluted Species (chds) model through the variable -rate in Reactions 1. The

variable rate, given in Definitions, is the product of the rate factor, k, and the concentration of

unmodified polymer chains, cP, as specified by the rate equation. The heat generated through

chemical reactions coupled with the Heat Transfer (ht) model, as defined in Heat Source 2.

A.2.2 Model of Polymer Biodegradation Process

This model is developed in a 2D axisymmetric domain to consider the biodegradation of a

circular polymer matrix. Material properties are given in Parameters under Global Definitions,

and variables are written in Definitions. The biodegradation rate is written in form of the

product of the rate constant and the concentration of species involved in a reaction. For

example, the biodegradation rate of the original amorphous chains is given as

k1*c_e*c_w+k2*c_e*c_w*c_m, where k1 is the rate constant for non-autocatalysis degradation,

k2 is the rate constant for autocatalysis degradation, c_e, c_w, and c_m refers to the

concentration of different species as described below. The difference in crystallinity along

thickness due to laser irradiation is expressed as area_cryst in Variables 1.

This model is composed of seven species: non-degraded amorphous chains (c_e), monomers

(c_m), non-degraded crystalline chains (c_c), degraded crystalline chains (c_cs), degraded

amorphous chains in stages 1, 2, and 3 (c_d1, c_d2, and c_d3). The changes of their

concentrations as a function of time and space due to biodegradation are calculated in the

Transport of Diluted Species (chds) model. The initial conditions are specified in Initial Values

1. The reaction rates for all species are specified in Reactions 1. For initial non-degraded

amorphous chains, the reaction rate is -rate_amor-rate_crystallization, which signifies that the

reduction of the initial non-degraded amorphous chains is due to degradation (-rate_amor) and

175

crystallization (-rate_crystallization). Degradation of initial amorphous chains generates

degraded amorphous chains in stage 1. The reaction rate of the degraded amorphous chains in

stage 1 is therefore given as rate_amor-rate_degraded_1-rate_crystallization_deg_1. The term,

rate_amor, is contributed from the degradation of the initial amorphous chains. The term,

-rate_degraded_1, represents the degradation of stage 1 amorphous chains, which generates

stage 2 amorphous chains. The term, -rate_crystallization_deg_1, accounts for the reduction of

stage 1 amorphous chains due to crystallization. Similar reaction equations are given for stage

2 amorphous chains as rate_degraded_1-rate_degraded_2-rate_crystallization_deg_2, and stage

3 amorphous chains as rate_degraded_2-rate_degraded_3-rate_crystallization_deg_3.

Degradation of stage 3 amorphous chains generates monomers, with the rate given as

rate_degraded_3. Concentration change of the crystalline chains is calculated as

rate_crystallization+rate_crystallization_deg_1+rate_crystallization_deg_2+rate_crystallizatio

n_deg_3-rate_cryst. The first three terms signify the contribution of crystalline chains from

amorphous chains. The last term considers the degradation of crystalline chains, which occurs

on crystal fold surfaces and generates the integral fold of crystalline chains.

A.2.3 Model of Drug Release Process

This model is developed as the Transport of Diluted Species (chds) model in a 2D axisymmetric

domain to consider the biodegradation and drug release of a circular polymer matrix. This

model is an extension of the model described in Sec. A.2.2 in that drug molecules and porosity

due to polymer biodegradation are considered, while the governing equations are similar. In

addition to the seven species mentioned in Sec. A.2.2, the eighth species is added to the current

model to consider the addition of drug molecules. The concentration of drug molecules is

c_drug, which is a function of time and space. The reaction rate of drug, given in Reaction 1, is

176

0 assuming no reaction between drug and other species. Polymer degradation generates a

porous structure in the matrix. Assuming the polymer molecular weights forming pores follow

a normal distribution, the porosity, por, is written as 1-(erf((MW-MWr)/sdv)+1)*0.5*bulk as

given in Definitions, where MW is the molecular weight as a function of time and space and

MWr is the molecular weight at which pore starts to form. The diffusivity of drug, Dc_drug, is

proportional to the porosity as specified in Convection and Diffusion 1.

A.3 Purification Process of Drug Loaded PLLA Matrix for GPC measurements

In drug release tests, PLLA matrix is loaded with the model drug, rhodamine B. To measure the

molecular weights of drug loaded PLLA, GPC measurements are performed. The model drug,

with a small molecular weight of 479 g/mol, may block GPC columns, which are designed to

separate macromolecules. Therefore, the drug molecules need to be removed.

To conduct the purification process, around 5 mg of drug loaded matrix is dissolved in 0.1 mL

chloroform through sonication for 5 min. Both drug and PLLA can be dissolved in chloroform,

and a clear solution is obtained. 20 mL methanol is then added into the solution. The solution

is sonicated for 5 min. Drug is dissolved in methanol. PLLA, on the other hand, is not soluble

in methanol and forms the suspension. To collect PLLA suspension, centrifugation is

conducted which separates materials based on mass. The centrifugation is operated at 3000

rpm for 15 min, and PLLA suspension and drug/methanol solution are separated in the centrifuge

tube. The drug/methanol solution, located in the upper part of the tube, is removed, leaving

behind the solid PLLA. This process is repeated up to three times until remained drug is not

visually observable. The collected PLLA is dried in a hood for 48 hours to remove chloroform

and methanol, and then measured by GPC.

177

A.4 Publications Under Candidature

Hsu, S.-T., Tan, H, and Yao, Y. L., 2012, “Effect of Laser Induced Crystallinity

Modification on Biodegradation Profile of Poly(L-lactic Acid),” Journal of Manufacturing

Science and Engineering, submitted.

Also in Proceedings of 31st International Congress on Applications of Lasers and

Electro-Optics (ICALEO), Sept. 2012, M204.

Hsu, S.-T., Tan, H, and Yao, Y. L., 2012, “Effect of Excimer Laser Irradiation on

Crystallinity and Chemical Bonding of Biodegradable Polymer,” Polymer Degradation and

Stability, Vol. 97, No. 1, pp. 88-97.

Also in Proceedings of 30th International Congress on Applications of Lasers and

Electro-Optics (ICALEO), Oct. 2011, M1306.

Hsu, S.-T., Wang, H., Satoh, G., and Yao, Y. L., 2011, “Application of Surface Structuring

with Lasers,” Proceedings of 30th International Congress on Applications of Lasers and

Electro-Optics (ICALEO), M1201, Invited Paper.

Hsu, S.-T. and Yao, Y. L., 2011, “Effect of Film Formation Method and Annealing on

Morphology and Crystal Structure of Poly(L-lactic Acid) Films,” Journal of Manufacturing

Science and Engineering, accepted.

Also in Proceedings of 6th International Manufacturing Science and Engineering

Conference (MSEC), June 2011, MSEC2011-50205.

Hsu, S.-T. and Yao, Y. L., 2013, “Effect of Drug Loading on Laser Modified Polymer

Biodegradation,” Materials Letters, submitted.

178

Hsu, S.-T. and Yao, Y. L., 2013, “Effect of Drug Loading and Laser Surface Melting on

Drug Release Profile from Biodegradable Polymer,” Journal of Applied Polymer Science,

submitted.

Wang, H., Hsu, S.-T., Tan, H., Yao, Y. L., Chen, H., Azer, M. N., 2012, “Predictive

Modeling for Glass-Side Laser Scribing of Thin Film Photovoltaic Cells,” Journal of

Manufacturing Science and Engineering, submitted.

Also in Proceedings of 40th North American Manufacturing Research Conference

(NAMRC), June 2012, NAMRC40-7774.