SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

124
The Pennsylvania State University The Graduate School Department of Materials Science and Engineering SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE COPOLYMERS AND BLENDS A Thesis in Materials Science and Engineering by Mantana Kanchanasopa © 2004 Mantana Kanchanasopa Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy May 2004

Transcript of SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

Page 1: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

The Pennsylvania State University

The Graduate School

Department of Materials Science and Engineering

SOLID-STATE STRUCTURE AND DYNAMICS OF

LACTIDE COPOLYMERS AND BLENDS

A Thesis in

Materials Science and Engineering

by

Mantana Kanchanasopa

© 2004 Mantana Kanchanasopa

Submitted in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

May 2004

Page 2: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

The thesis of Mantana Kanchanasopa has been reviewed and approved* by the following

James P. Runt Professor of Polymer Science Thesis Advisor Chair of Committee Ian R. Harrison Professor Emeritus of Polymer Science Evangelos Manias Assistant Professor of Materials Science and Engineering Christopher A. Siedlecki Assistant Professor of Surgery and Bioengineering Gary L. Messing Distinguished Professor of Ceramic Science and Engineering Head of the Department of Materials Science and Engineering

* Signatures are on file at the Graduate School

Page 3: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

iii

ABSTRACT

This dissertation focuses on polymers synthesized from L-lactide and its

stereoisomers, i.e., poly L-lactide (L-PLA), L-lactide/meso-lactide random copolymers and

L-PLA/D-PLA blends. Height and phase images from tapping mode AFM experiments,

acquired at various tapping forces, reveal that the morphology and lamellar thickness of the

L-lactide/meso-lactide copolymers are strongly influenced by meso-lactide content and

crystallization conditions. Good agreement was observed between mean lamellar

thicknesses derived from AFM and previous small-angle X-ray scattering experiments. The

considerable decrease in lamellar thickness with increasing comonomer content suggests

that meso-lactide units are substantially excluded from the L-lactide crystal lattice, at least

under the isothermal crystallization conditions used here.

A dielectric study of amorphous and crystalline samples of the L-lactide/meso-

lactide copolymers demonstrates that the relaxation strength and mean relaxation time (and

its distribution) of the segmental process are modified by the presence of the crystalline

phase. The mean segmental relaxation time is longer and its distribution is broader in the

samples with higher degrees of crystallinity. The significant difference in the fraction of

mobile segments (determined from the relaxation strength) and the amorphous fraction

derived from DSC) indicates the existence of the so-called “rigid amorphous phase”. This

conclusion is also supported by the observation of a difference in temperature dependent

behavior of the relaxation strength of amorphous and crystalline samples. Differences in

the details of the relaxation processes, as well as the fraction of rigid amorphous phase of

Page 4: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

iv

these copolymers, suggest that the chain dynamics of amorphous segments are influence by

both degree of crystallinity and lamellar organization.

For D-PLA/L-PLAs blends, stereocomplex crystallites are observed along with

homopolymer crystals. The content of each crystal type is found to depend strongly on the

composition and molecular weight of the component polymers. These parameters have

only a small influence on the melting temperature of the stereocomplex but significantly

affect those of homopolymer crystallites. It is suggested by the results of real-time

dielectric crystallization experiments, along with WAXD and DSC measurements after

crystallization, that crystallization of the stereocomplex is faster than homopolymer

crystallites. Consequently, the homopolymer crystals developed under these conditions are

influenced significantly by constraints imposed by existing racemic crystallites and,

therefore, have lower melting temperatures than those of the neat polymers.

Page 5: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

v

TABLE OF CONTENTS

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

LIST OF TABLES ........................................................................................................ xv

ACKNOWLEDGEMENTS .........................................................................................xvi

CHAPTER 1 Introduction ............................................................................................... 1

1.1 Background ................................................................................................... 1

1.2 Polylactide stereoisomers ............................................................................. 5

1.3 L-PLA, D-PLA stereocomplex................................................................... 11

1.4 Chain conformation and crystal structure................................................... 16

1.5 Hydrolysis and enzymatic degradation ...................................................... 18

1.6 Objective and scope of this research .......................................................... 21

1.7 References .................................................................................................. 24

CHAPTER 2 Materials and Experimental procedure ................................................... 29

2.1 L-lactide/meso-lactide copolymers ............................................................. 29

2.1.1 Materials ....................................................................................... 29

2.1.2 Crystalline microstructure by tapping mode AFM....................... 29

2.1.3 Amorphous phase dynamics......................................................... 32

A. Sample preparation..................................................................32

Page 6: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

vi

B. Wide-angle X-ray diffraction ...................................................32

C. Differential scanning calorimetry............................................33

D. Dielectric measurements .........................................................33

2.2 Enantiomeric blends of L-PLA and D-PLA................................................ 34

2.2.1 Materials and sample preparation................................................. 34

2.2.2 Real-time study of crystallization from the glassy state............... 35

2.2.3 Differential scanning calorimetry................................................. 36

2.2.4 Wide-angle X-ray diffraction (WAXD) ....................................... 36

2.3 References ................................................................................................... 37

CHAPTER 3 Crystal morphology of Lactide Stereocopolymers by

Tapping mode AFM ...................................................................................................... 38

3.1 Overall morphology .................................................................................... 38

3.2 Contrast in height and phase images ........................................................... 42

3.3 Determination of lamellar thicknesses ........................................................ 45

3.3.1 Methodology ................................................................................ 45

3.3.2 L-PLA........................................................................................... 50

3.3.3 Meso-lactide copolymers ............................................................. 52

3.4 Bulk morphology......................................................................................... 55

3.5 Summary .................................................................................................... 57

3.6 References ................................................................................................... 58

Page 7: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

vii

CHAPTER 4 Amorphous Chain Dynamics of Amorphous and Crystalline L-PLA and

Stereocopolymers .......................................................................................................... 60

4.1 Relaxation processes by dielectric relaxation spectroscopy........................ 60

4.2 Sub-glass process ........................................................................................ 60

4.3 Segmental relaxation ................................................................................... 63

4.3.1 Crystallization from the glassy state.............................................63

4.3.2 Segmental dynamics of amorphous and crystallized samples......66

A. Temperature-dependent dielectric response ............................66

B. Influence of the Crystalline Phase ...........................................70

C. The rigid amorphous phase .....................................................76

4.4 Higher temperature relaxations ................................................................... 79

4.5 Summary ..................................................................................................... 84

4.6 References ................................................................................................... 85

CHAPTER 5 L-PLAs and D-PLA Blends .................................................................... 88

5.1 Crystallization from the glassy state............................................................ 89

5.2 Real-time crystallization investigated using DRS....................................... 90

5.2.1 Crystallization mechanism ...........................................................92

5.2.2 Rigid amorphous phase ................................................................98

5.2.3 Influence of crystalline phase on amorphous segments ...............99

5.3 Summary ................................................................................................... 100

Page 8: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

viii

5.4 References ................................................................................................. 101

CHAPTER 6 Summary and Suggestions for Future Research ................................... 102

6.1 Summary ................................................................................................... 102

6.2 Suggestions for future research ................................................................. 104

6.3 References ................................................................................................. 107

Page 9: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

ix

LIST OF FIGURES

Figure 1.1 Process steps in lactic acid polymer production ........................................... 2

Figure 1.2 (a) R-configuration and (b) S-configuration of lactic acid............................ 3

Figure 1.3 Stereoisomers of lactide ................................................................................ 3

Figure 1.4 Lamellar thicknesses (measured from AFM edge profiles) of L-PLA solution

grown crystals after annealing at different temperatures ................................................ 6

Figure 1.5 Melting temperatures as a function of degree of supercooling of PLLA

(= L-PLA), L-lactide/meso-lactide copolymers (B, C and D) and L-lactide/D-lactide

copolymers (E, F and G) ................................................................................................. 8

Figure 1.6 Spherulitic growth rates as a function of crystallization temperature of

PLLA, L-lactide/meso-lactide copolymers (B, C and D) and L-lactide/D-lactide

copolymers (E, F and G) ................................................................................................. 9

Figure 1.7 Bulk crystallinity as a function of degree of supercooling of PLLA,

L-lactide/meso-lactide copolymers (B, C and D) and L-lactide/D-lactide copolymers

(E, F and G) ................................................................................................................... 10

Figure 1.8 (a) DSC thermograms of blends of L-PLA and D-PLA. The blend ratio of

L-PLA to D-PLA is noted by the appropriate curve. (b) X-ray diffraction profiles for

a D-PLA homopolymer and L-PLA/D-PLA blends: (-⋅-) D-PLA homopolymer;

() 50/50 blend; (---) 25/75 L-PLA/D-PLA ................................................................ 12

Figure 1.9 Heats of fusion for (●) homopolymer and (○) stereocomplex crystallites

as a function of mole fraction of D-PLA (XD) for (a) solution cast films of

L-PLA/D-PLA blends (Mv ~ 2.6 × 104 for both homopolymers), melt crystallized

Page 10: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

x

at 140 oC and (b) the precipitate of L-PLA/D-PLA blends (Mv ~ 3.6 × 105 for both

homopolymers).............................................................................................................. 13

Figure 1.10 Heat of fusion of (●) homopolymer and (○) stereocomplex crystallites

as a function of molecular weight (XD = 0.5) for solution cast films of

L-PLA/D-PLA blends (the component polymers having similar molecular weight)

melt crystallized at 140 oC ............................................................................................. 14

Figure 1.11 DSC thermograms of films (prepared by solution casting) of blends of (A)

a D-rich PLA and (B) a L-rich PLA. Mole fractions of D-lactide (XD) in each of these

polymers are as indicated on the figure......................................................................... 15

Figure 1.12 Projections (a) along the helical axis of the 10/3 (on the left) and 3/1

(on the right) helical conformation and (b) perpendicular to the helical axis of

10/3 (top) and 3/1 (bottom) helical conformation of L-PLA ........................................ 18

Figure 3.1 AFM phase images of L-PLA isothermally crystallized at different

temperatures: (a)-(b) at 120 °C, (c)-(d) at 132 °C, and (e)-(h) at 147 °C...................... 39

Figure 3.2 Phase images of 6% meso-lactide copolymer crystallized at different

temperatures: (a)-(b) at 114 °C, (c)-(d) at 120 °C ......................................................... 41

Figure 3.3 Height and phase images of L-PLA crystallized at 126 °C: (a)-(b) at

a set-point amplitude ratio of rsp ~ 0.78 and (c)-(d) of rsp ~ 0.69, respectively ............. 44

Figure 3.4 (a) Height image and sectional profile and (b) phase image and sectional

profile of L-PLA crystallized at 120 °C ........................................................................ 46

Figure 3.5 (a) Height image and sectional profile and (b) phase image and sectional

profile of 3% meso-lactide copolymer crystallized at 114 °C....................................... 47

Page 11: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

xi

Figure 3.6 (a) Tip geometry from SEM and (b) effect of tip convolution on the

characterized feature...................................................................................................... 49

Figure 3.7 Average thickness of L-PLA lamellae determined from AFM height and

phase imaging, as well as that reported previously from SAXS,

as a function of Tc.......................................................................................................... 52

Figure 3.8 Mean lamellar thickness for PLLA, 3% and 6% meso-lactide copolymers

as a function of degree of supercooling from AFM phase images. The error bars

present the 95%confidence limits of these average values. .......................................... 53

Figure 3.9 Lamellar thickness distributions of (a)-(b) PLLA, (c)-(d) 3% meso-lactide

copolymer and (e)-(f) 6% meso-lactide copolymer as a function of Tc; data determined

in (a) and (c) from height images, (b),(d) and (f) from phase images and (e) from both

height and phase images of the 6% meso-lactide crystallized at 114 °C.

(x-axis: thickness (nm), y-axis: Tc or imaging mode, and z-axis: fraction of lamellar

thickness distribution) ................................................................................................... 54

Figure 3.10 (a)-(c) Height images of PLLA (Tc = 132 °C) microtomed at -25 °C and (d)

phase image of the same area as (c) .............................................................................. 56

Figure 4.1 Isochronal plots of (a) dielectric constant and (b) dielectric loss of L-PLA

crystallized at 141 oC in frequency range from 10 Hz to 10 kHz.................................. 61

Figure 4.2 log fm vs. 1/T of (a) the α-process (closed symbols) and β-process (open

symbols) of amorphous L-PLA ( ) and 3% meso-lactide ( ), 6% meso-lactide ( ) and

12% meso-lactide copolymers ( ). (b) The β-process of amorphous PLLA ( ), L-PLA

crystallized at 90 oC ( ) and the crystallized 12% meso-lactide copolymer crystallized

Page 12: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

xii

at 105oC (□) ................................................................................................................... 62

Figure 4.3 (a) Dielectric loss and (b) dielectric constant spectra during isothermal

crystallization of the 3% meso-lactide copolymer at 80 oC as a function of tc ............. 65

Figure 4.4 (a) Dielectric constant and (b) loss for amorphous L-PLA as a function of

frequency at the indicated temperatures ........................................................................ 67

Figure 4.5 (a) Dielectric constant and (b) loss for L-PLA crystallized at 120 oC as a

function of frequency at the indicated temperatures ..................................................... 68

Figure 4.6 Relaxation strengths of amorphous samples of L-PLA and the meso-lactide

copolymers at the indicated temperatures ..................................................................... 69

Figure 4.7 Mean relaxation times of amorphous L-PLA and meso-lactide copolymers

....................................................................................................................................... 70

Figure 4.8 Comparison between the relaxation time of the amorphous and crystallized

samples of (a) L-PLA, (b) 3%, (c) 6% and (d) 12% meso-lactide copolymer .............. 71

Figure 4.9 ε″/ε″m vs. log (f/fm) for (a) amorphous and crystallized L-PLA, and

(b) amorphous and crystallized 12% meso-lactide copolymer...................................... 75

Figure 4.10 HN broadening parameter, a, of amorphous (closed symbols) and L-PLA,

3% and 6% meso-lactide copolymers crystallized at Tc = 120 oC, and 12% meso-lactide

copolymer crystallized at Tc = 117 oC (open symbols used for crystallized samples).

( ) L-PLA, ( ) 3% meso-lactide, ( ) 6% meso-lactide and ( ) 12% meso-lactide .... 76

Figure 4.11 Fraction of mobile phase as a function of temperature. Numbers in the

figure legend correspond to the sample number listed in Table 4.1..............................77

Figure 4.12 Mobile segment fraction vs. crystalline fraction for all samples investigated

.......................................................................................................................................78

Page 13: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

xiii

Figure 4.13 Plots of (a) log ε″ (closed symbols) and log ε″der (open symbols) and

(b) experimental ε′ vs. frequency at 130 ( ) and 155 oC ( ) for the amorphous

12% meso-copolymer. Data of ε″ were multiplied by 10 to offset its plots from

the ε″der plots..................................................................................................................81

Figure 4.14 Plots of ε″der vs. frequency at 125 ( ), 135 ( ) and 145 oC ( ) for (a)

crystalline L-PLA (Tc = 90 oC) and (b) 12% meso-lactide copolymer (Tc = 105 oC).

Experimental ε´ of 12% meso-lactide copolymer (Tc = 105 oC) is shown in (c) ..........83

Figure 5.1 Experimental Tgs of L-PLA1/D-PLA ( ) and L-PLA2/D-PLA ( ) as

a function of D-PLA composition. Predictions from a simple Fox-Flory mixing rule

are shown by the solid lines. ........................................................................................88

Figure 5.2 DSC heating thermograms illustrating cold crystallization and melting

processes of initially amorphous samples; (a) neat D-PLA, L-PLAs, (b) HB blends

and (c-d) LB blends. ...................................................................................................... 89

Figure 5.3 Dielectric loss spectra acquired during isothermal crystallization of

the 70:30 L-PLA1/D-PLA blend at 60 °C as a function of tc. Crystallization times

(in min) are shown in the figure legend......................................................................... 91

Figure 5.4 Fitting results of the dielectric spectra of LB73 crystallized at Tiso = 60 °C

as a function of tc; (a) mean relaxation time, (b) broadening parameter “a”, skewness

parameter “b” and segmental relaxation strength (∆ε). ................................................. 91

Figure 5.5 Rigid segment fractions as a function of crystallization time of

(a) homopolymers L-PLA2 ( ) and D-PLA ( ) and (b) L-PLA2/D-PLA blends

Page 14: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

xiv

(HB55 ( ), HB64 ( ), HB73 ( ) and L-PLA2 ( )); at Tiso = 70 °C (open symbols)

and 80 °C (closed symbols)........................................................................................... 92

Figure 5.6 Rigid fractions of D-PLA and LB blends as a function of crystallization

time at a similar TDiff (= 10˚C); Tiso = 70, 65, 60 and 57.5 °C for DPLA, LB37, LB55

and LB73, respectively. ................................................................................................. 93

Figure 5.7 ∆Hm of LB blends after crystallization in real-time DRS experiments

at Tiso = 70 °C; Closed and open symbols represent the stereocomplex and

the homopolymer crystals, respectively. ....................................................................... 95

Figure 5.8 DSC thermograms presenting the melting process (Tm1) of homopolymer

crystallites of (a) neat D-PLA, and the L-PLAs (at Tiso = 70 °C) and (b) HB64

(1st and 2nd curves from the top) and HB73 blends (3rd and 4th from the top)

crystallized at 70 (1st and 3rd curve) and 80 °C (2nd and 4th curve) ...............................96

Figure 5.9 Melting points of stereocomplex crystals (Tm2) formed in HB (open symbols)

and LB blends (closed symbols) crystallized at 70 ( ) and 80 °C ( ). ...................... 97

Figure 5.10 DSC thermograms presenting the melting process (Tm1) of homopolymer

crystallites of LB blends at Tiso = 70 °C........................................................................ 98

Figure 5.11 Difference in Tg between crystallized and amorphous samples as

a function of total percent crystallinity determined from WAXD for ( ) LB37,

( ) LB55, ( ) LB73, ( ) HB55, ( ) D-PLA and ( ) L-PLA2. The line in

this figure indicates the relationship between ∆Tg and crystalline content for

LB55 and HB55 in which only stereocomplex crystals are present............................ 100

Figure 6.1 DSC thermograms of LB55 crystallized from the melt at (a) 140 oC,

(b) 160 oC and (c) 170 oC for different tc .................................................................... 106

Page 15: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

xv

LIST OF TABLES

Table 1.1 Stereo-specificity of lactic acid isomers derived from various organisms

in the genus Lactobacilli ................................................................................................. 4

Table 1.2 Nominal comonomer content, R-stereoisomer content and molecular weights

of lactide copolymers ...................................................................................................... 8

Table 1.3 Unit cell dimensions of the α-form of L-PLA.............................................. 16

Table 2.1 Characteristics of L-PLA and L-lactide/meso-lactide copolymers .............. 30

Table 2.2 Isothermal crystallization temperatures and times used in this study;

(identical to those used in ref. [1])................................................................................. 31

Table 3.1 Mean lamellar thickness and distribution from AFM phase and

height images................................................................................................................. 51

Table 4.1 Crystalline, mobile and RAP fractions along with VFT parameters and

fragilities for PLLA and the meso-lactide copolymers, isothermally crystallized at

the designated temperatures .......................................................................................... 72

Table 5.1 Glass transition temperature, crystal fractions of stereocomplex (Xc2)

and homopolymer crystallites (Xc1) determined by WAXD and DSC, as well as

the rigid fraction (frigid) estimated from DRS for quenched and crystallized HB

and LB blends................................................................................................................ 94

Page 16: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

xvi

ACKNOWLEDGEMENTS

I would like to thank my advisor, Prof. James Runt for his guidance,

encouragement and kindness during my time here at Penn State. I am grateful to Prof. Ian

Harrison, Evangelos Manias and Christopher Siedlecki for serving as thesis committee.

Special thanks go to Dr. Manias and Ken Strawhecker for their helps and invaluable

suggestions on AFM technique. I would also like to thank many students and post-docs in

Polymer Science Program at Penn State for their useful comments and supports, especially

James Garrett, Jin Xing, Shihai Zhang and Jaedong Cho.

I appreciate friendship and supports from many Thai friends; Nopparat

Plucktaveesak, Atitsa Petchsuk, Rattana Tantatherdtam, Ratda Suhataikul, Cattleya

Kongsupapsiri, Preuchsuda Suphantharida, Wallop Promsen, Pornpen Atorngitjawat, and

Pakorn Opaprakasit. They have made my life at Penn State delightful and memorable. I

would like to acknowledge the Royal Thai Government for the financial support.

Last but not least, I am indebted to my parents, brothers and sisters for perpetually

cherishing, encouraging and supporting me throughout my whole life. I dedicate this work

to my father and mother, Anuchit and Kanjana Kanchanasopa.

Page 17: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

Chapter 1

Introduction

1.1 Background

Polylactic acids or polylactides are well known as hydrodegradable and

biocompatible thermoplastics. They have typically been named by attaching the prefix

“poly” to the name of monomer from which they are derived. Thus, polylactic acid is the

polymer derived from lactic acid, and a polylactide is synthesized from a lactide (i.e., a

cyclic dimer of lactic acid), although sometimes a lactide polymer is referred to as

polylactic acid, since both have the same constituent repeating unit H-[OCH(CH3)CO]n-

OH.

Lactic acid based polymers are of interest in medical applications such as surgical

sutures [1], implants for bone fixation [2,3] and controlled release systems [4,5]. Many

studies [6-9] have revealed that hydrolysis of poly (L-lactide) (L-PLA) is catalyzed by

specific enzymes, e.g. proteinase K (a fungal protease from the mold Tritirachium album).

Their enzymatic degradability and mechanical properties comparable to those of

polyethylene and polystyrene [8] have attracted interest in using these degradable polymers

in agriculture as mulch film or in everyday life as a packaging material. Perhaps most

importantly, this family of polymers has received attention since the monomers can be

derived from carbohydrates, which are constituents of renewable agricultural sources such

as corn, potato starch, cane, cheese whey, etc [10].

Page 18: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

2

Figure 1.1 Process steps in lactic acid polymer production [10]

A typical production process for the synthesis of lactic acid based polymers is

presented schematically in Figure 1.1. The lactic acid monomers are produced via

fermentation by microorganisms. The genus Lactobacillus is an industrially well-known

strain used for lactic acid production. Due to the presence of a chiral carbon atom, lactic

acid occurs in two different configurations, S and R-lactic acid, as shown in Figure 1.2. As

seen in Table 1.1, it is obvious that choice of microorganism has a profound affect on the

isomer obtained. Since lactic acids possess both hydroxyl and carboxyl groups, they are

feasibly polymerized via a polycondensation reaction. However, only relatively low

molecular weight polymers result from this reaction scheme. High molecular weight

polymers can be efficiently synthesized by means of catalytic ring-opening polymerization

of lactides. This process can be conducted by cationic [11], anionic [12,13] and non-ionic

insertion mechanisms [11,14,15]. Lactides are cyclic acid dimers and the possible lactide

stereoisomers are presented in Figure 1.3.

Fermentation

Lactic acid recovery and purification

Lactide formation and purification

Polymerization

Sugars Nutrients

Polylactide, Poly (lactic-acid)

Page 19: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

3

(a)

CHOH

CH3

COOH

COH

HCOOH

CH3

1 2

3

(b)

CHOH

CH3

COOH

C

HOHHOOC

CH3

12

3

Figure 1.2 (a) R-configuration and (b) S-configuration of lactic acid

L - lactide D – lactide meso - lactide

Figure 1.3 Stereoisomers of lactide

O

O O

O CH3

CH3R

R

O

O O

O CH3

CH3S

R

O

O O

O CH3

CH3S

S

Page 20: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

4

Table 1.1 Stereo-specificity of lactic acid isomers derived from various organisms in the

genus Lactobacilli [10]

Lactic acid isomer a Organisms

R

S

R/S

L. delbrueckii subsp. delbrueckii, L. delbrueckii subsp. bulgaricus,

L. jensenii, L. coryifomis subsp. corniformis

L. amylophilus, L. salivarius, L. agilis, L. bavaricus, L. casei

subsp. casei, L. maltaromicus

L. acidophilus, L. amylovorus, L. helveticus, L. curvatus, L.

homohiochii, L. plantarum

a R or S : the isomer reportedly makes up 90% or more of the total lactic acid

R/S : 25-75% of the total lactic acid is S(+)-lactate

The possibility of racemization of L-lactide during the polymerization reaction has

been reported. Kricheldorf, et al. [12,16] found that it preferably occurs in both cationic

and anionic polymerizations due to initiator-induced racemization of the monomer.

However, racemization occurs much less frequently in the non-ionic insertion mechanism

in the presence of weak Lewis acids such as aluminum, tin, and zinc alkoxides, as well as

SnCl2, SnBr2, SnCl4, or SnBr4.

This brief introduction clearly illustrates that the family of polylactides

encompasses polymers with various chemical structures depending on feed composition,

and polymerization conditions. Details of the impact of the chemical structures on physical

and chemical properties of lactide stereocopolymers will be discussed further in section

1.2. In addition, Ikada, et al. have shown [17] that a racemic crystal structure

(stereocomplex) with melting temperature approximately 50 °C higher than that of either

lactide homopolymer is observed in physical mixtures of poly L-lactide (L-PLA) and poly

Page 21: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

5

D-lactide (D-PLA). Due to their higher temperature properties, the stereocomplex has

received some attention as a potential high performance biodegradable material. Thus

some background knowledge of L-PLA/D-PLA blends is important and included in section

1.3. Chain conformation and crystal structure of the crystalline phases in non-blended and

blended polylactides are summarized in section 1.4. Since degradability of polylactides is

relevant to a number of their important applications, a very large number of papers have

been published on this topic. Therefore, a summary of hydrolysis and enzymatic

degradation of polylactide is included in section 1.5, although the main objective of the

research in this thesis (section 1.6) is not associated with this issue.

1.2 Polylactide stereoisomers

Extensive studies of crystallization behavior and morphology of optically pure

polylactide has been carried out from the bulk and solution. It was suggested in an early

study that the equilibrium melting temperature (Tmo) of L-PLA is approximately 215 °C,

estimated using a Hoffmann - Weeks extrapolation method [18]. Iannace, et al. [19]

conducted an investigation of the bulk crystallization kinetics at various crystallization

temperatures (Tc). Applying the Avrami equation, Xr(t) = 1 - exp(-Ktn), to experimental

relative crystallinity Xr(t) vs crystallization time data (t), the mode of nucleation along with

the geometry of the crystal growth was identified. Their results indicate that the maximum

isothermal crystallization rate occurs at 105 °C and the Avrami exponent at all Tc (90 - 130

°C) is close 3, consistent with spherulitic growth geometry and instantaneous nucleation.

By analyzing spherulite growth rates in the context of the Lauritzen-Hoffman

crystallization model [20] molecular weight was proposed to have a significant influence

Page 22: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

6

on crystallization regime [21]. A transition from regime II to regime I growth was observed

at 163 °C for L-PLA with Mv ~ 1.5 × 105, together with a change in growth morphology

(from spherulites to axialites) at the same temperature. For high molecular weight L-PLAs

(Mv ~ 2.6 - 6.9 × 105) the analysis suggests crystallization regime II for all Tc investigated.

Miyata, et al. [22] have reported that the fold surface free energy (σe) of low molecular

weight L-PLA (Mw 5.0 x 104) is smaller than that of crystals formed from higher molecular

weight L-PLA. These authors argue that this implies that the fold surface associated with

high molecular weight crystals is composed predominately of loose-loop folds.

Figure 1.4 Lamellar thicknesses (measured from AFM edge profiles) of L-PLA solution

grown crystals after annealing at different temperatures [23].

Lamellar single crystals have been grown by isothermal crystallization of L-PLA

from dilute solution [18,23]. Lamellar thicknesses, discerned from the length of shadows

in electron micrographs or from edge profiles using atomic force microscopy (AFM),

confirm that lamellae crystallized from solution at higher temperatures are thicker, as are

those annealing in the solid state at higher temperatures (see Figure 1.4 for example).

Page 23: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

7

In contrast to the polylactide homopolymer, the investigation of crystallization

behavior and lamellar microstructure of polylactide stereocopolymers has been conducted

only infrequently. Solution grown single crystals of L-lactide/meso-lactide

stereocopolymers were investigated ~ 30 years ago by Fischer, et al. [24]. Small angle X-

ray scattering (SAXS) and DSC experiments revealed that the long spacing (L), crystal

thickness (lc) and melting temperature of these single crystals are influenced by meso-

lactide content, in addition to Tc. At constant degrees of supercooling (∆T), only small

variations of L were reported between the various random copolymers, whereas lc

decreased systematically with increasing XD (up to 15% comonomer unit). ∆T was

operationally defined in this paper as Ts-Tc, where Ts is the observed dissolution

temperature of the crystals. Fischer, et al. also reported that the melting temperatures

decreased with increasing comonomer content and that the unit cell lattice parameters of

these single crystals were not influenced by the presence of the meso-lactide comonomer.

It was concluded in this investigation that R stereoisomers are distributed in the amorphous

phase as well as partially incorporated in the crystalline phase. However, it was also

reported that rejection of R-counits from the crystalline phase increased significantly with

increasing ∆T.

In previous research conducted by our group [25-28], isothermal melt-

crystallization of random L-lactide/D-lactide and L-lactide/meso-lactide copolymers (Table

1.2) was performed in order to investigate the impact of the comonomer units on

crystallization kinetics, lamellar microstructure and melting behavior.

Page 24: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

8

Table 1.2 Nominal comonomer content, R-stereoisomer content and molecular weights of

lactide copolymers [27]

Figure 1.5 Melting temperatures as a function of degree of supercooling of PLLA (= L-

PLA), L-lactide/meso-lactide copolymers (B, C and D) and L-lactide/D-

lactide copolymers (E, F and G) [27]

As seen in Figure 1.5, the experimental melting temperatures of the copolymers

decrease with increasing comonomer content. Combined with crystal thicknesses (lc)

derived from small-angle x-ray scattering (SAXS) results, experimental melting points

Page 25: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

9

were used along with the Gibbs - Thomson expression to estimate equilibrium melting

temperatures and fold surface free energies. The results reported in ref. 26 - 27 clearly

show that Tmo and σe decreased dramatically with increasing comonomer content. It was

proposed that the reduction in σe is associated with increasing defect concentration at the

crystal interfaces.

Figure 1.6 Spherulitic growth rates as a function of crystallization temperature of PLLA,

L-lactide/meso-lactide copolymers (B, C and D) and L-lactide/D-lactide

copolymers (E, F and G) [27]

Page 26: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

10

Figure 1.7 Bulk crystallinity as a function of degree of supercooling of PLLA, L-

lactide/meso-lactide copolymers (B, C and D) and L-lactide/D-lactide

copolymers (E, F and G) [27]

Figures 1.6 and 1.7 present spherulite growth rates and bulk crystallinities of these

copolymers, respectively. L-PLA exhibits a much higher growth rate and crystallinity than

the copolymers. In addition, the temperature at which the maximum growth rate occurs is

shifted somewhat to lower temperatures with increasing meso- and D-lactide content.

Comparison at equivalent comonomer contents of L-lactide/D-lactide copolymers with L-

lactide/meso-lactide copolymers, i.e., copolymers F vs. B and G vs. C in the figures,

reveals that bulk crystallinities and spherulitic growth rates of copolymers with bulkier D-

lactide defects are lower. Similar to the results from solution grown single crystals [24],

lamellar thicknesses (from SAXS) of crystals isothermally melt-crystallized at a similar

degree of supercooling were found to decrease significantly with increasing comonomer

content. Experimental Tmo’s were compared to values predicted from several copolymer

melting models, in order to provide insight into the location of the stereochemical defects.

The result showed that predicted equilibrium melting points from the Flory exclusion [29]

Page 27: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

11

and uniform inclusion models [30] were much larger than the experimental Tmo’s.

However, the experimental values were in very good agreement with predictions from the

Wendling - Suter model [31] in the limit of complete comonomer exclusion. This finding

suggests that D- and meso-lactide co-units are significantly rejected from L-lactide

crystals.

1.3 L-PLA, D-PLA stereocomplex

Complexes are known to form between two complementing stereoregular

polymers, enhanced by electrostatic forces, hydrogen bonding, or van der Waals

interactions. Thus, a stereocomplex generally has different physicochemical properties

compared to the parent polymers. Most of such complexes are insoluble in solvents that

dissolve the stereospecific parent polymers, and they possess melting temperatures that are

higher than those of the individual component polymers.

Reports of stereocomplex formation have appeared for blends of isotactic and

syndiotactic polymers (e.g. poly (methyl methacrylate) [32]) and between two opposite

enantiomeric polymers (e.g. poly (α-methyl-α-ethyl-β-propiolactones) [33], poly (γ-benzyl

glutamate) [34], poly (olefin-carbon monoxide)s [35], poly (epichlorohydrin) [36], as well

as polylactide [17]). Stereocomplexes have also been reported between polymers with

opposite stereoregular configurations from different families, e.g. L-polypeptides and D-

polylactide [37]. However, not all optically active polymers are capable of stereocomplex

formation, such as isotactic poly (β-hydroxybutyrate) and poly (β-hydroxyvalerate) [36],

and isotactic poly ((R,S)-propyleneoxide)s) [38].

Page 28: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

12

The racemic crystal (stereocomplex) formed from D- and L-PLA was first reported

by Ikada, et al. [17]. In this initial paper, each enantiomeric polylactide was dissolved

separately in methylene chloride (at concentration of 1.0 g/dl), and the solutions were then

mixed in different volume ratios. DSC and WAXD evaluation of the precipitates from

these mixed solutions confirmed the presence of stereocomplex crystals. Figure 1.8a shows

that the melting endotherm of both homopolymers appears at ~180 °C, whereas a new

melting endotherm at ~230 °C was observed for the blends. The X-ray diffraction pattern

of the 50:50 blend (Figure 1.8b), in which the stereocomplex solely developed, is very

different from that of the homopolymers. In accord with the DSC results, the diffraction

pattern of the 25/75 L-PLA/D-PLA mixture exhibits the characteristics of both the D-PLA

homopolymer and 50:50 blend.

(a) (b)

Figure 1.8 (a) DSC thermograms of blends of L-PLA and D-PLA. The blend ratio of L-

PLA to D-PLA is noted by the appropriate curve. (b) X-ray diffraction

profiles for a D-PLA homopolymer and L-PLA/D-PLA blends: (-⋅-) D-PLA

homopolymer; () 50/50 blend; (---) 25/75 L-PLA/D-PLA [17].

Page 29: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

13

Since the publication of ref 17, a series of studies based on solution casting [39,40]

and precipitation of mixed L-PLA/D-PLA solutions with nonsolvent [41,42] have been

conducted to understand how stereocomplex formation is influenced by blending ratio,

molecular weight and optical purity. Research on melt crystallization of L-PLA/D-PLA

blends obtained from solution cast films has been carried out as well [43-45].

(a) (b)

Figure 1.9 Heats of fusion for (●) homopolymer and (○) stereocomplex crystallites as a

function of mole fraction D-PLA (XD) for (a) solution cast films of L-PLA/D-

PLA blends (Mv ~ 2.6 × 104 for both homopolymers), melt crystallized at 140

°C [43] and (b) the precipitate of L-PLA/D-PLA blends (Mv ~ 3.6 × 105 for

both homopolymers) [41].

As seen in Figure 1.9, stereocomplex formation was found to be favored in blends

with a component ratio near 1:1, regardless of molecular weight and crystallization

conditions (precipitation and melt crystallization of solution cast films). It is interesting to

note that solely racemic crystallites are formed in blends of low molecular weight polymers

at XD = 0.4 and 0.6 (Figure 1.9a). The authors suggested that the ‘excess’ homopolymer

Page 30: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

14

chains are trapped at the boundary of the racemic crystallites without sufficient space for

homopolymer crystallization. The content of stereocomplex and homopolymer crystallinity

in L-PLA/D-PLA blends at XD = 0.5 for melt-crystallized blends of homopolymers of

similar molecular weight is shown in Figure 1.10. Clearly, these crystallinities are strongly

dependent on molecular weight. Similar behavior has been observed for solution cast films,

the only difference is that the critical molecular weight below which only the

stereocomplex is formed, is ~4 × 104 from solution casting [40]. This is much higher than

that of the melt crystallized blends (Figure 1.10). It was reported that racemic

crystallization takes place during the drying process in solution casting [41]. In blends of

higher molecular weight polymers, microscopic phase separation into L-PLA and D-PLA

rich regions is likely during solvent evaporation, and homopolymer crystallization is

therefore favored.

Figure 1.10 Heats of fusion of (●) homopolymer and (○) stereocomplex crystallites as a

function of molecular weight (at XD = 0.5) for solution cast films of L-

PLA/D-PLA blends (the component polymers having similar molecular

weight) melt crystallized at 140 °C [43].

Page 31: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

15

Figure 1.11 DSC thermograms of films (prepared by solution casting) of blends of (A) a

D-rich PLA and (B) a L-rich PLA. Mole fractions of D-lactide (XD) in each of

these polymers are as indicated on the figure [46].

Figure 1.11 shows the influence of optical purity on stereocomplex formation. It

was found that the stereocomplex is also formed in blends composed of D-rich PLA and L-

rich PLA [46]. The melting endotherm associated with racemic crystallites becomes

smaller and shifts to lower temperatures as XD of component A and B approach 0.5, and

finally disappears in the blend of polymers with XD(A) = 0.83 and XD(B) = 0.14. The study

of isothermal melt-crystallization of these blends revealed that the temperature range over

which crystallization can be achieved is extended to higher temperatures in the blends [44].

This lead to the idea of enhancing the crystallization rate of the polylactide homopolymer

by adding a small amount of its enantiomer: it has been reported that stereocomplex

crystals develop at temperatures between Tm of the homopolymer and racemic crystallites

are potential nucleation sites for homopolymer crystallization [47,48]. Their nucleation

Page 32: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

16

efficiency was found to be higher than that of talc, commonly used as a nucleating agent

for polymers.

1.4 Chain conformation and crystal structure

The unit cell structure of the lactide homopolymers, L-PLA and D-PLA have been

widely investigated. These studies have been carried out using wide angle X-ray diffraction

techniques on fibers [49-52] and films [53]. Electron diffraction of solution growth single

crystals has also been utilized [54,55]. Unit cell types and dimensions reported for L-PLA

are summazied in Table 1.3. It should be noted that L-PLA and D-PLA homopolymers

crystallize analogously, resulting in an identical diffraction pattern and crystal structure.

Table 1.3 Unit cell dimensions of the α-form of L-PLA

Unit cell dimensions (Å) Authors

a b c

Unit cell Ref.

Samples

De Santis and Kovacs

Hoogsteen et al.

Kobayashi et al.

Marega et al.

Miyata et al.

Kalb et al.

10.70

10.60

10.50

10.70

10.78

10.34

6.45

6.10

6.10

6.13

6.04

5.97

27.8

28.8

28.8

28.9

Pseudo-orthorhombic

Pseudo-orthorhombic

Pseudo-orthorhombic

Pseudo-orthorhombic

Orthorhombic

Pseudo-orthorhombic

50

51

52

53

54

55

Fiber

Fiber

Fiber

Film

Single crystal

Single crystal

De Santis and Kovac [50] built a molecular model of the L-PLA helical structure

based on a conformational analysis. This model was verified by comparing a fiber

diffraction pattern to the predicted diffraction pattern calculated by taking the Fourier

transform of an energy minimized atomic arrangement. Even though some additional

reflections are observed in the diffraction pattern of the fiber, the experimental and

Page 33: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

17

predicted patterns were found to be in satisfactory agreement. The unit cell (the so-called α

form) contains two chains exhibiting 10/3 helical conformations. This structure has been

confirmed in a number of other publications [51-53].

A β-crystal structure has also been reported for L-PLA spun fibers [49]. The

diffraction patterns of melt-spun fibers are found to correspond to the usual α-crystal form,

while both α and β-forms are present in the solution-spun fibers. The details of the crystal

structure and chain conformation of the β-form has been investigated by Hoogsteen, et al.

[51]. The ratio of α and β-phase depends strongly on the draw ratio and temperature.

Based on the fiber diffraction pattern, it was concluded that the β-phase has an

orthorhombic unit cell with a = 10.31, b = 18.21, and c = 9.00 Å. The unit cell contains six

chains in a 3/1 helical conformation. The 10/3 and 3/1 helical projections along the helical

axis are shown in Figure 1.12a and those perpendicular to the helical axis are present in

Figure 1.12b. Recently, a new crystal phase, the γ-form, was reported by Cartier, et al. [56]

for L-PLA single crystals, epitaxial crystallized on hexamethylbenzene.

Differences in morphology, crystal structure and chain conformation of

homopolymer and racemic crystallites have been reported by Brizzolara, et al. [57]. Single

crystals of the stereocomplex and homopolymer exhibit a triangular shape and a lozenge

(hexagonal) form, repectively. On the basis of molecular simulations and experimental

diffraction patterns, it was proposed that the racemic crystallite is formed by packing 3/1

helices of opposite absolute configuration alternately side by side. It was reported earlier

[58] that the L- D-PLA stereocomplex crystallizes in a triclinic unit cell with parameters a

= 9.16, b = 9.16, and c = 8.70 Å, α = β = 109.2o and γ = 109.8o. The 3/1 helical

Page 34: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

18

conformation present in the stereocomplex has been confirmed by vibrational spectroscopy

(IR and Raman) as well [59].

(a) (b)

Figure 1.12 Projections (a) along the helical axis of the 10/3 (on the left) and 3/1 (on the

right) helical conformation and (b) perpendicular to the helical axis of 10/3

(top) and 3/1 (bottom) helical conformation of L-PLA [51].

1.5 Hydrolysis and enzymatic degradation

Hydrolytic degradation of polylactides has been studied in many laboratories. It

was found that the non-enzymatic degradation mechanism of these polymers depends on

material dimensions and environmental conditions. Based on in vitro hydrolysis under

phosphate-buffered aqueous conditions (pH 7.4), heterogeneous degradation or faster

internal degradation was observed in thicker (thickness ≥ 2 mm) specimens of L-lactide/D-

lactide copolymers (XD ∼ 0 to 0.50) [60,61], while a homogeneous bulk-erosion

Page 35: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

19

mechanism was reported for thinner films (thickness < 150 µm) [62-64]. In contrast, the

change in molecular weight distribution and surface morphology of L-PLA films during

hydrolysis revealed that hydrolysis in dilute alkaline solution proceeded mainly via a

surface erosion mechanism [65].

In the case of heterogeneous degradation, the polymer matrix degrades

macroscopically homogeneous at the very beginning. However, the situation becomes

markedly different when soluble oligomeric compounds are generated in the matrix by the

hydrolytic cleavage of ester bonds. The soluble oligomers that are close to the surface can

escape from the matrix before complete degradation, while it is very difficult for those

located in the interior of the specimen to diffuse out. This results in greater acidity inside

the specimen than at the surface. When the aqueous medium is buffered at neutral (pH

7.4), the neutralization of carboxylic groups present at the surface also contributes to the

decrease of the surface acidity. Therefore autocatalysis in which new carboxylic end

groups form via ester bond clevage accelerates the hydrolytic reaction of the remaining

ester bonds in the bulk compared to the rate at the surface.

Tsuji, et al. conducted systematic studies to investigate the effects of stereoisomer

content and morphology on homogeneous bulk-erosion. These authors showed [64] that

the autocatalytic hydrolyzabilities of amorphous polylactides decreased in the following

order: D-lactide/L-lactide copolymers > L-PLA, D-PLA > L-PLA/D-PLA (1:1) blend. The

greater hydrolyzability of the copolymer compared with those of the homopolymer was

ascribed to the lower tacticity of the former, which decreases intramolecular interactions,

and therefore the copolymer chains are more susceptible to attack from water. Likewise,

these authors argued that the retarded hydrolysis of the blend is a consequence of the

Page 36: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

20

strong interactions between L-PLA and D-PLA chains. To understand the influence of

morphology, the hydrolysis of fully amorphous (only “free” amorphous regions), partially

and fully crystallized (no “free” amorphous regions) samples of optically pure L-PLA

(specimens 50 µm thick) were followed by monitoring weight loss and changes in

molecular weight as a function of hydrolysis time [62-63]. Results suggested that

hydrolysis in partially crystallized samples occurred predominantly in the amorphous

region between the crystallites at a rate in excess of that of the fully amorphous specimen.

Additionally, higher hydrolysis rate was observed in the fully crystallized samples (at

higher Tc) with higher crystallinity. These results were explained by proposing that more

defects are introduced into the non-crystalline regions of higher crystallinity materials [62]

and these defects enhance water penetration into amorphous regions resulting in more

rapid hydrolysis (i.e., shorter induction period and higher hydrolysis rate). The accelerated

hydrolysis of higher crystallinity polylactides has also been attributed to the enhanced

catalytic effect and water uptake facilitated by the increased density of terminal groups

(hydroxyl and carboxylic acid) in the amorphous phase [63].

Unlike non-enzymatic hydrolysis, enzymatic degradation (e.g. by proteinase K)

proceeds mainly via a surface erosion mechanism and primarily in ‘free’ amorphous

regions [8]. Moreover, the enzymatic hydrolysis rate decreases with increasing the initial

Xc [8,66-68]. Studies of non-crystalline lactide copolymers (with high comonomer

contents) indicated that proteinase K preferentially degrades S-S linkages, followed by S-R

(or R-S), bonds as opposed to R-R ones [9,69].

Page 37: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

21

1.6 Objective and scope of this research

As noted in the previous sections, the physical properties of the semi-crystalline

polylactides can be tailored by modifying their chemical and morphological structures.

Fundamental knowledge of these matters is crucial for successful design of these materials

for specific applications. Consequently, the primary objective of the research in this thesis

is to develop an understanding of the influence of crystallization conditions on the solid-

state structure of L-lactide/meso-lactide copolymers and enantiomeric L-PLA/D-PLA

blends by means of atomic force microscopy (AFM), DSC and WAXD. In addition, the

influence of the crystalline phase on the chain dynamics of these polymers is investigated

by dielectric relaxation spectroscopy. Details of experimental procedures are presented in

Chapter 2.

In previous SAXS studies of L-lactide/meso-lactide and L-lactide/D-lactide

copolymers by our research group [26,27], lamellar thicknesses were found to decrease

significantly with increasing comonomer content. However, determination of

microstructural parameters from reciprocal space SAXS data requires a model structure

(e.g. ‘infinite’ lamellae stacks) and there is a degree of skepticism regarding the

transformation of SAXS data to 1D correlation or interface distribution functions.

Although care was taken in the analyses in these earlier studies, independent confirmation

of these earlier findings is of importance.

Studies of lamellar structure have been limited by the availability of high

magnification experimental probes, which can directly image at the nanometer length

scale. Since AFM has comparable spatial resolution to TEM and no special sample

preparation is required, this technique, particularly tapping mode AFM, has received

Page 38: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

22

considerable attention recently in the study of the lamellar morphology of semi-crystalline

polymers [70-72]. In Chapter 3 (published in Biomacromolecules 2003, 4, 1203), an

investigation of the lamellar microstructure of L-PLA and two random L-lactide/meso-

lactide copolymers by tapping mode AFM is presented. Mean lamellar thickness and

distribution, as they are influenced by comonomer content and Tc, are characterized and

comparison is made between the results of the present work and those from earlier SAXS

experiments [26].

In Chapter 4, the influence of crystalline content and crystallization conditions on

the amorphous phase dynamics of these polymers is reported (published in

Macromolecules 2004, 37, 863). Differences in the characteristics of amorphous chain

segments in fully amorphous and semi-crystalline samples are discussed in the context of

the dielectric relaxation strength, and mean relaxation time and its distribution. Moreover,

the presence of the so-called rigid amorphous phase is revealed from analysis of the

fraction of mobile phase as a function of degree of crystallinity. Conclusions are arrived at

based on results from a variety of semi-crystalline samples (with crystallinity ranging from

30 – 70%) by varying meso-lactide content and crystallization conditions.

Among the literature reviewed earlier in this chapter, studies of blends of L-PLA

and D-PLA of dissimilar molecular weight are relatively rare. Chapter 5 focuses on blends

of relatively high molecular weight D-PLA and two different molecular weight L-PLAs, in

order to investigate how their crystallization behavior is influenced by molecular weight,

composition and crystallization conditions. Real-time crystallization of these blends from

the glassy state is monitored by dielectric relaxation spectroscopy. The fractions of

stereocomplex and homopolymer crystallites developed in the DRS experiments are

Page 39: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

23

examined by DSC and WAXD. Modification of the melting and glass transition

temperatures as a function of developed crystal phase is investigated as well.

Finally, a summary and some suggestions for future research will be included in

Chapter 6.

Page 40: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

24

1.7 References

1. Pennings, J. P., Dijkstra, H. and Pennings, A. J. Polymer 1993, 34, 942.

2. Leenslag, J. W., Penning, A. J., Bos, R. M., Rozema, F. R. and Boering, G.

Biomaterials 1987, 8, 311.

3. Winet, H. and Bao, J. Y. J. Biomed. Mater. Res. 1998, 40, 567.

4. Herrmann, J. and Bodmeier, R. J. Control. Release 1995, 36, 63.

5. Miyajima, M., Koshika, A., Okada, J. and Ikeda, M. J. Control. Release 1999, 61, 295.

6. Reeve, M., McCarthy, S., Downey, M. and Gross, R. Macromolecules 1994, 27, 825.

7. MacDonald, R., McCarthy, S. and Gross, R. Macromolecules 1996, 29, 7356.

8. Tsuji, H. and Miyauchi, S. Polym. Degrad. Stab. 2001, 71, 415.

9. Li, S., Tenon, M., Garreau, H., Braud, C. and Vert, M. Polym. Degrad. Stab. 2000, 67,

85.

10. Mobley, D. P. Plastic from microbes; Microbial synthesis of polymers and polymer

precursors; Hanser publishers: New York, 1994

11. Kricheldorf, H. R. and Kreiser, I. Makromol. Chem. 1986, 187, 1861.

12. Kricheldorf, H. R. and Kreiser-Saunders, I. Makromol. Chem. 1990, 191, 1057.

13 Jedlinski, Z., Walach, W., Wojciech, K., Kurcok, P. and Adamus, G. Makromol. Chem.

1991, 192, 2051.

14. Feng, X. D., Song, C. X. and Chen, W. Y. J. Polym. Sci., Polym. Lett. Ed. 1983, 21,

593.

15. Vion, J. M., Jerome, R., Teyssie, M., Aubin, M. and Prud’home, R.E. Macromolecules

1986, 19, 1828.

16. Kricheldorf, H. R., Berl, M. and Scharnagl, N. Macromolecules 1988, 21, 286.

Page 41: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

25

17. Ikada, Y., Jamshidi, K., Tsuji, H. and Hyon, S. H. Macromolecules 1987, 20, 904.

18. Kalb, B. and Penings, A. J. Polymer 1980, 21, 607

19. Iannace, S., Ambrosio, L. and Nicolais, L. Frontiers in biomedical polymer application

1998, 1, 287.

20. Hoffman, J. D., Davis, G. T., Lauritzen, J. I., Jr. in Treatise on Solid State Chemistry

(Hanny, N. B., Ed.), Vol. 3., Plenum Press: New York, 1976, Chapter 7.

21. Vasanthakumari, R. and Pennings, A. J. Polymer 1983, 24, 175.

22. Miyata, T. and Masuko, T. Polymer 1998, 39, 5515.

23. Miyata, T. and Masuko, T., Polymer 1997, 38, 4003.

24. Fisher, E. W., Sterzel, H. J. and Wegner, G. Kolliod Z. Z. Polym. 1973, 251, 980.

25. Huang, J., Buehler, N., Hall, E., Kean, R., Kolstad, J., Wu, L. and Runt, J. ACS

Polymer Preprints 1996, 37, 370.

26. Huang, J., Lisowski, M. S., Runt, J., Hall, E. S., Kean, R. T., Buehler, N. and Lin J. S.

Macromolecules 1998, 31, 2593.

27. Baratian, S., Hall, E. S., Lin, J. S., Xu, R. and Runt, J. Macromolecules 2001, 34,

4857.

28. Cho, J., Baratian, S., Kim, J., Yeh, F., Hsiao B. S. and Runt J. Polymer 2003, 44, 711.

29. Flory, P. J. Trans. Faraday. Soc. 1955, 51, 848.

30. Sanchez, I. C. and Eby, R. K. Macromolecules 1975, 8, 638.

31. Wendling, J., Suter, U. W. Macromolecules 1998, 31, 251

32. Fox, T. G., Garrett, B. S., Goode, W. E., Gratch, S., Kincaid, J. F., Spell, A. and

Stroupe, J. D. J. Am. Chem. Soc. 1958, 80, 1768.

Page 42: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

26

33. He, Z. R. and Prud’homme, R. E. Macromolecules 1999, 32, 7655.

34. Baba, Y., and Kagemoto, A. Macromolecule 1977, 10, 458.

35. Jiang, Z. Z., Boyer, M. T., and Sen, A. J. Am. Chem. Soc. 1995, 117, 7037.

36. Singfield, K. L. and Brown, G. R. Macromolecules 1995, 28, 1290.

37. Slager, J. and Domb, A. J. Biomacromolecules 2003, 4, 1308.

38. Okihara, T., Tsuji, M., Kawaguchi, A., Katayama, K. I., Tsuji, H., Hyon, S. H. and

Ikada, Y. J. Macromol. Sci. Phys. 1991, B30, 119.

39. Tsuji, H., Horii, F., Hyon, S. H. and Ikada, Y. Macromolecules 1991, 24, 2719.

40. Tsuji, H., Hyon, S. H. and Ikada, Y. Macromolecules 1991, 24, 5651.

41. Tsuji, H., Hyon, S. H. and Ikada, Y. Macromolecules 1991, 24, 5657.

42. Tsuji, H., Hyon, S. H. and Ikada, Y. Macromolecules 1992, 25, 2940.

43. Tsuji, H. and Ikada, Y. Macromolecules 1993, 26, 6918.

44. Tsuji, H. and Ikada, Y. Macromol. Chem. Phys. 1996, 197, 3483.

45. Brochu, S., Prud’homme, R. E., Barakat, I. and Jerome, R. Macromolecules 1995, 28,

5230.

46. Tsuji, H. and Ikada, Y. Macromolecules 1992, 25, 5719.

47. Schmidt, S. C. and Hillmyer, M. A. J. Polym. Sci, Polym. Phys 2001, 39, 300.

48. Yamane, H. and Sasai, K. Polymer 2003, 44, 2569.

49. Eling, B., Gogolewski, S. and Pennings, A. J., Polymer 1982, 23, 1587.

50. De Santis, P. and Kovacs, A. J. Biopolymers 1968, 6, 299.

51. Hoogsteen, W., Postema, A. R., Pennings, A. J., Ten Brinke, G. and Zugenmaier, P.

Macromolecules 1990, 23, 634.

Page 43: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

27

52. Kobayashi, J., Asahi, T., Ichiki, M., Oikawa, A., Suzakui, H., Watanabe, T. and

Shikinami, Y. J. Appl. Phys. 1995, 77, 2957.

53. Marega, C., Marigo, A., Di Noto, V., Zannetti, R., Martorana, A. and Paganetto, G.

Makromol. Chem. 1992, 193, 1599.

54. Miyata, T. and Masuko, T. Polymer 1997, 38, 4003.

55. Kalb, B. and Penings, A. J. Polymer 1980, 21, 607

56. Cartier, L., Okihara, T., Ikada, Y., Tsuji, H., Puiggali, J. and Lotz, B. Polymer 2000,

41, 8909.

57. Brizzolara, D., Cantow, H. J., Diederichs, K., Keller, E. and Domb, A. J.

Macromolecules 1996, 29, 191.

58. Okihara, T., Tsuji, M., Kawaguchi, A., Katayama, K. I., Tsuji, H., Hyon, S. H. and

Ikada, Y. J. Macromol. Sci. Phys. 1991, B30, 119.

59. Kister, G., Cassanas, G. and Vert, M. Polymer 1998, 39, 267.

60. Li, S. and Vert, M. Macromolecules 1994, 27, 3107.

61. Li, S. J. Biomed. Mater. Res. 1999, 48, 342.

62. Tsuji, H. and Ikada, Y. Polym. Degrad. Stab. 2000, 67, 179.

63. Tsuji, H., Mizuno, A. and Ikada, Y. J. Appl. Polym. Sci. 2000, 77, 1452.

64. Tsuji, H. Polymer 2002, 43, 1789.

65. Tsuji, H. and Ikada, Y. J. Polym. Sci., Polym Chem. 1998, 36, 59.

66. MacDonald, R. T., McCarthy, S. P. and Gross, R. A. Macromolecules 1996, 29, 7356.

67. Reeve, M. S., McCarthy, S. P., Downey, M. J. and Gross, R. A. Macromolecules 1994,

27, 825.

Page 44: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

28

68. Li, S. and McCarthy, S. Macromolecules 1999, 32, 4454.

69. Li, S., Girard, A., Garreau, H. and Vert, M. Polym. Degrad. Stab. 2001, 71, 61.

70. Trifonova, D., Varga, J. and Vancso, G. J. Polym. Bull. 1998, 41, 341.

71. Ivanov, D., Amolou, Z. and Magonov, S. Macromolecules 2001, 2, 1007.

72. Schonherr, H., Bailey, L. E and Frank, C. W. Langmuir 2002, 18, 490.

Page 45: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

Chapter 2

Materials and Experimental Procedure

This chapter provides information on the materials, sample preparation, and

characterization techniques utilized throughout this research. For clarity, the experiment

detail is described in two parts. Section 2.1 relates to the study of microstructure and

dynamics of L-lactide/meso-lactide copolymers. Section 2.2 focuses on enantiomeric

blends of L- and D-PLA.

2.1 L-lactide/meso-lactide copolymers

2.1.1 Materials

L-PLA and L-lactide/meso-lactide copolymers (Table 2.1) were synthesized via

ring-opening polymerization in a presence of tin (II) octanoate [1]. Previous NMR studies

have shown that these copolymers, crystallizable and rich in L-lactide, are essentially

random [2]. Molecular weights were determined by gel permeation chromatography, using

tetrahydrofuran as the mobile phase and polystyrene molecular weight fractions as

calibration materials. R stereoisomer content was determined by chiral liquid

chromatography and reported in mole [or, equivalently, weight] percent. As-received

polymers were dried at 80 °C for 24 hr under vacuum and stored in a desiccator until used.

2.1.2 Crystalline microstructure by tapping mode AFM

A 1% solution of the desired (co)polymer in chloroform was deposited on a glass

cover slip and the solvent allowed to evaporate overnight at room temperature. Dried

Page 46: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

30

samples were melted at a temperature 30 °C above their nominal melting point for 3 min in

a Mettler hot-stage (model FP-82), then transferred to a second hot-stage pre-set at the

crystallization temperature (Tc) and isothermally crystallized for tc min (Table 2.2). Film

thicknesses used in the study of surface morphology were ≤ 5 µm.

The morphologies of the free surfaces of the cast films were characterized using a

Nanoscope III MultiMode AFM (Digital Instruments) in tapping mode. The free amplitude

(A0) used in this study was between 15 - 30 nm. Most samples were characterized using a

set-point amplitude ratio (rsp) between 0.7 - 0.5. However, for PLLA crystallized at 120

and 126 °C, the 3% meso-lactide copolymer crystallized at 114 and 120 °C, and the 6%

meso-lactide copolymer crystallized at Tc = 114 °C, rsp was 0.9 - 0.8. [rsp is defined as the

ratio of the cantilever’s oscillating amplitude (in contact) to its freely oscillating amplitude

(out of contact)]. Set-point amplitudes near 1.0 correspond to very light normal forces

(soft tapping), and the lower rsp, the higher the tapping force. Acquisition of images less

than ~20 × 20 µm was carried out at 0.5 Hz. To obtain initial information on larger spatial

regions, a higher scan rate of 1 Hz was occasionally used.

Table 2.1 Characteristics of L-PLA and L-lactide/meso-lactide copolymers [1]

Material R (%) Mn Mw Tg (°C)a Tmo (°C)b

L-PLA

3% meso-lactide copolymer

6% meso-lactide copolymer

0.4

2.1

3.4

65,500

65,800

63,900

127,400

122,600

119,100

56

56

56

215

200

187

a Glass transition temperature of quenched (amorphous) samples

b Equilibrium melting point [1]

Page 47: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

31

Table 2.2 Isothermal crystallization temperatures and times used in this study; (identical to

those used in ref [1]).

tc (min) Tc (°C)

L-PLA 3%meso 6% meso

114 18 65

117 27 80

120 12 40 95

126 20 100 240

132 60 300 600

135 90 540 840

138 210 840

141 360

147 900

All samples were interrogated with non-contact silicon cantilevers having resonant

frequencies ranging from 280 - 380 kHz. Samples for imaging the bulk morphology

(thickness ~1 mm, prepared by hot pressing) were isothermally crystallized, then cross-

sectioned with glass knives at room temperature and at –25 °C (i.e., ~85 °C below Tg)

using a Leica UCT Ultramicrotome. The surface of the cross-section was consecutively

sectioned until a smooth surface (to the eye) was obtained. Only the cross-sectional surface

of the sample block was studied further – the thin sections were discarded.

Page 48: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

32

2.1.3 Amorphous phase dynamics

A. Sample preparation

For dielectric and DSC studies, films were prepared in a Carver laboratory hot

press (model 2699). Films were sandwiched between two glass slides (covered with

aluminum foil) and isothermally crystallized in a Mettler (model FP-82) hot-stage. The

surfaces of amorphous and crystalline films (with dimensions of ~2.2 × 2.2 cm and ~70 -

150 µm in thickness) were then gold-sputtered.

The isothermal crystallization conditions described in section 2.1.2 were adopted to

prepare crystalline samples. To obtain amorphous films, molten films were quenched into

ice water for a short time (< 1-2 min), wept dry then maintained in a dessicator overnight

before beginning dielectric experiments. X-ray diffraction was utilized to insure that all

quenched films were non-crystalline.

B. Wide-angle X-ray diffraction

WAXD patterns were obtained using a Scintag Pad V diffractometer, operated with

CuKα radiation (λ = 0.154 nm) at 35 kV and 30 mA. Samples were scanned continuously

using a scanning rate of 1 o/min from 2θ = 5 to 30 o. Data were collected at every 0.02o.

Using Origin software, crystal reflections and the amorphous halo were deconvoluted

simultaneously by fitting with Lorentz and Gaussian functions, respectively. Crystalline

fractions were estimated by dividing the area under the appropriate crystalline reflections

with the total area (crystal peaks plus amorphous halo).

Page 49: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

33

C. Differential scanning calorimetry

Degrees of crystallinity were also determined using a Perkin Elmer DSC 7.

Calibration was performed using indium as a standard. Approximately 3-5 mg of a

particular sample was placed in an aluminum pan and scanned from 30 to 200 °C at a

heating rate of 10 °C/min. The degree of crystallinity (Xc) was calculated by dividing the

measured heat of fusion (∆Hf) by 100 J/g, which is an estimate of ∆Hfo for completely

crystalline PLLA [1].

D. Dielectric measurements

Dielectric studies were carried out using a Novocontrol GmbH Concept 40

broadband dielectric spectrometer (DRS). Frequency sweeps from 6 MHz to 0.01 Hz were

conducted isothermally. Temperature was controlled by a Novocontrol Quatro Cryosystem.

While data collection was carried out continuously from 105 (or 90) to –100 °C for the

crystalline samples, it was separated into 2 scans for amorphous samples: dielectric spectra

were first recorded from 40 to 74 °C to avoid the influence of crystallization at

temperatures above Tg. Experiments at lower temperatures were then performed from 40 to

–100 °C. Experiments up to higher temperatures (i.e., up to 155 °C) were conducted using

the samples from the above experiments.

In a separate series of experiments, relaxation spectra for initially amorphous

samples were collected during crystallization (from the glassy state) at 80 °C (from 6 MHz

to 10 Hz) at ~100 s intervals.

Page 50: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

34

D.1. Analysis of dielectric spectra

Dielectric loss spectra in the temperature range of the glass transition were fitted

with the Havriliak-Negami (HN) function [3]:

ε∗(ω) = ε′(ω) − iε″(ω) (1)

= ε∞ − iσ0 /(ε0ω)s + ∆ε / [1+(τmω)a]b (2)

∆ε is the relaxation strength (i.e., equal to the interval between the relaxed and unrelaxed

dielectric constants, i.e., ε∞ - ε0). The second term in eq. 2 represents the contribution from

dc conduction to the dielectric loss, where σ0 is the dc conductivity constant in units of

S/cm. The distribution of relaxation times (τ) is described by the central relaxation time

(τm), a broadening parameter (a) and a parameter characterizing the skewness of the

distribution (b). It is well known that the segmental (α process) relaxation time

distributions for crystalline and amorphous polymers are symmetric and asymmetric,

respectively [4]. Consequently, the parameter b was fixed at 1 for fitting of the segmental

process of the crystalline samples. In all preliminary fittings of the α process of amorphous

samples, the parameter b was always found to be near 0.5. Therefore, it was fixed at 0.5 for

all subsequent curve fitting.

2.2 Enantiomeric blends of L-PLA and D-PLA

2.2.1 Materials and sample preparation

Relatively low molecular weight L-PLA (L-PLA1, Mv = 3.2 x 103) was purchased

from Birmingham Polymers. L-PLA2 (Mv = 1.4 x 104) and D-PLA (Mv = 8.7 x 104)

Page 51: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

35

(PURASORB) were kindly supplied by Cargill Dow Polymers and PURAC Biochem,

respectively. The viscosity-average molecular weight (Mv) was calculated from [η] = 5.45

× 10- 4 Mv0.73 [5], where [η] is the intrinsic solution viscosity in chloroform at 25 °C.

Each polymer was separately dissolved in dichloromethane to form 6 wt%

solutions. The L-PLA solution was mixed with the D-PLA solution in desired proportions

and stirred vigorously for at least 24 h. Mixtures were cast onto glass slides covered with

aluminum foil and dried overnight at room temperature. Blend films were further dried in a

vacuum oven at 40 °C for two weeks to remove trapped solvent. Amorphous samples of

these films were prepared in a Mettler (model FP-82) hot-stage by quenching the molten

films into ice water. Hereafter, L-PLA1/D-PLA and L-PLA2/D-PLA blends are referred as

LB and HB followed by the blend composition. For example, LB73 is the blend between

70% L-PLA1 with 30% D-PLA.

2.2.2 Real-time study of crystallization from the glassy state

Real-time isothermal crystallization experiments on selected blends were carried

out with a Novocontrol GmbH Concept 40 broadband dielectric spectrometer. Relaxation

spectra for initially amorphous samples were collected at isothermal temperatures (Tiso)

from 6 MHz to 1 Hz at various crystallization times (tc). The collection time for each tc was

~2 min. Dielectric-loss spectra at particular tc were fitted with the empirical Havriliak-

Negami (HN) function (section 2.1.3.D1). Final crystal structure, crystallinity and

transition temperatures of these crystallized samples were characterized by DSC and

WAXD.

Page 52: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

36

2.2.3 Differential scanning calorimetry

For crystalline samples, glass transition temperatures, Tms and degrees of

crystallinity were determined using a TA instruments DSC. Calibration was performed

using indium as a standard. Approximately 2-3 mg of a particular sample was placed in an

aluminum pan and scanned from -30 to 250 °C (or 220 °C for the homopolymers) at a

heating rate of 10 °C/min. The crystallinities associated with homocrystals and

stereocomplex crystals were calculated by dividing the measured heats of fusion (∆Hm)

with their ∆Hmo, which are 100 and 142 J/g for the L-PLA [1] and the stereocomplex [6],

respectively.

For amorphous samples, the same procedure was applied in order to obtain their Tg.

Since these samples crystallize at temperatures > Tg, non-isothermal crystallization

characteristics (i.e., cold crystallization temperature (Tc,c), Tm and crystallinity) were

obtained on heating as well.

2.2.4 Wide-angle X-ray diffraction (WAXD)

The samples used in the dielectric experiments were verified to be originally

amorphous by WAXD. The crystal structures and crystal fractions of the crystallized

samples were also revealed using this technique. WAXD patterns were obtained and

analysed using the same procedures as described in section 2.1.3.B.

Page 53: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

37

2.3 References

1. Huang, J., Lisowski, M. S., Runt, J., Hall, E. S., Kean, R. T., Buehler, N. and Lin J. S.

Macromolecules 1998, 31, 2593.

2. Thakur, K. A. M., Kean, R. T., Hall, E. S., Kolstad, J. J., Lindgren, T. A., Doscotch, M.

A., Siepmann, J. I. and Munson, E. J. Macromolecules 1997, 30, 2422.

3. Havriliak, S. and Negami, S. J. Polym. Sci. Polym. Symp. 1966, 14, 99.

4. Boyd R.H. Polymer 1985, 26, 323; Boyd R.H. Polymer 1985, 26, 1123.

5. Schindler, A., Harper, D. J. Polym. Sci., Polym. Chem. Ed. 1979, 17, 2593.

6. Loomis, G. L., Murdoch, J. R., Gardner, K. H. Polymer Preprints 1990, 31(2), 55.

Page 54: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

Chapter 3

Crystal morphology of Lactide Stereocopolymers

by Tapping mode AFM

3.1 Overall morphology

Crystal habits exhibiting sheaf-like features are mostly observed on the surfaces of

the polylactide polymers crystallized at moderate degrees of supercooling, ∆T (Figures 3.1

and 3.2). Figures 3.1a - h are representative AFM phase images of L-PLA crystallized

isothermally at 120, 132 and 147 °C, imaged at various magnifications. Figures 3.1b and

3.1d were acquired from the boxed areas of Figures 3.1a and 3.1c, respectively. Figures

3.1e and 3.1f are images of ‘spherulites’ acquired from two different areas of the specimen

crystallized at 147 °C. Higher magnification images of areas marked as g and h in Figure

3.1e are presented in Figures 3.1g and 3.1h, respectively. Lamellae seen in these images

orient relative to the surface in different ways, i.e., edge-on, flat-on and at some angle to a

surfaces.

(a) (b)

031a.005

a

( c) 031a.005

a

( c)b 031a.005

a

( c) 031a.005

a

( c)b

Page 55: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

39

(c) (d)

(e) (f)

(g) (h)

Figure 3.1 AFM phase images of L-PLA isothermally crystallized at different

temperatures: (a)-(b) at 120 °C, (c)-(d) at 132 °C, and (e)-(h) at 147 °C

Page 56: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

40

Features in certain portions of Figure 3.1a (for example, as indicated by the arrow)

represent flat-on views of lamellae, while in other areas lamellae are edge-on, or nearly so.

Many lamellae are edge-on in Figure 3.1b, particularly in the lower left portion. Lamellar

sheaf-like structures like those seen in Figure 3.1b are generally considered to be an

intermediate stage in the development of a 3D spherulitic superstructure. The structure

consists of organized, interconnected lamellae. Planar lamellae (an example of which is

indicated with arrow A in Figure 3.1b) are typically observed in central portions of the

superstructures. These ‘dominant’ lamellae are seen to splay (an example of which is

indicated by arrow B in Figure 3.1b) and branch (see arrow C in Figure 3.1b) at later stages

of growth. It has been proposed that the splaying results from repulsive forces between

lamellae, possibly originating from cilia or un-crystallized segments protruding from fold

surfaces [1,2].

As seen in Figure 3.1, the mean size of the L-PLA superstructure at 120 °C is

smaller than that crystallized at higher Tc’s, i.e., at 132 and 147 °C. This is also true for the

3% and 6% meso-lactide copolymers, and is in good agreement with previous work in

which primary nucleation density was found to decrease at lower ∆T [3]. Compact

spherulitic structures are observed on the surface of L-PLA crystallized at 120 °C, while

more open structures are seen in L-PLA crystallized at higher Tc. For example, Figure 3.1c

shows an image of a portion of a L-PLA spherulite crystallized at 132 °C. Comparison

between the high magnification images of L-PLA crystallized at 120 °C (Figure 3.1b) and

at 132 °C (Figure 3.1d), shows lamellae to be packed more closely in the former. Regions

between edge-on lamellar stacks filled with spiral terraces (flat on), indicative of screw

dislocations, are seen in Figure 3.1c and one such region is marked with an arrow. Screw

Page 57: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

41

dislocations are one of the factors that lead to space-filling in spherulites [4]. At higher Tc,

the edge-on lamellae appear to be more separated (Figure 3.1f). Larger numbers of screw

dislocations are observed between edge-on lamellae, as well as in regions far from the

center of spherulites (Figure 3.1h). Lamellar orientation on the surface of L-PLA

crystallized at 147 °C is different from those crystallized at the lower Tcs: they are mostly

parallel to the surface.

(a) (b)

(c) (d)

Figure 3.2 Phase images of 6% meso-lactide copolymer crystallized at different

temperatures: (a)-(b) at 114 °C, (c)-(d) at 120 °C

Page 58: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

42

High magnification images of different regions of the 6% meso-lactide copolymer

crystallized at 120 °C are shown in Figure 3.2c and d. Figure 3.2c shows a view of the

central portion of a spherulite. Many areas in Figure 3.2d exhibit edge-on lamellae,

oriented mostly parallel to each other within sheaf-like structures. Note also the absence of

branch points for edge-on lamellae. Sheaf-like structures with more complex lamellar

organization are observed for L-PLA crystallized at the same Tc (Figures 3.1a and 3.1b). A

comparison of Figures 3.1b and 3.2d suggests that increasing comonomer content leads to

a reduction in the average number of lamellae per stack. The effect of crystallization

temperature on crystalline microstructure of the 3% and 6% meso-lactide copolymers is

similar to that observed for L-PLA.

3.2 Contrast in height and phase images

Tapping force, as influenced by A0 and rsp, is an important experimental variable

and its effect on features observed in AFM height and phase images has been discussed in

a number of publications [5-10]. Low force (i.e., high rsp with moderate A0) is normally

desirable for obtaining true surface topography: tip - sample surface interactions are weak

in this case. At higher tapping forces (i.e., lower rsp and moderate/large A0) the image is

influenced by the strong tip - sample interactions and the indentation force is dominated by

sample stiffness and tip - sample contact area. For heterogeneous surfaces (e.g. those

consisting of crystalline and amorphous regions), contrast is often improved at higher

tapping forces [8]. Larger tapping forces are also necessary for probing a component lying

beneath a surface [6]. However, higher forces can result in surface indention and distortion

(broadening) of the apparent lateral dimensions of features. In the current work, an

Page 59: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

43

important objective is to quantitatively determine lamellar thickness as a function of

crystallization conditions and comonomer content. Therefore, experiments were initially

conducted at low tapping force. If the results from a given low force experiment were

deemed unsatisfactory for quantitative analysis, one or more higher-tapping force

experiments were performed.

Height images are obtained at constant amplitude, where the cantilever tip height is

controlled via a feedback routine so as to maintain a constant set-point amplitude (and thus

a constant average normal force). Therefore, information in height images, taken at

moderate A0 and moderate/ low rsp, is related to sample topography and stiffness. Phase

images are obtained concurrently with height images by monitoring the phase lag between

the oscillating tip (output) and the driving signal of the piezo element (input). The contrast

in this imaging mode is dramatically modified by tip - sample interactions (attractive forces

(adhesion) and repulsive forces (indentation) at higher and lower rsp, respectively). The

polylactides under investigation all have glass transition temperatures Tg ~ 60 °C (see

Table 2.1), significantly above ambient. There is a difference in stiffness between the

glassy amorphous and crystalline regions and, although not large, is sufficient for contrast

development. Contrast modification with varying tapping force is exhibited in Figures 3.3a

- d. The images in Figure 3.3 were acquired from the same area of a L-PLA specimen

crystallized at 126 °C, but at two different rsp. Increasing tapping force remarkably

improves the contrast between the crystalline and amorphous regions in the phase images,

but there is little change in the corresponding height images. As a result, this indicates that

the features observed in height images at these scanning conditions are dominated by the

surface topography, while the phase images are determined by the indentation forces.

Page 60: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

44

(a) (b)

(c) (d)

Figure 3.3 Height and phase images of L-PLA crystallized at 126 °C: (a)-(b) at a

set-point amplitude ratio of rsp ~ 0.78 and (c)-(d) of rsp ~ 0.69, respectively.

Page 61: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

45

3.3 Determination of lamellar thicknesses

3.3.1 Methodology

Given the excellent spatial resolution achievable with tapping mode AFM, the

question of using this technique for quantitative analysis of domain sizes has naturally

arisen. Three important factors influence AFM spatial resolution: scan size [11], the effect

of tip geometry, and sample deformation [10,12,13]. Analyses in this work were performed

on images with dimensions between 0.8 x 0.8 and 1.2 x 1.2 µm2 to rule out the influence of

scan size on determining feature widths on the order of 10 nm (500 pixels provide

sufficient resolution). Only a few images with large dimensions, i.e. ~1.6 x 1.6 µm2, were

taken into the consideration. As seen in the previous section, lamellar orientation varies

from area to area. A relatively wide area was imaged first in order to view a region covered

with sheaf-like structures. Subsequently, regions with a comparatively high concentration

of planar edge-on lamellae, e.g. near a center of the sheaf-like structure, were chosen and

imaged at higher magnifications. Samples were characterized initially at low tapping force

to avoid surface deformation, but sufficient contrast could not always be obtained under

such conditions, requiring the use of somewhat higher tapping forces.

The widths of edge-on lamellae from height or phase images (perpendicular to the

long axes of the lamellae) were determined using the ‘section analysis’ portion of the

Nanoscope AFM software. Since the possibility of some small inclination of lamellae to

the surface cannot be ruled out, the mean lamellar thicknesses (lc) and their distributions

should strictly be considered ‘apparent’ [14]. Only height images of L-PLA samples

crystallized at 120 °C (largest ∆T) exhibit contrast as good as observed in phase images, as

demonstrated in Figure 3.4. There are however a modest number of other samples for

Page 62: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

46

which the height images exhibit sufficient contrast so that quantitative analysis could be

conducted.

(a) Height image

(b) Phase image

Figure 3.4 (a) Height image and sectional profile and (b) phase image and sectional

profile of L-PLA crystallized at 120 °C.

Page 63: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

47

(a) Height image

(b) Phase image

Figure 3.5 (a) Height image and sectional profile and (b) phase image and sectional

profile of 3% meso-lactide copolymer crystallized at 114 °C.

As seen in Figures 3.3-3.5, it is clear that crystalline lamellae are associated with

brighter regions in the phase images, as expected for the set-points (rsp) used [8,10,15].

Brighter regions in height images however can be either crystalline lamellae or amorphous

regions: but this assignment can be made by reference to the corresponding phase image.

For the 3% meso-lactide copolymer (Figure 3.5), the lamellar structure is clearly observed

only in the phase image (Figure 3.5b): thus it suggests that the surface of the copolymer

specimens is covered with un-crystallized melt. This obviously creates problems in

determining lamellar thicknesses from these height images. For example, it is not possible

Page 64: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

48

to delineate the lamellar structure of the 3% copolymer (crystallized at 114 °C) in the

marked area of the height image (Figure 3.5a). However, the phase image of the same

region clearly shows the lamellar morphology (Figure 3.5b). This permits characterization

of the microstructure without having to stain, etch, or otherwise modify the polymers in

order to reveal the crystalline regions lying beneath the surface.

The tip geometry of the silicon probe can have a significant influence on the image

acquired. Analysis of selected AFM tips used in the experiments by scanning electron

microscopy (Figure 3.6a) shows that the tips are cone-shaped with θ = 15˚ and θ′ = 19˚. In

order to determine the thickness of edge-on lamellae, deconvolution of the tip profile was

taken into consideration as also demonstrated in Figure 3.6b. The apparent lamellar

thickness was found to be overestimated by ~0.4 - 1 nm, which constitutes a correction of

only ~3 - 10% in the measured value.

Quantitative interpretation of data from phase images must be done carefully,

particularly since some details of phase image formation are not completely understood

[8,10,15]. Nevertheless, it is possible to determine feature sizes from phase images, e.g. in

Ref. 14,16, although there is an uncorrected effect of the tip profile on the image

(presumably relatively small) along with uncertainties due to broadening from high normal

force and from the possibility of lamellar tilt. Lamellar thicknesses were estimated from

phase images using the width at half-height of well-defined peaks in section profiles of

edge-on lamellae (Figure 3.5b).

Page 65: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

49

(a)

(b)

θ θ’

x

h1 h2

x’

θ’

w

h1 tan θ = x h2 tan θ’ = x’ Edge width = W-(x+x’)

AFM image

Real feature

θ θ’

x

h1 h2

x’

θ’

w

h1 tan θ = x h2 tan θ’ = x’ Edge width = W-(x+x’)

AFM image

Real feature

Figure 3.6 (a) Tip geometry from SEM and (b) effect of tip convolution on the

characterized feature.

Page 66: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

50

3.3.2 L-PLA

Figure 3.7 shows a plot of the mean lamellar thicknesses as a function of

crystallization temperature for PLLA, determined from AFM height (where possible) and

phase images. These values are also tabulated in Table 3.1, along with the corresponding

95% confidence intervals around the means. Fewer lamellae in height images can be

confidently used for quantitative analysis, as seen in the last column of Table 3.1. This

results from the contrast limitation in height images noted above. Also displayed in Figure

3.7 are values determined in the previous SAXS study of the identical L-PLA and

copolymers [3]. The SAXS values were determined from a ‘finite’ lamellar stack model

(the principle evidence for this is the relationship between the bulk and linear

crystallinities). Note that the AFM and SAXS values are quite similar considering the

experimental error, confirming the scattering model used in the interpretation of the SAXS

data. In addition, these results indicate that lamellar thicknesses extracted from surfaces of

films by AFM are representative, at least in this case, of the mean bulk lc as determined by

SAXS.

In the work of Abe et al. [17], ‘lamellar periodicities’ estimated from profiles of

flat-on PLLA lamellae crystallized at 120 and 140 °C are in the range of 14-16 and 16-18

nm, respectively. These values are approximately 2-5 nm larger than the lamellar

thicknesses determined here. However, this would be expected since both the crystalline

and amorphous layers are measured for the flat-on lamellae. Even though the experiments

in reference 18 and in the current research were performed using different AFM methods

(i.e., dynamic force vs. tapping mode), there is consistency in the results of these two

studies.

Page 67: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

51

Table 3.1 Mean lamellar thickness and distribution from AFM phase and height images

Sample

Method Material Tc (oC) Mean (nm)

95%confidence interval (±)

Number of lamellae

120 11.4 0.9 58 126 12.1 0.4 158 132 15.2 0.4 182 138 15.4 0.6 84

L-PLA

141 16.5 0.5 210 114 12.4 1.0 56 120 12.2 1.4 45 126 13.5 0.4 143 132 11.9 0.8 58

3%-meso

135 13.3 0.5 185 114 11 0.6 109 117 9.5 0.6 55 120 9.7 0.8 52

Phase Image

6%-meso

126 9.9 0.7 57 120 10.4 0.6 39 132 12.4 0.7 73 L-PLA

141 12.2 0.9 49 126 11.3 0.6 59 3%-meso 135 11.2 0.7 60

Height Image

6%-meso 114 10.5 0.9 24

Page 68: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

52

Crystallization temperature ( oC)

110 120 130 140 150

Thi

ckne

ss (n

m)

0

5

10

15

20

25

30

SAXSPhase imageHeight image

Figure 3.7 Average thickness of L-PLA lamellae determined from AFM height and

phase imaging, as well as that reported previously from SAXS [3], as a function

of Tc.

3.3.3 Meso-lactide copolymers

As expected (see Table 3.1 and Figure 3.7), mean lamellar thicknesses increase

with increasing Tc for L-PLA (from ~11 to 17 nm, from phase images). In general,

however, the influence of Tc on the crystal thickness of the copolymers (3 and 6%meso) is

difficult to distinguish due to the experimental uncertainty associated with individual data

points. However, lamellar thicknesses for the copolymers are generally smaller than PLLA,

particularly at lower ∆T, in agreement with results from the previous SAXS experiments.

As discussed earlier, such a reduction with comonomer content is consistent with

substantial R stereoisomer exclusion from crystalline regions. Mean lc values derived from

phase images are shown in Figure 3.8. Thickness distributions for all samples from height

and phase images are presented in Figure 3.9. Lamellar thicknesses from phase images are

consistently higher than those from height images (although within experimental

Page 69: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

53

uncertainty for some samples). In addition, the thickness distributions from phase images

tend to be skewed somewhat to larger sizes (note that the size scale changes from Figure

3.9a to b and 3.9c to d). There are several possible reasons for this. As noted earlier, the

contrast in phase images is formed from complex tip-sample interactions, and the influence

of the tip profile on these images is uncorrected. In addition, in order to define lamellar

thicknesses reasonably, slightly different procedures were used for height (width of

constant height region: Figure 3.4a) vs. phase images (width at half-height of well-defined

peaks in section profiles: Figures 3.4b and 3.5b). In the present analysis, lamellae

distinguishable in height images represent only ~20–50% of the population resolvable in

the phase image of the same spatial region that was analyzed. This difference in sampling

could also contribute to differences in mean thickness values and distributions determined

from height and phase images.

Degree of supercooling (oC)

50 60 70 80 90 100 110

Thi

ckne

ss (n

m)

6

8

10

12

14

16

18

20

PLLA3%meso6%meso

Figure 3.8 Mean lamellar thickness for PLLA, 3% and 6% meso-lactide copolymers

as a function of degree of supercooling from AFM phase images. The error bars

present the 95%confidence limits of these average values.

Page 70: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

54

(a) (b)

6 10 14 18

120 C

132 C

141 C

0.0

0.2

0.4

0.6

4 8 12 16 20 24

120 C

126 C

132 C

138 C

141 C

0.0

0.2

0.4

0.6

(c) (d)

4 8 12 16

126 C

135 C

0.0

0.2

0.4

0.6

4 8 12 16 20 24

114 C

120 C

126 C

132 C

135 C

0.0

0.2

0.4

0.6

(e) (f)

4 8 12 16 20

Height

Phase0.0

0.2

0.4

0.6

2 6 10 14 18 22

114 C

117 C

120 C

126 C

0.0

0.2

0.4

0.6

Figure 3.9 Lamellar thickness distributions of (a)-(b) PLLA, (c)-(d) 3% meso-lactide

copolymer and (e)-(f) 6% meso-lactide copolymer as a function of Tc;

Page 71: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

55

data determined in (a) and (c) from height images, (b),(d) and (f) from phase

images and (e) from both height and phase images of the 6% meso-lactide

crystallized at 114 °C.

(x-axis: thickness (nm), y-axis: Tc or imaging mode, and z-axis: fraction of

lamellar thickness distribution)

3.4 Bulk morphology

As an extension of the work described in the previous sections, the bulk

morphology was investigated by evaluating “microtomed” samples, i.e. internal surfaces of

the samples were exposed by microtoming away the outer parts of specimens and

subsequently imaged by AFM. Knife marks are clearly visible in AFM images (not shown)

of samples microtomed at room temperature (~40 °C below Tg), inhibiting evaluation of

the bulk microstructure. Microtoming at lower temperature (-25 °C) produced specimens

with much smoother surfaces.

Figure 3.10 presents height and phase images of a cross-section of L-PLA

crystallized at 132 °C. Spherulites are revealed in low magnification images (Figure

3.10a). Areas marked as B and C were probed at higher magnification (Figure 3.10b and

3.10c) in an attempt to obtain information on the lamellar subunits. Lamellae similar to

those seen in the “free” surface experiments are observed, as seen in higher magnification

images like Figure 3.10c and d, albeit with reduced clarity compared to the ones obtained

at the “free” surface. The qualitative similarity between these lamellae and the ones

observed at the polymer surface, combined with the quantitative agreement of the lamellae

thickness as measured by AFM at the surface and by SAXS in the bulk, strongly suggest

Page 72: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

56

that the morphologies observed at the polymer surface closely reflect the bulk crystal

morphology of these polymers.

(a) (b)

(c) (d)

Figure 3.10 (a)-(c) Height images of PLLA (Tc = 132 °C) microtomed at -25 °C and (d)

phase image of the same area as (c).

Page 73: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

57

Lamellar organization in spherulitic superstructures is complex, with lamellae

oriented randomly in three dimensions and disordered regions intervening. Considering

that sectioning through this structure occurs randomly, as does selection of an area for

scanning, the probability of observing a large number of edge-on lamellae for thickness

measurement is low. Moreover, surface roughness, even when microtoming at –25 °C,

leads to a reduction in resolution. As a consequence, the images of cross-sections are not

suitable for quantitative analysis of lamellar thickness.

3.5 Summary

High resolution tapping mode AFM provides a promising way of performing

qualitative and quantitative studies of lamellar microstructure of crystalline polymers,

without the requirement of complex sample preparation methods. In this research the

lamellar morphology of L-PLA and two poly(L-lactide-co-meso-lactide) random

copolymers is investigated. Mean lamellar thicknesses are lower for the random

copolymers compared to L-PLA, particularly at lower ∆T. Along with the fact that the

degree of crystallinity decreases by about a factor of two upon introducing ~3% randomly

placed R co-units [3], this result strongly suggests that a significant portion of the R

stereounits are rejected from crystalline lamellae. The most important finding of the

present study is that lamellar thicknesses determined from AFM phase and height images

are in good agreement with those determined in our previous SAXS study of the identical

materials.

Page 74: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

58

3.6 References

1. Bassett, D. C. and Olley, R. Polymer 1984, 25, 935.

2. Bassett, D. C. and Vaughen, A. S. Polymer 1985, 26, 717.

3. Huang, J., Lisowski, M. S., Runt, J., Hall, E. S., Kean, R. T., Buehler, N. and Lin J. S.

Macromolecules 1998, 31, 2593.

4. Bassett, D. and Hodge, A. Proc. R. Soc. Lond. 1981, A377, 61.

5. Schonherr, H., Bailey, L. E. and Frank, C. W. Langmuir 2002, 18, 490.

6. Magonov, S. N., Cleveland, J., Elings, V., Denley, D. and Whangbo, M. H. Surf. Sci.

1997, 389, 201.

7. Van Noort, J. T., Van der Werf, K. O., De Grooth, B. G. and Van Hulst, N. F.,

Ultramicroscopy 1997, 69, 117.

8. Bar, G., Thomann, Y., Brandsch, R., Cantow, H. J. and Whangbo, M. H. Langmuir

1997, 13, 3807.

9. Pickering, J. P. and Vancso, G. J. Polym. Bull. 1998, 40, 549.

10. Knoll, A., Magerle, R. and Krausch, G. Macromolecules 2001, 34, 4159.

11. Scan size affects the image resolution since by design the instrument collects 512 x 512

points for each scanned image; thus, each “pixel”/point in a 10 µm scan corresponds to

a 19.5 x 19.5 nm2 sample area, and for a 1.6 µm scan line each point corresponds to

about 3 nm.

12. Sheyko, S. Adv. Polym. Sci. 2000, 151, 61.

13. Weihs, T., Nawaz, Z., Jarvis, S. and Pethica, J. Appl. Phys. Lett. 1991, 59, 3536.

14. Ivanov, D., Amolou, Z. and Magonov, S. Macromolecules 2001, 34, 8944.

Page 75: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

59

15. Schmitz, I., Schreiner, M., Friedbacher, G. and Grasserbauer, M. Appl. Surf. Sci. 1997,

115, 190.

16. Basire, C. and Ivanov, D. A., Phys. Rev. Lett. 2000, 85, 5587.

17. Abe, H., Kikkawa, Y., Inoue, Y. and Doi, Y., Biomacromolecules 2001, 2, 1007.

Page 76: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

Chapter 4

Amorphous Chain Dynamics of Amorphous and

Crystalline L-PLA and Stereocopolymers

4.1 Relaxation processes by dielectric relaxation spectroscopy

Two relaxation processes, α and β, are observed in the dielectric relaxation spectra

of amorphous and crystallized samples in the temperature range between -100 and 105 °C.

Figure 4.1 presents isochronal plots of dielectric constant (ε') and loss (ε″) of L-PLA

crystallized at 141 °C. The broad local β process occurs at much lower temperatures

compared to the relatively strong α segmental relaxation. The β-relaxation is believed to

originate from local twisting motions of main chains [1].

4.2 Sub-glass process

Figure 4.2a presents peak positions (fm) of the α and β-processes of the amorphous

L-PLA and copolymers as a function of temperature. These data follow the Vogel-Fulcher-

Tammann (VFT) and Arrhenius forms, respectively. As seen in Figure 4.2b, the presence

of the crystalline phase has no effect on the average relaxation time of the β process.

Activation energies for the non-correlated ester motion were determined to be in the range

of 8-11 kcal/mol for all samples. These values compare very well with those reported for

aliphatic polyesters in the literature (9-13 kcal/mol) [1-3].

Page 77: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

61

(a)

2.7

2.8

2.9

3.0

3.1

3.2

-100 -50 0 50 100Temperature (oC)

ε'

10 kHz1 kHz100 Hz10 Hz

α

β

(b)

0.00

0.02

0.04

0.06

-100 -50 0 50 100Temperature (oC)

ε"

10 kHz1 kHz100 Hz10 Hz

β

α

Figure 4.1 Isochronal plots of (a) dielectric constant and (b) dielectric loss of L-PLA

crystallized at 141 °C in frequency range from 10 Hz to 10 kHz.

Page 78: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

62

(a)

1.E-02

1.E+00

1.E+02

1.E+04

1.E+06

2.8 3.0 3.2 3.4 3.6 3.8 4.0

1000/T (K-1)

Log(

f m) β-process

α-process

(b)

1.E+02

1.E+04

1.E+06

3.0 3.2 3.4 3.6 3.8 4.0

1000/T (K-1)

Log(

f m)

β-process

1.E+02

1.E+04

1.E+06

3.0 3.2 3.4 3.6 3.8 4.0

1000/T (K-1)

Log(

f m)

β-process

Figure 4.2 log fm vs. 1/T of (a) the α-process (closed symbols) and β-process (open

symbols) of amorphous L-PLA ( ) and 3% meso-lactide ( ), 6% meso-

lactide ( ) and 12% meso-lactide copolymers ( ). (b) The β-process of

amorphous PLLA ( ), L-PLA crystallized at 90 °C ( ) and the crystallized

12% meso-lactide copolymer crystallized at 105°C (□).

Page 79: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

63

4.3 Segmental relaxation

4.3.1 Crystallization from the glassy state

In order to illustrate general trends in the segmental relaxation upon crystallization,

the findings from a real time crystallization experiment from the glassy state of an

amorphous 3% meso-lactide copolymer at 80 °C are first presented. Loss spectra (Figure

4.3a) exhibit significant changes in breadth and relaxation strength when tc exceeds 17

min. The absence of changes at tc < 17 min is not associated with an induction time, as

crystallization is very rapid under these conditions, beginning within a few minutes of the

sample reaching Tc. The fraction of mobile amorphous segments participating in the α

process can be estimated from the ratio of the relaxation strength (ε′∞- ε′0) at tc to that at tc

= 0 min (Figure 4.3b). The portion of mobile relaxing segments consumed at the point at

which fm begins to change significantly is ~35% for this sample (i.e., at tc ~ 27 min).

Similar observations have been reported in real-time dielectric crystallization studies of

poly(ethylene terephthalate) [PET] [4], poly(ether ether ketone) [5] and hydroxybutyrate -

hydroxyvalerate copolymers [3]. There is no clear indication of two separate segmental

processes in the loss spectra in Fig 4.3a, as has been observed during crystallization of PET

[6,7].

Up to tc = 17 min, a large fraction of the relaxing segments are unconstrained,

having mobility similar to that of segments of the completely amorphous sample. These

include segments between growing spherulites as well as those between and within nascent

lamellar stacks. However, the presence of the crystalline phase early in the crystallization

process results in a broadening of the segmental relaxation process, as the motion of a

small portion of the relaxing segments becomes perturbed. The results of simultaneous

Page 80: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

64

wide- and small-angle x-ray scattering experiments on these same polylactides during

crystallization [8] suggest that crystallization of thinner lamellae or lamellar stacks occurs

between dominant lamellae as crystallization proceeds and, together with the findings of

‘static’ SAXS experiments [9], support a ‘finite stack’ model for the final microstructure.

Thus, amorphous segments between dominant lamellae are either ‘consumed’ during

crystallization, or their motion becomes quite constrained. In fact, a number of authors

have argued that essentially all non-crystalline segments within relatively highly crystalline

lamellar stacks are restricted from undergoing the usual segmental process [5,10,11]. As

crystallization proceeds, the remaining mobile amorphous segments become constrained to

varying degrees, giving rise to a net reduction in fm and a broad relaxation resulting from a

broad range of environments. Reduction in the relaxation strength with time arises from

‘consumption’ of relaxing segments in the formation of lamellar crystals, and in the

formation of rigid amorphous segments, which will be discussed in more detail below.

Real-time DRS experiments of isothermal crystallization of the polylactides under

investigation here can only be carried out within a limited temperature range due to

relatively rapid crystallization from the glassy state. Consequently, investigation of the

influence of crystallinity and microstructure on chain dynamics was primarily performed

by comparing dielectric spectra of amorphous and ‘fully-crystallized’ samples.

Page 81: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

65

(a)

Frequency (Hz)

ε″

0.0

0.2

0.4

1.E+01 1.E+03 1.E+05 1.E+07

0 1722 2737 4757 97347

Frequency (Hz)

ε″

0.0

0.2

0.4

1.E+01 1.E+03 1.E+05 1.E+07

0 1722 2737 4757 97347

(b)

2.5

3.0

3.5

4.0

4.5

5.0

1.E+01 1.E+03 1.E+05 1.E+07

0 1722 2737 4757 97347

Frequency (Hz)

ε′

2.5

3.0

3.5

4.0

4.5

5.0

1.E+01 1.E+03 1.E+05 1.E+07

0 1722 2737 4757 97347

Frequency (Hz)

ε′

Figure 4.3 (a) Dielectric loss and (b) dielectric constant spectra during isothermal

crystallization of the 3% meso-lactide copolymer at 80 °C as a function of tc.

Page 82: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

66

4.3.2 Segmental dynamics of amorphous and crystallized samples

A. Temperature-dependent dielectric response

The dielectric spectra of amorphous and semi-crystalline L-PLA are displayed in

Figures 4.4 and 4.5, respectively. The presence of the crystalline phase alters the segmental

relaxation in two significant ways. First, as discussed in the previous section, the relaxation

strength declines dramatically, as anticipated from previous studies on crystalline polymers

[e.g.12]. In addition, the temperature dependence of ∆ε for the crystallized samples is

opposite to that of the fully amorphous ones, in which ∆ε decreases with increasing

temperature. Similar observations have been reported for other semi-crystalline polymers

[13-15].

For the fully amorphous samples, the temperature-dependent behavior of ∆ε is in

good agreement with expectations from the Kirkwood-Onsager-Fröhlich equation: i.e., ∆ε

is proportional to Nµ2g/T, where µ, N and g are the dipole moment, number density of

movable dipoles and correlation factor, respectively [16]. If changes in g and N with

temperature are negligible, this expression predicts that ∆ε will decrease with increasing

temperature (i.e., as a result of randomized dipoles with increasing temperature).

The idea of a rigid amorphous phase (RAP) has been introduced in order to explain

the behavior of the crystalline materials. RAP segments are non-crystalline (unrelaxed at

Tg) that have lower mobility compared to those of mobile amorphous segments (that relax

near Tg) [17,18]. RAP has been associated with segments in order – disorder interphases at

crystal surfaces and, as noted above, with non-crystalline segments in interlamellar

Page 83: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

67

regions. It is likely that the mobility of these chain segments is enhanced at higher

temperatures (above Tg) and ∆ε would therefore be expected to increase with temperature.

(a)

PLLA

2.5

3.5

4.5

5.5

6.5

1.E-03 1.E-01 1.E+01 1.E+03 1.E+05 1.E+07Frequency (Hz)

74 C 72 C70 C 68 C66 C 64 C62 C 60 C

ε′

(b)

PLLA

0.0

0.2

0.4

0.6

0.8

1.E-03 1.E-01 1.E+01 1.E+03 1.E+05 1.E+07Frequency (Hz)

74 C 72 C70 C 68 C66 C 64 C62 C 60 C

ε″

Figure 4.4 (a) Dielectric constant and (b) loss for amorphous L-PLA as a function of

frequency at the indicated temperatures.

Page 84: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

68

(a)

PLLA: Tc=120oC

2.5

2.7

2.9

3.1

3.3

3.5

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06 1.E+08Frequency (Hz)

ε'

90 C 85 C80 C 75 C70 C

(b)

PLLA: Tc = 120oC

0.00

0.02

0.04

0.06

0.08

0.10

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06 1.E+08Frequency (Hz)

ε"

90 C 85 C80 C 75 C70 C

Figure 4.5 (a) Dielectric constant and (b) loss for L-PLA crystallized at 120 °C as a

function of frequency at the indicated temperatures.

Figure 4.6 presents the mean ∆ε for the segmental process, based on the results of

three specimens of each amorphous material. The error bars represent the standard

deviation of each data point. As seen in Figure 4.4, only a portion of the L-PLA segmental

process is observed between 60 and 64 °C and the uncertainty associated with curve fitting

is therefore larger. From the results at higher temperatures, the strength of the segmental

Page 85: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

69

process is seen to depend on the chemical composition of the polymers. The influence of

regio-regularity on the chain conformations and flexibility of poly (lactic acid) polymers

has been studied using Raman spectroscopy, light scattering [19-21] and RIS Metropolis

Monte Carlo calculations [22]. The characteristic ratios (C∞) of lactic acid polymers have

been found to decrease with increasing D-lactic acid content [21], indicating that the

copolymer chains are more flexible than those of the L-PLA homopolymer. This

conclusion is also supported by the present study: the somewhat higher relaxation strength

(Figure 4.6) and lower τm (Figure 4.7) for the higher meso-lactide content copolymers.

Temperature (oC)

58 60 62 64 66 68 70 72 74 76

∆ε

1.4

1.6

1.8

2.0

2.2

2.4

2.6

2.8

3.0

3.2

12%meso6%meso3%mesoPLLA

Figure 4.6 Relaxation strengths of amorphous samples of L-PLA and the meso-lactide

copolymers at the indicated temperatures.

Page 86: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

70

2.86 2.88 2.90 2.92 2.94 2.96 2.98 3.00 3.02

-10

-8

-6

-4

-2

0

2

12%meso6%meso3%mesoPLLA

ln(τ

m)

1000/T (K-1)2.86 2.88 2.90 2.92 2.94 2.96 2.98 3.00 3.02

-10

-8

-6

-4

-2

0

2

12%meso6%meso3%mesoPLLA

ln(τ

m)

1000/T (K-1) Figure 4.7 Mean relaxation times of amorphous L-PLA and meso-lactide copolymers.

B. Influence of the Crystalline Phase

The influence of the crystalline phase on the segmental τm is evident in Figure 4.8.

DSC experiments reveal that, as expected, the crystallinities of the L-PLA samples are

higher than those of the copolymers, and the higher the meso-lactide content, the lower the

crystallinity (Table 4.1). For L-PLA, the presence of the relatively high crystallinity has an

important effect on the average relaxation time of the α-process. The mean τm of the still

mobile, yet perturbed, amorphous segments is much longer than that of the fully

amorphous L-PLA. Compared to L-PLA, a smaller difference between τm of crystalline

and amorphous samples is observed for copolymers with the higher meso-lactide contents.

Eventually, there is no difference in τm between crystalline and amorphous samples for the

12% meso-lactide copolymer (Figure 4.8d). The latter results for these lowest crystallinity

samples are similar to those reported recently for L-PLA by Mijovic and Sy [23].

Therefore, our results suggest that crystalline content is the origin of the different

Page 87: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

71

observations for L-PLA in the present study and those reported in ref 23. In addition, the

data in ref 23 were acquired at 90 - 160 °C, compared with 75 - 90 °C in the present study:

at higher temperatures, it is likely that constraints imposed on restricted amorphous

segments would be relaxed.

(a) (b)

PLLA

-16

-12

-8

-4

0

4

2.6 2.7 2.8 2.9 3.0 3.11000/T (K-1)

Ln( τ

m)

amorphousTc = 90CTc = 120CTc = 132CTc = 141C

3%meso-lactide-16

-12

-8

-4

0

4

2.6 2.7 2.8 2.9 3.0 3.1

1000/T (K-1)

Ln( τ

m)

amorphousTc = 120CTc = 132CTc = 135C

(c) (d)

6%meso-lactide-16

-12

-8

-4

0

4

2.6 2.7 2.8 2.9 3.0 3.1

1000/T (K-1)

Ln( τ

max

)

amorphousTc = 120CTc = 126CTc = 132C

12%meso-lactide-16

-12

-8

-4

0

4

2.6 2.7 2.8 2.9 3.0 3.11000/T (K-1)

Ln( τ

m)

amorphousTc = 105CTc = 111CTc = 117C

Figure 4.8 Comparison between the relaxation time of the amorphous and crystallized

samples of (a) L-PLA, (b) 3%, (c) 6% and (d) 12% meso-lactide copolymer.

Page 88: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

72

Table 4.1 Crystalline, mobile and RAP fractions along with VFT parameters and fragilities

for PLLA and the meso-lactide copolymers, isothermally crystallized at the designated

temperatures

Sample Material Tc (°C)

fcrystal fmobilea,b fRAP

a T0c

(°C) Bc Tref

d (°C)

Amorphous

1

2

3

4

L-PLA

-

90

120

132

141

0

0.51

0.60

0.66

0.68

-

0.25

0.27

0.18

0.18

-

0.24

0.13

0.16

0.14

21

38

33

32

34

1300

1060

1160

1140

1140

56

67

64

63

65

Amorphous

5

6

7

3% meso

-

120

132

135

0

0.48

0.61

0.58

-

0.31

0.21

0.21

-

0.21

0.18

0.21

20

23

23

22

1300

1310

1290

1310

55

59

58

58

Amorphous

8

9

10

6% meso

-

120

126

132

0

0.47

0.48

0.55

-

0.35

0.24

0.24

-

0.18

0.28

0.21

19

21

20

19

1320

1320

1320

1350

55

57

56

55

Amorphous

11

12

13

12% meso

-

105

111

117

0

0.34

0.34

0.38

-

0.33

0.27

0.22

-

0.33

0.39

0.40

17

16

19

16

1340

1380

1320

1370

54

54

54

53

a Based on calculation at T ~ T0 + 50 °C

b The fraction of mobile component determined from DRS was converted to

mass fraction by multiplying by ρa/ρs, where ρa and ρs are the fully

amorphous and sample densities, respectively. Densities of fully crystalline

and amorphous PLLA have been reported to be 1.378 and 1.258 g/cm3,

respectively [24]. Sample densities were estimated from [ρc × fcrystal] + [ρa ×

1-fcrystal].

c Standard deviation of VFT parameters: within ~5% for B and ~ 1 °C for T0

d Temperature at which τmax = 100 s

Page 89: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

73

A careful examination of Figure 4.8a leads to the conclusion that the segmental

dynamics of crystallized L-PLA is also determined by crystallization conditions. The

somewhat longer τm for the sample crystallized at 90 °C implies that on average, more

significant constraints (e.g. arising from the relatively large fraction of amorphous

segments residing in tie-chains between lamellae) are imposed on mobile amorphous

segments in this sample, compared with those crystallized at the higher temperatures. This

is not due to crystalline content since the crystallinity of this sample is lower than the

others. Other findings [25] also support the idea that the crystal microstructure developed

at low Tc during very rapid crystallization is an important factor influencing segmental

dynamics. At Tc = 90 °C, although the spherulite growth rate is below its maximum value

(observed at Tc ~ 126-129 °C for L-PLA [4]), the primary nucleation rate is quite high. In

fact, crystallization on cooling from the melt is undoubtedly non-isothermal at this Tc due

to very rapid crystallization. The crystal microstructure developed under these conditions is

likely composed of ‘mobile’ amorphous segments that are relatively constrained due to the

rapid solidification.

The experimental τm were fit with the VFT equation [26]:

τm = τ0 exp(B/(T-T0)) (3)

where τ0 is a prefactor correlated to the time scale at which the molecules are attempting to

overcome an energy barrier, B is related to the strength for glass-forming, and T0 is a

temperature below Tg at which the segments become frozen. Taking τ0 to be 10-14 s [27], B

and T0 were calculated for the amorphous and crystallized materials and these are

summarized in Table 4.1. The parameter B is related to the fragility, m, a measure of the

rapidity with which a material changes its mean relaxation time in the vicinity of Tg [28]:

Page 90: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

74

20 )1(10ln

refref T

TT

Bm−∗∗

= (4)

Where Tref is a reference temperature, defined here as the temperature at which τmax = 100

s, and normally takes on value close to Tg obtained from standard scanning measurements

[29]. Within experimental uncertainity, the calculated fragilities for all amorphous and

crystalline samples are the same (m ~ 150, defining Tref as above). The invariance of m in

the presence and with degree of crystallinity is in keeping with previous findings [23,30].

Values for Tref are summarized in Table 4.1 for the various materials under

investigation in this study. There is no difference between Tref of the amorphous L-PLA

and meso-lactide copolymers, consistent with our previous measurement of Tgs of these

materials by DSC [4]. However, Tref increases by about 10 °C for crystallized L-PLA, not

an uncommon observed for crystallizable polymers, but changes very little (if at all) for the

lower crystallinity copolymers. In general, a slightly higher Tg has been observed in other

studies for samples crystallized at lower Tc (both cold and melt-crystallization), as well as

from non-isothermal crystallization at higher cooling rates [e.g. 21,31], and it does appear

that Tref is slightly higher for the L-PLA crystallized at 90 °C.

Normalized loss plots [ε″/ε″m vs. log (f/fm)] are depicted in Figures 4.9a and 4.9b

for the amorphous and crystallized samples of L-PLA and the 12% meso-lactide

copolymer, respectively. These figures present dielectric loss spectra at comparable

temperatures; i.e., at T = T0 + 50 °C and T0 + 55 °C for L-PLA and 12% meso-lactide

copolymer, respectively. Compared with the amorphous samples, it is apparent that the

relaxations are broader and more symmetrical when crystallinity is present. The fitted

Page 91: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

75

broadening parameters (a) are plotted in Figure 4.10 (for samples crystallized at or near

120 °C) as a function of temperature, demonstrating the influence of crystallinity. Smaller

values of a (i.e., broader segmental processes) are observed for crystalline L-PLA, in

contrast to the lower crystallinity copolymers.

(a)

PLLA0.0

0.2

0.4

0.6

0.8

1.0

ε"/ ε

" m

amorphousTc = 90 CTc = 120 CTc = 132 CTc = 141 C

(b)

12%meso-lactide

0.0

0.2

0.4

0.6

0.8

1.0

1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03log (f/fm)

ε"/ ε

" m

amorphousTc = 105 CTc = 111 CTc = 117 C

Figure 4.9 ε″/ε″m vs. log (f/fm) for (a) amorphous and crystallized L-PLA, and (b)

amorphous and crystallized 12% meso-lactide copolymer.

Page 92: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

76

Figure 4.10 HN broadening parameter, a, of amorphous (closed symbols) and L-PLA, 3%

and 6% meso-lactide copolymers crystallized at Tc = 120 °C, and 12% meso-

lactide copolymer crystallized at Tc = 117 °C (open symbols used for

crystallized samples). ( ) L-PLA, ( ) 3% meso-lactide, ( ) 6% meso-lactide

and ( ) 12% meso-lactide.

C. The rigid amorphous phase

Since the strength of the α relaxation is characteristic of the mobile phase, the

fraction of mobile amorphous segments (fmobile) in the crystallized samples that relax at a

given temperature can be written as [14]:

fmobile (T) = ∆εsc(T) / ∆εam(T) (5)

where ∆εsc and ∆εam are the relaxation strength of the crystallized and fully amorphous

samples, respectively. The rigid segment fraction (frigid) is therefore simply frigid(T) = 1 -

fmobile(T). In order to obtain fmobile, values of ∆εsc and ∆εam at the same temperature are

required. This is not feasible in the present case since crystallization occurs at temperatures

above 74 °C for the amorphous samples. Since ∆εam was found (Figure 4.6) to be constant

a

0.0

0.2

0.4

0.6

0.8

1.0

50 60 70 80 90 100Temperature (oC)

PLLA 3%meso 6%meso 12%meso

AmorphousCrystalline

Page 93: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

77

above 68 °C for the different polymers, ∆εam at temperatures above 74 °C was assumed to

take on these ‘limiting’ values. Finally, since the dielectric relaxation strength is

proportional to the number of dipoles per unit volume of a sample, in order to compare

fmobile with fcrystal determined from DSC or WAXD (and to calculate fRAP consistently),

fmobile (T) should be converted to mass fractions. Details of this conversion are reported

under Table 4.1. The changes in fmobile are relatively small, i.e., on the order of ~0.01.

(a)

0.0

0.2

0.4

0.6

60 70 80 90 100Temperature ( oC )

fmob

ile

1 2 3 4 5 6 7

(b)

0.0

0.2

0.4

0.6

60 70 80 90 100Temperature ( oC )

fmob

ile

8 9 10 11 12 13

Figure 4.11 Fraction of mobile phase as a function of temperature. Numbers in the figure

legend correspond to the sample number listed in Table 4.1.

Page 94: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

78

Values of fmobile (mass fraction) as a function of temperature for the crystallized

samples are displayed in Figure 4.11 (the samples are identified by a number

corresponding to Table 4.1). fmobile appears to increase slightly with temperature for all but

one sample (sample No.1), consistent with the idea of progressive loss of constraint for

rigid amorphous segments. However, caution must be exercised here since the uncertainty

in fmobile is estimated to be on the order of several percent.

0.0

0.2

0.4

0.6

0.8

1.0

0.0 0.2 0.4 0.6 0.8 1.0fcrystal

PLLA3%meso6%meso12%meso

Mobile Fraction

fmob

ile

Figure 4.12 Mobile segment fraction vs. crystalline fraction for all samples investigated.

The lowest temperature at which the α process can be resolved without significant

uncertainty due to overlap from dc conduction, is in the range of ~75 – 67.5 °C (from L-

PLA to the copolymers). Therefore comparison of fmobile and fRAP between the crystalline

samples was made at T0 + 50 °C (Table 4.1). The relationship between fmobile and fcrystal is

presented in Figure 4.12. The data for PLLA and the copolymers clearly deviate from the

prediction for a simple two-phase model (solid line). As noted earlier, the crystallinities in

Table 4.1 and Figure 4.12 were determined using DSC and ∆Hfo = 100 J/g [4]. Since other

values for ∆Hfo for PLLA have appeared in the literature [25,32,33], we confirmed the

crystallinities using wide-angle x-ray diffraction (WAXD) measurements. This was

Page 95: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

79

accomplished by dividing the area under the crystalline x- reflections in the 2θ range from

5 - 30˚ by the total area (crystalline reflections + amorphous halo). In fact, the WAXD

values and those determined from DSC using 100 J/g for ∆Hfo were found to be in very

good agreement.

For semi-flexible poly(ether ketone ketone) [31] and poly(phenylene sulfide) [34]

copolymers, the difference between the Tgs of amorphous and crystalline samples is lower

for copolymers with higher comonomer content – as observed here. This was correlated

with lower rigid amorphous phase content, which is reduced as more flexible meta-

linkages are incorporated into the para-connected homopolymers. In contrast, however, the

impact of meso-lactide comonomer on a reduction of fRAP is not observed for the

polylactide copolymers. This is good evidence that the effectiveness of the crystal phase in

hindering the segmental relaxation does not simply depend on the crystallinity and RAP

content, but also on an effective size of the non-crystalline phase modified during the

crystallization process [35,36].

4.4 Higher temperature relaxations

In addition to the primary focus on the segmental process of L-PLA and the

copolymers, the dielectric relaxation behavior at higher temperatures was investigated as

well. Since losses due to dc conduction are very strong in this higher temperature range, an

alternative representation of the experimental loss factor data is employed. The first

derivative of the real part of the dielectric constant (ε′) has been shown to be a good

Page 96: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

80

approximation of the Kramers – Kronig transform of ε′ to ε″, and to provide an ‘ohmic

conduction-free’ dielectric loss, ε″der. [37]:

ff

der ln)('

2"

∂∂

−=επε (6)

In a series of model calculations, ε″der and ε″ have been demonstrated to exhibit the same

peak frequencies, and as long as the relaxations are relatively broad (as they are in our

case), the relaxation strength of a derivative spectrum is a very good approximation to that

of ε″ [38].

In order to avoid crystallization during the course of the DRS experiment, spectra

were acquired for the amorphous 12% meso-lactide copolymer in the temperature range

from 130 – 155 °C (see Figure 4.13). The log ε″ and log ε″der. spectra at two temperatures

are presented in Figure 4.13a. Many characteristics of the high loss at lower frequency are

what are expected from electrode polarization [38]: very high dielectric constant (within

the range of 102 - 106, Figure 4.13b), and scaling of the slopes of the experimental and

derivative loss spectra at higher frequencies with ω-1 and ω-2, respectively.

Careful examination of the ε″der spectra (on both log and linear scales) suggests a

small shoulder at higher frequencies (fm ~ ca. 104 Hz) of lower strength. The relaxation

time associated with global chain dynamics (αn) is well known to depend strongly on

molecular weight, i.e., τnormal ∝ Mnx where x ~ 3.1 - 3.5 [2,39]. Based on the VFT

parameters reported in ref 2, αn is estimated to be near fm ~ 20 – 100 Hz, in the temperature

range of 125 – 155 °C, i.e., within ~2 decades of the observed process. Therefore, given

the approximate nature of the estimate, the similarity of the relaxation strength with that

Page 97: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

81

reported in ref 11, and that the sample is in the melt at these temperatures, this process is

assigned to the normal mode of the 12% meso-lactide copolymer chains.

(a)

1.E-01

1.E+01

1.E+03

1.E+05

1.E+07

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06

Frequency (Hz)

ε″

Amorphous130oC155oC

EP

1.E-01

1.E+01

1.E+03

1.E+05

1.E+07

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06

Frequency (Hz)

ε″

Amorphous130oC155oC

EP

(b)

1.E+00

1.E+01

1.E+02

1.E+03

1.E+04

1.E+05

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06Frequency (Hz)

ε′

155oC

130oC1.E+00

1.E+01

1.E+02

1.E+03

1.E+04

1.E+05

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06Frequency (Hz)

ε′

155oC

130oC

Figure 4.13 Plots of (a) log ε″ (closed symbols) and log ε″der (open symbols) and (b)

experimental ε′ vs. frequency at 130 ( ) and 155 °C ( ) for the amorphous

12% meso-copolymer. Data of ε″ were multiplied by 10 to offset its plots

from the ε″der plots.

DRS spectra (log ε″der) at three temperatures between 125 and 145 °C for L-PLA

crystallized at 90 °C are shown in Figure 4.14a. A process is clearly observed in the range

of ca. 1 – 10 Hz. This could originate from an αc process, arising from local motion in the

crystalline phase [9], Maxwell – Wagner – Sillars (MWS) interfacial polarization [40], or

Page 98: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

82

possibly electrode polarization. It is difficult to distinguish between these assignments,

although the peak position is comparable to that reported for the αc relaxation in

mechanical loss spectra of PLLA [24,41]. A process at a similar frequency is observed in

the 125 and 135 °C spectra of the 12% meso-lactide copolymer crystallized at 105 °C

(Figure 4.14b). In the 145 °C DRS spectrum, the loss at low frequencies rises appreciably.

Again, given the very high dielectric constant (>102, Figure 4.14c) and the scaling of the

‘high frequency’ slopes of the log ε″ and log ε″der representations in this frequency range

this is undoubtedly associated with electrode polarization [38]. It is important to note that

145 °C > Tm of the 12% copolymer (~140 °C), hence the similarity to the spectra in Figure

4.13. Note the appreciable intensity near 104 Hz in the 145 °C spectrum, similar to the

spectra of the amorphous 12% copolymer in Figure 4.13. This result supports the

assignment of the relaxation in this region to the normal mode (αn).

Page 99: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

83

(a)

(b)

(c)

ε´

Frequency (Hz)

Figure 4.14 Plots of ε″der vs. frequency at 125 ( ), 135 ( ) and 145 °C ( ) for (a)

crystalline L-PLA (Tc = 90 °C) and (b) 12% meso-lactide copolymer (Tc =

105 °C). Experimental ε´ of 12% meso-lactide copolymer (Tc = 105 °C) is

shown in (c).

1.E+00

1.E+02

1.E+04

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06

125C

135C

145C

Crystalline 12% meso

1.E-02

1.E+00

1.E+02

1.E+04

1.E+06

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06

125C

135C

145C

Crystalline PLLA

Frequency (Hz)

ε″

1.E-02

1.E+00

1.E+02

1.E+04

1.E+06

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06

125C

135C

145C

Crystalline PLLA

Frequency (Hz)

ε″

1.E-02

1.E+00

1.E+02

1.E+04

1.E+06

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06

125C

135C

145C

Crystalline 12% meso

Frequency (Hz)

ε″

1.E-02

1.E+00

1.E+02

1.E+04

1.E+06

1.E-02 1.E+00 1.E+02 1.E+04 1.E+06

125C

135C

145C

Crystalline 12% meso

Frequency (Hz)

ε″

Page 100: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

84

4.5 Summary

The dynamics of amorphous segments in fully amorphous and semi-crystalline

PLLA and three L-lactide/meso-lactide copolymers were investigated using DRS. The

difference in temperature-dependent behavior of the relaxation strength of the α process

for amorphous and crystalline samples is consistent with the presence of a rigid amorphous

component. Higher degrees of crystallinity generally result in lower ∆ε longer τm, and

broader the relaxation time distributions. Although it initially appears that the crystallinity

is the overriding factor controlling the characteristics of the α-process, a detailed analysis

reveals that the crystalline microstructure also play an important role.

A process is observed at higher temperatures (125 – 155 °C) for the amorphous

12% meso-lactide copolymer at 102–104 Hz. Evidence at hand suggests that this is

associated with the global chain relaxation (αn). The very pronounced process observed at

lower frequencies (10-1-102 Hz) in these samples undoubtedly arises from electrode

polarization. The origin of the process in this temperature range in crystallite samples

remains uncertain, although it occurs near the in the temperature range at which the αc

process has been observed in mechanical experiments.

Page 101: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

85

4.6 References

1. Starkweather, H. W. and Avakian, P. Macromolecules 1993, 26, 5084.

2. Mierzwa, M., Floudas, G., Dorgan, J., Knauss, D. and Wegner, J. J. Non-Cryst. Solids

2002, 307-310, 296.

3. Nogales, A., Ezquerra, T. A., Garcia, J. M. and Balta-Calleja, F. J. J. Polym. Sci.

Polym. Phys. 1999, 37, 37.

4. Ezquerra, T. A., Balta-Calleja, F. J. and Zachmann, H. G. Polymer 1994, 35, 2600.

5. Nogales, A., Ezquerra, T. A., Denchev, Z, Sics, I., Balta-Calleja, F. J. and Hsiao, B. S.

J. Chem. Phys. 2001, 115, 3804.

6. Alves, N. M., Mana, J. F., Balaguer, E., Mesequer Duenas, J. M. and Gomez Ribelles,

J. L Polymer 2002, 43, 4111.

7. Ezquerra, T. A. Unpublished results.

8. Cho, J. D., Baratian, S., Kim, J., Yeh, F., Hsiao, B. S. and Runt, J. Polymer 2003, 47,

711.

9. Huang, J., Lisowski, M. S., Runt, J., Hall, E. S., Kean, R. T., Buehler, N. and Lin J. S.

Macromolecules 1998, 31, 2593.

10. Lin, J., Shenogin, S. and Nazaranko Polymer 2002, 43, 4733.

11. Ivanov, D. A., Pop, T., Yoon, D. Y. and Jonas, A. M. Macromolecules 2002, 53, 9813.

12. Boyd R. H. Polymer 1985, 26, 323; Boyd R. H. Polymer 1985, 26, 1123.

13. Schlosser, E. and Schonhals, A. Colloid. Polym. Sci. 1989, 267, 963.

14. Huo, P. and Cebe, P. Macromolecules 1992, 25, 902; Huo, P. and Cebe, P. Polymer

1992, 30, 239.

Page 102: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

86

15. Nogales, A., Denchev, Z., Sics, I. and Ezquerra, T. A. Macromolecules 2000, 33, 9367.

16. McCrum, N. G., Read, B. E. and Williams, G. Anelastic and dielectric effects in

polymeric solids; Dover Publications; New York, 1991.

17. Schick, C., Wurm, A. and Mohammed, A. Thermochimica Acta 2003, 396, 119.

18. Pak, J., Pyda, M. and Wunderlich, B. Macromolecules 2003, 36, 495.

19. Kang, S. and Hsu, S. L. Macromolecules 2001, 34, 4542.

20. Yang, X., Kang, S., Hsu, S. L., Stidham, H. D., Smith, P. B. and Leugers, A.

Macromolecules 2001, 34, 5037.

21. Kang, S., Zhang, G., Aou, K., Hsu, S. L., Stidman, H. and Yang, X. J. Chem. Phys.

2003, 118, 3430.

22. Blomqvist, J. Polymer 2001, 42, 3515.

23. Mijovic, J. and Sy, J.W. Macromolecules 2002, 35, 6370.

24. Urayama, H., Kanamori, T. and Kimura, Y. Macromol. Mater. Eng. 2001, 286, 705.

25. Iannace, S. and Nicolais, L. J. Appl. Polym. Sci. 1997, 64, 911.

26. Ferry, J. D. Viscoelastic Properties of Polymers; Wiley: New York, 1980.

27. Jin, X., Zhang, S. H. and Runt, J. Macromolecules 2003, 36, 8033.

28. Angell, C. A. Science, 1995, 267, 1924.

29. Angell, C. A. J Non-Cryst Solids 1991, 131-133, 13.

30. Ngai, K. L. and Roland, C. M. Macromolecules 1993, 26, 6824.

31. Kalika, D. S. and Krishnaswamy, R.K. Macromolecules 1993, 26, 4252;

Krishnaswamy, R. K. and Kalika, D. S. Polymer 1996, 37, 1915.

32. Fischer, E. W., Sterzel. and Wegner, H. J. Kolloid-Z. Z. Polym. 1973, 251, 980.

Page 103: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

87

33. Cohn, D., Younes, H. and Marom, G. Polymer 1987, 28, 2018.

34. Kalika, D. S. and Wu, S. S. J. Macromol. Sci.-Phys. 1996, B35, 179.

35. Schick, C. and Nedbal, J. Progr. Colloid. Polym. Sci. 1988, 78, 9.

36. Huo, P. and Cebe, P. Colloid Polym. Sci. 1992, 270, 840.

37. Steeman, P. A. M. and van Turnhout, J. Macromolecules, 1994, 27, 5421.

38. Wübbenhorst M. and van Turnhout, J. J. Non-Cryst. Solids 2002, 305, 40.

39. Ren, J., Urakama, O. and Adachi, K. Macromolecules 2003, 36, 210.

40. Hayward D., Pethrick R. A. and Siriwittayakorn T. Macromolecules 1992, 25, 1480.

41. Pluta, M. and Galeski, A. J. Appl. Polym. Sci 2002, 86, 1386.

Page 104: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

Chapter 5

L-PLAs and D-PLA Blends

Experimental Tgs of amorphous L-PLA1/D-PLA (LB) blends as a function of

D-PLA (XD) content are presented in Figure 5.1, as well as Tgs estimated using a simple

Fox mixing rule, 1/Tg (blend) = W1/Tg(1) + W2/Tg(2) where W1 and W2 = weight fraction

of each component. A single Tg is observed for all LB blends, intermediate between those

of the component polymers. This indicates miscibility between these enantiomeric

polylactides at a length scale of ~10 nm, the approximately experimental probe size of

DSC. Even though similar behavior is observed for the HB blends, the component Tgs are

too close to one other to deduce the phase behavior.

40

45

50

55

60

65

70

0 0.2 0.4 0.6 0.8 1Fraction of D-PLA

Tg (o C

)

LB

HB

Fox eq.

Figure 5.1 Experimental Tgs of L-PLA1/D-PLA ( ) and L-PLA2/D-PLA ( ) as

a function of D-PLA composition. Predictions from a simple Fox-Flory

mixing rule are shown by the solid lines.

Page 105: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

89

5.1 Crystallization from the glassy state

Figure 5.2 shows the thermograms on heating at 10 °C for the different

homopolymers and blends, from the glassy state to above Tm. Comparison of the cold

crystallization exotherms (Tc,c) for both blend series demonstrate that crystallization

depends strongly on blend composition. Crystallization on heating D-PLA rich blends

(Figure 5.2c) occurs at temperatures significantly lower than that for neat D-PLA. This is

also the case for L-PLA rich blends (Figure 5.2b and d).

(a) (b)

D-PLA

L-PLA2

L-PLA1Tm1Tc,c

70 90 110 130 150 170 190Temperature (oC)

44 J/g

54 J/g

49 J/g

Hea

t flo

w (W

/g),

endo D-PLA

L-PLA2

L-PLA1Tm1Tc,c

70 90 110 130 150 170 190Temperature (oC)

44 J/g

54 J/g

49 J/g

Hea

t flo

w (W

/g),

endo

Hea

t flo

w (W

/g),

endo

HB55HB64HB73

Tm2

Tm1

Tc,c

80 120 140 160 180 200Temperature (oC)

100 220 240

46 J/g48 J/g11 J/g

8 J/g 42 J/g

L-PLA2

54 J/g

Hea

t flo

w (W

/g),

endo

HB55HB64HB73

Tm2

Tm1

Tc,c

80 120 140 160 180 200Temperature (oC)

100 220 240

46 J/g48 J/g11 J/g

8 J/g 42 J/g

L-PLA2

54 J/g

(c) (d)

LB28LB37LB46LB55

Tc,c Tm2

60 80 120 140 160 180 200Temperature (oC)

100 220 240

34 J/g51 J/g62 J/g67 J/g

D-PLA44 J/g Tm1

Hea

t flo

w (W

/g),

endo

LB28LB37LB46LB55

Tc,c Tm2

60 80 120 140 160 180 200Temperature (oC)

100 220 240

34 J/g51 J/g62 J/g67 J/g

D-PLA44 J/g Tm1

Hea

t flo

w (W

/g),

endo

Temperature (oC)

LB64

LB73

LB82

Tc,c

Tm2Tm1

60 80 120 140 160 180 200100 220 240

63 J/g

49 J/g

35 J/g5 J/gHea

t flo

w (W

/g),

endo

49 J/g

L-PLA1

Temperature (oC)

LB64

LB73

LB82

Tc,c

Tm2Tm1

60 80 120 140 160 180 200100 220 240

63 J/g

49 J/g

35 J/g5 J/gHea

t flo

w (W

/g),

endo

49 J/g

L-PLA1

Figure 5.2 DSC heating thermograms illustrating cold crystallization and melting

processes of initially amorphous samples; (a) neat D-PLA, L-PLAs,

(b) HB blends and (c-d) LB blends.

Page 106: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

90

The subsequent melting endotherms of all blends are clearly observed at ~215 - 217

°C (Tm2, in Figure 5.2b – 5.2d) revealing that the cold crystallization of these blends is

primarily associated with stereocomplex formation. For the homopolymers, a single

melting endotherm associated with homopolymer crystallites (Tm1) is observed near 160 °C

for L-PLA1 (Mn = 1.1 x 104, PDI = 1.6) and 178 °C for L-PLA2 (Mn = 5.5 x 104, PDI = 2)

and D-PLA (Figure 5.2a). These results demonstrate that crystallization of stereocomplex

crystallites is favored over homopolymer crystals under these crystallization conditions.

5.2 Real-time crystallization investigated using DRS

Figure 5.3 presents dielectric loss spectra of the segmental relaxation monitored

during a real time crystallization experiment from the glassy state of an initially amorphous

LB73 blend at Tiso = 60 °C. These spectra were fit with the HN function [1] and τm, ∆ε and

the distribution shape parameters are presented in Figure 5.4. These parameters are clearly

modified during the crystallization, similar to the observations reported for other semi-

crystalline polymers [2-4]. The reduction in ∆ε with tc (Figure 5.4b) results from

consumption of mobile amorphous segments during the crystallization process. As

crystallization proceeds, the remaining mobile amorphous segments become constrained to

varying degrees, giving rise to a longer τm (lower fm) (Figure 5.4a). The distribution of

relaxation times becomes broader upon crystallization due to the broad range of

environments experienced by the remaining mobile amorphous segments.

Page 107: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

91

0.05

0.15

0.25

0.35

0.45

1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06Frequency (Hz)

Die

lect

ric lo

ss

0 24 68 1838 64114 314674 894

Figure 5.3 Dielectric loss spectra acquired during isothermal crystallization of the 70:30

L-PLA1/D-PLA blend at 60 °C as a function of tc. Crystallization times (in min)

are shown in the figure legend.

(a)

(b)

0.0

0.4

0.8

1.2

1.6

2.0

2.4

0.0 0.5 1.0 1.5 2.0log tc (min)

a, b

, ∆ε

a

b

∆ε

0.0

0.4

0.8

1.2

1.6

2.0

2.4

0.0 0.5 1.0 1.5 2.0log tc (min)

a, b

, ∆ε

a

b

∆ε

Figure 5.4 Fitting results of the dielectric spectra of LB73 crystallized at Tiso = 60 °C as a

function of tc; (a) mean relaxation time, (b) broadening parameter “a”,

skewness parameter “b” and segmental relaxation strength (∆ε).

1.E-04

1.E-03

0.0 1.0 2.0

log (τm)

Page 108: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

92

5.2.1 Crystallization mechanism

The amount of remaining mobile segments at any given tc is proportional to

∆εt/∆ε0, where ∆εt and ∆ε0 are the relaxation strengths at tc and zero time, respectively.

Although rigid segments are not completely composed of crystalline phase, frigid = 1 - fmobile

= fcrystal + fRAP, results in Table 5.1 suggest that crystalline segments constitute the majority

of the rigid segments (discussed later). Consequently, the formation of a rigid segment

fraction (frigid), estimated from 1- (∆εt/∆ε0), as a function of tc is a measure of the

crystallization kinetics. As expected, the induction time and crystallization rate of the

homopolymers and HB blends are influenced by Tiso (Figure 5.5). Crystallization from the

glassy state is enhanced at higher Tiso (larger Tiso – Tg = TDiff).

(a) (b)

0.0

0.2

0.4

0.6

0.8

0.0 1.0 2.0 3.0Log tc (min)

Rig

id fr

actio

n

PLA2_70 PLA2_80DPLA_70 DPLA_80

0.0

0.2

0.4

0.6

0.8

0.0 1.0 2.0 3.0Log tc (min)

Rigi

d fra

ctio

n

HB55_70 HB64_70 HB73_70 PLA2_70HB55_80 HB64_80 HB73_80 PLA2_80

Figure 5.5 Rigid segment fractions as a function of crystallization time of

(a) homopolymers L-PLA2 ( ) and D-PLA ( ) and (b) L-PLA2/D-PLA

blends (HB55 ( ), HB64 ( ), HB73 ( ) and L-PLA2 ( )); at Tiso = 70 °C

(open symbols) and 80 °C (closed symbols)

For the neat homopolymers (Figure 5.5a), the crystallization rate of L-PLA2 is

comparable to that of D-PLA. Unfortunately, the crystallization kinetics of low molecular

weight L-PLA1 could not be investigated by dielectric experiments due to difficulty in film

Page 109: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

93

preparation (noted previously in Chapter 2). However, the much lower Tc,c of L-PLA1

(Figure 5.2a) compared to that of L-PLA2 and D-PLA strongly suggests that L-PLA1

crystallizes much more rapidly than the higher molecular weight homopolymers.

Figure 5.5b demonstrates that the induction time (t0) and crystallization half-time,

t1/2, of the HB blends are considerably shorter than those of neat polymers at Tiso = 70 °C.

However at Tiso = 80 °C, crystallization of the HB blends occurs at comparable rates to that

of the parent homopolymers. Since Tgs of HB and LB blends are considerably different, it

is most appropriate to compare the behavior of these blends at similar TDiff. At TDiff ~ 10, it

is obvious that t0 and t1/2 for the LB blends (Figure 5.6) are much smaller than those of HB

blends (i.e., at Tiso = 70 °C for the latter, Figure 5.5b). In order to rationalize these findings,

the crystalline structure of the fully crystallized samples (i.e., after the in situ dielectric

experiments) was investigated by WAXD and DSC. The fractions of homopolymer and

stereocomplex crystallites determined by these techniques, as well as the rigid fraction

estimated from DRS at the conclusion of crystallization, are summarized in Table 5.1.

Figure 5.6 Rigid fractions of D-PLA and LB blends as a function of crystallization time at

a similar TDiff (= 10˚C); Tiso = 70, 65, 60 and 57.5 °C for DPLA, LB37, LB55

and LB73, respectively.

0.0

0.2

0.4

0.6

0.8

0.0 1.0 2.0 3.0Log tc (min)

Rigi

d fra

ctio

n

DPLA_70

LB37_65

LB55_60

LB73_57.5

Page 110: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

94

Table 5.1: Glass transition temperature, crystal fractions of stereocomplex (Xc2) and

homopolymer crystallites (Xc1) determined by WAXD and DSC, as well as the rigid

fraction (frigid) estimated from DRS for quenched and crystallized HB and LB blends

Sample Tiso (°C) Tg ∆Hm2/∆Hm1

J/g

Xc2 /Xc1

DSC

Xc2 /Xc1

WAXD

frigid

DRS

D-PLA Amorphous

70

80

60

68

70

0/48

0/43

0/0.48

0/0.43

0/0.47

0/0.53

0.58

0.67

L-PLA2 Amorphous

70

80

57

65

70

0/58

0/60

0/0.58

0/0.60

-

-

0.57

0.74

HB55 Amorphous

70

80

58

61

64

48/6

53/14

0.34/0.03

0.37/ 0.02

0.23/0.03

0.40/0

0.37

0.43

HB64 Amorphous

70

80

58

61

64

40/11

39/20

0.28/0.11

0.27/0.20

0.16/0.09

0.24/0.17

0.37

0.48

HB73 Amorphous

70

80

58

60

62

36/13

42/17

0.25/0.13

0.29/0.17

0.20/0.08

0.28/0.18

0.38

0.48

LB37 Amorphous

65

70

56

57

61

52/0

54/1

0.37/0

0.38/0.01

0.35/0

0.43/0

0.44

0.53

LB55 Amorphous

60

70

50

56

58

64/0

65/0

0.45/0

0.46/0

0.47/0

0.57/0

0.47

0.59

LB73 Amorphous

57.5

70

47

53

52

52/9

52/10

0.37/0.09

0.37/0.10

0.21/0.10

0.47/0.06

0.47

0.42

Page 111: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

95

Only racemic crystallites are developed in the 1:1 blends (HB55 and LB55),

regardless of L-PLA molecular weight. This clearly demonstrates that the component

homopolymers of HB55 (and LB55) are miscible. The influence of blend composition on

racemic and homopolymer crystal fractions of non-equimolar blends (Figure 5.7) is

generally similar to that reported previously for blends of similar molecular weight [5-7].

A reduction in racemic crystallite content is observed for XD > 0.5 or XD < 0.5, and vice

versa for homopolymer crystallites.

0

20

40

60

80

0 0.2 0.4 0.6 0.8 1

DPLA fraction

∆H m

, J/g

of p

olym

er

Figure 5.7 ∆Hm of LB blends after crystallization in real-time DRS experiments at

Tiso = 70 °C; Closed and open symbols represent the stereocomplex and the

homopolymer crystals, respectively.

Insight into the mechanism of crystallization of the blends can be gained by

comparing the crystallization kinetics to the crystalline fractions and structures (Table 5.1

and Figures 5.8 - 5.10). At Tiso = 70 °C, the findings for HB55 (only racemic crystallites

developed) and the neat homopolymers suggest that stereocomplex formation, by packing

L-PLA and D-PLA chains alternately [8], has a higher nucleation and crystallization rate

than homopolymer crystallization of the non-blended samples. Surprisingly, t0 and t1/2 of

non-equimolar HB blends where both stereocomplex and homopolymer crystallites are

Page 112: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

96

developed are similar to those of HB55. This implies that compared to the neat polymers,

the shorter induction period for crystallization in all HB blends (Figure 5.5b) is due to the

faster nucleation rate of the stereocomplex. Additionally, it is proposed that homopolymer

crystallites in non-equimolar HB blends are developed after stereocomplex formation is (or

nearly) completed at Tiso = 70 °C. Such behavior has also been reported for L-PLA/D-PLA

blend crystallization from the melt (i.e., at Tiso = 140 °C) [6-7].

(a) (b)

L-PLA1

L-PLA2D-PLA

156 oC

174 oC

176 oC

Tiso = 70 oC

70 110 170Temperature (oC)

19090 130 150

Hea

t flo

w (W

/g),

endo

L-PLA1

L-PLA2D-PLA

156 oC

174 oC

176 oC

Tiso = 70 oC

70 110 170Temperature (oC)

19090 130 150

Hea

t flo

w (W

/g),

endo

HB73

165 oC

40 60 80 120

Hea

t flo

w (W

/g),

endo Tg

100 160140 180Temperature (oC)

HB64

HB73

165 oC

40 60 80 120

Hea

t flo

w (W

/g),

endo Tg

100 160140 180Temperature (oC)

HB64

Figure 5.8 DSC thermograms presenting the melting process (Tm1) of homopolymer

crystallites of (a) neat D-PLA, and the L-PLAs (at Tiso = 70 °C) and (b) HB64

(1st and 2nd curves from the top) and HB73 blends (3rd and 4th from the top)

crystallized at 70 (1st and 3rd curve) and 80 °C (2nd and 4th curve)

Page 113: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

97

205

215

225

0 0.2 0.4 0.6 0.8 1DPLA fraction

Tm (o C

)

LB(70) LB(80) HB(70) HB(80)

Figure 5.9 Melting points of stereocomplex crystals (Tm2) formed in HB (open symbols)

and LB blends (closed symbols) crystallized at 70 ( ) and 80 °C ( ).

It was found that while Tm1 of the L-PLA2-rich samples (Figure 5.8b, at Tiso = 70

°C) is considerably lower than that of neat L-PLA2 (Figure 5.8a), Tm2 of all HB blends

(Figure 5.9, at Tiso = 70 °C) is similar (i.e., between 215-217 °C). The influence of

crystallization of HB blends at Tiso = 80 °C and LB blends on the Tm1 and Tm2 (Figure 5.8 -

5.10) is similar to that observed in the HB blends at Tiso = 70 °C. This suggests that

homopolymer crystals in these blends are formed under constraints imposed by existing

stereocomplex crystallites, limiting the lamellae thickness of homopolymer crystallites.

Similar results have been reported in melt-miscible, crystalline/crystalline blends of poly

(butylene terephthalate, PBT) and polyarylates (PAr) [9,10]. PBT and PAr crystals coexist

in 1:1 blends. The Tm of PBT crystals in the samples where PBT crystals developed after

PAr crystallization is lower than in the case where crystallization of PBT and PAr is in the

reverse sequence [9].

Page 114: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

98

1

0

1

2

3

LB82

40 60 80 120H

eat f

low

(W/g

), en

do Tg

100 160140

Temperature (oC)

LB28

180

LB37

LB73

Tm1

1

0

1

2

3

LB82

40 60 80 120H

eat f

low

(W/g

), en

do Tg

100 160140

Temperature (oC)

LB28

180

LB37

LB73

Tm1

Figure 5.10 DSC thermograms presenting the melting process (Tm1) of homopolymer

crystallites of LB blends at Tiso = 70 °C

5.2.2 Rigid amorphous phase

Table 5.1 shows that the content of homopolymer and stereocomplex crystallites

obtained from DSC are comparable to those from WAXD within experimental uncertainty.

In addition, the difference between the total crystal fraction (Xc1+Xc2) and frigid of the

blends is much smaller than that of the homopolymers. The results reported in chapter 4

(Table 4.1) for the L-lactide/meso-lactide copolymers demonstrate that frigid is ~ 0.2 - 0.4

higher than fcrystal for samples having crystallinities similar to those of the L-PLA/D-PLA

blends. Although the results suggest that the amount of rigid amorphous phase (fRAP, = frigid

- fcrystal) is much smaller for stereocomplex-rich samples, further conclusions about the

extent of RAP (or lack thereof) in L-PLA/D-PLA blends are not possible at this time. This

is because experiments could only be carried out at a few Tisos for each blend and the

uncertainty of the findings at ‘high’ Tiso (i.e., Tiso = 70 for LB blends), as discussed in the

next paragraph.

At Tiso = 70 °C, frigid of LB blends estimated from the measured ∆ε0 is considerably

lower than fcrystal derived from DSC and WAXD. Since it is not possible that frigid < fcrystal,

Page 115: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

99

the results indicate that the first dielectric spectrum recorded, when the temperature is

stabilized at 70 °C, is influenced by crystallization during the heating process and/or during

the stabilization time at Tiso. In order to correct for this, frigid at this Tiso is estimated from

∆ε0 from lower Tiso, where crystallization prior to acquiring the first dielectric spectrum

does not occur. Although the corrected frigid of these blends (shown in Table 5.1) are

comparable to the degree of crystallinity from DSC and WAXD, it is clearly some

uncertainty associated with this procedure. This was the principle reason that DRS

experiments at higher Tiso were not carried out.

5.2.3 Influence of crystalline phase on amorphous segments

The difference between Tg of the amorphous and crystallized samples (∆Tg) for the

LB series blends, HB55, D-PLA and L-PLA2 are plotted as a function of degree of

crystallinity in Figure 5.11. The straight line in this figure illustrates the relationship

between ∆Tg and crystalline content for LB55 and HB55, in which only stereocomplex

crystals are present. The results in this figure indicate that constraints imposed by the

crystalline phase on amorphous segments of D-PLA and L-PLA blends strongly depend on

the crystal content and are insensitive to the ratio of stereocomplex to homopolymer

crystallites.

Page 116: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

100

0

5

10

15

0 10 20 30 40 50 60 70% crystallinity

∆T g

(o C)

LB37LB55LB73HB55D-PLAL-PLA2

Figure 5.11 Difference in Tg between crystallized and amorphous samples as a function of

total percent crystallinity determined from WAXD for ( ) LB37, ( ) LB55,

( ) LB73, ( ) HB55, ( ) D-PLA and ( ) L-PLA2. The line in this figure

indicates the relationship between ∆Tg and crystalline content for LB55 and

HB55 in which only stereocomplex crystals are present.

5.3 Summary

It is evident from the results of non-isothermal DSC experiments that crystallization

of all L-PLA/D-PLA blends is faster than that of their component homopolymers. It is

suggested from results of crystallization kinetics experiments and investigation of the

crystalline unit cell structure that stereocomplex formation is favored in equimolar and

non-equimolar L-PLA/D-PLA blends. Crystallites developed from the excess

homopolymer in such blends are observed to have a much lower melting temperature than

the neat component polymers, and it is proposed that the lower Tm1 arises due to

crystallization of the homopolymer under constraint from existing racemic crystallites. The

highest stereocomplex content was observed at XD = 0.5 and decreased as excess

homopolymer increased. The modification of Tg as a function of degree of crystallinity

reveals that constraints imposed on mobile amorphous segments is, as expected, strongly

Page 117: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

101

dependent on crystalline content, but not the ratio of stereocomplex to homopolymer

crystallites.

5.4 References

1. Havriliak, S. and Negami, S. J. Polym. Sci. Polym. Symp. 1966, 14, 99

2. Ezquerra, T. A., Balta-Calleja, F. J. and Zachmann, H. G. Polymer 1994, 35, 2600.

3. Nogales, A., Ezquerra, T. A., Denchev, Z., Sics, I., Balta-Calleja, F. J. and Hsiao, B. S.

J. Chem. Phys. 2001, 115, 3804.

4. Kanchanasopa, M. and Runt, J. Macromolecules 2004, 37, 863.

5. Tsuji, H., Hyon, S. H. and Ikada, Y. Macromolecules 1991, 24, 5651.

6. Tsuji, H. and Ikada, Y. Macromolecules 1993, 26, 6918.

7. Brochu, S., Prud’homme, R. E., Barakat, I. and Jerome, R. Macromolecules 1995, 28,

5230.

8. Brizzolara, D., Cantow, H. J., Diederichs, K., Keller, E. and Domb, A. J.

Macromolecules 1996, 29, 191.

9. Liu, A. S., Liau, W. B. and Chiu, W. Y. Macromolecules 1998, 31, 6593.

10. Liau, W. B., Liu, A. S. and Chiu, W. Y. Macromol. Chem. Phys 2002, 203, 294.

Page 118: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

Chapter 6

Summary and Suggestions for Future Research

6.1 Summary

In this thesis, solid-state structure and chain dynamics of polylactide

stereocopolymers and enantiomeric blends were studied using a variety of experimental

tools. Morphology and lamellar microstructure of L-PLA and random L-lactide/meso-

lactide copolymers isothermally crystallized at selected temperatures were investigated

by tapping mode AFM. This was the first study on the morphology of polylactide

stereocopolymers using a real-space probe [1]. Qualitative and quantitative analysis of

lamellar structure were performed using height and phase images. Image features

acquired at various set-point amplitude ratios suggested that the surfaces of copolymer

samples were covered with amorphous segments. More open spherulites with an

abundance of screw dislocations between edge-on lamellar stacks were observed in

samples crystallized at higher temperatures. Mean lamellar thicknesses are lower for the

random copolymers compared to L-PLA, particularly at lower ∆T. Mean lamellar

thicknesses derived from AFM are in very good agreement with those determined

previously from SAXS. Surfaces of microtomed specimens were also studied to

investigate the bulk crystal morphology. Although quantitative analysis was not feasible

due to knifemarks induced by microtoming process, lamellar organization similar to that

seen in the free surface experiments was observed at high magnifications.

A broadband dielectric investigation of crystalline and amorphous samples of L-

PLA and L-lactide/meso-lactide copolymers revealed the influence of crystalline

Page 119: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

103

microstructure on chain dynamics. Two relaxations, the segmental (α) and sub-glass (β)

processes, were observed in the temperature range between –100 and 105 oC. Unlike the

α-process, the β local relaxation is not affected by the presence of the crystalline phase.

Modification of the characteristics of the α process, i.e., the relaxation strength, mean

relaxation time and shape parameters, as a function of crystal content was investigated on

‘fully crystallized’ samples as well as in real-time crystallization experiments. As

expected, the strength of the α relaxation was smaller and its distribution broader for

materials with higher degrees of crystallinity. The relaxation strength of the α process of

the amorphous samples increased with temperature, while that of the crystalline materials

changed very little or in the opposite direction with temperature. Combined with the

deviation from the relationship between crystal fraction and mobile fraction for a simple

two phase model, this behavior was explained by the existence of a rigid amorphous

component. Changes in mean relaxation time and the Vogel temperature for the

copolymers crystallized at various temperatures are influenced by crystalline

microstructure, as well as by crystalline content. A relaxation is also observed in higher

temperature (125 – 155 oC) experiments on amorphous and crystalline materials. The

evidence at hand suggests that these processes are associated with the normal mode and

αc process, respectively.

For enantiomeric blends of D-PLA with two different molecular weight L-PLAs,

crystallization rate was found to be higher than that of the homopolymers D-PLA and L-

PLA. Results from DSC (supported by those from WAXD) reveal the presence of two

crystalline structures (i.e., stereocomplex and homopolymer crystallites,) whose

proportions are dependent on blend composition, molecular weight and crystallization

Page 120: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

104

conditions. Homopolymer crystallites, developed in the presence of the stereocomplex,

melt at lower temperatures than the neat homopolymers, indicating thinner lamellae in

the former. It was postulated that this results from a decrease in average crystallizable

length, as well as special constraints imposed by stereocomplex crystals, on the portions

of homopolymer chains remaining after stereocomplex formation. The melting

temperature of stereocomplex crystallites remains unchanged with blend composition

and crystallization conditions considered in this study. The Tgs of these blends suggest

that the influence of the stereocomplex on mobility of amorphous segments is similar to

that of homopolymer crystallites. In general, the influence of constraints imposed by the

crystalline phase on the chain dynamics of the enantiomeric blends is similar to that

observed for the lactide stereocopolymers.

6.2 Suggestions for future research

Real-time crystallization of L-PLA/D-PLA blends by DRS (chapter 5) revealed

that crystallization of these blends from the glassy state is very fast and likely to begin

before reaching the desired crystallization temperature, especially for higher temperature

experiments. This, combined with limitation of the range of measurement frequencies,

hinders further experiments at higher crystallization temperatures. Some preliminary

studies on isothermally crystallized LB55 blends (from the molten state) in a hot-stage

were undertaken. Figures 6.1a-c present DSC results of samples crystallized at Tc = 140,

160 and 170 oC, respectively. These thermograms demonstrate that multiple melting

endotherms are developed under these conditions, except at Tc = 170 oC. In addition, the

Tms and relative proportion of each melting endotherm depend on tc and Tc. This is in

Page 121: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

105

contrast to the well-defined single melting peak observed for stereocomplex crystallites

throughout the investigation of fully crystallized samples from the glassy state (Chapter

5). Additional studies on the morphology, crystallization and melting processes of these

blends are necessary to explain these results and to fully understand the crystallization

and physical properties of these blends.

Time-resolved simultaneous WAXD/SAXS is a powerful technique for

investigating the development of crystal microstructure, i.e. crystal form and lamellar

morphology, during crystallization and melting [2-3]. Applied to the study of

crystallization of L-PLA/D-PLA blends, it is not only possible to examine crystallization

kinetics over a broad temperature range, but it can also provide insight into the

mechanism of homopolymer and stereocomplex crystallization throughout the

crystallization process. The melting processes can be followed in real time using this

experimental approach as well. In addition, the study of blends of enantiomeric

polylactides having different optical purity (e.g. D-PLA and L,meso-lactide copolymers)

should also be considered. It is likely that the crystallization behavior of these blends is

different from that of L-PLA/D-PLA blends.

Moreover, information from high resolution AFM will be very helpful for

understanding the organization of stereocomplex and homopolymer lamellae within the

crystal microstructures of L-PLA/D-PLA blends.

Page 122: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

106

(a)

Hea

t Flo

w (W

/g)

50 100 150 200 250

Temperature (°C)

59 J/g

61 J/g

76 J/g

78 J/g

73 J/g

202 oC

203 oC

218 oC

217 oC

220 oC

Tc = 140 oC for 5, 30 min, 2.5, 3.5 and 9.5 hrs (bottom to top)

(b)

Hea

t Flo

w (W

/g)

50 100 150 200 250

Temperature (°C)

77 J/g

72 J/g

71 J/g

92 J/g

215 oC

214 oC

217 oC

218 oC

Tc = 160 oC for 5, 10, 20 min and 12 hrs (bottom to top)

(c)

Hea

t Flo

w (W

/g)

50 100 150 200 250

Temperature (°C)

71 J/g

73 J/g

67 J/g 217 oC

212 oC

213 oC

28 J/g

22 J/g

Tc = 170 oC for 5, 10 and 20 min (bottom to top)

Figure 6.1 DSC thermograms of LB55 crystallized from the melt at (a) 140 oC,

(b) 160 oC and (c) 170 oC for different tc.

Page 123: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

107

6.3 References

1. Kanchanasopa, M., Manias, E. and Runt, J. ACS Polym. Prepr. 2001, 42(2), 300.

2. Cho, J. D., Baratian, S., Kim, J., Hsiao, B. S. and Runt, J. Polymer 2003, 44, 711.

3. Ania, F., Flores, A. and Calleja, F. J. B. J. Macromol. Sci. Phys. 2003, 42, 653.

Page 124: SOLID-STATE STRUCTURE AND DYNAMICS OF LACTIDE …

VITA

Mantana Kanchanasopa was born in Thailand on July 12, 1972. In 1994, she

received B.S. degree (second honors) in Materials Science from Chulalongkorn University.

She was then awarded a full scholarship from National Science and Technology

Development Agency (NSTDA) for continuing her M.S. degree in Polymer science at

Petroleum and Petrochemical College, Chulalongkorn University. After her graduation in

1996, she had worked with both government organizations and industrial sector; NSTDA,

Petroleum Authority of Thailand (PTT), and Thai Petrochemical Industry (TPI). In Fall

1999, she was granted a scholarship from The Royal Thai Government to pursue her Ph.D.

studies in Department of Materials Science and Engineering (Polymer science program),

The Pennsylvania State University.