+ All Categories
Home > Documents > Photochemical Crosslinking Reactions in Polymers

Photochemical Crosslinking Reactions in Polymers

Date post: 31-Dec-2016
Category:
Upload: vanthu
View: 227 times
Download: 2 times
Share this document with a friend
248
Photochemical Crosslinking Reactions in Polymers Nicholas Carbone Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy In the Graduate School of Arts and Sciences Columbia University 2012
Transcript
Page 1: Photochemical Crosslinking Reactions in Polymers

Photochemical Crosslinking Reactions in Polymers

Nicholas Carbone

Submitted in partial fulfillment of the

requirements for the degree of

Doctor of Philosophy

In the Graduate School of Arts and Sciences

Columbia University

2012

Page 2: Photochemical Crosslinking Reactions in Polymers

©2011

Nicholas Carbone

All rights reserved

Page 3: Photochemical Crosslinking Reactions in Polymers

ABSTRACT

Photochemical Crosslinking Reactions in Polymers

Nicholas Carbone

The post-synthesis modification of polymer properties has very broad applications in

industry. It is employed to produce products that are impossible to directly synthesize, modify

biomolecules for medical use, and provide compounds for industrial and academic research.

Modifying the polymer molecular weight distribution through crosslinking is one of the simplest

methods of achieving the desired properties. The crosslinking of polymers to form gels has

been used for decades in the automotive industry to produces tires. More recently, polymer

crosslinking has been applied to environmental cleanup, wound healing materials, consumer

products, artificial organs, self-healing coatings, and the micro-patterning of surfaces.

Photocrosslinking using additives is one of the safest and most robust methods as it allows

precise control of the reaction in space and time.

This thesis explores the photocrosslinking of polymers and relates it to structure and

mechanism: the properties and intermiscibilities of four crosslinker chemistries and five

polymers are rationally designed in Chapter 3. The influence of additive functionality on the

reaction is explored in depth with a single additive chemistry in Chapter 4. The mechanism and

reaction location is examined in Chapter 5, and a survey of the efficacy of additional

crosslinking chemistries is performed in Chapter 6.

Page 4: Photochemical Crosslinking Reactions in Polymers

Multiple polymers and additives are examined in Chapter 3 and their reactions and

mechanisms are examined to predict efficacy and utility. It is shown that multiple reaction

chemistries allow crosslinking, and that the manipulation of functionality and polymer allows

the exploration of specific reaction mechanisms. The differential refractive indices of the

polymers are measured by experiment, and the intermiscibility of polymer-additive systems are

calculated using group contribution techniques.

A hydrogen abstraction induced radical crosslinking mechanism is explored in depth in

Chapter 4. Benzophenone-derived additives are used to study photocrosslinking in thin films

while varying multiple parameters: the irradiation time, additive to polymer molar ratio,

additive functionality, and polymer mobility. Bi-functionality is found to increase the density of

radicals in glassy and rubbery systems. The macroradical recombination and scission reactions

are modeled and shown to conform to experiment. Analysis of the model results shows that

the functionality of the additive is only important above a molar ratio threshold. Below the

threshold combination reactions are binary and there are no macroradical bridging reactions in

the bi-functional system. Above the threshold the density of radicals is so high that the

combination reactions are pseudo first order and macroradical bridging causes differences in

the behavior of mono- and bi-functional additive systems. The changes in the molecular weight

distributions with reaction extent are tracked using size exclusion chromatography (SEC) with

multiple detectors.

Chapter 5 studies the hydrogen abstraction reaction of Chapter 4 using electron

paramagnetic resonance (EPR) to confirm the predicted reaction location. Spin-trap

Page 5: Photochemical Crosslinking Reactions in Polymers

experiments demonstrate that radicals primarily form on the expected tertiary carbon,

confirming the hydrogen abstraction mechanism employed in Chapter 4. Unexpected peaks in

the EPR spectra point towards a potentially new reaction between benzophenone and the spin-

trap. The EPR experiments are also used to verify that no other radical reactions are occurring

in the systems of Chapter 4.

The experimental space is expanded in Chapter 6 to other polymers and other

crosslinking chemistries. Full characterization of all the potential reactions was impossible, but

many polymer-additive combinations are shown to react in the predicted ways. Certain

crosslinker chemistries and functionalities combine to allow the study of macromolecule

combination without scission and these chemistries are suggested for further use in the

experimental study of the early stages of crosslinking and gelation.

This thesis finds that at low molar ratios, the additive functionality does not matter;

functionality is only important at high molar ratios of additive to polymer due to high radical

density and a transition to pseudo-first order kinetics. The expectations in Chapter 3 are shown

to be accurate and the choice of additive chemistry and polymer allows the preferential

selection of desired reactions. Hydrogen abstraction can be forced to occur either from the

pendant group or the chain backbone, leading to systems in which chain scission is not possible

and combination can be exclusively selected. Non-radical based crosslinking chemistries can

also be used to produce crosslinks without risk of chain scission. These findings have a wide

applicability in many fields and implications for the improved design of radical-based

crosslinking systems.

Page 6: Photochemical Crosslinking Reactions in Polymers

i

Table of Contents

1. CHAPTER 1: MOTIVATION AND INTRODUCTION .......................................................................................... 1

1.1. MOTIVATION ........................................................................................................................................... 1

1.2. INTRODUCTION TO PHOTOCHEMISTRY ............................................................................................................ 3

1.2.1. Radicals ........................................................................................................................................... 3

1.2.2. Photochemical Generation of Radicals (80, 81) ................................................................................. 4

1.3. POLYMERS AND THEIR PROPERTIES .............................................................................................................. 10

1.3.1. Individual Chain Properties ............................................................................................................ 10

1.3.2. Ensemble Properties ...................................................................................................................... 14

1.4. REFERENCES .......................................................................................................................................... 17

2. CHAPTER 2: EXPERIMENTAL TECHNIQUES.................................................................................................. 20

2.1. ELECTRON PARAMAGNETIC RESONANCE ....................................................................................................... 20

2.2. HIGH PRESSURE LIQUID CHROMATOGRAPHY .................................................................................................. 23

2.2.1. Size Exclusion Chromatography(2).................................................................................................. 25

2.2.2. HPLC Detectors .............................................................................................................................. 30

Refractive Index Detectors .......................................................................................................................................31

Ultraviolet Absorption Detectors ..............................................................................................................................33

Static Light Scattering Detectors ...............................................................................................................................34

2.2.3. Calibration of the HPLC Instrument ................................................................................................ 37

Light Scattering Detector Calibration ........................................................................................................................38

SEC Column Calibration ............................................................................................................................................38

Refractive Index Detector Calibration .......................................................................................................................41

Ultraviolet Light Detector Calibration .......................................................................................................................42

2.3. PREPARATION AND CHARACTERIZATION TECHNIQUES ....................................................................................... 42

2.3.1. Spin Coating .................................................................................................................................. 42

Page 7: Photochemical Crosslinking Reactions in Polymers

ii

2.3.2. Ellipsometry ................................................................................................................................... 43

2.3.2.1. Characterization of the Thin Films ........................................................................................................44

2.4. REFERENCES .......................................................................................................................................... 46

3. CHAPTER 3: RATIONAL DESIGN AND PRELIMINARY INVESTIGATIONS ........................................................ 47

3.1. INTRODUCTION ...................................................................................................................................... 47

3.2. POLYMER SELECTION ............................................................................................................................... 48

3.2.1. Polystyrene (PS) ............................................................................................................................. 48

3.2.2. Poly(�-methyl styrene) (PamS) ....................................................................................................... 51

3.2.3. Polyisobutylene (PIB) ..................................................................................................................... 52

3.2.4. Polymethylmethacrylate (PMMA) .................................................................................................. 54

3.2.5. Poly(normal-butylacrylate) (PnBA) ................................................................................................. 55

3.2.6. 1,4-polybutadiene (PBD) ................................................................................................................ 57

3.2.7. Conclusion ..................................................................................................................................... 59

3.3. CROSSLINKER SELECTION .......................................................................................................................... 59

3.3.1. Benzophenone (BP) ........................................................................................................................ 60

3.3.2. Xanthone (XAN) ............................................................................................................................. 62

3.3.3. Phthalimide (PTH) .......................................................................................................................... 63

3.3.4. Phenylazide (FEN) .......................................................................................................................... 64

3.3.5. Conclusion ..................................................................................................................................... 65

3.4. DIFFERENTIAL REFRACTIVE INDEX ................................................................................................................ 66

3.4.1. Experimental ................................................................................................................................. 67

3.4.2. Results ........................................................................................................................................... 68

3.5. MARK-HOUWINK-SAKURADA CONSTANTS .................................................................................................... 70

3.6. SOLUBILITY AND INTERACTION PARAMETER. .................................................................................................. 71

3.7. REFERENCES .......................................................................................................................................... 78

Page 8: Photochemical Crosslinking Reactions in Polymers

iii

4. CHAPTER 4: PHOTOCHEMICALLY INDUCED RADICAL REACTIONS IN POLYSTYRENE AND POLY(NORMAL-

BUTYLACRYLATE) ................................................................................................................................................. 80

4.1. INTRODUCTION ...................................................................................................................................... 80

4.2. MODELING OF SCISSION AND COMBINATION REACTIONS ................................................................................... 83

4.2.1. Modeling random macroradical �-scission ..................................................................................... 83

4.2.2. Modeling random macroradical recombination .............................................................................. 85

4.2.3. Linking the model to HPLC-SEC data ............................................................................................... 86

4.3. MATERIALS ........................................................................................................................................... 88

4.3.1. Substrate Preparation .................................................................................................................... 88

4.3.2. Solution Preparation ...................................................................................................................... 89

4.3.3. Sample Preparation ....................................................................................................................... 89

4.3.4. GPC analysis .................................................................................................................................. 89

4.4. RESULTS ............................................................................................................................................... 90

4.5. DISCUSSION........................................................................................................................................... 93

4.5.1. Higher functional species generate more radicals ........................................................................... 94

Simple lattice model of the thin film .........................................................................................................................98

Effects of diffusion distance on radical formation.................................................................................................... 101

Effect of chain mobility on radical formation .......................................................................................................... 104

4.5.2. Insights from macroradical scission and recombination modeling ................................................. 105

The role of functionality in glassy binary systems .................................................................................................... 105

Best fits and the probabilistic transition .................................................................................................................. 113

4.5.3. The lattice model and the probabilistic transition ......................................................................... 119

4.6. CONCLUSIONS ...................................................................................................................................... 120

4.7. FURTHER DIRECTIONS ............................................................................................................................ 121

4.8. REFERENCES ........................................................................................................................................ 122

5. CHAPTER 5: ELECTRON PARAMAGNETIC RESONANCE OF BENZOPHENONE AND POLYSTYRENE IN TOLUENE

124

Page 9: Photochemical Crosslinking Reactions in Polymers

iv

5.1. INTRODUCTION .................................................................................................................................... 124

5.2. MATERIALS ......................................................................................................................................... 126

5.2.1. Sample Preparation ..................................................................................................................... 126

5.2.2. EPR Measurements ...................................................................................................................... 127

5.3. RESULTS AND DISCUSSION....................................................................................................................... 128

5.3.1. Controls and the Intrinsic PBN Peak.............................................................................................. 128

5.3.2. Benzophenone and PBN: A Potential New Reaction ...................................................................... 132

5.3.3. Polystyrene, Benzophenone and PBN: Preferential Hydrogen Abstraction ..................................... 139

5.4. CONCLUSION ....................................................................................................................................... 142

5.5. FURTHER DIRECTIONS ............................................................................................................................ 143

5.6. REFERENCES ........................................................................................................................................ 144

6. CHAPTER 6: PRELIMINARY INVESTIGATIONS OF OTHER POLYMER AND CROSSLINKER SYSTEMS ............. 145

6.1. MATERIALS ......................................................................................................................................... 146

6.1.1. Substrate Preparation .................................................................................................................. 146

6.1.2. Solution Preparation .................................................................................................................... 147

6.1.3. Sample Preparation ..................................................................................................................... 147

6.1.4. GPC analysis ................................................................................................................................ 148

6.2. POLYSTYRENE AND OTHER CROSSLINKER CHEMISTRIES ................................................................................... 149

6.2.1. Polystyrene and Phenylazide ........................................................................................................ 149

6.2.2. Polystyrene and Phthalimide ........................................................................................................ 150

6.3. POLY(NORMAL-BUTYLACRYLATE) AND OTHER CROSSLINKER CHEMISTRIES ........................................................... 151

6.4. POLY(1,4-BUTADIENE) AND CROSSLINKERS ................................................................................................. 154

6.5. POLYMETHYLMETHACRYLATE AND CROSSLINKERS .......................................................................................... 155

6.5.1. Polymethylmethacrylate and Benzophenone ................................................................................ 156

6.5.2. Polymethylmethacrylate and Phthalimide .................................................................................... 157

6.6. CONCLUSIONS ...................................................................................................................................... 158

Page 10: Photochemical Crosslinking Reactions in Polymers

v

6.7. FURTHER DIRECTIONS ............................................................................................................................ 161

6.8. REFERENCES ........................................................................................................................................ 162

7. CHAPTER 7: CONCLUSIONS ...................................................................................................................... 163

8. CHAPTER 8: IMPLICATIONS FOR FUTURE WORK ....................................................................................... 165

9. APPENDIX ................................................................................................................................................. 167

Page 11: Photochemical Crosslinking Reactions in Polymers

vi

List of Figures

FIGURE 1.1.1: PHOTOCHEMICAL FORMATION OF A FREE RADICAL THROUGH CARBONYL EXCITATION AND HYDROGEN ABSTRACTION FROM

THE BACKBONE OF AN ALKYL POLYMER. THIS SCHEMA IS REPRESENTATIVE OF THE MAJORITY OF SYSTEMS IN THIS STUDY. .............. 9

FIGURE 1.1.2: THE STABILITY OF RADICAL CENTERS AND THEIR RELATION TO HYDROGEN ABSTRACTION............................................. 9

FIGURE 2.1: ENERGY DIAGRAM OF UNPAIRED ELECTRONS IN AN INDUCED MAGNETIC FIELD. ........................................................ 21

FIGURE 2.2: DIAGRAM OF A BASIC EPR INSTRUMENT. A MICROWAVE SOURCE SENDS MICROWAVES THROUGH AN ATTENUATOR INTO A

CIRCULATOR THAT GUIDES THEM INTO THE SAMPLE CHAMBER AND THEN BACK OUT TO THE DETECTOR. THE SAMPLE CHAMBER SITS

BETWEEN ELECTROMAGNETS. .............................................................................................................................. 22

FIGURE 2.3: DIAGRAM OF AN HPLC SYSTEM. THE PUMP TAKES THE ELUENT FROM THE RESERVOIR AND PUMPS IT AT HIGH PRESSURE

THROUGH THE INJECTOR, COLUMN(S), AND DETECTOR(S). SAMPLES ARE INTRODUCED INTO THE INJECTOR AND ARE SEPARATED INTO

THEIR ANALYTES IN THE COLUMN(S). THE DETECTOR(S) MEASURE THE PROPERTIES OF THE ANALYTES. THE COLUMN OVEN

MAINTAINS AN ISOTHERMAL SYSTEM. .................................................................................................................... 24

FIGURE 2.4: HPLC MOBILE PHASE. THE FLOW IS LEFT TO RIGHT. CUTAWAY SHOWS NOT-TO-SCALE REPRESENTATION OF THE PACKED

STATIONARY PHASE. FIGURE ADAPTED FROM (2). ..................................................................................................... 26

FIGURE 2.5: HPLC STATIONARY PHASE. SOLID PARTICLE (BLACK LINE) IS SURROUNDED BY A STAGNANT BOUNDARY (DOTTED LINE)

THROUGH WHICH ONLY MASS TRANSFER OCCURS. PORE IS NOT TO SCALE AND REPRESENTS THE PORE VOLUME OF THE PARTICLE

(WHITE) AND THE PORE VOLUME ACCESSIBLE TO ATTACHMENT THROUGH AFFINITY OR OTHER MEANS (GREY). FIGURE ADAPTED

FROM (2) ....................................................................................................................................................... 26

FIGURE 2.6: BASIC OPERATION OF A REFRACTIVE INDEX DETECTOR. A BEAM OF LIGHT PASSES THROUGH A CELL CONTAINING BOTH THE

REFERENCE AND THE SAMPLE. THE DIFFERENCE IN REFRACTIVE INDEX BETWEEN THE TWO LEADS TO A CHANGE IN THE ANGLE OF THE

REFRACTED LIGHT, FROM WHICH THE INSTRUMENT CALCULATES THE DIFFERENCE IN REFRACTIVE INDEX. ................................. 31

FIGURE 2.7: BASIC HARDWARE OF AN ULTRAVIOLET ABSORPTION DETECTOR. A BEAM OF LIGHT PASSES THROUGH A MONOCHROMATOR

TO SELECT THE DESIRED UV WAVELENGTH, AND THEN INTO A BEAM SPLITTER. ONE OF THE RESULTING BEAMS PASSES THROUGH THE

SAMPLE FLOW CELL AND TO A PHOTODIODE AND THE OTHER BEAM PASSES THROUGH A REFERENCE (USUALLY AIR) TO ANOTHER

PHOTODIODE. ................................................................................................................................................. 33

Page 12: Photochemical Crosslinking Reactions in Polymers

vii

FIGURE 2.8: BASIC HARDWARE OF THE MULTIANGLE LASER LIGHT SCATTERING DETECTOR. A LASER PASSES THROUGH A POLARIZER TO

MINIMIZE OUT-OF-PLANE SCATTERING, AND THEN INTO THE SCATTERING VOLUME OF THE SAMPLE CELL. RADIALLY ARRANGED

PHOTODIODES DETECT THE SCATTERED LIGHT AS A FUNCTION OF ANGLE, WHILE A PHOTODIODE AT ZERO DEGREES MEASURES THE

LIGHT INTENSITY. ............................................................................................................................................. 35

FIGURE 2.9: SEC COLUMN CALIBRATION CURVES FOR BOTH THE UV (LEFT) AND RI (RIGHT) DETECTORS. EACH PLOT INCLUDES THE

CALIBRATION CURVES FROM FOUR DIFFERENT TIMES OVER THE COURSE OF THE EXPERIMENTS TO SHOW THE CONTINUITY OF THE

COLUMN CALIBRATION....................................................................................................................................... 40

FIGURE 2.10: BASIC SETUP OF A SPECTROSCOPIC ELLIPSOMETER. A BROADBAND LIGHT SOURCE IS ELLIPTICALLY POLARIZED AND REFLECTS

OFF THE SAMPLE ON A SUBSTRATE IN THE PLANE OF INCIDENCE. THE REFLECTED LIGHT, WHICH HAS BEEN MODIFIED BOTH PARALLEL

(�-POLARIZED) AND PERPENDICULAR (�-POLARIZED) TO THE PLANE OF INCIDENCE, PASSES THROUGH ANOTHER POLARIZER

(REFERRED TO AS THE ANALYZER) AND INTO THE DETECTOR. ........................................................................................ 44

FIGURE 2.11: DETERMINATION OF THE RELATIONSHIP BETWEEN SPIN SPEED AND FILM THICKNESS FOR POLYSTYRENE AND POLY(N-

BUTYLACRYLATE) THIN FILMS SPUN COATED FROM 20MG/ML SOLUTIONS OF TOLUENE. EACH DATA POINT IS AN AVERAGE OF THREE

SAMPLES WITH ERROR BARS THE SUM OF THE ELLIPSOMETER-REPORTED ERRORS IN THE THREE MEASUREMENTS. DOTTED LINE IS THE

� ∝ �− �� RELATIONSHIP FROM EQUATION (2.17) SCALED TO THE PS DATA.............................................................. 45

FIGURE 3.1: LAP SYNTHESIS OF PS................................................................................................................................ 48

FIGURE 3.2: PS RADICAL GENERATION THROUGH HYDROGEN ABSTRACTION BY BP. ................................................................... 49

FIGURE 3.3: LAP SYNTHESIS OF PAMS. .......................................................................................................................... 51

FIGURE 3.4: LCP OF POLYISOBUTYLENE. ......................................................................................................................... 53

FIGURE 3.5: THE ATRP SYNTHESIS OF POLYMETHYLMETHACRYLATE. ..................................................................................... 54

FIGURE 3.6: LAP OF POLY(NORMAL-BUTYLACRYLATE). ...................................................................................................... 56

FIGURE 3.7: LAP OF 1,4 POLYBUTADIENE. ..................................................................................................................... 58

FIGURE 3.8: BENZOPHENONE (LEFT) AND THE BI-FUNCTIONAL BENZOPHENONE (RIGHT) USED IN THESE STUDIES. .............................. 61

FIGURE 3.9: THE MONO-FUNCTIONAL XANTHONE (LEFT) AND THE BI-FUNCTIONAL XANTHONE (RIGHT) USED IN THESE STUDIES. ........... 62

FIGURE 3.10: PHTHALIMIDE (LEFT) AND THE BI-FUNCTIONAL PHTHALIMIDE (RIGHT) USED IN THESE STUDIES. ................................... 63

FIGURE 3.11: RADICAL GENERATION THROUGH HYDROGEN ABSTRACTION AFTER A SET IN PHTHALIMIDE(26) .................................. 64

FIGURE 3.12: PHENYLAZIDE (LEFT) AND THE BI-FUNCTIONAL FORM (RIGHT) USED IN THESE STUDIES............................................... 65

Page 13: Photochemical Crosslinking Reactions in Polymers

viii

FIGURE 3.13: NITRENE INSERTION REACTION BETWEEN A PHENYLAZIDE AND A C-H BOND. .......................................................... 65

FIGURE 3.14: REPRESENTATIVE DC/DC PLOTS FROM THE RID. A) PAMS IN THF AT 40OC B) PMMA IN THF AT 40

OC. .................... 68

FIGURE 3.15: STRUCTURES OF STYRENE AND POLYSTYRENE ................................................................................................. 74

FIGURE 4.1: SCHEMATIC DESCRIPTIONS OF THE PHOTOCHEMICALLY INDUCED RADICAL REACTIONS OF BP AND BP-BP LEADING TO THE

PRODUCTION OF MACRORADICALS. ....................................................................................................................... 82

FIGURE 4.2: SCHEMATIC DESCRIPTION OF MACRORADICAL CROSSLINKING REACTIONS, INCLUDING THE RADICAL RECOMBINATION-BRIDGING

REACTION UNIQUE TO BI-FUNCTIONAL BP-BP. ........................................................................................................ 82

FIGURE 4.3: �-SCISSION OF ALLYLIC POLYMER BACKBONE. .................................................................................................. 83

FIGURE 4.4: PHOTOACTIVE CROSSLINKER Α-META-,Ω-PARA-BIS-BENZOPHENONE (BP-BP) .......................................................... 88

FIGURE 4.5: RESULTS OF EXPERIMENTS WITH POLYSTYRENE AND ADDITIVES, BOTH BP AND BP-BP. A) NUMBER FRACTION OF CHAINS IN

32:1 BP:PS FILMS AFTER 0, 3, 6, AND 9 HOURS OF IRRADIATION. INSET GRAPH SHOWS THE CONTROL AFTER 0 AND 9 HOURS OF

IRRADIATION. B) NUMBER FRACTION OF CHAINS IN 16:1 BP-BP:PS FILMS AFTER 0, 3, 6, AND 9 HOURS OF IRRADIATION. C)

NUMBER FRACTION OF CHAINS IN 0:1, 1:1, 2.4:1, 8:1 AND 32:1 BP:PS FILMS AFTER 9 HOURS OF IRRADIATION. D) NUMBER

FRACTION OF CHAINS IN 0:1, 0.5:1, 1.2:1, 4:1 AND 16:1 BP-BP:PS FILMS AFTER 9 HOURS OF IRRADIATION. ALL AXES ARE

EQUAL. .......................................................................................................................................................... 92

FIGURE 4.6: RESULTS OF EXPERIMENTS WITH POLY(N-BUTYL-ACRYLATE) AND ADDITIVES, BOTH BP AND BP-BP. A) NUMBER FRACTION OF

CHAINS IN 2.4:1 BP:PNBA FILMS AFTER 0, 3, 6, AND 9 HOURS OF IRRADIATION. INSET GRAPH SHOWS THE CONTROL AFTER 0 AND

9 HOURS OF IRRADIATION. B) NUMBER FRACTION OF CHAINS IN 1.2:1 BP-BP:PNBA FILMS AFTER 0, 3, 6, AND 9 HOURS OF

IRRADIATION. C) NUMBER FRACTION OF CHAINS IN 0:1, 1:1, 2.4:1, 8:1 AND 32:1 BP:PNBA FILMS AFTER 3 HOURS OF

IRRADIATION. D) NUMBER FRACTION OF CHAINS IN 0:1, 0.5:1, 1.2:1, AND 16:1 BP-BP:PNBA FILMS AFTER 3 HOURS OF

IRRADIATION. ALL AXES ARE EQUAL. ..................................................................................................................... 93

FIGURE 4.7: A) NUMBER FRACTION OF CHAINS IN POLYSTYRENE FILMS AFTER 9 HOURS OF IRRADIATION WITH PURE PS (LINE),

BENZOPHENONE AND PS 32:1 MOLAR RATIO (DASHED), AND BI-FUNCTIONAL BENZOPHENONE AND PS 16:1 MOLAR RATIO

(DOTTED). B) NUMBER FRACTION OF CHAINS IN POLY(N-BUTYL-ACRYLATE) FILMS AFTER 3 HOURS OF IRRADIATION WITH PURE

PNBA (LINE), BENZOPHENONE AND PNBA 2.4:1 MOLAR RATIO (DASHED), AND BI-FUNCTIONAL BENZOPHENONE AND PNBA 1.2:1

MOLAR RATIO (DOTTED). ALL DISTRIBUTIONS HAVE AN AREA OF 1. ............................................................................... 96

FIGURE 4.8: REACTIONS THAT COMPETE WITH MACRORADICAL SCISSION AND COMBINATION. ...................................................... 98

Page 14: Photochemical Crosslinking Reactions in Polymers

ix

FIGURE 4.9: COMPARISON OF LATTICE CELL SIZES FOR BP (TOP) AND BP-BP (BOTTOM). LATTICE VOLUMES ARE SCALED SUCH THAT THE

MOLAR RATIO OF BP IS TWICE THAT OF BP-BP TO MIRROR EXPERIMENTS. REPRESENTATIVE INTER-BENZOPHENONE MOIETY

DISTANCES ARE MARKED. ................................................................................................................................. 100

FIGURE 4.10 NEAREST NEIGHBOR DISTANCE VS MOLAR RATIO. THIS PLOT COMPARES THE NEAREST NEIGHBOR DISTANCE OF BP (GREY) TO

BP-BP (BLACK) AT 1:1, 2.4:1, 8:1, AND 32:1 MOLAR RATIOS OF BENZOPHENONE MOIETIES TO PS MOLECULES OF LENGTH MN.

THERE ARE TWO BENZOPHENONE MOIETIES PER BP-BP. DISTANCE IS THE LATTICE DISTANCE, WITH THE MOLAR VOLUME OF BP THE

LATTICE CELL VOLUME. ALKYL LINKER IN BP-BP IS IGNORED. DOTTED LINE IS DIFFUSION DISTANCE FOR ILLUSTRATIVE PURPOSES

ONLY........................................................................................................................................................... 103

FIGURE 4.11: ARBITRARILY SCALED SUBTRACTION (LHS EQUATION (1.2)) (LINE), MODEL OF SCISSION (RHS EQUATION (1.2)) (DASHED)

AND CONTROL DATA (DOTTED) FROM WHICH THE SCISSION WAS CALCULATED. DATA IS FROM BP-BP:PS 16:1 AFTER 9 HOURS

IRRADIATION. PLOT IS FOR COMPARISON ONLY....................................................................................................... 106

FIGURE 4.12: FITTING OF THE MODEL DISTRIBUTIONS TO DATA. LEFT COLUMN BP:PS 32:1 AFTER 9 HOURS. RIGHT COLUMN BP-BP:PS

16:1 AFTER 9 HOURS. TOP PLOTS ARE FITS WITH THE MODEL SCISSION USED IN THE COMBINATION MODELING. BOTTOM PLOTS ARE

FITS WITH THE DATA SCISSION USED IN THE COMBINATION MODELING. DATA (THICK LINE), OVERALL FIT (CIRCLES), AND ARBITRARILY

SCALED CONTROL (DOTTED), SCISSION (SQUARES), END-LINKED (DASH-DOTTED), 3-ARM STAR (DASHED) AND 4-ARM STAR (LINE)

POLYMERS ARE INCLUDED. INSET GRAPHS SHOW THE SAME PLOTS FULL SCALE. .............................................................. 107

FIGURE 4.13: BEST FITS (THICK LINE) TO THE LHS OF EQUATION (4.9) USING A PEAK FITTING ALGORITHM THAT DOES NOT INCORPORATE

END-LINKING (LEFT) AND ONE THAT DOES (RIGHT). DATA (POINTS) IS FROM BP:PS 32:1 AFTER 3 (TOP), 6 (MIDDLE) AND 9

(BOTTOM) HOURS OF IRRADIATION. CURVES INCLUDE THE END-LINKING COMBINATION (DASH-DOTTED), THE 3-ARM STAR (DASHED)

AND THE 4-ARM STAR (LINE). ALL AXES ARE EQUAL. ............................................................................................... 109

FIGURE 4.14: BEST FITS (THICK LINE) TO THE LHS OF EQUATION (1.1) USING A PEAK FITTING ALGORITHM THAT DOES NOT INCORPORATE

END-LINKING (LEFT) AND ONE THAT DOES (RIGHT). DATA (POINTS) IS FROM BP-BP:PS 16:1 AFTER 3 (TOP), 6 (MIDDLE) AND 9

(BOTTOM) HOURS OF IRRADIATION. CURVES INCLUDE THE END-LINKING COMBINATION (DASH-DOTTED), THE 3-ARM STAR (DASHED)

AND THE 4-ARM STAR (LINE). ALL AXES ARE EQUAL. ............................................................................................... 110

FIGURE 4.15: BEST FITS (THICK LINE) TO THE LHS OF EQUATION (4.9) USING A PEAK FITTING ALGORITHM THAT DOES NOT INCORPORATE

END-LINKING (LEFT) AND ONE THAT DOES (RIGHT). DATA (POINTS) IS FROM BP:PS 8:1 AFTER 3 (TOP), 6 (MIDDLE) AND 9

Page 15: Photochemical Crosslinking Reactions in Polymers

x

(BOTTOM) HOURS OF IRRADIATION. CURVES INCLUDE THE END-LINKING COMBINATION (DASH-DOTTED), THE 3-ARM STAR (DASHED)

AND THE 4-ARM STAR (LINE). ALL AXES ARE EQUAL. ............................................................................................... 111

FIGURE 4.16: BEST FITS (THICK LINE) TO THE LHS OF EQUATION (4.9) USING A PEAK FITTING ALGORITHM THAT DOES NOT INCORPORATE

END-LINKING (LEFT) AND ONE THAT DOES (RIGHT). DATA (POINTS) IS FROM BP-BP:PS 4:1 AFTER 3 (TOP), 6 (MIDDLE) AND 9

(BOTTOM) HOURS OF IRRADIATION. CURVES INCLUDE THE END-LINKING COMBINATION (DASH-DOTTED), THE 3-ARM STAR (DASHED)

AND THE 4-ARM STAR (LINE). ALL AXES ARE EQUAL. ............................................................................................... 112

FIGURE 4.17: COMPARISON OF LINEAR (DOTTED) AND SQUARED (LINE) FITS FOR BEST FIT PEAKS IN BP-BP:PS 4:1 (LEFT) AND BP-BP:PS

16:1 (RIGHT). FITS ARE END-LINKING (TRIANGLES), 3-ARM STARS (CIRCLES) AND 4-ARM STARS (SQUARES). THE CORRELATION

COEFFICIENTS FOR THE FITS ARE INCLUDED. Y-AXES ARE NOT THE SAME. ...................................................................... 114

FIGURE 4.18: COMPARISON OF BEST FIT LINES FOR BI-FUNCTIONAL ADDITIVE (DASHED) AND MONO-FUNCTIONAL ADDITIVE (LINE) AT FOUR

RATIOS OF BENZOPHENONE MOIETIES TO PS: 1:1 (UPPER LEFT), 2.4:1 (UPPER RIGHT), 8:1 (LOWER LEFT) AND 32:1 (LOWER

RIGHT). Y-AXES ARE NOT EQUAL. BOTTOM TWO FITS IN EACH PLOT ARE 4-ARM STAR FIT; TOP TWO ARE 3-ARM STAR (1:1, 2.4:1,

8:1) OR END-LINKING (32:1). ........................................................................................................................... 115

FIGURE 4.19: MAP OF THE RADICAL REACTIONS WITH MONO-FUNCTIONAL ADDITIVE AND THE SPECIES PROBABILITIES. ..................... 117

FIGURE 4.20: MAP OF THE RADICAL REACTIONS WITH BI-FUNCTIONAL ADDITIVE AND THE SPECIES PROBABILITIES. .......................... 118

FIGURE 5.1: REACTION OF THE SPIN TRAP N-TERT-BUTYL-∝-PHENYL-NITRONE (PBN) WITH A RADICAL. A TRANSIENT RADICAL BECOMES

ESSENTIALLY PERMANENT WHEN REACTED WITH A SPIN TRAP. .................................................................................... 125

FIGURE 5.2: MOLECULAR STRUCTURE OF THE CHEMICALS USED IN THESE INVESTIGATIONS. A) TOLUENE B) POLYSTYRENE C)

BENZOPHENONE D) N-TERT-BUTYL-∝-PHENYL-NITRONE. ........................................................................................ 126

FIGURE 5.3: EPR SPECTRA OF A SAMPLE CONSISTING SOLELY OF PS IN TOLUENE BEFORE, DURING, AND AFTER CONCURRENT UV

IRRADIATION. THERE ARE NO DETECTABLE RADICALS. .............................................................................................. 129

FIGURE 5.4: PBN BACKGROUND SIGNAL BEFORE, DURING, AND AFTER IRRADIATION. WHEN ANY TWO OF THE THREE SIGNALS ARE

SUBTRACTED, THE BOTTOM PLOT IS PRODUCED. ..................................................................................................... 130

FIGURE 5.5: EPR OF PURE PBN IN TOLUENE (DOTTED) AND A SIMULATION OF THE EXPECTED PEAKS FORMED BY A SINGLE SPIN 1 ATOM

(SOLID). THE HYPERFINE INTERACTION OF THE PEAKS IS NOTED. THE SUBTRACTION BELOW SHOWS NOTHING BUT NOISE, THOUGH

THERE ARE HINTS OF ANOTHER SET OF PEAKS DISCUSSED LATER. ................................................................................. 131

Page 16: Photochemical Crosslinking Reactions in Polymers

xi

FIGURE 5.6: COMPARISON OF THE EPR SPECTRA OF PBN IN TOLUENE AND PS AND PBN IN TOLUENE SAMPLES DURING IRRADIATION. THE

PLOT OF THE SUBTRACTION SHOWS NO REACTION HAS TAKEN PLACE. ........................................................................... 132

FIGURE 5.7: COMPARISON OF BP AND PBN IN TOLUENE BEFORE (SOLID) AND AFTER (DARK SOLID) THIRTY MINUTES OF IRRADIATION

WITHIN THE EPR. TOP PLOT IS BP AND PBN BEFORE IRRADIATION............................................................................. 134

FIGURE 5.8: EPR OF BP AND PBN IN TOLUENE (DOTTED) AFTER 30 MINUTES OF CONCURRENT IRRADIATION AND A SIMULATION OF THE

INTERACTION BETWEEN THE RADICAL AND A SINGLE SPIN 1 ATOM AND A SINGLE SPIN ½ ATOM (SOLID). HYPERFINE INTERACTIONS

ARE LABELED. ................................................................................................................................................ 135

FIGURE 5.9: THE TWO POSSIBLE STRUCTURES OF PBN AFTER TRAPPING THE RADICAL THAT FORMS THE TRIPLET OF DOUBLETS. THE PHENYL

RADICAL (LEFT) COMES FROM TOLUENE WHILE THE DIPHENYLMETHYL RADICAL (RIGHT) COMES FROM BP. ............................ 136

FIGURE 5.10: EPR OF BP AND PBN IN TOLUENE (DOTTED) AFTER 30 MINUTES OF CONCURRENT IRRADIATION AND A SIMULATION OF THE

INTERACTION BETWEEN THE RADICAL AND A SINGLE SPIN 1 (SOLID). HYPERFINE INTERACTIONS ARE LABELED. ........................ 137

FIGURE 5.11: PLAUSIBLE MECHANISM FOR THE CREATION OF A CARBONYL ON THE ALPHA CARBON OF THE NITROXIDE. .................... 138

FIGURE 5.12: EPR OF BP AND PBN IN TOLUENE (DOTTED) AFTER 30 MINUTES OF CONCURRENT IRRADIATION AND THE SUM OF THE

PREVIOUSLY DISCUSSED SIMULATIONS (SOLID). ...................................................................................................... 138

FIGURE 5.13: EPR OF PS, BP, AND PBN IN TOLUENE (DOTTED) AND THE SIMULATION FITS. A) AND B) ARE THE CONSTITUTIVE

SIMULATIONS FROM THE BP AND PBN IN TOLUENE SYSTEM. THE WEIGHTED SUM OF THE FITS IN A) AND B) IS FITTED TO THE NEW

SYSTEM IN C). SUBPLOTS SHOW THE RESIDUAL DATA AFTER SUBTRACTION. .................................................................. 140

FIGURE 5.14: DIRECT OVERLAY OF THE RAW DATA WITH (THIN LINE) AND WITHOUT (THICK LINE) PS IN SOLUTION WITH THE BP AND PBN

IN TOLUENE AFTER THIRTY MINUTES OF CONCURRENT IRRADIATION. ............................................................................ 141

FIGURE 6.1: PS AND FEN (LEFT) AND PS AND FEN-FEN (RIGHT) AFTER 0 (DOTTED), 3 (DASHED), AND 6 (SOLID) HOURS OF IRRADIATION.

INSET GRAPHS ARE THE FULL PEAK OF THE RESPECTIVE PLOT. ..................................................................................... 149

FIGURE 6.2: PS AND PTH AND PS AND PTH-PTH AFTER 0 (DOTTED), 3 (DASHED) AND 6 (SOLID) HOURS OF IRRADIATION. .............. 150

FIGURE 6.3: PNBA AFTER 0 (DOTTED), 3 (DASHED) AND 6 (SOLID) HOURS OF IRRADIATION ...................................................... 152

FIGURE 6.4: PNBA CONTROL (DOTTED), PNBA AND FEN AFTER 6 HOURS (DASHED) AND PNBA CONTROL AFTER 6 HOURS (SOLID). ... 153

FIGURE 6.5: PNBA CONTROL (DOTTED), PNBA AND PTH (DASHED), AND PNBA AND PTH-PTH (SOLID) AFTER 6 HOURS IRRADIATION.

................................................................................................................................................................. 153

FIGURE 6.6: TRANSESTERIFICATION. ............................................................................................................................ 154

Page 17: Photochemical Crosslinking Reactions in Polymers

xii

FIGURE 6.7: PBD CONTROL (DOTTED) AND AFTER 6 HOURS (SOLID), PBD AND BP (DASHED) AND PBD AND BP-BP (DASHED) AFTER 6

HOURS. ....................................................................................................................................................... 155

FIGURE 6.8: PMMA CONTROL BEFORE (DOTTED) AND AFTER (SOLID) 6 HOURS OF IRRADIATION................................................. 156

FIGURE 6.9: PMMA AND BP-BP (DASHED), PMMA AND BP (BOLD LINE), AND PMMA (LINE) AFTER 6 HOURS IRRADIATION. DOTTED

LINE IS CONTROL. ........................................................................................................................................... 157

FIGURE 6.10: PMMA CONTROL (DOTTED) AND CONTROL, PMMA AND PTH (BOLD LINE), AND PMMA AND PTH-PTH (DASHED) AFTER

6 HOURS IRRADIATION. .................................................................................................................................... 158

Page 18: Photochemical Crosslinking Reactions in Polymers

xiii

List of Tables

TABLE 1.1.1: ORBITAL TRANSITIONS AND OPTICAL ABSORPTION(80)A .................................................................................... 4

TABLE 1.1.2: SELECTED CHROMOPHORES (80)A ................................................................................................................ 6

TABLE 2.1: MOST ABUNDANT PARAMAGNETIC ISOTOPES OF C,H,O AND N AND THEIR SPINS. ...................................................... 22

TABLE 2.2: NUMBER OF EPR PEAKS GENERATED BY THE COMBINATION OF SPIN AND NUMBER OF ATOMS. ...................................... 22

TABLE 3.1: STRUCTURE AND HYDROGEN BOND DISSOCIATION ENERGIES OF POLYSTYRENE. ........................................................... 50

TABLE 3.2: STRUCTURE AND HYDROGEN BOND DISSOCIATION ENERGIES OF POLY(∝-METHYLSTYRENE). .......................................... 52

TABLE 3.3: STRUCTURE AND HYDROGEN BOND DISSOCIATION ENERGIES OF POLYISOBUTYLENE. ..................................................... 53

TABLE 3.4: STRUCTURE AND HYDROGEN BOND DISSOCIATION ENERGIES OF POLYMETHYLMETHACRYLATE. ....................................... 55

TABLE 3.5: STRUCTURE AND HYDROGEN BOND DISSOCIATION ENERGIES OF POLY(NORMAL-BUTYLACRYLATE). .................................. 57

TABLE 3.6: STRUCTURE AND HYDROGEN BOND DISSOCIATION ENERGIES OF POLY(1,4 BUTADIENE). ............................................... 59

TABLE 3.7: BOND DISSOCIATION ENERGIES OF BENZOPHENONE FOR EQUATION (1.10). ............................................................. 61

TABLE 3.8: CONTROL AND POLYMER SOLUTION CONCENTRATIONS USED IN DIFFERENTIAL REFRACTIVE INDEX EXPERIMENTS. ................ 67

TABLE 3.9: DIFFERENTIAL REFRACTIVE INDEX INCREMENT OF POLYMERS USED IN THIS STUDY ....................................................... 69

TABLE 3.10: MARK-HOUWINK-SAKURADA CONSTANTS OF THE POLYMERS IN THF. .................................................................. 71

TABLE 3.11: HILDEBRAND SOLUBILITY PARAMETERS OF THE SOLVENTS BY GROUP CONTRIBUTION ............................................... 75

TABLE 3.12: HILDEBRAND SOLUBILITY PARAMETERS OF THE POLYMERS AND THEIR MONOMERS BY GROUP CONTRIBUTION ................ 75

TABLE 3.13: HILDEBRAND SOLUBILITY PARAMETERS OF THE ADDITIVES BY GROUP CONTRIBUTION ............................................... 76

TABLE 4.1: PERCENT OF ORIGINAL DISTRIBUTION THAT HAS REACTED IN FIGURE 4.7 ................................................................ 97

TABLE 4.2: RADIUS OF GYRATION AND OVALITY OF ADDITIVE MOLECULES1 ............................................................................ 102

TABLE 4.3: MOLAR RATIO OF ADDITIVE TO PS AND THE LINEAR LATTICE DISTANCE THAT CORRESPONDS TO. .................................. 120

TABLE 5.1: SAMPLES USED IN EPR STUDIES .................................................................................................................. 127

Page 19: Photochemical Crosslinking Reactions in Polymers

xiv

Acknowledgements

This Ph.D. thesis took the support of a large and varied group of people; only a small

fraction can be acknowledged here. First and foremost I’d like to thank my advisor Professor

Jeffrey Koberstein for his support, guidance, and timely insight. Without him this would not

have been possible. Professor Turro provided invaluable discussions and lab space; his students

chemicals and suggestions. I would like to thank all the members of the Koberstein and Turro

labs, past and present, for their support and help in and out of the lab. Hernán Rengifo, Chris

Grigoras, Alberto Moscatelli, Greg Carroll, Jeffrey Lancaster, Ellane Park, and Ben Dach deserve

special mention for their assistance. They asked the questions that pushed this project further.

Jose Pelaez, Tomi Herceg and Mary Dickson helped me establish the foundations on which this

work is based, and Madalina Ene provided indispensible practical and conceptual help in the

lab. I had the excellent administrative support of Teresa Colaizzo and Mary Ko to thank for the

smooth operation of the lab; I don’t know how you do it, but you do it wonderfully. Thank you.

On the personal side, I want to thank my family for the years of love, support and

understanding that brought me through both the good times and the rough. To the friends

who helped maintain my sanity: you know who you are. Thank you. Last and most importantly

I’d like to thank Jessica Freeman-Slade for her love and understanding.

Page 20: Photochemical Crosslinking Reactions in Polymers

xv

For My Family

Frank and Britani Carbone

Harry and Ann Bartelme

Frank and Josephine Carbone

Michael Carbone

Page 21: Photochemical Crosslinking Reactions in Polymers

xvi

This Page Intentionally Left Blank

Page 22: Photochemical Crosslinking Reactions in Polymers

1

1. Chapter 1: Motivation and Introduction

1.1. Motivation

Polymers bridge both the gap between small molecules and large biological molecules and

the gap between molecular-scale interactions and bulk interactions. Small molecule properties

are constant irrespective of the number of molecules in the bulk. Large biological molecules

such as proteins require stabilization to extend their individual properties to the aggregate (1).

Polymers are long chain molecules made up of one or more simple chemical building blocks

whose molecular-scale interactions can have a large influence on bulk properties through short-

and long-range physical and chemical interactions. The a priori design of polymer systems from

desired properties is an ongoing challenge (2-10).

Polymer property manipulation begins at synthesis. The length of the chains (11-14),

the width of the length distribution (15, 16) , the presence of other monomers and their order

on the chain (17-20), the addition of functional groups within or at the end of the chain (21-24),

the incorporation of multifunctional monomers (25, 26) or crosslinking agents (27-35), and the

synthesis method (36-39) all affect the physical, chemical, and thermodynamic properties of

the resulting material. The industrial production of polymers with specific properties

introduces high volume and high throughput requirements to polymer synthesis that can make

production infeasible(40, 41).

Post-synthesis processing offers an alternative route to the desired properties at both

industrial and laboratory scales. Chemical modification of the polymer monomer(42) ,

Page 23: Photochemical Crosslinking Reactions in Polymers

2

reactions with functional groups added during synthesis (24, 43, 44), and crosslinking allow the

further modification of synthetic polymers.

Crosslinking, either to promote chain branching (45-50) or full network formation (17,

51-54), is a common method for polymer modification because of its profound influence on

polymer mechanical properties (45, 55, 56) and other characteristics such as mobility, solubility,

and diffusion. Covalent crosslinks introduced during polymer synthesis generally produce

thermoset polymers that retain the shape in which they were polymerized and are insoluble in

all solvents. Crosslinks are introduced after synthesis by the addition of a crosslinking additive,

by exposure to high energy radiation, or by a combination of the two processes. Examples of

crosslinking additives include sulfur (i.e. vulcanization) (57-59) and peroxides (60-62) that

crosslink upon heating, and glutaraldehyde, a common fixative for proteins (63-68). Radiative

methods are typically based on the formation of radicals by exposure to high energy radiation

such as gamma rays (52, 69) and electron beams (28, 70-72).

Alternatively, photocrosslinking of preformed polymers can be achieved by use of a

photosensitizer and visible or UV radiation. The ease and safety of photochemical crosslinking

makes it an attractive option. Most photosensitizers crosslink through radical chemistry;

hydrogen abstraction reactions are often used to generate the radicals. These reactions will be

discussed in more detail later. Polymer crosslinking through hydrogen abstraction is used in the

synthesis of grafted polymers (73), the initiation of thiol-ene crosslinking (74), and macroradical

recombination (24, 43, 75-77).

Page 24: Photochemical Crosslinking Reactions in Polymers

3

This thesis explores the photocrosslinking of preformed polymers using primarily

hydrogen abstraction methods. The properties of four photoactive crosslinkers, both mono-

functional and bi-functional, are discussed in Chapter 3 as well as the properties of the

polymers in which they are being studied. Their mutual miscibility is also discussed. Chapter 4

examines one additive, Benzophenone, in detail in both glassy and rubbery thin films and

investigates the effects of additive functionality on the changing molecular weight distributions.

Chapter 5 investigates the location of the radical on the polymer and where reactive preference

lies in the systems of Chapter 4. Chapter 6 surveys the results of preliminary investigations into

other combinations of additive and polymer and links the results to the properties discussed in

Chapter 3.

1.2. Introduction to Photochemistry

Photochemistry is the study of the chemical effects of light. While light can have many

effects on molecules, we focus here on the photochemical generation of radicals.

1.2.1. Radicals

The term radical is usually used in one of two ways. It can be used to refer to an atom

or molecule with an unpaired electron (25), or it can be used to refer to the unpaired electron

itself. In this work we will use the second definition. The ‘benzophenone radical’, for example,

will refer to the unpaired electron on the benzophenone molecule. If there is more than one

radical on a molecule the specific radical will be referenced with further identifying

information.

Page 25: Photochemical Crosslinking Reactions in Polymers

4

Radicals are generally produced through homolysis of covalent bonds, induced by

pyrolysis, thermal decomposition, or photolysis(78). Some oxidation and reduction reactions

can also generate single radicals(79). Ketones, aldehydes and azo- compounds are often used

to generate alkyl radicals(78) thermally or photochemically.

1.2.2. Photochemical Generation of Radicals (80, 81)

The photochemical generation of radicals occurs through the absorption of a photon. A

photon of energy h� is absorbed by a molecule if the molecule contains a subunit

(chromophore) in which the difference in energy between the excited state (��) and ground

state (��) is equal to the photon energy as shown in equation (1.1).

ℎ�= �� − �� (1.1)

Photon absorption promotes an electron from a σ, π, or n orbital to a �∗ or �∗ orbital. Each

transition is associated with photons of different energies; the transitions and their energies are

shown in Table 1.1.1.

Table 1.1.1: Orbital Transitions and Optical Absorption(80)A

Orbital Transition Absorption region

(nm) Extinction coefficient

(� ����� ����) � → �∗ 100-200 103 � → �∗ 150-250 102-103 � → �∗

Isolated Conjugated

180-250

200-Infrared 102-104

� → �∗ Isolated

Conjugated

220-320

250-Infrared 1-400

A) This table is taken from reference (80), Table 1.2, pg 9.

Page 26: Photochemical Crosslinking Reactions in Polymers

5

Photoexcitation leads to two singly occupied orbitals. If the excitation occurred without an

electron spin inversion, then there are no unpaired spins in the molecule and a singlet state

results. This is associated with the � → �∗ transition. An excitation with spin inversion leads to

two unpaired spins and the triplet state, associated with the � → � ∗ transition. Spin inversion

during excitation is forbidden (82) but the magnitude of spin-orbital coupling in the specific

molecule or chromophore can bias the system towards an intersystem crossing to the triplet

state immediately after excitation. The energies of the photons required for orbital transitions

indicate that orbital energies increase as � < � < � < � ∗ < � ∗. The specific orbital energies,

transition energies, and excitation wavelengths to generate a specific state are highly

dependent upon the local environment. Interorbital and interatomic interactions all have an

effect all of these criteria and the specific chromophore that the photon interacts with is very

important. Selected chromophores and their orbital transitions are shown in Table 1.1.2. ����

is the wavelength at which there is an absorption peak, and ���� is the molar extinction

coefficient from the Beer-Lambert law, equation (1.2):

� = log(�� �⁄ )= ��� (1.2)

where A is the optical absorbance, �� is the incident light flux, � is the transmitted light flux, � is

the solute concentration, � the instrument path length, and � the molar extinction coefficient, a

species-specific constant.

Page 27: Photochemical Crosslinking Reactions in Polymers

6

Table 1.1.2: Selected Chromophores (80)A

Chromophore ����

(�� )(�) ����

(� ����� ����)(�) Orbital Transition

193 104 � → �∗

187 271

103 15

� → �∗ � → �∗

347 5 � → �∗

(A) This table is an excerpt from reference (80), Table 1.1 pg. 6. (b) Wavelength of maximum optical absorption (c) Molar extinction coefficient (���(�� �⁄ )= ���) at ����

Table 1.1.2 shows three different chromophores that are important for the molecules

used in our studies and some of their properties. The alkenyl group, with a carbon-carbon

double bond, requires a high energy photon for its � → �∗ singlet transition and has a

commensurate high extinction coefficient because there is significant spatial overlap between

the π and �∗ orbitals(82). A high extinction coefficient, which is associated with high

absorbance values, means that a small number of alkenyls in a system will have a deceptively

large effect on the system’s chemistry, assuming light of the appropriate wavelength is present

in the system. Conjugation of the � bond is also possible depending upon R1 to R4. Conjugated

� bonds provide both energy stabilization (83) and lower the energy required for excitation

(84). The carbonyl group, with a carbon-oxygen double bond, has two different states. Low

wavelength photons have sufficient energy for a � → �∗ singlet transition. This transition also

has a relatively high extinction coefficient and therefore a large influence on the chemistry of

the system. Higher energy states have a higher probability of homolysis, where the bond

C CR1

R2 R3

R4

C OR2

R1

R1 N N R2

Page 28: Photochemical Crosslinking Reactions in Polymers

7

dissociates into two molecules, because the excited resonance has a higher energy and

therefore produces larger oscillations (85, 86). Longer wavelength photons have less energy

and are therefore lead predominantly to the lower energy � → � ∗ triplet transitions. The

triplet state has relatively low molar extinction, indicating decreased concentration sensitivity

in comparison to the higher energy transitions and higher stability with a lower probability of

homolysis. This is an advantage in our systems because the lower rate of triplet formation with

time will allow us to investigate the initial stages of the reaction and assume a quasi steady

state concentration of triplets and reactants. The conjugation of the carbonyl � bond is less

likely than that of the alkenyl both because of the 2 fewer potential bonds and the highly polar

nature of the carbon-oxygen bond (87). The third and final group in Table 1.1.2 is an azo group,

with a nitrogen-nitrogen double bond. This is also a � → � ∗ transition, and has the smallest

extinction coefficient as well as the longest wavelength requirement. With ���� = 347�� ,

the shoulder of the wavelength sensitivity approaches visible light, which makes the

compounds much more difficult to work with, for obvious reasons.

From the above information we can therefore predict that the � → � ∗ triplet transition

of the carbonyl will probably produce additives for our systems that are easy to work with

because the wavelength is too short for visible light sensitivity. Carbonyl-derivatives will also

lead to fewer side and dissociation reactions because they do not react with molecularly

destructive high energy photons and most polymer systems we are interested in do not contain

chromophores that are sensitive in that range (80).

Page 29: Photochemical Crosslinking Reactions in Polymers

8

Molecules in the singlet or triplet state have two unpaired electrons which can

recombine into their ground state by giving up their energy through fluorescence (from the

singlet), phosphorescence (from the triplet) or molecular vibrations (85). A radical is formed

when a chemical reaction occurs between one of the excited electrons and another molecule,

leaving a single unpaired electron in the molecule. In most of our systems, the radical is

generated on the additive’s carbonyl through hydrogen abstraction from the polymer backbone

as shown in Figure 1.1.1. In this representative reaction schema, an ultraviolet photon excites

the carbonyl chromophore to the singlet state, which immediately transitions to a triplet

through an intersystem crossing (not shown, as the quantum yield is 1 (85, 88)). The triplet,

which has a lifetime of 10-4 seconds (85), then encounters the polymer backbone and its

multiple carbon-hydrogen bonds. Assuming it is sterically favored, the triplet will abstract a

hydrogen from the polymer backbone, producing free radicals on both the carbonyl

chromophore and the polymer backbone. The free radical is of lower energy than the triplet

state (85, 89) and is energetically preferred. It is also more stable and therefore will last longer

in the system.

C

O

R1 R2

hvC

O

R1 R2

Excitation

Ground State Triplet State

C

O

R1 R2

Triplet State

Hydrogen Abstraction

H2C

CH

R3

Alkyl Polymer Backbone

C

OH

R1 R2

Free Radical

H2C

C

R3

Polymeric Free Radical

nn

Page 30: Photochemical Crosslinking Reactions in Polymers

9

Figure 1.1.1: Photochemical formation of a free radical through carbonyl excitation and hydrogen abstraction from the backbone of an alkyl polymer. This schema is representative of the majority of systems in this study.

Figure 1.1.1 shows how two free radicals are formed from the excitation of a carbonyl

chromophore by light and the excited triplet state’s interaction with the polymer backbone.

The hydrogen on the polymer backbone that is abstracted by the triplet state is determined by

how stable the resulting radical will be. In general, the stability of a radical center is

determined by how substituted it is. A tertiary radical is more stable than a secondary radical,

which is more stable than a primary radical (Figure 1.1.2). The stability of the radical center

increases with substitution because there is an inverse correlation between substitution and

radical energy. The orbital overlap in highly substituted centers allows more freedom of

movement to the radical and there is the possibility of radical migration with enough energy,

both of which lower the energy of the radical. We would therefore expect a carbonyl triplet to

hydrogen abstract first from a tertiary carbon. This is discussed in more detail in Chapter 3.

Figure 1.1.2: The stability of radical centers and their relation to hydrogen abstraction.

We have seen how photons excite molecules, producing singlet or triplet excited states

in chromophores that can, in interaction with other molecules, produce free radicals. These

radicals are the source of the reactions discussed in this thesis. They will be investigated in

polymer thin films and solutions. In thin films they are investigated indirectly by tracking the

CHH

H

CHR1

H

CR2R1

H

CR2R1

R3

Radical StabilityProbability of Hydrogen Abstraction

Radical Energy

Page 31: Photochemical Crosslinking Reactions in Polymers

10

molecular weight distribution changes in the polymer; in solutions they are investigated directly

through electron paramagnetic resonance.

1.3. Polymers and their Properties

1.3.1. Individual Chain Properties

A linear homopolymer is made up of a large number of covalently bonded bivalent

monomers. The molecular weight of the chain is defined as the number of monomers, � times

the molecular weight of the monomer, �. Equation (1.3) shows this quantity.

� = �� (1.3)

Polymers are long chain molecules whose secondary structure is constantly in flux due

to thermal energy. Simple physical properties of the molecule such as size, volume, and shape

can never be absolutely defined. Instead, a series of analogous properties are used. Flory(31)

and de Gennes(90) have excellent derivations and discussions of these properties in chapters

10 and 1, respectively. The square end to end distance of the polymer chain, ��, is found to be

�� = ��� (1.4)

with � the number of monomers in the chain and � the linear size of the mer along the

backbone of the polymer. The radius of gyration, ��, of the polymer is one sixth of Equation

(1.4).

��� =

��

6=���

6 (1.5)

This assumes the polymer is a freely jointed chain with no external flow field and no

inter- or intra- chain interactions that behaves as a random walk. Equation (1.5) is an accurate

Page 32: Photochemical Crosslinking Reactions in Polymers

11

form of the radius of gyration for a polymer molecule treated as an aphysical, purely

mathematical construct. Equation (1.5) is a good first approximation for polymer properties

and can be modified to increase its physical accuracy to the degree desired.

All covalent chemical bonds connect at an angle proscribed by the interactions of the

individual valence shells. Equation (1.6) introduces fixed bond angles to the radius of gyration,

with � the interior angle between two backbone molecules.

��� =

���

6�1 + ����

1− ����� (1.6)

There are energetic barriers to bond rotation as the valence shells of pendant atoms

interact with each other during rotation. The energetic minima during rotation are the gauche+,

gauche-, and trans positions. These energetic minima bias the free rotation of the bonds in the

chain. Equation (1.7) introduces restricted rotation to the fixed bond angle radius of gyration in

Equation (1.6), with ⟨����⟩ the ensemble average of the allowed rotational angles as defined

by the rotational energy landscape.

��� =

���

6�1 + ����

1 − ������1 + ⟨����⟩

1 − ⟨����⟩� (1.7)

Equations (1.5)- (1.7) are increasingly accurate radii of gyration for an individual polymer

chain developed a priori from mathematical random walks. They neglect both physical and

chemical interactions. Equations (1.5)- (1.7) are valid for polymer chains that behave like

Gaussian random walks; only individual polymer chains in certain solvents have the freedom of

motion to access a Gaussian random walk configuration.

Page 33: Photochemical Crosslinking Reactions in Polymers

12

Van der Waals, Hydrogen bonding, and solvation effects all change the conformation

and size of a polymer chain in solution. If we assume only London dispersion forces affect the

solvent-polymer interactions and the validity of mean field theory in a polymer-solvent system,

then the Gibbs free energy of mixing a polymer molecule into solvent (for a binary mixture

only) as derived in (31) is shown in equations (1.8)-(1.9).

∆�� = ��[����(��)+ ����(��)+ �������] (1.8)

��� =(�� − ��)

�����

(1.9)

Where R is the gas constant, T the absolute temperature, �� and �� the number of moles and

volume fraction of species �, �� and �� the Hildebrand solubility parameter of the polymer and

solvent, respectively, and ��is the molar volume of a polymer monomer.

The specific and nonspecific interactions of a single polymer molecule in solution

generally change the physical dimensions of the polymer chain. If the favorable interactions

perfectly balance out the unfavorable interactions, then the polymer chain is said to be in a

Theta solvent, and the chain configuration is that of a random walk. In the theta state,

equations (1.5)- (1.7) are essentially valid and perfectly describe the radius of gyration, with

accuracy increasing with increasing equation number for the reasons discussed in their section.

In a bad solvent, the unfavorable interactions dominate and the polymer condenses into itself

and may precipitate. In a good solvent the favorable interactions dominate, the polymer

swells, the radius of gyration increases, and excluded volume effects begin to dominate.

Excluded volume considerations appear when the long range polymer-polymer and

polymer-solvent interactions become more important than the short range interactions. This is

Page 34: Photochemical Crosslinking Reactions in Polymers

13

caused both by previous occupancy of a space by a polymer segment far down the chain and

the formation of a solvent “sheath”, where each monomer surrounds itself with solvent

molecules which stiffen the chain and shield polymer-polymer interactions. The chain, in

essence, straightens and expands. Flory and Krigbaum showed (91) that despite the lack of

defined shape for the polymer in solution, the radius of gyration for a non-ideal polymer in

solution can be determined through the addition of an expansion factor, �.

��� =

���

6�1 + ����

1− ������1 + ⟨����⟩

1 − ⟨����⟩��� (1.10)

�� − �� =27

2(2�)����1 2� − ���√�

where �� is the Flory-Huggins interaction parameter defined in equation (1.9), and � is the

number of monomers.

The expansion factor ∝ incorporates the polymer-polymer and polymer-solvent

interaction thermodynamics. We simplify equation (1.10) by defining the statistical segment

length, β, which is the length of the chain that, from far away, behaves exactly like an ideal

Gaussian random walk. All length and size affecting terms are collected in this variable and we

redefine equation (1.5):

��� = ��� (1.11)

� =����

6�1 + ����

1 − ������1 + ⟨����⟩

1 − ⟨����⟩�

Page 35: Photochemical Crosslinking Reactions in Polymers

14

The renormalization of the length of the chain requires a renormalization of the number

of monomers in the chain, N. We define � = ��� , with � the number of monomers per

statistical segment β, also known as the Kuhn length. The Kuhn length is a simple

approximation for the stiffness of the chain, as it quantifies the distance at which chain motion

is no longer correlated.

Within the statistical segment length β, the size of the chain scales as the number of

segments to a power that reflects the quality of the solvent, � ∝ �� (90). The radius of

gyration of a molecule is proportional to the number of monomers and the statistical segment

length to a solvent-dependent power.

��� ∝ ���� ∝

���� ∝ ������ ∝ ����

��� (1.12)

where 1 2� ≤ � ≤ 35� as the system varies from a theta �� = 1

2� � to a good �� =3 5� �

solvent.

We have defined the radius of gyration of a single polymer molecule both as an ideal

and in solution. This simplest of properties – the size of a molecule in solution – has surprising

depths that require both complicated assumptions and detailed knowledge of the polymer and

its properties. The extension of polymer properties to multiple chains is equally complicated.

1.3.2. Ensemble Properties

It is currently impossible to synthesize polymers samples in which every polymer chain is

exactly the same length (31, 90, 92). Polymer samples are defined by their distributions.

Equation (1.3) defines a single chain; summing it over all chains in a distribution provides the

Page 36: Photochemical Crosslinking Reactions in Polymers

15

number average molecular weight, ��, the number of monomers per chain averaged over all

chains in the distribution.

�� =∑ �����

∑ ��� (1.13)

The second moment of the distribution, the weight average molecular weight �� ,

quantifies the mass distribution of the polymer sample:

�� =∑ ����

��

∑ ����� (1.14)

This is the quantity most often associated with the bulk properties of a polymeric material

because large molecular weight species play a disproportionately large role in the properties of

a polymer. Dividing the �� by �� produces the polydispersity index, the measure of the width

of the distribution. The theoretical minimum polydispersity is 1; the lowest achieved

polydispersity is around 1.01. The theoretical maximum polydispersity is infinite. Experimental

polydispersities above 2 or 3 are not uncommon.

The z-average and z+1 average molecular weights are defined below and are generally

associated with rheological properties.

�� =∑ ����

��

∑ �����

(1.15)

���� =∑ ����

��

∑ �����

(1.16)

Each theoretical distribution is generally related to a different experimental method.

Fractionalization, where the change in polymer solubility with size is exploited to fractionate

Page 37: Photochemical Crosslinking Reactions in Polymers

16

the polymer into lower and lower PDIs, measures the number average molecular weight, as the

separation is dependent upon the size of the molecule. Turbidity measurement techniques

extract the weight average molecular weight, while viscometric methods extract the z average

molecular weights. The main polymer analysis technique used in this work is high pressure

liquid chromatography because it measures the entire distribution even at low concentrations.

Page 38: Photochemical Crosslinking Reactions in Polymers

17

1.4. References

1. V. Ragoonanan, A. Aksan, Transfusion Medicine and Hemotherapy 34, 246 (2007). 2. K. Ariga, Q. M. Ji, J. P. Hill, A. Vinu, Soft Matter 5, 3562 (2009). 3. J. E. Puskas et al., Macromol. Mater. Eng. 286, 565 (Oct, 2001). 4. K. Raemdonck, J. Demeester, S. De Smedt, Soft Matter 5, 707 (2009). 5. R. P. Wool, Soft Matter 4, 400 (2008). 6. J. T. Koberstein, “Molecular engineering of thin polymer films prepared from functionally-

terminated oligomers” (Inst. Mater. Sci.,Connecticut Univ.,Storrs,CT,USA., 1995). 7. J. T. Koberstein, Journal of Polymer Science Part B-Polymer Physics 42, 2942 (Aug, 2004). 8. J. T. Koberstein, D. Cho, D. Wong, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 46, 458

(2005). 9. J. T. Koberstein et al., J. Adhes. 66, 229 (1998). 10. J. R. Lancaster et al., Chem. Mater. 20, 6583 (2008). 11. G. T. Dee, B. B. Sauer, Trends in Polymer Science 5, 230 (Jul, 1997). 12. G. T. Dee, B. B. Sauer, Advances in Physics 47, 161 (Mar-Apr, 1998). 13. X. Quan, I. Gancarz, J. T. Koberstein, G. D. Wignall, Macromolecules 20, 1431 (1987). 14. V. K. Raghavendran, M. C. Waterbury, V. Rao, L. T. Drzal, Journal of Adhesion Science and

Technology 11, 1501 (1997). 15. C. F. Laub, J. T. Koberstein, “Effect of brush polydispersity on the interphase between end-

grafted brushed and polymeric matrixes” (Inst. of Materials Science,Connecticut Univ.,Storrs,CT,USA., 1994).

16. J. van der Gucht, N. A. M. Besseling, G. J. Fleer, Macromolecules 35, 6732 (Aug, 2002). 17. J. P. Baetzold, I. Gancarz, X. Quan, J. T. Koberstein, Macromolecules 27, 5329 (1994). 18. A. C. Balazs, M. Gempe, C. W. Lantman, Macromolecules 24, 168 (Jan, 1991). 19. M. A. Biesalski, S. Duman, K. Shroff, J. Ruehe, PMSE Prepr. 96, 17 (2007). 20. A. F. Galambos, T. P. Russell, J. T. Koberstein, “Origin of multiple melting in polyurethane block

copolymers” (Inst. Mater. Sci.,Univ. Connecticut,Storrs,CT,USA., 1990). 21. C. Jalbert, J. T. Koberstein, A. Hariharan, S. K. Kumar, Macromolecules 30, 4481 (Jul, 1997). 22. J. T. Koberstein, D. H. T. Lee, J. F. Elman, P. M. Thompson, I. Yilgor, Polym. Prepr. (Am. Chem.

Soc., Div. Polym. Chem.) 32, 265 (1991). 23. P. A. O. R. Muisener, J. T. Koberstein, S. Kumar, Proc. Annu. Meet. Adhes. Soc. 23rd, 288 (2000). 24. J. Ruhe, K. Seidel, R. Toomey, Abstracts of Papers, 227th ACS National Meeting, Anaheim, CA,

United States, March 28-April 1, 2004, COLL (2004). 25. S. Murayama, S. Kuroda, Z. J. Osawa, Polymer 34, 2845 (1993). 26. G. Mabilleau et al., Journal of Biomedical Materials Research Part A 77A, 35 (2006). 27. H. R. Kricheldorf, H. H. Thiessen, Polymer 46, 12103 (Dec, 2005). 28. G. W. Meyer et al., Polymer 37, 5077 (Oct, 1996). 29. S. Cuney et al., Journal of Applied Polymer Science 65, 2373 (1997). 30. B. Unal, R. C. Hedden, Polymer 47, 8173 (2006). 31. P. J. Flory, Principles of Polymer Chemistry. (Cornell University Press, Ithica, New York, 1953). 32. W. Radke, G. Litvinenko, A. H. E. Muller, Macromolecules 31, 239 (Jan, 1998). 33. O. Prucker, B. Peng, S. Golze, H. Murata, J. Ruhe, Polym. Prepr. (Am. Chem. Soc., Div. Polym.

Chem.) 44, 418 (2003). 34. M. Doytcheva, R. Stamenova, V. Zvetkov, C. B. Tsvetanov, Polymer 39, 6715 (1998). 35. M. Nagata, S. Hizakae, Macromol. Biosci. 3, 412 (Aug, 2003). 36. M. Castonguay, J. T. Koberstein, Z. Zhang, G. Laroche, Tissue Eng. Intell. Unit 6, 1 (2001).

Page 39: Photochemical Crosslinking Reactions in Polymers

18

37. D. Cunliffe, J. E. Lockley, J. R. Ebdon, S. Rimmer, B. J. Tabner, Macromolecules 34, 3882 (Jun 5, 2001).

38. J. A. Johnson, M. G. Finn, J. T. Koberstein, N. J. Turro, Macromolecules (Washington, DC, U. S.) 40, 3589 (2007).

39. X. Wu, R. H. Pelton, A. E. Hamielec, D. R. Woods, W. McPhee, Colloid and Polymer Science 272, 467 (Apr, 1994).

40. F. N. Cogswell, Polymer Melt Rheology: a guide for industrial practice. (Wiley, New York, 1981). 41. A. Ram, Fundamentals of Polymer Engineering. (Plenum Press, New York, 1997). 42. R. Fayt et al., Macromolecules 20, 1442 (Jun, 1987). 43. J. Ruhe, D. Madge, Abstracts of Papers, 225th ACS National Meeting, New Orleans, LA, United

States, March 23-27, 2003, PMSE (2003). 44. J. A. Johnson, J. T. Koberstein, N. J. Turro, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 50,

No pp given (2009). 45. S. N. Vouyiouka, E. K. Karakatsani, C. D. Papaspyrides, Progress in Polymer Science 30, 10 (2005). 46. B. Blottiere, T. C. B. McLeish, A. Hakiki, R. N. Young, S. T. Milner, Macromolecules 31, 9295

(1998). 47. W. Hu, J. T. Koberstein, J. Polym. Sci., Part B: Polym. Phys. 32, 437 (1994). 48. J. T. Koberstein, T. P. Russell, D. J. Walsh, L. Pottick, Macromolecules 23, 877 (1990). 49. S. T. Milner, T. C. B. McLeish, R. N. Young, A. Hakiki, J. M. Johnson, Macromolecules 31, 9345

(1998). 50. J. F. Rontani, Trends Photochem. Photobiol. 4, 125 (1997). 51. J. P. Baetzold, J. T. Koberstein, Macromolecules 34, 8986 (2001). 52. C. B. Bucknall, V. L. P. Soares, J. Polym. Sci., Part B: Polym. Phys. 42, 2168 (2004). 53. P. J. Cole, J. L. Lenhart, J. A. Emerson, J. T. Koberstein, C.-Y. Hui, Proc. Annu. Meet. Adhes. Soc.

27th, 470 (2004). 54. J.-F. Fu et al., Polym. Adv. Technol. 19, 1597 (2008). 55. R. J. Gaymans, J. Schuijer, ACS Symp. Ser. 104, 137 (1979). 56. R. J. Gaymans, T. E. C. Van Utteren, J. W. A. Van den Berg, J. Schuyer, J. Polym. Sci., Polym. Chem.

Ed. 15, 537 (1977). 57. I. S. Zemel, C. M. Roland, Polymer 33, 3427 (1992). 58. N. J. Morrison, M. Porter, Rubber Chemistry and Technology 57, 63 (1984). 59. M. R. Krejsa, J. L. Koenig, Rubber Chemistry and Technology 66, 376 (1993). 60. N. G. Gaylord, M. Mehta, R. Mehta, Journal of Applied Polymer Science 33, 2549 (May, 1987). 61. G. E. Hulse, R. J. Kersting, D. R. Warfel, J. Polym. Sci. Pol. Chem. 19, 655 (1981). 62. W. Brostow, T. Datashvili, K. P. Hackenberg, Polymer Composites 31, 1678 (Oct, 2010). 63. Z. Chen, Z. Liu, F. Liu, L. Yang, Zhongguo Shengwu Gongcheng Zazhi 23, 99 (2003). 64. G. DeSantis, J. B. Jones, Current Opinion in Biotechnology 10, 324 (1999). 65. G. G. Guilbault, Bio/Technology 7, 349 (1989). 66. A. Jayakrishnan, S. R. Jameela, Biomaterials 17, 471 (1996). 67. L. Lin et al., Zhongguo Yixue Kexueyuan Xuebao 25, 735 (2003). 68. J. Xu et al., Huagong Jinzhan 29, 494 (2010). 69. B. D. Shirodkar, R. P. Burford, Radiat. Phys. Chem. 62, 99 (2001). 70. N. S. Allen, D. Lo, M. S. Salim, P. Jennings, Polym. Degrad. Stabil. 28, 105 (1990). 71. J. Cheng et al. (International Business Machines Corporation, USA., US, Application: US 2009),

pp. 25pp , Cont of U S Ser No 13,444. 72. V. K. Sharma, J. Mahajan, P. K. Bhattacharyya, Radiat. Phys. Chem. 45, 695 (1995). 73. D. Dyer, R. Jordan, Ed. (Springer Berlin / Heidelberg, 2006), vol. 197, pp. 47-65. 74. C. R. Morgan, F. Magnotta, A. D. Ketley, J. Polym. Sci. Pol. Chem. 15, 627 (1977).

Page 40: Photochemical Crosslinking Reactions in Polymers

19

75. M. D. Millan, J. Locklin, T. Fulghum, A. Baba, R. C. Advincula, Polymer 46, 5556 (2005). 76. G. T. Carroll, M. E. Sojka, X. Lei, N. J. Turro, J. T. Koberstein, Langmuir 22, 7748 (2006). 77. G. T. Carroll, L. D. Triplett, A. Moscatelli, J. T. Koberstein, N. J. Turro, Journal of Applied Polymer

Science 122, (2010). 78. J. A. Kerr, in Free Radicals, J. K. Kochi, Ed. (John Wiley & Sons, New York,, 1973), vol. 1, pp. 2 bd. 79. W. A. Pryor, Free Radicals. (McGraw-Hill Book Company, New York, ed. 1, 1966). 80. W. Schnabel, Polymers and light : fundamentals and technical applications. (Wiley, Chichester,

2007), pp. xiv, 382 p. 81. N. J. Turro, Modern molecular photochemistry. (Benjamin/Cummings Pub. Co., Menlo Park,

Calif., 1978), pp. 628 p. 82. in McGraw-Hill Science and Technology Encyclopedia. (McGraw-Hill, 2005). 83. I. Fernandez, G. Frenking, Chem.-Eur. J. 12, 3617 (Apr, 2006). 84. C. H. Yuan, R. West, Appl. Organomet. Chem. 8, 423 (Jul, 1994). 85. N. J. Turro, V. Ramamurthy, J. C. Scaiano, Modern molecular photochemistry of organic

molecules. (University Science Books, Sausalito, Calif., 2010), pp. xxxiii, 1084 p. 86. J. K. Kochi, Ed., Free Radicals, vol. 2 (John Wiley and Sons, New York, 1973), vol. 2. 87. S. H. Hamid, Ed., Handbook of Polymer Degradation, (Marcel Dekker, Inc., New York, 2000). 88. J. R. Lancaster, Columbia University (2011). 89. W. A. Pryor, Free radicals. McGraw-Hill series in advanced chemistry (McGraw-Hill, New York,,

1966), pp. xii, 354 p. 90. P.-G. de Gennes, Scaling Concepts in Polymer Physics. (Cornell University Press, Ithaca, New

York, 1979). 91. P. J. Flory, Krigbaum, J. Chem. Phys. 18, 1086 (1950). 92. A. Goto, T. Fukuda, Progress in Polymer Science 29, 329 (Apr, 2004).

Page 41: Photochemical Crosslinking Reactions in Polymers

20

2. Chapter 2: Experimental Techniques

Electron Paramagnetic Resonance (EPR) and High Pressure Liquid Chromatography-Size

Exclusion Chromatography (HPLC-SEC) are the major analytical techniques used in the current

study. Preparation and Characterization techniques used include Spin Coating and

Ellipsometry. In this chapter, a brief introduction to these techniques is given.

2.1. Electron Paramagnetic Resonance

Electron Paramagnetic Resonance (EPR) is a powerful technique for studying radicals in

materials. Unpaired electrons are paramagnetic: they can be induced into aligning with the

magnetic field of a strong magnet but are not normally aligned. An EPR induces the unpaired

electrons in a material to line up with a magnetic field, much like atomic nuclei are aligned in

Nuclear Magnetic Resonance (NMR). Microwave radiation is input into the EPR as the magnetic

field is varied. The unpaired electrons resonate with the microwave energy at a specific

magnetic field strength that is unique to their orbital environment, allowing the determination

of their neighboring atoms.

Electrons are spin ½ with a magnetic quantum number of ms=±½. Unpaired electrons

align with the EPR’s magnetic field, and when resonance occurs, the electrons oscillate between

the antiparallel (ms=+½) and the parallel (ms=-½) energy levels, corresponding to two different

energies, as shown in Figure 2.1. This state flip changes the energy flowing through the system,

producing a signal. Only unpaired electrons respond to EPR; this allows EPR to study the

Page 42: Photochemical Crosslinking Reactions in Polymers

21

radicals in orbitally complicated molecules without interference from the rest of the molecule,

as in NMR. Atoms with multiple unpaired electrons can have compound spins; the most

abundant paramagnetic isotopes in organic molecules are listed in Table 2.1.

Figure 2.1: Energy diagram of unpaired electrons in an induced magnetic field.

The flip between the parallel and antiparallel state occurs at a specific energy level for

every radical. Equation (2.1) is the energy difference between the two states:

∆� = ℎ� = ��� (2.1)

Where ℎ is Planck’s constant, � is the frequency of the microwave radiation, � is the �-factor,

β is the Bohr magneton, and � is the magnetic field. The �-factor is a measure of the intrinsic

magnetic moment of the electron and is unitless.

The interaction of the unpaired electron with the spins of nearby nuclei through

magnetic coupling (in a similar way to spin-spin coupling in NMR) opens up additional energy

states to the electron. The coupling can occur either between the unpaired electron and its

own nucleus, which allows for energy states linked to the paramagnetic spin (Table 2.1), or

between the unpaired electron and nearby nuclei, which opens up energy states whose EPR

Page 43: Photochemical Crosslinking Reactions in Polymers

22

intensities are linked to the number of nearby atoms through Pascal’s triangle. The additional

states generate additional resonances that produce spectra that are unique to the local

environment and are a combination of the intra- and inter-nuclear magnetic interactions. Each

interaction generates a number of peaks based upon how many nuclei are nearby (Table 2.2).

Table 2.1: Most abundant paramagnetic isotopes of C,H,O and N and their spins.

Atom Spin 13C ½ 1H ½

17O 5 2�

14N 1

Table 2.2: Number of EPR peaks generated by the combination of spin and number of atoms.

Number of Atoms: 1 2 3 4 Spin Number of Peaks

½ 2 3 4 5 1 3 5 7 9 32� 5 11 17 23

52� 6 11 16 21

Figure 2.2: Diagram of a basic EPR instrument. A microwave source sends microwaves through an attenuator into a circulator that guides them into the sample chamber and then back out to the detector. The sample chamber sits between electromagnets.

An EPR instrument is diagrammed in Figure 2.2. The microwaves are generated in a

source, usually a klystron, and follow waveguides through an attenuator and other systems into

the sample chamber. If unpaired electrons are resonating at the current magnetic field

strength, then some of the microwave energy will be absorbed in electron spin flips. The

Page 44: Photochemical Crosslinking Reactions in Polymers

23

energy follows the waveguide out of the resonator and into the detector, where the spectra are

recorded by a computer.

2.2. High Pressure Liquid Chromatography

High Pressure Liquid Chromatography (HPLC) is a technique developed in the early

1900s by the Russian botanist Mikhail S. Tswett, and is one of the most powerful analytical

tools used in polymer science today. Tswett filled a glass column with powdered chalk or

alumina, and then poured his sample of plant pigments in solvent into the column. Gravity

pushed the solution through the packing, and the differential attraction between the packing

and the various pigment compounds in the sample led to striped bands of pure pigment

compounds. Tswett was thus able to separate and individually test the compounds for their

properties and structure(1). His initial gravimetric liquid chromatography technique has been

upgraded with today’s modern equipment and technology, and is critical to the manufacture of

modern pharmaceuticals.

All modern HPLC systems consist of four key components: a high pressure pump, an

injector, chromatography column(s), and one or more detectors. The HPLC system used in this

study is diagrammed in Figure 2.3. The pump pushes high pressure solvent (known as eluent)

through the system while the sample is injected into the eluent. The sample passes through

the column(s), where separation occurs through one of many different techniques, and then

the separated sample is characterized using one or more detectors that analyze different

properties depending upon the application. Together, the column(s) and the detector(s) are

Page 45: Photochemical Crosslinking Reactions in Polymers

24

the heart of any HPLC system and will be discussed in more detail later. The columns separate

the sample into its constituents based on the chemistry or property desired, while the

detectors characterize the separated sample in the desired way. HPLC can therefore be

optimally tuned to the particular sample and information desired.

Figure 2.3: Diagram of an HPLC system. The pump takes the eluent from the reservoir and pumps it at high pressure through the injector, column(s), and detector(s). Samples are introduced into the injector and are separated into their analytes in the column(s). The detector(s) measure the properties of the analytes. The column oven maintains an isothermal system.

The columns separate the sample into its constituents because some constituents have

more interaction with the column packing material than others, and therefore exit the column

later. There are many different types of columns, each designed for separation based upon a

different chemical or physical interaction. For example, polar packing materials separate the

constituents based upon polarity; ion-exchange packing materials separate by charge; affinity

packing separates based upon specific interactions such as hydrogen bonding or enzyme-

substrate interactions, and size exclusion packing materials separate by molecule size. The

most widely used HPLC column used in polymer science is the size exclusion column, as it

provides details on the size and distribution of polymer chains. When size exclusion columns

are used, people often refer to HPLC as Size Exclusion Chromatography (SEC) or Gel Permeation

Chromatography (GPC). In order to fully understand SEC and how it will be used in this work to

quantify molecular weight changes, we must understand how SEC separates based upon size.

Page 46: Photochemical Crosslinking Reactions in Polymers

25

2.2.1. Size Exclusion Chromatography(2)

As discussed above, SEC separates polymer molecules based upon their size. But the

“size” of a polymer molecule is hard to define, as polymers in solution behave both as freely-

deformable spheres and freely-moving chains. As we will see, SEC separates polymer molecules

based upon a “size” that can and must be correlated to their molecular weight through

absolute measurement or calibration of the SEC.

All HPLC columns operate on the same basic principles as the first column developed by

Tswett. They consist of a tube (now stainless steel) filled with a stationary phase that separates

the analytes and a mobile phase that moves the analytes through the system. While Tswett’s

packing material was powdered chalk or alumina, modern stationary phases are highly

engineered application-specific materials that are packed in proprietary ways to insure as

optimal a packing distribution as possible. The stationary phase in SEC is usually polystyrene-

divinylbenzene copolymer beads that have been synthesized with a known pore-size

distribution that is optimal for analytes to only partially diffuse into and out of the pores. Thus,

smaller molecules can diffuse farther into the pores than larger molecules, making the volume

of the column a function of the molecule size and separating the molecules such that the larger

ones elute from the columns first. The mobile phase moves the analytes through the column

and limits bulk diffusion times. A model of an HPLC column therefore has two regimes: the

stationary, intraparticle regime and the mobile, interparticle regime.

Page 47: Photochemical Crosslinking Reactions in Polymers

26

Figure 2.4: HPLC mobile phase. The flow is left to right. Cutaway shows not-to-scale representation of the packed stationary phase. Figure adapted from (2).

The mobile phase and its variables are shown in Figure 2.4. The flow is left to right. The

sample inters from the left with concentration ��(�) and velocity Q. The interstitial velocity

� =��

����� is the mobile phase velocity between the packing, and is related to the inner

diameter of the column, � and the void fraction ��. The column has length L and is oriented on

the Z-axis. The mobile phase concentration of component � is ��� while its radial diffusivity is

��� .

Figure 2.5: HPLC stationary phase. Solid particle (black line) is surrounded by a stagnant boundary (dotted line) through which only mass transfer occurs. Pore is not to scale and represents the pore volume of the particle (white) and the pore volume accessible to attachment through affinity or other means (grey). Figure adapted from (2)

Page 48: Photochemical Crosslinking Reactions in Polymers

27

The stationary phase and its variables are shown in Figure 2.5. Mass transfer of species

� through the stagnant film boundary (dotted line) with rate �� is the only source to the particle

of radius ��. The particle has a porosity of ��, and species � has an intrapore concentration of

���, a porosity independent effective diffusivity of ���, and a concentration within the solid

phase of ���∗ .

For the development of a model, which follows (2), we will assume:

1. The system is isothermal (experimentally maintained by a column oven)

2. The packing is spherical and uniform.

3. Radial concentration gradients are negligible.

4. There is no convective flow within pores.

5. There is an instantaneous equilibrium between the particle surface and the interior of

the pores.

6. Film mass transfer is the mechanism mediating the interfacial mass transfer between

the mobile and stationary phases.

7. Diffusion and mass transfer parameters are constant and independent of mixing effects.

Based upon the assumptions above, we can derive the differential mass balance on species � in

both the mobile (Equation (2.2)) and stationary phase (Equation (2.3)) where the variables have

been defined in Figure 2.4 and Figure 2.5.

−�����������

+ �������

+������

+3��(1 − ��)

�������� − ���,���� � = 0

(2.2)

�1 − �������

��+ ��

������

− ����� �1

���

�����

������

��= 0 (2.3)

Page 49: Photochemical Crosslinking Reactions in Polymers

28

The first term of the mobile balance (Equation (2.2)) is the Fickian diffusion in the Z-direction;

the second term the convective transport; the third term the temporal concentration gradient;

and the fourth term the interfacial transport between the stationary and mobile phase. The

first term of the stationary balance (Equation (2.3)) is the concentration with time of species �

attached through affinity or other interactions to the solid; the second term is the

concentration with time in the pores; and the third term is the diffusion into and out of the

pores.

By defining the following dimensionless constants,

��� =������

, ��� =������

, ���∗ =

���∗

���, � =

��

�, � =

��, � =

�,

���� =��

���, ��� =

���������

, �� =������

����, �� =

3�����(1 − ��)

��,

we can transform the model equations into the dimensionless equations (1.1) and (1.2)

−1

����

��������

+������

+������

+ ������ − ���,����= 0 (2.4)

����1 − ������

∗ + ������− �� �1

���

�����

������

��= 0 (2.5)

with initial conditions

� = 0, ��� = ���(0,�), ��� = ���(0, �, �)

and boundary conditions

Page 50: Photochemical Crosslinking Reactions in Polymers

29

� = 0, ������

= ���� ���� −���(�)

����

� = 0, ������

= 0

� = 1, ������

= ������� − ���,����, ��� = ���,���

where ��� is the maximum value of ���(�).

Equations (1.1) and (1.2) are the general differential mass balances for any

multicomponent HPLC separation. We incorporate size exclusion effects by defining the

accessible porosity, 0 < ���� < ��, the volume fraction of the pore volume accessible to a

species �, which replaces �� in �� and ���, and the ����� term of Equation (1.2). If we further

assume that there are no specific interactions between the analyte and the solid surface of the

packing, then we can also neglect all ���∗ related terms.

The multicomponent HPLC SEC separation equations, with the above assumptions and

constraints, show that the sample separation in a SEC column is therefore only dependent upon

the accessible porosity, ���� , for each species in the sample. A large accessible porosity implies

the molecules are comparatively small and will therefore have a larger volume to traverse

before eluting from the column. A small accessible porosity implies the molecules are large and

have a relatively small volume to traverse before eluting. Molecules smaller than the smallest

pore can traverse the entire volume. They will not separate, and will elute last in an

undifferentiated group. Molecules larger than the largest pore cannot enter the pores and will

flow directly to the column exit, eluting first in an undifferentiated mass.

Page 51: Photochemical Crosslinking Reactions in Polymers

30

The difficulty in defining the size of a polymer molecule has already been discussed. The

changing shape and size of a polymer makes absolute calculation of interactions between

polymer and SEC pores incredibly difficult. We can gain some understanding, though, by

approximating the pores as cylindrical tubes and the polymer molecules as solid spheres of

radius ��. The accessible porosity for a species � is then proportional to the volume of the

polymer molecule, as the relative volume of the pore and polymer determine the volume

percent accessible to the polymer in solution.

���� ∝ ��

� ∝ ���� ���

���� (2.6)

Substituting in the definition of single chain molecular weight from equation (1.12) we

find that the key separation variable in SEC-HPLC is proportional to the molecular weight of

each species to the 3 2� power though the overall power incorporates both the molecular

weight and the solvent quality exponents.

���� ∝ ��

��� �

���� ���

���� (2.7)

We will therefore expect the calibration curve to have a power law dependence on the

molecular weight. The accessible porosity in equation (1.3) is that of only a single molecular

weight; this is the source of the SEC-HPLC’s ability to separate polymer molecules: each

molecular weight traverses a different volume.

2.2.2. HPLC Detectors

There are many types of HPLC detectors, each designed to work on different principles

and properties of the potential samples. Commercial detectors are available that use

Page 52: Photochemical Crosslinking Reactions in Polymers

31

absorbance, fluorescence, electrochemical properties, refractive index, conductivity, mass

spectrometry, FT-IR spectrometry, dynamic and static light scattering, and other properties of

samples. The HPLC used in these studies was equipped with three detectors: a refractive index

detector, an ultraviolet absorbance detector, and a static light scattering detector. The

information from each complements the other two to provide a complete picture of the sample

under investigation.

Refractive Index Detectors

A refractive index detector (RID) uses the differential refractive index between the

sample and a reference to determine concentration. It has a comparatively high required

detection concentration, but has the advantage of being universally applicable to almost every

species. Only substances that are isorefractive with the eluent cannot be detected. The basic

hardware configuration of a RID is shown in Figure 2.6.

Figure 2.6: Basic operation of a Refractive Index Detector. A beam of light passes through a cell containing both the reference and the sample. The difference in refractive index between the two leads to a change in the angle of the refracted light, from which the instrument calculates the difference in refractive index.

A light source and optical bench produce a beam of light, which enters the sample cell. The

sample cell is usually quartz and has two compartments separated by a quartz pane oriented at

45 degrees to the incident light. One compartment is filled with the pure eluent as a reference,

while the stream from the HPLC columns passes through the other. Snell’s law, �����(��)=

�����(��), says that the angles of incidence, ��, and refraction, ��, are related by the ratio of

Page 53: Photochemical Crosslinking Reactions in Polymers

32

the refractive indices, �� and ��, of the two materials at the boundary. Thus, for the RID’s cell,

equation (2.8) holds.

�����(45)= ����(45− �) (2.8)

For small angles, ����(45 − �)= ����(45)− ����(�)= ����(45)− �� �� and therefore

� − ����

=�

����(45)= �������������� (2.9)

where we have combined the instrument constants into a calibration term and made the

assumption that � ≈ � � at low concentrations. The differential refractive index increment,

����� =

����

�, is the change in the refractive index of a solvent with changes in analyte

concentration. The signal from the RID will therefore be related to the concentration, � the

differential refractive index of the sample, �� ��� , and the calibration constant by equation

(2.10) .

�� =��������������

��

��

��� (2.10)

So two constants are required to calculate the concentration: the instrument specific

calibration constant, ��������������, which must be experimentally determined, and the

species-specific differential refractive index increment, �� ��� , which also must be

experimentally determined, though there are tables of values for most common species. The

concentration measured in a RID is independent of molecular weight above a degree of

polymerization of around 20. A RID detector measures the number average distribution of the

chains passing through it (3).

Page 54: Photochemical Crosslinking Reactions in Polymers

33

Ultraviolet Absorption Detectors

An ultraviolet absorption detector (UVD) uses the differential absorbance of the sample

at a specific wavelength to determine concentration. It can detect species at much lower

concentrations than a refractive index detector, but requires the species to absorb light at

specific wavelengths. The basic hardware configuration of a UVD is shown in Figure 2.7.

Figure 2.7: Basic hardware of an Ultraviolet Absorption Detector. A beam of light passes through a monochromator to select the desired UV wavelength, and then into a beam splitter. One of the resulting beams passes through the sample flow cell and to a photodiode and the other beam passes through a reference (usually air) to another photodiode.

An UV light source produces a beam of light, which passes through a monochromator to isolate

the wavelength to be used for probing the sample. The wavelength is chosen to maximize the

absorbance of the sample. The monochromatic light then passes through a beam splitter or

other device such that one part passes through the sample and the other passes through a

reference that allows �� to be measured. The sample beam passes through the sample cell of

width �, with a sample � of concentration �� and extinction coefficient �� inside. The Beer-

Lambert law for species � passing through the sample cell at time �, equation (2.11), defines the

absorbance.

��,� = ����� ������ � = �����

(2.11)

Page 55: Photochemical Crosslinking Reactions in Polymers

34

� is an easily determined length scale unique to the model of instrument, and �� is a

wavelength dependent constant unique to the species. So a monochromic UVD (the Beer-

Lambert law is only valid for monochromatic light) requires only the value of �� at its

wavelength of operation to accurately measure the concentration. This assumes, of course,

that the species absorbs at the chosen wavelength. The concentration measured by a UVD is

the concentration of UV absorbing species. Their connectivity is not measured. Therefore a

UVD measures the number of chromophores of a species (3). In a polymer with one

chromophore per monomer, this is equivalent to the weight average distribution.

Static Light Scattering Detectors

Light scattering is the only technique that directly measures the properties of the

molecule in question and as such it is uniquely suited to determine the properties of a

polymeric sample. In a light scattering detector (LCD), an incident beam of light passes through

a sample cell with a very dilute polymeric solution and the Rayleigh scattering from a polymer

chain is detected by photodiodes. A dilute solution is very important; one must be able to say

with some certainty that the laser is impacting one polymer chain at a time (there are models

and techniques for light scattering in more concentrated solutions, but they will not be treated

herein). Assuming that this criterion is met, the incident light will be absorbed by the polymer

chain and be immediately scattered away at multiple angles. The scattering is detected by one

or more photodiodes, and information is then extracted through either static or dynamic

methods.

Page 56: Photochemical Crosslinking Reactions in Polymers

35

Figure 2.8: Basic hardware of the Multiangle Laser Light Scattering Detector. A laser passes through a polarizer to minimize out-of-plane scattering, and then into the scattering volume of the sample cell. Radially arranged photodiodes detect the scattered light as a function of angle, while a photodiode at zero degrees measures the light intensity.

In the LCD shown in Figure 2.8, a monochromatic laser beam passes through a polarizer to

insure the proper polarization for maximum scattering in the plane of the photodiodes. It then

enters the sample cell, where it interacts with a polymer molecule in the scattering volume at

the center of the sample cell and the secondary scattering is detected by the photodiodes. Two

basic principles make light scattering a powerful technique: the intensity of the scattered light

is directly proportional to the molar mass times the concentration, and the angular dependence

of the scattered light is directly proportional to the molecule size. This is expressed in the

Rayleigh-Gans-Debye equation, equation (2.12).

�(�)=�(�)���������

��= � ∗���(�)[1− 2�����(�)] (2.12)

�� is the incident light, �(�)��������� the scattered light with angle. �(�) is therefore the

excess Rayleigh ratio: the ratio of scattered and incident light, corrected for size and distance of

scattering volume. � is the molar mass; � the solute concentration. �� is the second virial

coefficient, which incorporates the polymer-solvent interaction. �(�) is the scattering

Page 57: Photochemical Crosslinking Reactions in Polymers

36

function, which relates the angular variation in intensity to the radius of gyration of the

polymer. And

� ∗ =4����

��������

����

with �� the refractive index of the solvent, �� Avogaro’s number, �� the wavelength of the

incident light in vacuum, and ��

�� the specific refractive index increment, a measure of the

change in refractive index of the solution with concentration. The molar mass measured by

the LSD is the weight average molecular weight because the degree of scattering is directly

related to the molecular weight (3). The Rayleigh-Gans-Debye equation (equation (2.12)) is

valid under two conditions:

�� ��� − 1�≪ 1

�(4����)

��� ��� ��� ������ 2� �≪ 1

the refractive indices of the polymer and solvent must be close; and the size of the polymer

must be much less than the wavelength of the light.

By removing all the instrument- and polymer-independent constants, we can simplify

equation (2.12) to highlight the polymer specific variables:

�(�)∝ �� ���

����

(2.13)

Thus according to equation (2.13), with any two of the molar mass, the concentration, or the

refractive index increment, the third is immediately available from a LSD. The refractive index

Page 58: Photochemical Crosslinking Reactions in Polymers

37

increment is very important for this instrument, but any errors in the measurement are

doubled in calculations from light scattering data. LSDs are therefore usually used in

conjunction with refractive index detectors because the ratio of light scattering data to the

refractive index data eliminates the error doubling and outputs the molar mass directly, as

shown in equation (2.14).

�(�)

��=

���� ��

��������������� (2.14)

2.2.3. Calibration of the HPLC Instrument

We have discussed the four constitutive equations of the SEC-HPLC instrument for

practical experimentation. Each equation includes instrument-specific constants that must be

determined and understood before accurate measurements can be taken. The SEC columns

separate based upon equation (2.7); the relationship between the molecular weight of the

molecule and the degree of separation has to be experimentally determined. Equation (2.11)

determines the detection of analytes in a UVD; it contains an instrumental length scale, �, and a

species-specific extinction coefficient, ��. The RID is built upon equation (2.10); it requires an

instrument-specific calibration constant, ��������������, and a species-specific refractive index

increment, ��

��. The LSD, whose constitutive equation is equation (2.12), requires merely the

species specific refractive index increment, ��

��. This section describes in detail the calibration of

these instruments.

Page 59: Photochemical Crosslinking Reactions in Polymers

38

Light Scattering Detector Calibration

The light scattering detector used in these studies was a Wyatt Technology Corporation

miniDAWN TREOS with attached COMET ultrasonic cleaner. It is a three angle LS detector that

operates at 657nm and is operable in both in-line and batch modes. The instrument was

calibrated first by injecting 0.02μm filtered HPLC Toluene and measuring the voltage of the 90°

photodiode with the laser on and off. The known refractive index and transmission properties

of toluene allow for the calibration of photodiode voltage to photon counts. The other two

photodiodes were calibrated from the first measurement by injecting an isotropic scatterer (10

kDa monodisperse Polystyrene) into the chamber and adjusting the gains on the photodiodes

until all the signals were the same. There is no angular dependence to the scattering from an

isotropic scatterer; therefore the detected photons should be exactly the same in all

photodiodes. With both the toluene and isotropic scatterer calibrations, the LS detector could

convert voltages to scattering intensity. The toluene calibration was done in batch mode; the

isotropic scatterer in in-line mode. The light scattering detector was recalibrated in this way

every six months.

SEC Column Calibration

The light scattering instrument provides, in theory and practice, a direct measurement

of the molecular weight of the samples and, if the molecules are large enough, a direct

measurement of the radius of gyration through the Rayleigh-Gans-Debye equation, equation

(2.12). The calibration of the SEC columns provides both a confirming measurement and a tool

that allows us to exploit the differences in measurement techniques to investigate more subtle

aspects of the sample distributions. The simplest way of calibrating an SEC instrument is to

Page 60: Photochemical Crosslinking Reactions in Polymers

39

inject a sample of known constitution and then measure where the peak appears. This

provides an accurate and easy method of linking the elution time of a sample to its molecular

weight.

The instrument was calibrated using a commercial SEC calibration package (Varian

EasiVIAL™) which contains twelve well characterized, low polydispersity polystyrene samples in

batches of four, so a single run provides four data points for the calibration curve. The

calibration samples were prepared and filtered through a 0.2μm Teflon filter to remove

particulates and large aggregates. The retention time at which the calibration sample peaks

appeared were compiled into a table and then fitted using a third or fourth order polynomial.

Originally a fourth order fit was used, but the difference between that and a third order fit was

so small that the calibration was switched to third order for ease of subsequent calculation.

The polystyrene calibration curves for the SEC columns over the course of this work are shown

below in Figure 2.9. The curves for both the UV and RI detectors are shown, with each plot

containing four curves taken at different times during the course of the experiments. The

major deviations in the calibration curve over long times occur for molecular weights under

1000Da, a range in which we are not interested. Irrespectively, the SEC columns were

calibrated before each new batch of samples.

Page 61: Photochemical Crosslinking Reactions in Polymers

40

Figure 2.9: SEC Column Calibration curves for both the UV (left) and RI (right) detectors. Each plot includes the calibration curves from four different times over the course of the experiments to show the continuity of the column calibration.

The calibrations to link elution time to molecular weight were performed on commercially

available polystyrene standards (Varian EasiVIAL™). We know from equation (2.7) that the

accessible volume of the column in SEC is related to both the number of monomers in the

chain, �, and the polymer-solvent interactions, β. � is independent of the polymer, but β is

system dependent. To extend the calibration to polymers other than polystyrene, we utilize

the viscosity average, which is the �-th moment of the distribution (as opposed to the first,

equation (1.13), or the second, equation (1.14)). The �-th moment, often referred to as the

viscosity average, is defined in equation (2.15).

�� = �∑ ����

����

∑ �����

���

(2.15)

Staudinger (4) found the measurement of the viscosity of a dilute polymer solution was

a simple method of determining the molecular weight of the polymer because their presence,

10 12 14 16 18 20 22 24 26 28 3010

0

101

102

103

104

105

106

Elution Time (Minutes post sample injection)

Poly

sty

rene S

tandar

d M

ole

cula

r W

eig

ht (D

a)

SEC Calibration Curve for UV Data

12/2009

01/2010

03/2010

08/2010

10 12 14 16 18 20 22 24 26 28 3010

0

101

102

103

104

105

106

Elution Time (Minutes post sample injection)P

oly

sty

rene S

tandar

d M

ole

cula

r W

eig

ht (D

a)

SEC Calibration Curve for RID Data

Page 62: Photochemical Crosslinking Reactions in Polymers

41

even in very dilute quantities, greatly increased the viscosity of the solution. This defined the

intrinsic viscosity, the theoretical viscosity of a polymer solution as the concentration dropped

to zero. This theoretical value was found to be linear in log-log space when the molecular

weight is plotted against the viscosity of the solution. This relationship, shown in equation

(2.16), is the Mark-Houwink-Sakurada equation, with � and � the Mark-Houwink-Sakurada

constants, which are valid for an individual polymer-solvent-temperature system.

[�]= ��� (2.16)

The Mark-Houwink-Sakurada equation provides a simple method of converting SEC calibration

curves from one polymer to another and relates the viscosity of a solution to the polymer

molecular weight. Two polymer solutions with the same intrinsic viscosity behave the same; if

each solution has a different polymer we can use equation (2.16) to calculate the molecular

weight of either if we know the molecular weight for one polymer and Mark-Houwink-Sakurada

constants for both. We therefore calibrate the SEC with well characterized polystyrene

standards and use Mark-Houwink-Sakurada constants to calculate the molecular weights of

eluted polymers.

Refractive Index Detector Calibration

The refractive index detector used was a Shimadzu RID-10A refractive index detector

with active temperature control and a broad spectrum light source. The refractive index

detector was calibrated by injecting a sample of known refractive index, �� ��� , and

concentration into the detector and measuring the output. Equation (2.9) then allowed a direct

calculation of the refractive index detector’s calibration constant.

Page 63: Photochemical Crosslinking Reactions in Polymers

42

Ultraviolet Light Detector Calibration

The UV detector used was a Shimadzu UV-10AV UV detector with deuterium lamp and

variable wavelength selection. The UV detector was not directly calibrated; equation (2.11)

shows that it is not necessary. The only variable in the Beer-Lambert law is the concentration

of the analyte; the comparison of the sample beam to the reference beam automatically

calibrates the scale of the UV detector.

2.3. Preparation and Characterization Techniques

This work investigates the crosslinking properties of additives in polymer thin films. Size

Exclusion Chromatography was the main technique used in this work. The SEC was used to

study the reaction products and samples, and Electron Paramagnetic Resonance was used to

probe the radicals present in solution analogs of the polymer thin films. Other techniques were

needed to prepare and characterize the samples. The preparation and characterization of the

thin films required both Spin Coating to produce the films and Ellipsometry to characterize

them.

2.3.1. Spin Coating

Spin coating is a simple technique used to produce uniform thin films on substrates

from liquid solutions. The film substrate is held on a rotating platform whose rotational period

can be precisely controlled. The sample solution is placed on the substrate while the platform

rotates. The power of the technique is the uniform and predictable film heights achieved

through this simple method (5). The relationship between film height, ℎ, spin speed, � ,

Page 64: Photochemical Crosslinking Reactions in Polymers

43

molecular weight, �, Mark-Houwink-Sakurada exponent, �, and the solution concentration, �,

are set out in equation (2.17).

ℎ∝ � �� �� ���� � (2.17)

This simple technique allows one to quickly produce high quality films of a set thickness.

Confirming the thickness requires ellipsometry.

2.3.2. Ellipsometry

An ellipsometer is an instrument used to study the dielectric properties, including

complex refractive index, of thin films. This study used a spectroscopic ellipsometer to measure

the thickness of the polymer films and to determine the optimal spin coating conditions for film

preparation. A spectroscopic ellipsometer (Figure 2.10) uses a broad band light source to

investigate the complex refractive index of a thin film. The light passes through an elliptical

polarizer, which polarizes the light into the plane of incidence of the sample. The interaction of

the light with the thin film modifies the polarization of the reflected light both parallel (�-

polarized) and perpendicular (�-polarized) to the plane, which passes through a second

polarizer (referred to as an analyzer) and into the detector.

Page 65: Photochemical Crosslinking Reactions in Polymers

44

Figure 2.10: Basic setup of a spectroscopic ellipsometer. A broadband light source is elliptically polarized and reflects off the sample on a substrate in the plane of incidence. The reflected light, which has been modified both parallel (�-polarized) and perpendicular (�-polarized) to the plane of incidence, passes through another polarizer (referred to as the analyzer) and into the detector.

The complex reflectance ratio, ρ, is measured and decomposed into the � and � components.

The ratio of the amplitudes of the � and � components is parameterized into an amplitude, Ψ,

and a phase shift, ∆, as seen in equation (2.18).

� =����

= ��Δ���(Ψ) (2.18)

These quantities cannot directly measure the film properties except in special cases. Instead, a

mathematical model of a surface and substrate is created, and the calculated amplitude and

phase shift is compared to the experimental one. Some understanding of the constituents of

the thin film is therefore required.

2.3.2.1. Characterization of the Thin Films

The spin coater and ellipsometer were used in conjunction to determine the

relationship between film thickness and spin speed. Films were prepared from sample

solutions by spin coating onto piranha and UVOCS cleaned silicon wafers from a 20mg/ml

polymer solution. The speed was varied between 500 and 5000rpm, for both Polystyrene and

Page 66: Photochemical Crosslinking Reactions in Polymers

45

Poly(n-butylacrylate) solutions. Three samples were made at each speed and their thicknesses

were averaged to produce Figure 2.11. Error bars are the sum of the three ellipsometer-

reported errors in the thickness measurement. The dotted line demonstrates that both

polymers follow the expected ℎ ∝ ����� relationship. A spin speed of 2000rpm was chosen for

all subsequent experiments. This corresponds to a film thickness of 128±6 nm for Polystyrene

and 110±29 for Poly(normal-butylacrylate).

Figure 2.11: Determination of the relationship between spin speed and film thickness for polystyrene and poly(n-butylacrylate) thin films spun coated from 20mg/ml solutions of toluene. Each data point is an average of three samples with error bars the sum of the ellipsometer-reported errors in the three measurements.

Dotted line is the � ∝ ��� �� relationship from equation (2.17) scaled to the PS data.

0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000 55000

50

100

150

200

250

300

Spin Speed (rpm)

Film

Th

ickn

ess

(n

m)

Relationship Between Film Thickness and Spin Coater Speedfor Polystyrene and Poly(n-butylacrylate)

Page 67: Photochemical Crosslinking Reactions in Polymers

46

2.4. References

1. in What Is HPLC (High Performance Liquid Chromatography)? (Waters Scientific (Waters.com)), vol. 2010.

2. T. Gu, Mathematical Modeling and Scale-up of Liquid Chromatography. (Springer, New York, NY, 1995), pp. 120.

3. S. Mori, H. G. Barth, Size Exclusion Chromatography. (Springer, New york, 1999). 4. H. Staudinger, W. Heuer, BERICHTE DER DEUTSCHEN CHEMISCHEN GESELLSCHAFT 63, (1930). 5. D. W. Schubert, T. Dunkel, Material Research Innovations 7, (2003).

Page 68: Photochemical Crosslinking Reactions in Polymers

47

3. Chapter 3: Rational Design and Preliminary Investigations

3.1. Introduction

This chapter discusses the polymers and crosslinkers chosen for these studies. Each is

discussed individually. The important values and properties of the compounds are outlined and

the rational for choosing them is explained. Preliminary experiments and calculations are then

performed to understand their properties and interactions in preparation for understanding the

results of the experiments.

The polymers and crosslinkers we chose to investigate the hydrogen abstraction reactions

are discussed below. In choosing the polymers we focused on the following criteria:

Their use in previous studies of similar reactions.

Their mobility in thin films. We wished to explore these reactions in both static

and dynamic binary systems; we therefore needed to explore polymers both

above and below their glass transition temperature.

The ease and location of the hydrogen abstraction

Their miscibility in HPLC-SEC eluent and with the chosen additives

Here we explore the polymers chosen and discuss their predicted responses to hydrogen

abstraction.

Page 69: Photochemical Crosslinking Reactions in Polymers

48

3.2. Polymer Selection

The polymers that were chosen are discussed here along with the rationale behind the

choice and their accessibility to hydrogen abstraction. The polymers chosen were polystyrene,

poly(∝-methylstyrene), polyisobutylene, polymethylmethacrylate, poly(normal-butylacrylate),

and 1,4 polybutadiene.

3.2.1. Polystyrene (PS)

Previous investigations by our group and others of bi-functional benzophenone species

in polymers demonstrated the inhibition of dewetting (1), the formation of gels (2) , and other

property changes including crosslinking (3). Polystyrene is a commodity polymer that has been

studied extensively. Its physical and thermodynamic properties are readily available. The

prevalence of information on PS and its previous use in similar studies insured its use in these

studies.

Polystyrene is shown in Figure 3.1. Polystyrene consists of an alkyl backbone chain with

a pendant phenyl ring every two carbons. The ideal atacticity of the polystyrene is assumed in

this discussion. PS is synthesized using living anionic polymerization (LAP) with a sec-butyl-

lithium initiator; the LAP synthesis of polystyrene is shown in Figure 3.1. Polystyrene is a glass

at experimental temperatures.

Figure 3.1: LAP synthesis of PS

nAnionic Polymerization

THF/-78oC

n

Page 70: Photochemical Crosslinking Reactions in Polymers

49

Polystyrene is photoactive at low wavelengths because of the alkenyl groups in the

phenyl ring, which absorb at around 200nm. The conjugation of the π bonds in the phenyl ring

lowers the energy required for a � → �∗ transition to ���� ≈ 254�� . This excited singlet

state can migrate along the polystyrene chain through the formation of eximer singlets with

neighboring phenyl groups (4). This has been used in “photon harvesting”, in which the chain is

designed to funnel the energy through energy migration to specific reaction centers on the

chain (5). The excited singlet state can also quench through fluorescence or internal conversion

or through homolytic bond cleavage to form radical pairs. This restricts the available excitation

wavelengths for additives to those where PS is not photoactive.

Hydrogen abstraction was proposed as the mechanism by which benzophenone (BP)

and PS react (1-3, 6). The hydrogen abstraction reaction between PS and BP is below in Figure

3.2. The UV photon excites the BP into the singlet state. There is an immediate intersystem

crossing (I.S.C) to the triplet, which can hydrogen abstract from PS if they interact within the

triplet lifetime.

Figure 3.2: PS radical generation through hydrogen abstraction by BP.

The hydrogen that is abstracted is determined by how stable the resulting radical will be

(Figure 1.1.2). The lower the energy of the resulting species the more stable the radical is. The

bond dissociation energy (BDE) of the abstracted hydrogen provides a quantitative measure for

the ease of abstraction and the stability of the resulting radical (7). Determining the BDEs of

O

hvO O OH

I.S.C +

Ph

n

H

Singlet Triplet

Phn+

Page 71: Photochemical Crosslinking Reactions in Polymers

50

the C-H bonds of PS provides allows us to predict both the ease of hydrogen abstraction and

the location, as lower energy bonds will be preferentially abstracted from. BDE tables are

readily available (8-10), but do not have the BDEs of every bond on every substance. We

therefore use the BDEs of representative compounds with similar orbital configurations to the

compound we’re interested in. Table 3.1 shows the structure of PS with C-H bonds numbered

in increasing BDE, the representative compound for the numbered bond, and the BDE of that

bond. The specific C-H bond in the representative compound is bolded.

Table 3.1: Structure and hydrogen bond dissociation energies of polystyrene.

Polystyrene X-H Bond Number

Representative Compound

X-H BDE (kJ/mol) (8)

1

348.1

2

410.2

3

474.8

The PS BDEs indicate that the C-H bond on the tertiary backbone carbon is the most likely

hydrogen abstraction location. The second lowest BDE, that of the secondary carbon on the

backbone, has a 60 kJ/mol higher energy barrier. The tertiary carbon is where the hydrogen

abstraction will take place in PS. The radicals formed in PS can only be backbone radicals. This

does not take steric considerations into account.

21

3

n

CH

H2C

CH

Page 72: Photochemical Crosslinking Reactions in Polymers

51

3.2.2. Poly(�-methyl styrene) (PamS)

Polystyrene was chosen because it has an extensive literature and a history in similar

studies. BDEs predict that the majority of radicals will form on the tertiary carbon. We isolate

the secondary carbon radical and investigate it using a derivative of polystyrene: poly(∝-

methylstyrene), shown below in Figure 3.3.

PamS has the same basic structure as polystyrene, except there is a methyl group

attached to the alpha carbon where the tertiary radical forms in PS. The tertiary radical cannot,

therefore, form. The ideal atacticity of the PamS is assumed in this discussion. The synthesis of

PamS is the same as PS synthesis (Figure 3.3). PamS is a glass at experimental temperatures.

Figure 3.3: LAP synthesis of PamS.

Without the ability to hydrogen abstract from the tertiary carbon, PamS radicals are

theoretically both more difficult to form and in different locations. Table 3.2 quantifies the

potential abstraction locations and their BDEs. Where before the tertiary carbon had a 60

kJ/mol advantage in abstraction, in PamS the BDEs are equal. Both the secondary and primary

carbons can abstract with equal probability (again without considering steric effects). We

would therefore expect to see radicals in either location, and the radicals may migrate along

the chain by translating from bond number 2 to 1 to the subsequent monomer (11).

nAnionic Polymerization

THF/-78oC

n

Page 73: Photochemical Crosslinking Reactions in Polymers

52

Table 3.2: Structure and hydrogen bond dissociation energies of poly(∝-methylstyrene).

Poly(∝-methylstyrene) X-H Bond Number

Representative Compound

X-H BDE (kJ/mol) (8)

1

410.2

2

412.6

3

474.8

The PamS BDEs indicate that the primary and secondary C-H bonds are the most likely

hydrogen abstraction locations and the abstraction itself is more unlikely than in PS because

the BDE are 60 kJ/mol higher.

3.2.3. Polyisobutylene (PIB)

Polystyrene was used to investigate the tertiary radical; poly(∝-methyl styrene) was

used to investigate the secondary radical; and polyisobutylene was used to investigate the

secondary radical without the stabilization effect and steric hindrance of the phenyl group. The

phenyl group provides some stabilization to the β-carbon radical in polystyrene and Poly(∝-

methyl styrene) (11-13), so polyisobutylene, shown below in Figure 3.4, was used to investigate

radical formation and reactions without the stabilization and steric effects of the phenyl group.

Structurally, PIB is the same as PamS but with the phenyl group replaced with another

methyl group. We assume ideal atacticity for this discussion. It is synthesized through living

1

3

n

2

H2C

CH3H3C

CH3

CH

Page 74: Photochemical Crosslinking Reactions in Polymers

53

cationic polymerization (LCP); its synthesis is shown in Figure 3.4. PIB is a rubber at

experimental temperatures.

Figure 3.4: LCP of polyisobutylene.

Without the phenyl ring, we expect the radical stability to be much less than that of PS

and less than that of PamS. There are two possible locations for a radical on PIB: the primary

and secondary carbons. Table 3.3 lists the BDEs of the primary and secondary carbons; they are

close and very similar to the PamS BDEs. The secondary carbon is more preferred in PIB than in

PamS, but the 9 kJ/mol BDE difference is not significant enough to predict primarily secondary

radicals after hydrogen abstraction.

Table 3.3: Structure and hydrogen bond dissociation energies of polyisobutylene.

Polyisobutylene X-H Bond Number

Representative Compounds

X-H BDE (kJ/mol) (8)

1

410.2

2

419.7

The PIB BDEs indicate that the probability of hydrogen abstraction in PIB is about equal to that

in PamS for energetic reasons; there is no phenyl group to hinder the approach of the

abstracting molecule so we expect a higher probability due to steric considerations.

nTiCl4/RCl or (Et)2AlCl

(CH3)3CH2(CH3)2Cl

TiCl5

nCH3OH

nNH4OH

1

2

n

2

H2C

CH3

CH3

H3C

CH3

Page 75: Photochemical Crosslinking Reactions in Polymers

54

3.2.4. Polymethylmethacrylate (PMMA)

We used the previously discussed polymers to investigate the consequences of methyl

groups and phenyl groups on the reaction; we can use other polymers to investigate the effects

of alternate pendant groups. Poly(∝-methyl styrene) investigated the effects of the methyl

group in conjunction with the stabilization effects of the phenyl ring. Isobutylene investigated

the effects of the pure methyl group. We can use polymethylmethacrylate, shown in Figure

3.5, to investigate the effects of a new, more polar and chemically reactive pendant group on

hydrogen abstraction and radical reactions.

PMMA introduces an ester group with a pendant methyl in the place of one of the PIB

methyl groups. We assume ideal atacticity for this discussion. It is synthesized through ATRP;

the synthesis is in Figure 3.5. PMMA is a glass at experimental temperatures.

Figure 3.5: The ATRP synthesis of polymethylmethacrylate.

The ester group introduces new locations for hydrogen abstraction to occur. In PS,

PamS and PIB, the favored reaction locations were on backbone carbons. With PMMA, the

favored reaction location is the methyl pendant to the ester. The bonds and BDEs are shown in

Table 3.4. The BDEs of all three abstractable hydrogens are similar, with a 4 kJ/mol advantage

to the pendant methyl. The backbone methyl is about the same energy in the other direction

from the secondary carbon.

O

O

n

O

O

nCuBr

Page 76: Photochemical Crosslinking Reactions in Polymers

55

Table 3.4: Structure and hydrogen bond dissociation energies of polymethylmethacrylate.

Polymethylmethacrylate X-H Bond Number

Representative Compound

X-H BDE (kJ/mol) (8)

1

406.3

2

410.2

3

415.2

The PMMA BDEs indicate that the probability of hydrogen abstraction in PMMA is about equal

to that in PamS and PIB without taking sterics into account. Sterically, PMMA will have an

advantage of PamS. The polar carbonyl may have an effect as well on the abstraction

probabilities.

3.2.5. Poly(normal-butylacrylate) (PnBA)

The polymethylmethacrylate introduces new chemistries and new chemical centers. It

opens up the possibilities of pendant group radicals and reactions. All but one of the carbon

radical centers, though, are tertiary carbons (the sole secondary carbon is on the backbone).

We introduce pendant secondary carbons to the system using poly(normal-butylacrylate),

Figure 3.6.

PnBA replaces the methyl group pendent to the ester in PMMA with an n-butyl group,

introducing three secondary carbons between the ester and the methyl. It also removes the

methyl group on the PMMA backbone, reintroducing the tertiary carbon to the system. We

2

n

3

OO

1

OOH3C

H2C

CH3H3C

CH3

OHO

Page 77: Photochemical Crosslinking Reactions in Polymers

56

assume ideal atacticity for this discussion. It is synthesized through LAP; the synthesis is in

Figure 3.5. PnBA is a rubber at experimental temperatures.

Figure 3.6: LAP of poly(normal-butylacrylate).

PnBA reintroduces the tertiary carbon to potential hydrogen abstraction and introduces

secondary carbons on the pendant group. The reintroduction of the tertiary carbon makes it

the favored location of hydrogen abstraction. The bonds and BDEs of PnBA are in Table 3.5.

The tertiary backbone carbon has the lowest BDE, followed by the pendant secondary carbons.

Both have BDEs below 400 kJ/mol, and they are within on 4 kJ/mol of each other. We therefore

expect equal reactivity from both, though the pendant secondary carbons are favored for steric

reasons (7). The other possible locations for abstraction have approximately the same BDE as

PMMA, PIB, and PamS.

O

O

n

O

O

n1) RLi/Adduct initiator

2) Termination with CH3OH

Transesterification

BuOH/acid catalystO

O

n

Page 78: Photochemical Crosslinking Reactions in Polymers

57

Table 3.5: Structure and hydrogen bond dissociation energies of poly(normal-butylacrylate).

Poly(normal-butylacrylate) X-H Bond Number

Representative Compounds

X-H BDE (kJ/mol) (8)

1

388.3

2

392.1

3

410.2

4

411.1

5

421.3

The PnBA BDEs indicate that PnBA has a higher probability of hydrogen abstraction than all

other polymers other than PS, though the advantage is only 28 kJ/mol at best. The n-butyl

pendant group has been shown to be the preferred hydrogen abstraction site (7) because of its

high mobility. We would therefore expect little radical-induced scission in the backbone. The

radicals formed in PnBA can only form on the pendant side chain.

3.2.6. 1,4-polybutadiene (PBD)

All the previous polymers discussed allow us to elucidate the effects of various pendent

groups and radical formation locations on alkyl chain polymers. The radicals formed on carbon

centers on the backbone or pendant groups and the radical reactions extended from there. The

alkyl backbone, though, is one of many possible foundations on which a polymer can be

created. There are many inorganic polymers, such as polysiloxanes or polystannanes, in which

carbon is not present in the backbone. These have unique and interesting chemistries and

31

n

OO2

4

4

5

CH

OHO

OO

CH2

H2C

H2C

H3CCH3

Page 79: Photochemical Crosslinking Reactions in Polymers

58

properties that are beyond the scope of this text. One of the relatively simplest non-alkyl chain

backbones is a vinyl backbone. A vinyl backbone consists of single carbon-carbon bonds

alternating with double carbon-carbon bonds. Its relative simplicity comes out of the pure

carbon backbone; its complexity from the double bonds. We are investigating the effects of

functional groups and their placement on the reactions of the photocrosslinkers under

investigation. We can use one of the simplest vinyl polymers, 1,4-polybutadiene, shown in

Figure 3.7, to investigate the interactions of hydrogen abstractors with backbone double bonds.

PBD removes all pendant groups and forms a chain of pure secondary carbons. While

there are many potential reactions due to the double bond, we want to see if hydrogen

abstraction is feasible. Its LAP is shown in Figure 3.7. PBD is a rubber at experimental

temperatures.

Figure 3.7: LAP of 1,4 polybutadiene.

PBD consists solely of secondary carbons connected by double bonds. The double bond

increases the secondary carbon BDE by 17 kJ/mol (Table 3.6) over its single bonded

counterpart. We expect the least probability of hydrogen abstraction from PBD, and there is

the possibility of alternate reactions due to the double bond.

nCyclohexane or

Toluene

RLin

Page 80: Photochemical Crosslinking Reactions in Polymers

59

Table 3.6: Structure and hydrogen bond dissociation energies of poly(1,4 butadiene).

Poly(1,4 butadiene) X-H Bond Number

Representative Compounds

X-H BDE (kJ/mol) (8)

1

427.7

PBD will most likely not hydrogen abstract. Other work has shown that benzophenone, for

example, can only hydrogen abstract from bonds with BDEs of 427 kJ/mol and less (7).

3.2.7. Conclusion

Each chosen polymer explores a different aspect of the hydrogen abstraction reaction.

PS explores the backbone tertiary carbon, PamS blocks it off and forces secondary and primary

carbon abstraction, PIB is PamS without the steric interactions of the phenyl group, PMMA

introduces a polar pendant group which PnBA makes the preferred abstraction location, and

PBD explores backbone double bonds and secondary carbons. From the discussion above we

expect PS and PnBA to react the most, as they have the lowest BDEs. The PS reaction location

is on the backbone; unreacted backbone radicals may cause chain scission. PnBA, though,

preferentially forms the radical on the side chain, limiting chain scission. PS and PnBA are

glassy and rubbery, respectively, allowing the exploration of chain dynamics in systems that

highly favor hydrogen abstraction.

3.3. Crosslinker Selection

The crosslinkers that were chosen are discussed here along with the rationale behind their

choice and their primary crosslinking mechanism. The crosslinkers chosen were derivatives of

benzophenone, xanthone, phthalimide, and phenylazide. All bi-functional crosslinkers were

1

1 n

CHHC

Page 81: Photochemical Crosslinking Reactions in Polymers

60

synthesized by Jeffrey Lancaster after the methods of (7, 14). Crosslinker names used here are

not chemically rigorous; they are used for convenience. Structures are provided to make

rigorous names unnecessary.

3.3.1. Benzophenone (BP)

This thesis was initially motivated by previous investigation into the inhibition of dewetting

in polystyrene thin films by bi-functional BP species (14). Bi-functional BP species were being

studied as well in polymer degradation (3), and BP itself was used as a crosslinker as a pendant

functional group (15, 16) and as a photosensitizer (17, 18). BP has been extensively studied

both in polymeric systems and alone (4, 18, 19) and its physical and chemical properties are

readily available. The prevalence of information about BP and its previous use in similar studies

insured its use in these studies.

Benzophenone and the bi-functional benzophenone (BP-BP) used in these studies are

shown in Figure 3.8. The hydrogen abstraction mechanism of BP was previously discussed in

Figure 3.2. UV photons of 350nm are absorbed by the ketone, promoting electrons to the

singlet state. An immediate intersystem crossing to the triplet state occurs with a quantum

yield of unity (7, 11, 20). The two excited electrons abstract a hydrogen from the polymer,

forming a radical on the BP and one on the polymer. These radicals then participate in radical

reactions that in some cases lead to crosslinking or scission of the polymer chain.

Page 82: Photochemical Crosslinking Reactions in Polymers

61

Figure 3.8: Benzophenone (left) and the bi-functional benzophenone (right) used in these studies.

The BP triplet has an energy of 287 kJ/mol and a lifetime of 6.9 μs (20). The hydrogen

abstraction reaction must be exothermic to spontaneously occur (8); the energy of the excited

state and exiting bonds must be more than that of resulting bond, equation (2.16).

∆���� = ��������� − ��������� < 0

∆���� = [�({��}‡)+ ������]− [������ + ������]< 0 (3.1)

For ∆���� < 0 , with the BP triplet energy of 287 kJ/mol, the bond dissociation energy (BDE) of

the abstractable group ������ ≤ 481 ��

���� . The hydrogen abstraction step must also be

exothermic (11), so the upper limit of the bond energy is reduced to approximately 427 kJ/mol

(7) in benzophenone.

Table 3.7: Bond dissociation energies of Benzophenone for Equation (1.10).

�({��}‡) (20) ������ (21) ������ (22)

Benzophenone 287 442.7 (from Benzyl Alcohol)

325.1 (from Ethylbenzene)

The upper limit on the bond energy required for hydrogen abstraction in BP systems is

427 kJ/mol. The polymers under investigation must therefore have C-H bond energies below

this value for abstraction to take place; the lower the value the more likely the reaction. The bi-

functional BP species will have similar requirements; the attached ester chains will modify the

energetic of the molecule. A para-vinyl group reduces ������ by 30 kJ/mol, a meta-methyl

OO

O

OO

O

O

Page 83: Photochemical Crosslinking Reactions in Polymers

62

reduces ������ by 17 kJ/mol before the rate determining step is incorporated. Hydrogen

abstraction will therefore be less favored in bi-functional BP than in mono-functional.

3.3.2. Xanthone (XAN)

Benzophenone was chosen because its properties are well understood and it was previously

used in similar studies. Xanthone was chosen because it is structurally similar to benzophenone

and reacts in the same way but has a shorter lifetime and is slightly more polar. We study the

hydrogen abstraction reaction with a shorter lived species that better dissolves in polar

polymers such as PMMA and PnBA.

The mono-functional carboxyl-xanthone and its bi-functional species (XAN-XAN) are

shown in Figure 3.9. Its hydrogen abstraction mechanism is the same as that of BP; the triplet

ketone induces hydrogen abstraction. Its absorbance around 354nm is higher than that of BP,

generating more triplets, each of which has a lifetime of only 0.02μs in non-polar solvents and a

lifetime of 17.9 μs in polar solvents (20). It should preferentially crosslink more polar polymers.

Figure 3.9: The mono-functional xanthone (left) and the bi-functional xanthone (right) used in these studies.

XAN has a triplet energy of 310 kJ/mol; ������ ≤ 458 ��

���� from equation (2.16).

The ������ for hydrogen abstraction in XAN-XAN is lowered by 30 kJ/mol by a para-vinyl

group reduces and by 17 kJ/mol by a meta-methyl before the rate determining step is

O

O

O

O

O

OOO

O

O

OH

O

Page 84: Photochemical Crosslinking Reactions in Polymers

63

incorporated. Hydrogen abstraction will be less favored in XAN-XAN than in XAN, and more

favored in slightly polar polymers than non-polar ones. The yield of triplets will be higher than

in BP.

3.3.3. Phthalimide (PTH)

Benzophenone was chosen because of its previous uses in similar systems; Xanthone was

chosen because its hydrogen abstraction reaction is the same as BP but is favored in more polar

polymers. Both BP and XAN react in the triplet state. Phthalimide and its bi-functional form

were chosen because they react from the singlet state and triplet state (23, 24) and there

should only be a reaction in mono-functional systems.

The structure of phthalimide and its bi-functional species (PTH-PTH) are shown in Figure

3.10. The structure of PTH and its hydrogen abstracting reaction are very different from those

of BP and XAN.

Figure 3.10: Phthalimide (left) and the bi-functional phthalimide (right) used in these studies.

Phthalimide hydrogen abstracts, but only after a single electron transfer (SET) reaction (23-25)

that it can self-catalyze (26). The reaction is shown in Figure 3.11. A proton acceptor is

required for the first step, after which PTH is excited to the singlet state and reacts with an

unreacted PTH to generate a polymer radical and a dead PTH.

NH

O

O

N

O

O

N

O

O

Page 85: Photochemical Crosslinking Reactions in Polymers

64

Figure 3.11: Radical generation through hydrogen abstraction after a SET in phthalimide(26)

BP and XAN generate radicals on polymers and themselves at the same time; they may react

with the polymers themselves, eliminating the radicals and covalently bonding to the polymer.

This is a mechanism by which BP-BP forms a crosslink; in BP this mechanism permanently

removes the BP from the system. The generation of macroradicals by phthalimide removes the

PTH radical and prevents covalent bonding with the polymer. The density of radicals on the

polymer should therefore be higher with PTH than BP or XAN, assuming the same density of

photoexcitation. The alkyl chain in PTH-PTH prevents intermolecular reactions; it can react with

itself (23, 24). We therefore expect phthalimide to react in polymers with hydrogen acceptors

and only as a mono-functional species. In alkenes, it is not expected to photo-generate radicals

as it undergoes a cycloaddition reaction with alkenes (24).

3.3.4. Phenylazide (FEN)

Phenylazide was chosen because it reacts using a different mechanism from BP, XAN, and

PTH and only the bi-functional form can crosslink. FEN reacts through nitrene insertion. The

reaction is not as dependent upon BDEs and should be a more universal crosslinker in bi-

NH

O

O

N-

O

O

hvN-

O

O

+ NH

O

O

SET

-N

O

O

+N

O

O

-

RHHN

O

O

+R

Page 86: Photochemical Crosslinking Reactions in Polymers

65

functional form. It has no mechanism for radical formation and therefore no mechanism for

mono-functional crosslinking.

Phenylazide and its bi-functional analog (FEN-FEN) are shown in Figure 3.12. The aryl

azide group is the location of the reaction in both FEN and FEN-FEN. The nitrene insertion

reaction is diagrammed in Figure 3.13.

Figure 3.12: Phenylazide (left) and the bi-functional form (right) used in these studies.

In the nitrene insertion reaction, the phenylazide photogenerates nitrogen gas and a nitrene.

The nitrene inserts itself into a nearby C-H bond. This reaction does not generate macroradicals

and therefore only FEN-FEN can form crosslinks through a covalent bridging mechanism.

Figure 3.13: Nitrene insertion reaction between a phenylazide and a C-H bond.

FEN is used to investigate the formation of crosslinks through bi-functional methods exclusively,

using a different mechanism from BP, XAN, and PTH.

3.3.5. Conclusion

Mono-functional and bi-functional benzophenone, xanthone, phthalimide, and phenylazide

were the photochemical additive. BP and XAN react through hydrogen abstraction. BP has a

long triplet lifetime compared to XAN, but a lower absorption coefficient. XAN is also slightly

N3

OH

O

OO

N3

N3O

O

NN+

N-

hvR NR

R1HNR

R1

H

Page 87: Photochemical Crosslinking Reactions in Polymers

66

more polar than BP. PTH has very specific reactions that catalyze radical generation on

polymer, but only when mono-functional and the system can donate a proton. Phenylazide

reacts with polymers through nitrene insertion. It can only crosslink when bi-functional, and

does not generate radicals. Scission, macroradical recombination, and other radical side

reactions are therefore impossible and a pure understanding of crosslinking kinetics and

structures can be achieved.

3.4. Differential Refractive Index

As discussed in the section on light scattering detectors, the polymer’s refractive index

increment is an important requirement for determining the molecular weight of a sample from

light scattering data. The refractive index increment, dn/dc, measures the change in refractive

index of a polymer solution with concentration. The dn/dc allows the light scattering detector

to eliminate changes in detector intensity due to increased refraction and allows the calculation

of concentration from RID data. Polymer dn/dc values are available in literature (27-30), but

the reported values can vary widely for common polymer and be difficult to find for less

common ones. The dc/dc is dependent upon the specific polymer, solvent, wavelength, and

temperature combination being investigated. We therefore independently measured the

dn/dc values of the polymers used in this study in the conditions under which they will be

analyzed to insure quality and applicability. dn/dc was measured in two ways: the Shimadzu

RID-10A broad spectrum refractive index detector used in all studies in batch mode, and a

Brookhaven Instruments differential refractometer (Bi-DNDCW) at 535nm. The experiments

Page 88: Photochemical Crosslinking Reactions in Polymers

67

were performed separately from the same source solutions. Two different series of polymer

solutions were made, each with at least five different concentrations (Table 3.8).

3.4.1. Experimental

Polystyrene (PS) (Mn=46000, PDI=1.03, Waters Associates), poly(normal-butyl acrylate)

(PnBA) (Mn=60000 PDI=1.08, Polymer Source, Inc.), poly(∝-methyl-styrene) (PamS)

(Mn=32300, PDI=1.02, Polymer Source, Inc.), poly(1,4 butadiene) (PBD) (Mn=37500, PDI=1.03,

Polymer Source, Inc.), polyisobutylene (PIB) (Mn=47000, PDI=1.3, Polymer Source, Inc.),

carboxy-terminated polymethylmethacrylate (PMMA) (Mn=34300, PDI=1.06, Polymer Source,

Inc.), Potassium Chloride (KCl) (Sigma-Aldrich) and HPLC-grade Tetrahydrofuran (THF)

(Pharmco-Aaper) were used as received.

Two 6mL of 25mg/ml polymer in THF solutions were prepared for each polymer and two

6mL of 25mg/ml KCL in DI H2O solutions were prepared as a control. All solutions were shaken

for 30 minutes and then filtered through 0.2μm Teflon syringe filters in a clean hood to remove

large impurities. KCL and polymer solutions were then split into separate vials and diluted to

the desired concentrations. Table 3.8 shows the solution concentrations. Solutions were

0.2μm filtered again after preparation.

Table 3.8: Control and polymer solution concentrations used in differential refractive index experiments.

KCl Concentration (mg/ml) 25 12.5 5 1 0.5 0.1

Polymer Concentration (mg/ml) 15 10 5 1 0.5 N/A

Page 89: Photochemical Crosslinking Reactions in Polymers

68

Solutions were injected into a temperature controlled differential refractometer

(535nm, Bi-DNDCW, Brookhaven Instruments) (stabilized at 40oC) and the temperature

controlled refractive index detector (RID-10A, Shimadzu Scientific) (stabilized at 40oC) in batch

mode. Calibration was performed after cleaning the flow cells with THF, HPLC toluene,

Acetone, HPLC methanol, and DI H2O in that order to transition from organic solvents to polar.

After calibration, cleaning was performed in the opposite solvent order to transition back to

THF. Samples were injected in low-to-high concentration order, to minimize inaccuracies due

to residual polymer, and 20mL of HPLC THF was injected through between polymers to insure a

clean flow cell.

3.4.2. Results

The differential refractive indices of the studied polymers were measured to insure the

quality and applicability of the MW measurements in this study. They were measured both in a

differential refractometer and in the HPLC-SEC RID detector used to study all the samples in this

study.

Figure 3.14: Representative dc/dc plots from the RID. A) PamS in THF at 40oC B) PMMA in THF at 40oC.

0 1 2 3 4 5 6 7 8 9 100

200

400

600

800

1000

1200

1400

1600

1800

2000

Polymer Concentration (mg/mL)

RI

Diff

ere

nce

be

twe

en

Re

fere

nce

and

Sa

mp

le C

ell

0 1 2 3 4 5 6 7 8 9 100

100

200

300

400

500

600

700

800

900

Polymer Concentration (mg/mL)

RI

Diff

ere

nce

betw

ee

n R

efe

ren

ce a

nd

Sam

ple

Cell

Page 90: Photochemical Crosslinking Reactions in Polymers

69

Figure 3.14 shows two representative dn/dc plots from the RID detector. As the

concentration increases from zero to ten mg/ml, the refractive index difference between the

reference and sample cells in the RID increases. The plot on the left shows the increase in the

PamS system with the fit line whose slope is the dn/dc. Detector saturation can be seen in the

4, 5, and 10 mg/ml data; these are not included in the fit. The plot on the right shows the

increase in the PMMA system with the dn/dc fit line. The response is much less; saturation only

appears in the 10mg/ml solution. All samples provided at least five data points before

saturation occurred.

Plots similar to Figure 3.14 were produced for all polymers and both the RID detector

and the differential refractometer. The calculated dn/dc for the polymers in both instruments

are tabulated in Table 3.9 and compared with literature values.

Table 3.9: Differential refractive index increment of polymers used in this study

Polymer (in THF at 40oC) dn/dc (mL/g) λ=535nm

dn/dc (mL/g) λ=broad

spectrum

Representative Literature Value (mL/g)

Polystyrene 0.212 0.177 0.193 (546nm, 25oC) (29)

Poly(α-methylstyrene) 0.219 0.191 0.2044 (633nm, 25oC) (28)

Polyisobutylene 0.123 0.124 0.1131 (633nm,27oC) (31)

Poly(1,4 butadiene) 0.136 0.138 0.132 (546nm, 25oC) (29)

Poly(methylmethacrylate) 0.092 0.084 0.087 (633nm, 25oC) (29)

Poly(n-butyl-acrylate) 0.069 0.065 0.0651 (546nm, 30oC) (29)

The calculated values agree within variations due to wavelength dependence and temperature.

The broad spectrum source in the refractive index detector prevents direct measurements of

Page 91: Photochemical Crosslinking Reactions in Polymers

70

the dn/dc at the LS detector’s wavelengths, which increases the error in the molecular weight

as measured by the LS detector. A fixed wavelength source at the same wavelength as the LS

detector would be optimal; this was not available. Nevertheless we used the dn/dc values

measured by the broad spectrum refractive index detector in molecular weight calculations

because the LS detector uses the data from the RID for many of its calculations.

3.5. Mark-Houwink-Sakurada Constants

The Mark-Houwink-Sakurada equation (equation (2.16)) was discussed in section 2.2.3.

It uses the intrinsic viscosity to relate the molecular weights of different polymers. The Mark-

Houwink-Sakurada constants, � and �, are valid for an individual polymer-solvent-temperature

system.

The MHS constants used in these studies are tabulated in Table 3.10 along with their

sources. The difficulty in measuring the intrinsic viscosity of a material in a rigorous way,

combined with the constants’ dependence on temperature, solvent, and to some extent

molecular weight produces a wide range of literature values (19, 29, 31, 34, 35) for the same

polymers in the same conditions. The values used must be self-consistent. MHS constants

were used only when literature sources listed both the PS and other values.

Table 3.10 shows the MHS constants used in this study; the first listed PS values were

used in conjunction with the PnBA values, while the second listed PS values were used for all

other polymers. This maintained the source fidelity to insure proper relative conversion.

Page 92: Photochemical Crosslinking Reactions in Polymers

71

Table 3.10: Mark-Houwink-Sakurada constants of the polymers in THF.

Polymer Temp (oC) K (x10^3) ml/g a Source

PS 30

30

11.4

8.63

0.716

0.736

(19)

(35)

PamS 30 11.1 0.69 (35)

PIB 30 26.6 0.654 (35)

PMMA 30 8.97 0.71 (35)

PnBA 30 12.2 0.7 (19)

PBD 30 25.2 0.727 (35)

3.6. Solubility and Interaction Parameter.

Our systems are spun coated from a solution above C* into a non-equilibrium thin film.

The polymer chains interpenetrate and we have assumed that the additives are evenly

distributed throughout the thin film and in the solution and there are no preferred inter- or

intra- molecule interactions. The non-equilibrium thin film forces the polymer and additive

together initially, but the segmental motion in glassy systems and reputation in rubbery

systems allow limited motion that may lead to local phase separation. If separate additive and

polymer phases form over time, then the probability of local radical formation drops rapidly, as

most additives will be unable to react with polymer. We can gain insight into this by calculating

the solubility of the additives, polymers, and solvents. As discussed previously in section 1.3.1,

the Gibbs free energy of mixing a polymer molecule into solvent (for a binary mixture only) as

derived in (33) is shown in equations (3.2) and (3.3).

Page 93: Photochemical Crosslinking Reactions in Polymers

72

∆�� = ��[����(��)+ ����(��)+ �������] (3.2)

��� =(�� − ��)

�����

(3.3)

Where R is the gas constant, T the absolute temperature, �� and �� the number of moles and

volume fraction of species �, �� and �� the Hildebrand solubility parameter of the polymer and

solvent, respectively, and ��is the molar volume of a polymer monomer. The quality of mixing

between the two species is dependent upon their Hildebrand solubility parameters (Equation

(3.3)). The Hildebrand solubility parameter, the square root of the cohesive energy density, is a

parameter that was found to quantify the order of solubility of a solute in solvents (36). The

cohesive energy density is the amount of energy required to move a unit volume of molecules

from the bulk to infinity. The Hildebrand solubility parameter is therefore a measure of the

internal pressure of a molecule; molecules with similar solubility parameters will, in general, be

miscible. By comparing the solubility parameters of the polymers and additives we can

determine whether they will preferentially mix.

We use group contribution techniques to determine the solubility parameters of the

polymers and additives under investigation. Group contribution is a technique that is used to

calculate physical and thermodynamic properties from structure information in the absence of

experimental data (37). It is used most often in molecular design applications to provide first

order approximations for properties that are not readily available in literature (38). Group

contribution posits that physical and thermodynamic properties can be estimated by breaking

the compound of interest into specific atomic groups which contribute additively to the

property of interest. Group contribution is used in molecular design (39) to understand surface

Page 94: Photochemical Crosslinking Reactions in Polymers

73

segregation (40-43), interpolymer surface tension in polymer blends (44), and the transport

properties of crosslinked rubbers (45). There are many different group contribution methods

(36-38), each with separate group definitions and approaches. In this study, we use the

methods of Sheldon and Constantinou (37, 38, 46) and the DIPPR structure properties database

(47). The Hildebrand solubility parameter, in this group contribution system, is defined in

Equation (3.4).

� = ����� − ��

���

���

(3.4)

where ���� is the enthalpy of vaporization, � the gas constant, � the temperature, and �� the

molar volume. Equations (3.5) and (3.6) are those for ���� and ��.

���� − ℎ�� = ��ℎ����

+ ��ℎ����

(3.5)

�� − ��� = �������

+ �������

(3.6)

In these equations, � and � are the number of groups of each type, ℎ��� and ���� are the

contribution of the first order groups to the enthalpy and volume, respectively, ℎ��� and ����

are the contribution of the second order groups to the same, and ℎ�� and ��� are method

specific adjustable parameters. For the group contribution method used here, ℎ�� =

6.829 �� ���� (37) and ��� = 0.1211 ��

����� (46). In this method, the primary groups are

atoms or small functional groups and the secondary groups account for larger structures. The

structure of styrene (Figure 3.15) includes three primary groups (five aromatic C-H bonds

(atoms 1-5); an aromatic C (atom 6); one CH2=CH (atoms 7 and 8)) and one secondary group (a

Page 95: Photochemical Crosslinking Reactions in Polymers

74

6-membered ring, atoms 1-6) while the structure of polystyrene (Figure 3.15) is the same

except the CH2=CH primary group is replaced by individual CH2 and CH groups (atoms 7 and 8).

Figure 3.15: Structures of styrene and polystyrene

The Hildebrand Solubility parameters were calculated and compared to literature values

when available. Table 3.11 includes the solubility parameters of the solvents used in our

systems, Table 3.12 compares the solubility parameters of the polymers to their monomer, and

Table 3.13 compares the solubility parameters of the mono-functional additives to their bi-

functional forms. Monomer solubility was calculated as a validation tool; accurate values

within narrow ranges are available in literature while polymer solubility depends upon factors

including molecular weight and temperature which widen the range of reported solubility

parameters. The average absolute percent error (38) for the substances which we have

literature values for is 3.15±2.74%, which is comparable to the database error of 3.96±1.48%

when 664 experimental values were compared to calculated ones (38). The error for polymers

was defined as the distance outside the closest value of the literature range. This validates the

method for use with our additives, and we expect the additive solubility parameters to be

within about 5% of their true value.

Page 96: Photochemical Crosslinking Reactions in Polymers

75

Table 3.11: Hildebrand Solubility Parameters of the Solvents by Group Contribution

� �� ���� �

���

Calculated Literature (47) % Error

Toluene 19.29 18.25 5.69

THF 17.68 18.97 6.83

Table 3.12: Hildebrand Solubility Parameters of the polymers and their monomers by Group

Contribution

� �� ���� �

���

Calculated Literature(47) %

Error � �� ���� �

���

Calculated Literature

(36) %

Error

Styrene 19.01 19.02 0.03 Polystyrene 20.38 17.4-19.0 7.26

∝-methylstyrene

18.71 18.33 2.09 Poly(∝-

methylstyrene) 19.61 N/A

Normal-butylacrylate

18.03 17.92 0.62 Poly(Normal-butylacrylate)

19.64 18-18.5 6.16

Methyl-methacrylate

17.83 18.53 3.76 Poly(Methyl-

methacrylate) 19.47 18.6-26.4 0

Isobutylene 13.73 13.66 0.54 Polyisobutylene 15.51 16-16.6 3.06

1,3 butadiene 15.30 14.46 5.79 Poly(1,4

butadiene) 17.33 17-17.6 0

Page 97: Photochemical Crosslinking Reactions in Polymers

76

Table 3.13: Hildebrand Solubility Parameters of the additives by Group Contribution

� �� ���� �

���

Calculated Literature

(47) %

Error � �� ���� �

���

Calculated

Benzophenone 21.22 20.73 2.30 Bifunctional Benzophenone 21.38

Xanthone 18.74 N/A Bifunctional Xanthone 24.05

Phthalimide 20.04 N/A Bifunctional Phthalimide 19.19

Phenylazide 17.58 N/A Bifunctional Phenylazide 20.11

Similar solubility parameters imply miscibility (36). In trinary systems, preferential interaction

can increase the local density of species above the bulk average (33, 48-50). If the additives are

preferentially miscible in the polymer, the probably of reaction will increase significantly even if

residual solvent is present in the system.

Comparing Table 3.11, Table 3.12 and Table 3.13 we find that PIB and the bi-functional

xanthone additive are the least miscible of all the species; the large difference implies partial

solubility or insolubility both in each other and in the solvents. The bi-functional xanthone is

most likely immiscible in all other species, while only phenylazide is soluble in PIB. We

therefore expect few if any reactions in any system containing the bi-functional xanthone or

PIB. Phenylazide is quite miscible in PBD, which itself is at best only partially soluble with the

other additives. All the other polymers and additives are inter-miscible; PS is preferentially

miscible with all the additives other than the bi-functional xanthone and phenylazide, while

PamS, PMMA, and PnBA are equally miscible with toluene, the phthalimides and the bi-

functional phenylazide.

Page 98: Photochemical Crosslinking Reactions in Polymers

77

Group contribution methods allow the calculation of thermodynamic and physical

properties from structure. They provide property guidelines that help with molecular and

system design. They are not infallible; hydrogen bonding , dipole, and other specific

interactions can make substances that have the same solubility parameter immiscible, and

make apparently immiscible species miscible (36). Our group contribution calculations predict

that PS is the most miscible with the additives and that we should expect favorable interactions

with PamS, PMMA, and PnBA. The xanthone based additives will most likely microphase

separate rendering the majority of the molecules reactively impotent, and PIB is unlikely to

react at all. Despite these predictions, these studies were performed in non-equilibrium thin

films which force the species into close proximity so some reaction is to be expected.

Page 99: Photochemical Crosslinking Reactions in Polymers

78

3.7. References

1. G. T. Carroll, M. E. Sojka, X. Lei, N. J. Turro, J. T. Koberstein, Langmuir 22, 7748 (2006). 2. G. T. Carroll, L. D. Triplett, A. Moscatelli, J. T. Koberstein, N. J. Turro, Journal of Applied Polymer

Science 122, (2010). 3. M. D. Millan, J. Locklin, T. Fulghum, A. Baba, R. C. Advincula, Polymer 46, 5556 (2005). 4. W. Schnabel, Polymers and light : fundamentals and technical applications. (Wiley, Chichester,

2007), pp. xiv, 382 p. 5. S. E. Webber, Chemical Reviews (Washington, DC, United States) 90, 1469 (1990). 6. O. Prucker, J. Habicht, I.-J. Park, J. Ruhe, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 44,

470 (2003). 7. J. R. Lancaster, Columbia University (2011). 8. in Comprehensive Handbook of Chemical Bond Energies, Y. Luo, Ed. (CRC Press, 2007), pp. 19-

145. 9. in Handbook of Bond Dissociation Energies in Organic Compounds. (CRC Press, 2002). 10. J. Lalevee, X. Allonas, J.-P. Fouassier, J. Am. Chem. Soc. 124, 9613 (2002). 11. N. J. Turro, V. Ramamurthy, J. C. Scaiano, Modern molecular photochemistry of organic

molecules. (University Science Books, Sausalito, Calif., 2010), pp. xxxiii, 1084 p. 12. J. K. Kochi, Ed., Free Radicals, vol. 2 (John Wiley and Sons, New York, 1973), vol. 2. 13. W. A. Pryor, Free radicals. McGraw-Hill series in advanced chemistry (McGraw-Hill, New York,,

1966), pp. xii, 354 p. 14. G. T. Carroll, M. E. Sojka, X. G. Lei, N. J. Turro, J. T. Koberstein, Langmuir 22, 7748 (2006). 15. J. Ruhe, K. Seidel, R. Toomey, Abstracts of Papers, 227th ACS National Meeting, Anaheim, CA,

United States, March 28-April 1, 2004, COLL (2004). 16. J. Ruhe, R. Toomey, Abstracts of Papers, 226th ACS National Meeting, New York, NY, United

States, September 7-11, 2003, PMSE (2003). 17. N. S. Allen et al., Polym. Degrad. Stabil. 73, 119 (2001). 18. F. W. Deeg, J. Pinsl, C. Braeuchle, J. Phys. Chem. 90, 5715 (1986). 19. S. Beuermann, D. A. Paquet, J. H. McMinn, R. A. Hutchinson, Macromolecules 29, 4206 (Jun,

1996). 20. in Handbook of Photochemistry, Third Edition. (CRC Press, 2006), pp. 83-351. 21. in Comprehensive Handbook of Chemical Bond Energies. (CRC Press, 2007), pp. 255-368. 22. in Comprehensive Handbook of Chemical Bond Energies. (CRC Press, 2007), pp. 147-210. 23. G. McDermott, D. J. Dong, M. Oelgemoller, Heterocycles 65, 2221 (Sep, 2005). 24. U. C. Yoon, P. S. Mariano, Accounts Chem. Res. 34, 523 (Jul, 2001). 25. O. l. Michael, G. Axel, in CRC Handbookof Organic Photochemistry and Photobiology, Volumes 1

& 2, Second Edition. (CRC Press, 2003). 26. C. Sanchez-Sanchez, E. Perez-Inestrosa, R. Garcia-Segura, R. Suau, Tetrahedron 58, 7267 (Sep,

2002). 27. M. J. Leamen, N. T. McManus, A. Penlidis, J. Appl. Polym. Sci. 94, 2545 (2004). 28. N. T. McManus, A. Penlidis, J. Appl. Polym. Sci. 70, 1253 (1998). 29. J. Brandrup, E. H. Immergut, E. A. Grulke, Eds., Polymer Handbook, vol. 1 (John Wiley & Sons, Inc,

ed. 4, 1999), vol. 1, 4. 30. N. P. Cheremisinoff, Handbook of Polymer Science and Technology: Synthesis and properties.

(CRC Press, ed. 1, 1989), vol. 1, pp. 310. 31. R. R. Chance, S. P. Baniukiewicz, D. Mintz, G. Ver Strate, N. Hadjichristidis, International Journal

of Polymer Analysis and Characterization 1, 3 (1995).

Page 100: Photochemical Crosslinking Reactions in Polymers

79

32. H. Staudinger, W. Heuer, BERICHTE DER DEUTSCHEN CHEMISCHEN GESELLSCHAFT 63, (1930). 33. P. J. Flory, Principles of Polymer Chemistry. (Cornell University Press, Ithica, New York, 1953). 34. I. H. Craig, J. R. White, A. V. Shyichuk, I. Syrotynska, Polym. Eng. Sci. 45, 579 (Apr, 2005). 35. C. Jackson, Y. J. Chen, J. W. Mays, Journal of Applied Polymer Science 61, 865 (Aug, 1996). 36. D. W. v. Krevelen, P. J. Hoftyzer, Properties of Polymers, their estimation and correlation with

correlation with chemical structure. (Elsevier Scientific Publishing Co., New York, ed. 2nd, 1976), vol. 1, pp. 620.

37. L. Constantinou, R. Gani, AIChE J. 40, 1697 (Oct, 1994). 38. T. J. Sheldon, C. S. Adjiman, J. L. Cordiner, Fluid Phase Equilibria 231, 27 (Apr, 2005). 39. J. T. Koberstein, D. Cho, D. Wong, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.) 46, 458

(2005). 40. C. Jalbert, J. T. Koberstein, A. Hariharan, S. K. Kumar, Macromolecules 30, 4481 (1997). 41. P. Muisener et al., Macromolecules 36, 2956 (Apr, 2003). 42. P. A. V. O. R. Muisener et al., Macromolecules 36, 2956 (2003). 43. D. A. Wong, P. A. V. O'Rourke-Muisener, J. T. Koberstein, Macromolecules 40, 1604 (Mar, 2007). 44. P. A. Schacht, J. T. Koberstein, Polymer 43, 6527 (2002). 45. H. Q. Lin et al., Journal of Membrane Science 276, 145 (2006). 46. L. Constantinou, R. Gani, J. P. Oconnell, Fluid Phase Equilibria 103, 11 (Jan, 1995). 47. S. b. A. Design Institute for Physical Properties. (Design Institute for Physical Property

Research/AIChE). 48. P.-G. de Gennes, Scaling Concepts in Polymer Physics. (Cornell University Press, Ithaca, New

York, 1979). 49. J. T. Koberstein, Journal of Polymer Science Part B-Polymer Physics 42, 2942 (Aug, 2004). 50. J. T. Koberstein, W. C. Hu, D. Duch, C. J. Jalbert, Polym. Mater. Sci. Eng. 75, 395 (1996). 51. N. M. Ahmad, F. Heatley, P. A. Lovell, Macromolecules 31, 2822 (1998/05/01, 1998).

Page 101: Photochemical Crosslinking Reactions in Polymers

80

4. Chapter 4: Photochemically Induced Radical Reactions in Polystyrene and Poly(normal-butylacrylate)

(with Madalina Ene, Jeffrey Lancaster, and Gregory Carroll)

This chapter applies the insights developed in Chapter 4 to explore benzophenone-

induced radical reactions in polystyrene and poly(normal-butylacrylate). These two polymers

were chosen for this in-depth investigation because they are the two with the lowest bond

dissociation energies and therefore the highest probabilities of hydrogen abstraction. They

have few side reactions. Polystyrene is a glassy polymer at experimental temperatures and

poly(normal-butylacrylate) is rubbery. The reaction in two different diffusive regimes can

therefore also be explored.

4.1. Introduction

Crosslinking, either to promote chain branching (1-6) or full network formation (7-11), is a

common method for polymer modification because of its profound influence on polymer

mechanical properties (1, 12, 13) and other characteristics such as mobility, solubility, and

diffusion. Covalent crosslinks are usually introduced during polymer synthesis by the

incorporation of multifunctional monomers (14, 15) or crosslinking agents (16-19). The

resultant materials are generally thermoset polymers that retain the shape in which they were

polymerized and are insoluble in all solvents. Polymers may also be crosslinked after synthesis

by the addition of a crosslinking additive, by exposure to high energy radiation, or by a

combination of the two processes. Examples of crosslinking additives include sulfur (i.e.

vulcanization) (20-22) and peroxides (23-25) that crosslink upon heating, and glutaraldehyde, a

Page 102: Photochemical Crosslinking Reactions in Polymers

81

common fixative for proteins (26-31). Radiative methods are typically based on the formation

of radicals by exposure to high energy radiation such as gamma rays (9, 32) and electron beams

(17, 33-35). Alternatively, photocrosslinking of preformed polymers can be achieved by use of

a photosensitizer that forms radicals upon exposure to visible or UV radiation. The excited

photosensitizer generally abstracts a hydrogen atom from the polymer to form a macroradical.

Covalent crosslinks form when two macroradicals annihilate by recombination. A macroradical

on a chain backbone can also undergo chain scission by β-scission, a process which decreases

the molecular weight. The photosensitizer can be attached to the polymer during synthesis

(36-38) or can be incorporated as an additive (39-41).

Benzophenone and its derivatives are perhaps the most widely used photosensitizers

and have been studied extensively, both as additives (39, 40, 42-44) and incorporated in

functional polymer side chains (45, 46). Most recently, bi-functional bis-benzophenone

compounds have been used to induce photocrosslinking in preformed polymers (39, 47, 48) and

have been shown to be effective even for polymer glasses. The focus of the current study is to

compare the photocrosslinking properties and efficacy of benzophenone (BP) with those of a

bis-adduct (BP-BP) of benzophenone, both for a glassy polymer, polystyrene (PS), and a rubber,

poly(normal-butylacrylate) (PnBA). The spatial distribution of radical formation may differ in

the two systems because two benzophenones are coupled together to form BP-BP, and

because BP-BP introduces an additional bridging mechanism for crosslinking that is not possible

with BP. The photochemically-induced radical reactions leading to macroradicals in BP and BP-

BP are compared in Figure 4.1 and the macroradical crosslinking and chain scission reactions of

BP and BP-BP are compared in Figure 4.2.

Page 103: Photochemical Crosslinking Reactions in Polymers

82

Figure 4.1: Schematic descriptions of the photochemically induced radical reactions of BP and BP-BP leading to the production of macroradicals.

Figure 4.2: Schematic description of macroradical crosslinking reactions, including the radical recombination-bridging reaction unique to bi-functional BP-BP.

Page 104: Photochemical Crosslinking Reactions in Polymers

83

4.2. Modeling of scission and combination reactions

Figure 4.1 and Figure 4.2 graphically portray the radical generation and depletion reactions

present in the polymer-additive binary thin films discussed herein. Only two classes of

reactions affect the molecular weight distributions in the polymer thin film: macroradical

scission and macroradical recombination. Macroradical β-scission, represented in Figure 4.3,

occurs when a radical on the backbone of the polymer breaks the bond two carbons away

through electron transfer as shown by the arrows in Figure 4.3. One of the resulting chains

becomes a dead polymer and the other contains an end radical which can undergo further

radical reactions.

Figure 4.3: �-scission of allylic polymer backbone.

Macroradical recombination occurs when two radicals come together to form a covalent bond.

Macroradical scission and macroradical recombination are the only processes that change the

MWD in our systems. We can model these two reactions mathematically and compare the

models with our experiments to validate the hypothesized reactions in the system and gain an

understanding of the role of functionality and radical density in these systems.

4.2.1. Modeling random macroradical �-scission

We assume equal reactivity; no monomer on a polymer chain has a higher probability of

radical formation and subsequent scission than any other. Even if radical formation or scission

is more likely near the ends of a polymer chain, the number of monomers in the chain

R Rn/2 R R

+n-xR

xR

Page 105: Photochemical Crosslinking Reactions in Polymers

84

compared to at the ends makes the effects of end-scission very small. We assume pure linear

chains; this restricts our focus to the early stages of crosslinking before network formation. We

expect the model to become less accurate the further along the recombination reaction

progresses. We assume that there are significantly fewer scissions than the number of

structural units in a chain; this also restricts our focus to the early stages of scission. Large

numbers of scission events reduce all chains to a few monomers in length. We use the model

developed by Osamu Saito (49-52) and the formalisms of Hamielec (53) .

With these assumptions, we can write the differential mass balance of the weight

fraction of chains � with degree of polymerization � and degree of scission �:

��(�,�)

��= −��(�,�)+ 2��

�(�,�)

���

(4.1)

The first term on the right is the decrease of chains of length � due to scission; the second term

is the increase of chains of length � due to the scission of larger chains. We assume that chains

are large enough that a continuous distribution is an accurate approximation to the discrete

distribution. Saito found the analytical solution to equation (1.10). Equation (1.12) is his

solution.

�(�,�)= ��(�,0)+ ���2 + �(� − �)

��(�,0)��

�exp (−��) (4.2)

The first term on the right is the original weight fraction distribution. In these equations, the

degree of scission � is defined as the number of chain scission events per number of monomers

and is a time and kinetics independent radiation dose analog. � can therefore be derived from

experimental distributions as equation (1.12).

Page 106: Photochemical Crosslinking Reactions in Polymers

85

� =1

��,������−

1

��,������ (4.3)

where ��,������ is the number average chain length after degree of reaction � and ��,0����� is the initial

number average chain length.

4.2.2. Modeling random macroradical recombination

We assume equal reactivity; no monomer on a polymer chain has a higher probability of

radical formation or macroradical recombination than any other. Even if radical formation or

macroradical scission is more likely near the ends of a polymer chain, the number of monomers

in the chain compared to at the ends makes the effects of end-recombination very small. We

assume pure linear chains; this restricts our focus to the early stages of crosslinking before

network formation. We expect the model to become less accurate the further along the

recombination reaction progresses. We assume a single recombination per chain; this also

restricts our focus to the early stages of crosslinking. The solutions of the crosslinking models

of Saito (49, 50, 52) and McCoy (54, 55) are infinite sums of moment generating functions that

return increasingly accurate distributions through the modeling of the evolution of the number

average, weight average and higher moments of the polymer distribution. We employ a much

simpler model to achieve approximate results.

With these assumptions we investigate the probability of a single chain of length �

forming from the macroradical recombination of a single chain of length � with another chain:

�(�)= �(�)�(� − �) (4.4)

where � is the probability. Equation (2.1) merely states that the probability of a chain of length

� forming from two chains, one of length � and the other of length (� − �), is the probability of

Page 107: Photochemical Crosslinking Reactions in Polymers

86

each of the individual chains existing multiplied together. This applies the equal reactivity

assumption in an aphysical manner: every monomer has an equal chance of reacting with any

other monomer in the system irrespective of the distance between the monomers. This is not

true in an actual physical system, but is accurate if one considers all possible ways to connect a

saturated lattice of monomers into a distribution of chains. Equation (4.4) is valid for a single

chain formed from two individual chains; we sum over all � to produce the un-reduced

probability distribution:

�(�)= ��(�)�(� − �)

���

(4.5)

The assumption of equal reactivity, in which every monomer has an equal chance of

reacting, means that longer chains have a higher probability of reacting. The probability of a

chain of length � participating in a reaction is therefore �(�)∝ ��(�), where �(�) is the

number fraction. This is equal to the probability that a radical has formed on the chain. The

probability of chain length � after macroradical recombination is therefore proportional to the

weighted sum of the number fractions of the two polymer chains and is itself a number

fraction:

�(�)∝ �(�)= ��(� − �)�(�)�(� − �)

���

���

(4.6)

The entire recombination distribution is therefore the sum of equation (4.6) over all �.

4.2.3. Linking the model to HPLC-SEC data

The model uses number and weight fraction distributions of polymer chains of length �.

The experimental distributions from the HPLC-SEC must be converted into these distributions.

Page 108: Photochemical Crosslinking Reactions in Polymers

87

As discussed in the section on the HPLC-SEC detectors (section 2.2.2), RIDs measure the number

of chains of length � and UVDs measures the number of chromophores. All data used in this

paper was first converted to the number of chains of length �. If a different distribution was

needed for modeling, the data was converted to the needed distribution before calculations. If

� is the detector signal and � is the number fraction of chains of degree of polymerization, �,

then the following conversions were used before graphing and analysis:

����(�)=����(�) �⁄

∑(����(�) �⁄ )

����(�)=����(�)

∑ ����(�)

(4.7)

Once the data has been converted to the number fraction distribution of chains using equation

(2.2), we can interconvert between the number fraction and weight fraction using a conversion

factor ��, defined and proved in equation (4.8).

�� =�� ∑ ��∑ ����

�(�)= �(�)�� =��∑ ��

�� ∑ ��∑ ����

=����∑ ����

(4.8)

where � = �� is the degree of polymerization of species � and �� is the number of chains of that

species. We rearrange equation (4.2) and develop both a data analysis method and a method of

quantifying the extent of reaction:

�(�,�)−�(�,0)exp (−��)= ��exp (−��)�2 + �(� − �)

��(�,0)��

(4.9)

The left hand side (LHS) of equation (4.9) provides an alternate method of quantifying �: we

multiply the control distribution �(�,0) by the exponential and fit � to minimize the sum of

Page 109: Photochemical Crosslinking Reactions in Polymers

88

least square error of the LHS. This provides an excellent fit of the control to the data, and the

differences between them are therefore the changes due to irradiation.

4.3. Materials

Polystyrene (PS) (Mn=46000, PDI=1.03, Waters Associates), poly(normal-butyl acrylate)

(PnBA) (Mn=60000 PDI=1.08, Polymer Source, Inc.), HPLC-grade Toluene (Pharmco-Aaper) and

Benzophenone (BP) (Fischer Scientific) were used as received.

The bi-functional bis-benzophenone adduct (BP-BP) shown in Figure 4.4 was synthesized

following the method reported previously (39). Further details are given in (56).

Figure 4.4: Photoactive Crosslinker α-meta-,ω-para-bis-benzophenone (BP-BP)

4.3.1. Substrate Preparation

Silicon Wafers (Wafer World, Inc) were cut into 1cm2 squares and placed in Piranha

solution (3:1 H2SO4:H2O2 DANGER: Piranha is extremely dangerous) overnight. Wafers were

washed first in DI H2O and then in ethanol, and were subsequently dried with N2 in a clean

hood. After treatment in a UV-Ozone Cleaning System (UVOCS, ,Inc) for 20 minutes, the wafers

were again washed in DI H2O and Ethanol and dried with nitrogen in a clean hood. The cleaning

procedure removes all organic contaminants and produces a fresh oxide layer on the silicon.

Page 110: Photochemical Crosslinking Reactions in Polymers

89

4.3.2. Solution Preparation

Solutions of 20mg/ml PS and PnBA in Toluene were agitated with a vortexer for 30

minutes and split into nine scintillation vials containing approximately 0.5mL of solution each.

One vial was set aside as a control. BP was added to four vials in molar ratios of 32:1, 8:1, 2.4:1,

and 1:1, respectively, while BP-BP was added to the last four vials in molar ratios of 16:1, 4:1,

1.2:, and 0.5:1. The molar ratios for BP-BP were kept at half those of BP to keep the overall

content of the benzophenone moiety equal.

4.3.3. Sample Preparation

Polymer solutions were spin coated (Laurell Technologies Co. WS-400A-6NPP/LITE)

onto the silicon substrates at 2000 rpm for 1 minute. The film thicknesses measured by

ellipsometry (alpha-SE, J.A. Woolam, Co.) were 128±6 nm for Polystyrene and 110±29 nm for

poly(normal-butylacrylate). Samples were irradiated with a 350nm ultraviolet source (UVP, Inc,

200 �W/cm2) under an Argon blanket in quartz-topped chambers for 0min (control), 30 min, 60

min, 90 min, 3 hr, 4.5 hr, 6 hr, and 7.5hr. The source was run for 30 min before the start of

each experiment to eliminate startup transients. Specimens were irradiated on a rotating stage

to optimize the homogeneity of irradiation.

4.3.4. GPC analysis

After irradiation, specimens (i.e. films on cut silicon wafers) were stored in scintillation

vials in a dark drawer to eliminate the possibility of further reactions. No differences in GPC

response were observed between a fresh specimen and a specimen stored in this fashion for 5

days confirming the appropriateness of this storage method. EPR analysis showed no extant

radicals in the system post-irradiation.

Page 111: Photochemical Crosslinking Reactions in Polymers

90

Solutions for GPC analysis were prepared by placing 0.200 mL of HPLC toluene in each

vial and followed by shaking on a vortexer for 30 min. 0.05mL of a 1mg/mL solution of 2350

Dalton PS in tetrahydrofuran was added as an internal standard subsequent to the shaking.

Solutions were then filtered through a 0.2 micrometer PTFE syringe filter and injected into the

GPC (2 Polymer Laboratories’ ResiPore™ Columns in a Shimadzu HPLC system: LC-10ATvp

pump, CTO-10ACvp column oven, SPD-10a UV-Vis detector, RID-10A RI detector, SCL-10Avp

system controller and a Wyatt miniDAWN TREOS static light scattering detector). Analysis was

performed using custom MATLAB™ programs and Origin 8.0™. Molecular weights were

confirmed using a Wyatt miniDAWN light scattering detector (PS dn/dc: 0.185ml/g; PnBA

dn/dc: 0.08 ml/g). dn/dc was measured using Wyatt miniDAWN in batch mode at 657nm as

well as Brookhaven Instruments differential refractometer (BI-DNDCW) at 535nm to determine

wavelength dependence. GPC data were calibrated using linear PS standards (EasiVial PS-

M(2mL), Varian). Absolute PnBA molecular weights were determined from the PS calibrated

values by correcting for hydrodynamic volumes using tabulated Mark-Houwink constants (57).

4.4. Results

Polymer thin films containing BP and BP-BP were UV irradiated in an inert atmosphere to

affect photocrosslinking while varying multiple parameters: the irradiation time, additive to

polymer molar ratio, and polymer mobility. Polymer mobility was varied through the use of

glassy PS, in which motion is restricted to sub-chain length scales, and rubbery PnBA, in which

motion can occur on all possible length scales.

Page 112: Photochemical Crosslinking Reactions in Polymers

91

Figure 4.5a shows how the molecular weight distribution evolves with UV exposure time for

the polystyrene and BP system. The progression of both molecular weight increases and

decreases is easy to see, demonstrating the effectiveness of the benzophenone in polystyrene

thin films. Similar progressions are seen in samples with our bi-functional benzophenone

(Figure 4.5b). Figure 4.5c compares the distributions after 9 hours with different molar ratios of

benzophenone additive to polystyrene while Figure 4.5d compares the same with the BP-BP.

The molar ratios insure that the density of BP moieties in the system is the same throughout

the experiment. In the inset graph the pure PS control shows no change in molecular weight

when compared to the control PS after 9 hours of irradiation, demonstrating a lack of

photosensitivity at experimental wavelengths.

Page 113: Photochemical Crosslinking Reactions in Polymers

92

Figure 4.5: Results of experiments with polystyrene and additives, both BP and BP-BP. A) Number fraction of chains in 32:1 BP:PS films after 0, 3, 6, and 9 hours of irradiation. Inset graph shows the control after 0 and 9 hours of irradiation. B) Number fraction of chains in 16:1 BP-BP:PS films after 0, 3, 6, and 9 hours of irradiation. C) Number fraction of chains in 0:1, 1:1, 2.4:1, 8:1 and 32:1 BP:PS films after 9 hours of irradiation. D) Number fraction of chains in 0:1, 0.5:1, 1.2:1, 4:1 and 16:1 BP-BP:PS films after 9 hours of irradiation. All axes are equal.

Figure 4.6 contains the evolving molecular weight distributions of PnBA. The molecular

weight distributions of 1.2:1 BP:PnBA after up to 9 hours of irradiation are shown in Figure 4.6a.

The progression of molecular weight increases is easy to see, demonstrating the effectiveness

of the benzophenone in poly(n-butyl-acrylate) thin films. The inset graph again demonstrates

the lack of photosensitivity at experimental wavelengths. A similar progression is seen in

Figure 4.6b, which features the BP-BP system. More scission is visible than in Figure 4.6a.

Figure 4.6c compares the distributions after 3 hours with different molar ratios of BP while

Figure 4.6d compares the same with the BP-BP.

Page 114: Photochemical Crosslinking Reactions in Polymers

93

Figure 4.6: Results of experiments with poly(n-butyl-acrylate) and additives, both BP and BP-BP. A) Number fraction of chains in 2.4:1 BP:PnBA films after 0, 3, 6, and 9 hours of irradiation. Inset graph shows the control after 0 and 9 hours of irradiation. B) Number fraction of chains in 1.2:1 BP-BP:PnBA films after 0, 3, 6, and 9 hours of irradiation. C) Number fraction of chains in 0:1, 1:1, 2.4:1, 8:1 and 32:1 BP:PnBA films after 3 hours of irradiation. D) Number fraction of chains in 0:1, 0.5:1, 1.2:1, and 16:1 BP-BP:PnBA films after 3 hours of irradiation. All axes are equal.

4.5. Discussion

The results of the UV photocrosslinking of polymer thin films containing BP and BP-BP while

varying multiple parameters, Figure 4.5 and Figure 4.6, demonstrate that benzophenone and bi-

functional benzophenone work in both PS and PnBA to change the molecular weight

distribution in binary thin films.

Page 115: Photochemical Crosslinking Reactions in Polymers

94

4.5.1. Higher functional species generate more radicals

Figure 4.5 demonstrates that both the BP and BP-BP act to increase the molecular

weight of the polystyrene. The peak forming through macroradical recombination is primarily

twice the molecular weight of the original sample. Simulations discussed later show that chains

are not combining end-to-end; macroradicals are forming branch points. Single branch points

are therefore favored over multiple branch points per chain. Lower molecular weights also

increase with time and ratio due to macroradical scission (Figure 4.1). Area increases due to

macroradical recombination are equivalent at long irradiation times between BP and BP-BP;

macroradical scission occurs much more in the BP-BP systems. The evolution of the

macroradical recombination peak with time suggests that BP-BP recombination occurs more

rapidly than BP recombination at intermediate times. Once within reaction distance of the

backbone hydrogens, the hydrogen abstraction kinetics of BP and BP-BP are essentially the

same; the higher apparent speed of the reaction is therefore most likely due to a higher local

radical density in the BP-BP systems and therefore more opportunities for reaction. This is

shown later. Figure 4.5 demonstrates that benzophenone, whether BP or BP-BP, produces

molecular weight increases even in a solid, binary, glassy system consisting solely of PS and

additive.

Figure 4.6 demonstrates that both the BP and BP-BP act to increase the molecular

weight of the poly(normal-butylacrylate). The peak forming through macroradical

recombination is not primarily twice the molecular weight of the original. The number of

branch points per chain in the PnBA system vary much more than in the PS system. While there

are more single branch points than any other, the molecular weight distribution is much

Page 116: Photochemical Crosslinking Reactions in Polymers

95

broader and, with time and ratio, appears to be approaching a flat distribution. The formation

of infinite networks is suggested, though not proven as our experimental technique focuses on

the sol fraction. Macroradical recombination occurs frequently in the PnBA; macroradical

scission does not. NMR studies (56) of the reaction between a PnBA analog and BP proved that

PnBA preferentially forms radicals on the pendant group for steric and energetic reasons

(section 3.2.5). The PnBA branch points are therefore primarily on the pendant group; if

pendant group radicals are left unreacted the pendant group will undergo scission while

maintaining the macromolecule. PnBA therefore shows much less macroradical scission than

PS, in which the radicals form on the polymer backbone. Figure 4.6 demonstrates that

benzophenone, whether BP or BP-BP, produces molecular weight increases even in a solid,

binary, rubbery system consisting solely of PnBA and additive.

Figure 4.5c-d and Figure 4.6c-d prove that BP and BP-BP are directly responsible for the

reactions seen in both figures: as the molar ratio of additive to polymer is increased from 0:1 to

16:1 in the BP-BP system and 0:1 to 32:1 in the BP system, the size of the macroradical

recombination peaks and the extent of reaction increase. The molecular weight changes are

therefore directly affected by the density of BP or BP-BP generated radicals in the thin film.

Page 117: Photochemical Crosslinking Reactions in Polymers

96

Figure 4.7: A) Number fraction of chains in Polystyrene films after 9 hours of irradiation with pure PS (line), Benzophenone and PS 32:1 molar ratio (dashed), and bi-functional benzophenone and PS 16:1 molar ratio (dotted). B) Number fraction of chains in Poly(n-butyl-acrylate) films after 3 hours of irradiation with pure PnBA (line), Benzophenone and PnBA 2.4:1 molar ratio (dashed), and bi-functional benzophenone and PnBA 1.2:1 molar ratio (dotted). All distributions have an area of 1.

Figure 4.7 shows the molecular weight distributions of representative samples of PS and

PnBA with BP and BP-BP. Distributions have been normalized to an area of 1. The system is not

closed; chains too large or too small for effective SEC separation are not included in these plots

or in the normalization. We use the relative peak heights to estimate the percentage of chains

that have reacted. Results are shown in Table 4.1. After 9 hours, about 10% of the PS chains

have reacted in the BP system while 16% have reacted in the BP-BP system. For PnBA, 20% of

the chains have reacted after only 3 hours in the BP system and 63% have reacted after 3 hours

in the BP-BP system. As only radical reactions are feasible in these systems, we assert that BP-

BP generates more radicals than BP in both polymeric systems and more radicals are generated

in PnBA than in PS.

Page 118: Photochemical Crosslinking Reactions in Polymers

97

Table 4.1: Percent of Original Distribution that has Reacted in Figure 4.7

% Control Chains Reacted

BP

% Control Chains Reacted

BP-BP

PS (9 Hours) 9.79 15.66

PnBA (3 hours) 20.28 62.94

In both PS and PnBA there are more radicals generated by BP-BP than by BP. There is

therefore a fundamental difference between BP-BP and BP that leads to a difference in

hydrogen abstraction yield. The two benzophenones on BP-BP are photochemically and

sterically isolated from each other by an alkyl linkage; NMR experiments showed neither intra-

nor inter- BP radical reactions in CD3CN (56) and no resonance coupling. The BP-BP

chromophore includes an ester group that is not present in the BP. While this slightly changes

the absorbance of the molecule the change is too small to account for the large difference in

hydrogen abstraction yield. We can assume therefore that one bis-benzophenone basically

functions as two benzophenone molecules with the primary difference being that they are

connected in the bis adduct.

There are a number of reactions that compete with macroradical scission or

recombination in the binary thin films. As Figure 4.8 explores, after excitation but before

radical formation BP and BP-BP triplets can self quench: if two additive molecules are in close

proximity, then their triplet states can interact and they reduce to the ground state. NMR

studies (56) showed no radical reactions between BP moieties on the same BP-BP; this implies

there is no intra-molecule self-quenching and BP-BP self-quenching only occurs

intermolecularly. Once the triplet has reacted to generate a radical, the additive radicals can

Page 119: Photochemical Crosslinking Reactions in Polymers

98

covalently bond with macroradicals to generate dead polymer (Figure 4.8) in the BP case and

macroradicals in the BP-BP case. This is the precursor to the macroradical recombination-

bridging reaction in Figure 4.2 and produces a macroradical that cannot undergo scission, as the

radical is isolated from the polymer. It is not a truly competing reaction, but is one mechanism

that removes BP-BP radicals from the system.

Figure 4.8: Reactions that compete with macroradical scission and combination.

These competing reactions affect the number of radicals in the BP and BP-BP systems.

We build a simple lattice model of the thin film to visualize how.

Simple lattice model of the thin film

Throughout the experiments, the molar ratio of polymer to benzophenone moieties was

kept constant; the molar ratio of BP to polymer was twice the ratio of BP-BP to polymer to

account for the two benzophenones per BP-BP. We visualize the local density using a simple

Page 120: Photochemical Crosslinking Reactions in Polymers

99

lattice model. We assume a cubic lattice with a unit cell the volume of a benzophenone moiety.

We assume an ideal case in which the additives are uniformly distributed. Figure 4.9 is a planar

projection of the lattice around two adjacent additive molecules. The assumption of uniform

distribution allows Figure 4.9 to represent the local density at all locations in the polymer thin

film.

The lattice makes clear that the local density of benzophenone moieties is different in

BP and in BP-BP. BP, which is mono-functional, provides one photoactive molecule within each

average volume, assuming it is well distributed. BP-BP, which is bi-functional, also provides one

photoactive molecule within each average volume. But because it is bi-functional, there are

two potential photoexcitation locations; BP has one.

Page 121: Photochemical Crosslinking Reactions in Polymers

100

Figure 4.9: Comparison of lattice cell sizes for BP (top) and BP-BP (bottom). Lattice volumes are scaled such that the molar ratio of BP is twice that of BP-BP to mirror experiments. Representative inter-benzophenone moiety distances are marked.

The volume of a lattice cell is smaller in BP than in BP-BP, and the distance between BP

moieties in BP, ���, is smaller than the distance between BP-BP molecules, ������. The

nearest neighbor distance distribution (NND) in BP is a delta function; all moieties have six

neighbors at distance ���. The NND in BP-BP is more complicated; there are six moieties at

������, one at ������ − 1, one at ��������, and one at a lattice distance of one. The average

lattice distance in BP-BP is still approximately ������.

Page 122: Photochemical Crosslinking Reactions in Polymers

101

Effects of diffusion distance on radical formation

The competing reactions all require either two triplets or two radicals to come together.

If we assume that the thin films are ideal at time � = 0 and are allowed to grow increasingly

non-ideal with time, then the distance between species and the lifetimes of the species become

important.

The BP triplet has a lifetime of 10-4 seconds (58), while it can abstract a mol of H from a

secondary carbon within 2x10-6 seconds (58) and a mol can self-quench (in Benzene, which has

no abstractable hydrogens) in 3x10-6 seconds (58). H-abstraction and intermolecular self-

quenching are therefore equally probable in a saturated system. The NMR studies (56) showed

no radical reactions between BP moieties on the same BP-BP; this implies there is no intra-

molecule self-quenching. As shown in Figure 4.9, in an ideal distribution benzophenone

moieties are always surrounded by polymer. If non-idealities exist at � = 0 or there is a high

diffusion coefficient, aggregation will increase the local probability of self-quenching over

hydrogen abstraction. While the probabilities of intermolecular self-quenching and H-

abstraction are the same, H-abstraction and macroradical formation is always favored in our

systems assuming an ideal distribution.

The thin films under investigation are primarily polymer and additive; in the dense

network of physical entanglements small molecules have a significant diffusional advantage.

Table 4.2 tabulates the radius of gyration and ovality of BP and BP-BP in their most extended

conformations.

Page 123: Photochemical Crosslinking Reactions in Polymers

102

Table 4.2: Radius of gyration and ovality of additive molecules1

BP BP-BP

Radius Of Gyration (nm) 0.324 0.749

Ovality 1.37 1.75

1Values calculated in CS Chem3D Ultra from the most extended conformation’s energy minimum.

The significantly larger radius and ovality of the BP-BP decreases the effective diffusion

constant of BP-BP in the polymer thin film. The rotational flexibility of the alkyl linker in the BP-

BP allows for steric interactions that further restrict the volume swept out by BP-BP as

compared to BP. BP can therefore diffuse much farther than BP-BP, allowing BP molecules to

self-segregate and form aggregates that self-quench much more easily than BP-BP. The self-

quenching in BP aggregates due to non-idealities and a higher diffusion constant is one reason

why BP-BP generates more radicals than BP. The higher numbers of radicals in BP-BP systems

cause more radical reactions than in BP; the local distance between benzophenone moieties in

the thin films leads to more scission in BP-BP systems than BP.

The local distance between benzophenone in the system is very important to the

probability of reaction; the average distance is not. Figure 4.10 compares the NND of BP-BP

and BP for the four experimental molar ratios. At low ratios the NND of BP-BP is always farther

than that of BP. As the ratios increase, the NNDs of BP and BP-BP begin to overlap.

Once a macroradical has formed, it can undergo macroradical scission, recombination,

or reaction with a BP or BP-BP radical. Scission can occur spontaneously; macroradical

recombination and the reaction with benzophenone radicals require another radical. Two

Page 124: Photochemical Crosslinking Reactions in Polymers

103

radicals must be within a distance, ��, where τ is the lifetime of the radical before scission

occurs, for macroradical recombination to occur. Figure 4.10 includes a representative

diffusion distance at which ��� < �� and ������ > ��. Consideration of the average lattice

distance between benzophenone moieties in Figure 4.10 would therefore predict macroradical

recombination in BP, but not BP-BP. The bi-functionality of BP-BP insures there are always two

moieties within ��, though they have a higher probability of being on the same macromolecule

than BP radicals.

Figure 4.10 Nearest Neighbor distance vs Molar Ratio. This plot compares the nearest neighbor distance of BP (grey) to BP-BP (black) at 1:1, 2.4:1, 8:1, and 32:1 molar ratios of benzophenone moieties to PS molecules of length Mn. There are two benzophenone moieties per BP-BP. Distance is the lattice distance, with the molar volume of BP the lattice cell volume. Alkyl linker in BP-BP is ignored. Dotted line is diffusion distance for illustrative purposes only.

Page 125: Photochemical Crosslinking Reactions in Polymers

104

Figure 4.10 demonstrates, though, that the BP-BP systems will always have a greater

NND than BP systems. There will therefore be a greater distance for a radical to travel before it

can react with another radical. The probability of scission over recombination is therefore

always higher in BP-BP systems than BP systems from an average-distance perspective.

The lattice model interpretation of the polymer thin films demonstrates that the higher

average density and larger diffusion constant of BP promote general radical formation, but will

also lead to more microphase separation as BP molecules diffuse together and self-quench.

The higher local density of BP-BP and its lower diffusion rate guarantees less self-segregation

between BP-BP molecules and therefore more radical generation in BP-BP systems than BP

systems. The higher average distance in BP-BP systems also guarantees more scission than in

BP systems.

Effect of chain mobility on radical formation

Table 4.1 and Figure 4.7 further demonstrate that more radicals are generated in PnBA

than in PS. 10% of the PS has reacted in the presence of BP after 9 hours, while 20% of the

PnBA has reacted in the presence of BP after only 3 hours. This large difference in reaction

extent is mirrored in the BP-BP systems. This is because PS is in the glassy state in our systems

and PnBA is in the rubbery state. PS is restricted to � and higher motional modes in the glassy

state; segmental motion is therefore the primary mode for hole formation and additive motion.

PnBA, in the rubbery state, is not restricted to sub-modes; full chain mobility and reptation

provide a continuously evolving environment for chain and additive motion, increasing

significantly the probability of hydrogen abstraction and intermolecular covalent bonds.

Page 126: Photochemical Crosslinking Reactions in Polymers

105

In BP systems, the higher diffusion constants allow microphase separation and rapid

self-quenching of the triplet state, reducing the hydrogen abstraction yield compared to BP-BP

systems. The higher diffusion constants in rubbery systems, at the same time, allow more

macroradical recombination. The quantitative measurement of the diffusion constants and

kinetic parameters are beyond the scope of this work.

4.5.2. Insights from macroradical scission and recombination modeling

The models were not applied to the PnBA systems because the extent of reaction is too

high. Both models assume approximately one scission or one recombination per chain; the

PnBA data is indicative of many. The models were therefore only applied to PS data. Best fits

are defined by the minimization of the sum of squared error between the data and the fit.

The role of functionality in glassy binary systems

The results of the modeling of random scission, from equation (4.9), are compared with

data in Figure 4.11. The LHS of equation (4.9) calculated from the data is compared to the RHS

calculated from the control. The control is also present as a comparison. It is immediately

apparent that the fit is excellent in shape but poor in width. The congruence of the shape

confirms that the assumptions of the model are valid in our system and that the scission

present is random scission. The scission is therefore caused by the random formation of

radicals on the chain and subsequent β-scission. The disparity between the width of the model

and the data, in which the model extends to a degree of polymerization of about 500 and the

data to about 350, is caused primarily by the subtraction of the control peak in the LHS of

equation (4.9). The preferential reaction of large MW chains also plays a small role. The

scission model assumes no crosslinking or other radical processes; in our systems there are

Page 127: Photochemical Crosslinking Reactions in Polymers

106

both scission and macroradical recombination and the large MW chains react preferentially in

both these processes.

Figure 4.11: Arbitrarily scaled subtraction (LHS equation (1.2)) (line), model of scission (RHS equation (1.2)) (dashed) and control data (dotted) from which the scission was calculated. Data is from BP-BP:PS 16:1 after 9 hours irradiation. Plot is for comparison only.

The closest recombination fits are achieved using the scission calculated from equation

(4.9). Figure 4.12 compares the data and the fit for BP and BP-BP and both the model scission

and the scission from the data subtraction. Fits using the model scission to generate

recombination are closest. Inspecting the fits, we see that there is more scission in the BP-BP

systems than the BP systems as discussed in earlier sections. For the BP system, the best fit is

produced with equal amounts of end-linked, 3-arm, and 4-arm polymers. End-linked polymers

occur when scission products react with each other; the greater mobility of chain ends allows

them to react despite their low density in the polymer systems. For the BP-BP system in Figure

4.12, chain end macroradical recombination (Figure 4.2) (end-linking) is required for the best

fit.

100 200 300 400 500 600 700 800 900 1000 1100 1200

0

Number of Monomers per Chain (N)

Arb

itrary

Scalin

g (

AU

)

Page 128: Photochemical Crosslinking Reactions in Polymers

107

Figure 4.12: Fitting of the model distributions to data. Left column BP:PS 32:1 after 9 hours. Right column BP-BP:PS 16:1 after 9 hours. Top plots are fits with the model scission used in the combination modeling. Bottom plots are fits with the data scission used in the combination modeling. Data (thick line), overall fit (circles), and arbitrarily scaled control (dotted), scission (squares), end-linked (dash-dotted), 3-arm star (dashed) and 4-arm star (line) polymers are included. Inset graphs show the same plots full scale.

Figure 4.11 and Figure 4.12 establish that the scission in our system is random macromolecule

scission and that some systems require the assumption of end-to-end macroradical

recombination and other systems do not. We therefore fit our combination models to the LHS

of equation (4.9) both with and without end-linking to investigate the fits and extract

information from them.

Page 129: Photochemical Crosslinking Reactions in Polymers

108

We compare the data to the model fits in both BP-BP and BP systems at two different

molar ratios in Figure 4.13, Figure 4.14, Figure 4.15, and Figure 4.16. Figure 4.13 and Figure

4.14 compare BP:PS 32:1 and BP-BP:PS 16:1 fits, respectively, both without and with end-linking

as a mechanism. Figure 4.15 and Figure 4.16 compare BP:PS 8:1 and BP-BP:PS 4:1 fits,

respectively, both without and with end-linking as a mechanism. In the mono-functional

systems at both ratios, the recombination model without an end-linking mechanism best fits

the data, suggesting 3-arm and 4-arm stars are the primary recombination products in mono-

functional systems. In the bi-functional systems, the higher ratio (Figure 4.14) requires end-

linking for a good fit while the lower ratio (Figure 4.16) fits best without end-linking. This

continues to lower ratios: BP-BP:PS of 0.5:1, 1.2:1, and 4:1 fit best without end-linking while

16:1 fits best with end-linking. BP:PS ratios of 1:1, 2.4:1, 8:1, and 16:1 all fit best without end-

linking.

Page 130: Photochemical Crosslinking Reactions in Polymers

109

Figure 4.13: Best fits (thick line) to the LHS of equation (4.9) using a peak fitting algorithm that does not incorporate end-linking (left) and one that does (right). Data (points) is from BP:PS 32:1 after 3 (top), 6 (middle) and 9 (bottom) hours of irradiation. Curves include the end-linking combination (dash-dotted), the 3-arm star (dashed) and the 4-arm star (line). All axes are equal.

Page 131: Photochemical Crosslinking Reactions in Polymers

110

Figure 4.14: Best fits (thick line) to the LHS of equation (1.1) using a peak fitting algorithm that does not incorporate end-linking (left) and one that does (right). Data (points) is from BP-BP:PS 16:1 after 3 (top), 6 (middle) and 9 (bottom) hours of irradiation. Curves include the end-linking combination (dash-dotted), the 3-arm star (dashed) and the 4-arm star (line). All axes are equal.

Page 132: Photochemical Crosslinking Reactions in Polymers

111

Figure 4.15: Best fits (thick line) to the LHS of equation (4.9) using a peak fitting algorithm that does not incorporate end-linking (left) and one that does (right). Data (points) is from BP:PS 8:1 after 3 (top), 6 (middle) and 9 (bottom) hours of irradiation. Curves include the end-linking combination (dash-dotted), the 3-arm star (dashed) and the 4-arm star (line). All axes are equal.

Page 133: Photochemical Crosslinking Reactions in Polymers

112

Figure 4.16: Best fits (thick line) to the LHS of equation (4.9) using a peak fitting algorithm that does not incorporate end-linking (left) and one that does (right). Data (points) is from BP-BP:PS 4:1 after 3 (top), 6 (middle) and 9 (bottom) hours of irradiation. Curves include the end-linking combination (dash-dotted), the 3-arm star (dashed) and the 4-arm star (line). All axes are equal.

The macroradical scission model proves that the scission in our systems is a random process, as

would be expected from macroradical β-scission. The recombination model indicates that

Page 134: Photochemical Crosslinking Reactions in Polymers

113

mono-functional systems produce 3-arm and 4-arm stars, as expected. It also indicates that

there is a transition in the bi-functional systems between primarily 3-arm and 4-arm stars and

primarily end-linked and 4-arm stars when the ratio is between 4:1 and 16:1. We investigate

the transition in the following section.

Best fits and the probabilistic transition

The previous section demonstrates that there is a transition in the BP-BP system at an

additive:polymer molar ratio of between 4:1 and 16:1. Between these two ratios, the

macroradical recombination best fit model changes from one in which there is no end-linking of

chain end macroradicals to one in which end-linking is dominant.

Macroradical scission is a unitary process: only one radical is required for scission to

occur. Macroradical recombination is at the least a binary process: at least two radicals are

required for recombination. Photochemistry is the only source of radicals in the system. The

photon dose is therefore an analog to the density of radicals in the system. We therefore

expect the binary radical reactions to be second order in photon dose:

������ �� ������������ ������������� �������� ∝ �ℎ���� �����

The peak heights of the best fit peaks in Figure 4.13, Figure 4.14, Figure 4.15 and Figure 4.16 are

fitted using linear and second order fits in Figure 4.17. Figure 4.17a compares the best fit 3-arm

and 4-arm peaks in BP-BP:PS 4:1 with both linear and second order fits. Figure 4.17b compares

the best fit end-linked and 4-arm peaks in BP-BP:PS 16:1 with both linear and second order fits.

The correlation coefficient for each fit is marked. Below the transition, second order fits have

the highest ��. Above the transition, linear fits have the highest ��.

Page 135: Photochemical Crosslinking Reactions in Polymers

114

Figure 4.17: Comparison of linear (dotted) and squared (line) fits for best fit peaks in BP-BP:PS 4:1 (left) and BP-BP:PS 16:1 (right). Fits are end-linking (triangles), 3-arm stars (circles) and 4-arm stars (squares). The correlation coefficients for the fits are included. Y-axes are not the same.

The same fits are performed for all other PS ratios and both bi- and mono-functional

additive. The fits with the highest �� are plotted in Figure 4.18 for both BP:PS and BP-BP:PS

systems for the four benzophenone moiety molar ratios. Below the transition we identified,

the functionality of the additive does not matter. Both mono- and bi-functional additive best fit

peaks increase at the same rate within the same ratio. Above the transition ratio, there is a

first order dependence between photon dose and recombination reaction amplitude and there

is more radical recombination in bi-functional additive than mono-functional. The additives act

to increase the MWD of the glassy polystyrene solely through the density of radicals they

generate in the system below the threshold ratio; above it the functionality plays a role.

Page 136: Photochemical Crosslinking Reactions in Polymers

115

Figure 4.18: Comparison of best fit lines for bi-functional additive (dashed) and mono-functional additive (line) at four ratios of benzophenone moieties to PS: 1:1 (upper left), 2.4:1 (upper right), 8:1 (lower left) and 32:1 (lower right). Y-axes are not equal. Bottom two fits in each plot are 4-arm star fit; top two are 3-arm star (1:1, 2.4:1, 8:1) or end-linking (32:1).

The density of radicals at high molar ratios is large enough for functionality to play a role

in the reaction. Macromolecular covalent bridging (Figure 4.2) is the only mechanism that

depends on functionality. It therefore is expected to have significant probabilistic differences

from the other mechanisms that allow it to only occur in high radical density systems. We

assume the reaction rates in our system have two components: a kinetic and a probabilistic.

The kinetic contribution to the reaction rate is a function of the bulk diffusion of the chain,

which is linked to the chain size and conformation, and the local mobility of the radical.

Page 137: Photochemical Crosslinking Reactions in Polymers

116

Radicals near or on the ends of chains have more mobility and will react more often than those

at the center of the chain. The probabilistic contribution comes from the probability that the

chain is of a specific type. The probability of end-radical formation is different than that of

macroradical formation. Equation (4.10) describes this dependence.

����������� ∝ �(�,ℒ)�������� (4.10)

where �(�,ℒ) is the rate of reaction as a function of the bulk mobility of the chain, �, and the

location of the radical on the chain, ℒ. � incorporates the chain length, conformation, and

local environment. �������� is the probability of the species in question existing. We trace the species

probabilities in BP in Figure 4.19 and the probabilities in BP-BP in Figure 4.20. In both figures, a

photon excites a benzophenone moiety and a macroradical is generated with probability �. From

there, a fraction, �, of macroradicals undergo scission and a fraction, (1 − �), do not. The probability of

an end-radical is therefore �� and that of a macroradical is �(1− �). Assuming that the kinetic term

incorporates all mobility and diffusion considerations, the probability of reaction between two species is

then the product of the chances of them existing. The probability of an end-linked recombination is

therefore the probability of an end-radical squared.

Page 138: Photochemical Crosslinking Reactions in Polymers

117

Figure 4.19: Map of the radical reactions with mono-functional additive and the species probabilities.

The mono-functional probabilities are listed with their species in Figure 4.19. The bi-functional

systems require an extra photon. The probability of the new photon leading to a radical in Figure 4.20

is �, though the assumption of equal reactivity requires that � = � . They are kept separate in Figure

4.20 for clarity. Comparing the probability of a 4-arm star with that of a macromolecular bridging

reaction, we find that ������ ∝ �� while ������� ∝ ��. This is the reason bridging only occurs at high

radical densities and why functionality was only found to matter at a molar ratio above 4:1

additive:polymer.

Page 139: Photochemical Crosslinking Reactions in Polymers

118

Figure 4.20: Map of the radical reactions with bi-functional additive and the species probabilities.

The macroradical scission and recombination modeling has revealed that the

functionality of the hydrogen-abstracting additive in glassy polystyrene is only important at high

Page 140: Photochemical Crosslinking Reactions in Polymers

119

molar densities. The weight fraction of 3-arm and 4-arm stars formed in the thin films is the

same irrespective of additive functionality up to a point; above that point the density of radicals

is so high that covalent bridging begins to play a role.

4.5.3. The lattice model and the probabilistic transition

There is a transition molar ratio at which macroradical combination reactions become

first order and the density of radicals in the system allows covalent bridging. The lattice model

was used to define the NND between additive species in the thin film, with a lattice unit cell of

one benzophenone molar volume. Benzophenone has a molar volume of

���� = 168.077 ��

���� (59) while a PS chain with Mn=46000 ����� and ρ=1.05

����� has

a molar volume of ���� = 43809.5 ��

���� . That corresponds to 1.69 monomers per

lattice unit cell.

Taking the molar ratios of additive to polymer into account, Table 4.3 lists the

volumetric ratios of additive unit cells to polymer unit cells. In the 32:1 case, for example,

there is 1 additive molecule (in this case mono-functional) for every 8.1454 unit cells of

polymer (13.8 monomers). The bi-functional volumetric ratios are for BP-BP molecules,

not benzophenone moieties.

Page 141: Photochemical Crosslinking Reactions in Polymers

120

Table 4.3: Molar Ratio of additive to PS and the linear lattice distance that corresponds to.

Molar Ratio

BP

Volumetric Ratio (���)

�����

= ���

Molar Ratio BP-BP

Volumetric Ratio (������)

��������

= ������

0:1 N/A N/A 0:1 N/A N/A

1:1 260.651 6.39 0.5:1 521.303 8.05

2.4:1 108.605 4.77 1.2:1 217.209 6.01

8:1 32.5814 3.19 4:1 65.1628 4.02

32:1 8.1454 2.01 16:1 16.2907 2.54

The linear NND between mono-functional species is the cube root of the volumetric ratio

(equivalent to ��� and ���−�� in Figure 4.9) and is also shown in Table 4.3. At the highest ratio,

there is one additive for every two lattice cells. This is the cause of the transition: with every

other cell an additive there is a high probability that two radicals will form either in the same

lattice cell or in adjacent cells. Kinetics and chain motion will therefore play a very small role in

the radical recombination reactions compared to probabilities and the second order

combination reactions will become pseudo first order due to the excess radicals.

4.6. Conclusions

We have shown that the hydrogen abstractor benzophenone increases the density of

radicals in polymer thin films and this increase causes combination reactions that increase the

molecular weight of the polymer. This occurs in both glassy and rubbery polymers, represented

here with polystyrene and poly(n-butyl-acrylate). Increases in benzophenone molar

concentration increase the rate of reaction and, in rubbery poly(n-butyl-acrylate), increase the

potential extent of reaction. Bi-functional benzophenone produces more radicals than

Page 142: Photochemical Crosslinking Reactions in Polymers

121

benzophenone in both PS and PnBA because the higher mobility of benzophenone leads to

more self-quenching. The higher mobility also leads to more radicals in rubbery PnBA than in

glassy PS. The functionality of the benzophenone moiety is only a significant variable at high

molar densities for probabilistic reasons; physical entanglement and diffusive differences are

what lead to most reactive differences at low molar ratio.

4.7. Further Directions

Mathematical modeling of the full reaction network to extract kinetic constants.

Studies of the diffusion rate of the additives in the thin films.

Isolating individual reaction pathways through the use of different additive or polymer

chemistries.

Page 143: Photochemical Crosslinking Reactions in Polymers

122

4.8. References

1. S. N. Vouyiouka, E. K. Karakatsani, C. D. Papaspyrides, Progress in Polymer Science 30, 10 (2005). 2. B. Blottiere, T. C. B. McLeish, A. Hakiki, R. N. Young, S. T. Milner, Macromolecules 31, 9295

(1998). 3. W. Hu, J. T. Koberstein, J. Polym. Sci., Part B: Polym. Phys. 32, 437 (1994). 4. J. T. Koberstein, T. P. Russell, D. J. Walsh, L. Pottick, Macromolecules 23, 877 (1990). 5. S. T. Milner, T. C. B. McLeish, R. N. Young, A. Hakiki, J. M. Johnson, Macromolecules 31, 9345

(1998). 6. J. F. Rontani, Trends Photochem. Photobiol. 4, 125 (1997). 7. J. P. Baetzold, I. Gancarz, X. Quan, J. T. Koberstein, Macromolecules 27, 5329 (1994). 8. J. P. Baetzold, J. T. Koberstein, Macromolecules 34, 8986 (2001). 9. C. B. Bucknall, V. L. P. Soares, J. Polym. Sci., Part B: Polym. Phys. 42, 2168 (2004). 10. P. J. Cole, J. L. Lenhart, J. A. Emerson, J. T. Koberstein, C.-Y. Hui, Proc. Annu. Meet. Adhes. Soc.

27th, 470 (2004). 11. J.-F. Fu et al., Polym. Adv. Technol. 19, 1597 (2008). 12. R. J. Gaymans, J. Schuijer, ACS Symp. Ser. 104, 137 (1979). 13. R. J. Gaymans, T. E. C. Van Utteren, J. W. A. Van den Berg, J. Schuyer, J. Polym. Sci., Polym. Chem.

Ed. 15, 537 (1977). 14. S. Murayama, S. Kuroda, Z. J. Osawa, Polymer 34, 2845 (1993). 15. G. Mabilleau et al., Journal of Biomedical Materials Research Part A 77A, 35 (2006). 16. H. R. Kricheldorf, H. H. Thiessen, Polymer 46, 12103 (Dec, 2005). 17. G. W. Meyer et al., Polymer 37, 5077 (Oct, 1996). 18. S. Cuney et al., Journal of Applied Polymer Science 65, 2373 (1997). 19. B. Unal, R. C. Hedden, Polymer 47, 8173 (2006). 20. I. S. Zemel, C. M. Roland, Polymer 33, 3427 (1992). 21. N. J. Morrison, M. Porter, Rubber Chemistry and Technology 57, 63 (1984). 22. M. R. Krejsa, J. L. Koenig, Rubber Chemistry and Technology 66, 376 (1993). 23. N. G. Gaylord, M. Mehta, R. Mehta, Journal of Applied Polymer Science 33, 2549 (May, 1987). 24. G. E. Hulse, R. J. Kersting, D. R. Warfel, J. Polym. Sci. Pol. Chem. 19, 655 (1981). 25. W. Brostow, T. Datashvili, K. P. Hackenberg, Polymer Composites 31, 1678 (Oct, 2010). 26. Z. Chen, Z. Liu, F. Liu, L. Yang, Zhongguo Shengwu Gongcheng Zazhi 23, 99 (2003). 27. G. DeSantis, J. B. Jones, Current Opinion in Biotechnology 10, 324 (1999). 28. G. G. Guilbault, Bio/Technology 7, 349 (1989). 29. A. Jayakrishnan, S. R. Jameela, Biomaterials 17, 471 (1996). 30. L. Lin et al., Zhongguo Yixue Kexueyuan Xuebao 25, 735 (2003). 31. J. Xu et al., Huagong Jinzhan 29, 494 (2010). 32. B. D. Shirodkar, R. P. Burford, Radiat. Phys. Chem. 62, 99 (2001). 33. N. S. Allen, D. Lo, M. S. Salim, P. Jennings, Polym. Degrad. Stabil. 28, 105 (1990). 34. J. Cheng et al. (International Business Machines Corporation, USA., US, Application: US 2009),

pp. 25pp , Cont of U S Ser No 13,444. 35. V. K. Sharma, J. Mahajan, P. K. Bhattacharyya, Radiat. Phys. Chem. 45, 695 (1995). 36. O. Prucker, B. Peng, S. Golze, H. Murata, J. Ruhe, Polym. Prepr. (Am. Chem. Soc., Div. Polym.

Chem.) 44, 418 (2003). 37. M. Doytcheva, R. Stamenova, V. Zvetkov, C. B. Tsvetanov, Polymer 39, 6715 (1998). 38. M. Nagata, S. Hizakae, Macromol. Biosci. 3, 412 (Aug, 2003). 39. G. T. Carroll, M. E. Sojka, X. Lei, N. J. Turro, J. T. Koberstein, Langmuir 22, 7748 (2006).

Page 144: Photochemical Crosslinking Reactions in Polymers

123

40. N. S. Allen et al., J. Photochem. Photobiol., A 126, 135 (1999). 41. J. T. Koberstein, G. Carroll, J. Jahani, N. J. Turro, Polym. Prepr. (Am. Chem. Soc., Div. Polym.

Chem.) 48, 808 (2007). 42. M. Doytcheva et al., Journal of Applied Polymer Science 64, 2299 (1997). 43. S. P. Pappas, Prog. Org. Coat. 2, 333 (1974). 44. B. Qu, Y. Xu, L. Ding, B. Ranby, J. Polym. Sci., Part A: Polym. Chem. 38, 999 (2000). 45. J. Ruhe, D. Madge, Abstracts of Papers, 225th ACS National Meeting, New Orleans, LA, United

States, March 23-27, 2003, PMSE (2003). 46. J. Ruhe, K. Seidel, R. Toomey, Abstracts of Papers, 227th ACS National Meeting, Anaheim, CA,

United States, March 28-April 1, 2004, COLL (2004). 47. M. D. Millan, J. Locklin, T. Fulghum, A. Baba, R. C. Advincula, Polymer 46, 5556 (2005). 48. G. T. Carroll, L. D. Triplett, A. Moscatelli, J. T. Koberstein, N. J. Turro, Journal of Applied Polymer

Science 122, (2010). 49. O. Saito, J. Phys. Soc. Jpn. 13, 1451 (1958). 50. O. Saito, J. Phys. Soc. Jpn. 13, 198 (1958). 51. O. Saito, J. Phys. Soc. Jpn. 13, 1465 (1958). 52. O. Saito, Ed., Statistical Theories of Crosslinking, vol. 1 (Academic Press, New York, ed. 1, 1972),

vol. 1, 1, pp. 369. 53. V. J. Triacca, P. E. Gloor, S. Zhu, A. N. Hrymak, A. E. Hamielec, Polym. Eng. Sci. 33, 445 (Apr,

1993). 54. Y. Kodera, B. J. McCoy, AIChE J. 43, 3205 (Dec, 1997). 55. M. Wang, J. M. Smith, B. J. McCoy, AIChE J. 41, 1521 (Jun, 1995). 56. J. R. Lancaster, Columbia University (2011). 57. S. Beuermann, D. A. Paquet, J. H. McMinn, R. A. Hutchinson, Macromolecules 29, 4206 (Jun,

1996). 58. N. J. Turro, V. Ramamurthy, J. C. Scaiano, Modern molecular photochemistry of organic

molecules. (University Science Books, Sausalito, Calif., 2010), pp. xxxiii, 1084 p. 59. S. b. A. Design Institute for Physical Properties. (Design Institute for Physical Property

Research/AIChE).

Page 145: Photochemical Crosslinking Reactions in Polymers

124

5. Chapter 5: Electron Paramagnetic Resonance of Benzophenone and

Polystyrene in Toluene (with Dr. Alberto Moscatelli)

5.1. Introduction

Electron Paramagnetic Resonance (EPR) allows the direct study of radicals in both

solution and solid media (1). In is used in a wide variety of fields to better understand the

sources, the sinks, and the reactions in radical-containing systems. Its uses include: in medicine

the study of the effects of titanium dioxide exposure on nucleic acids (2); in inorganic chemistry

the study of the mechanism behind photoluminescence in ZnO phosphors (3); and in organic

chemistry it is used extensively in polymer science.

EPR is used in polymer science to study many different aspects of the synthesis,

properties, and degradation of polymers in solid and liquid form. Yamada (4) and Kamachi (5)

each wrote an excellent review of the uses of EPR in radical polymerization. It is used to

determine the mechanisms for initiation, propagation, termination, and chain transfer in a

variety of polymers, both solid and liquid, as well as the study of the kinetics and intermediates

(6) of the same. EPR can be used to study the photodegradation and stabilization of polymers

(7) post synthesis, included the effect of weathering on automotive coatings (8). It is also used

to study the segmental dynamics (9, 10) and activation volumes for the local glass transition

(11) of polymers.

In most studies, a spin trap is used because the radical species under investigation has

too short a lifetime or too low a concentration (12). This allows the investigation of

intermediate structures, as in some cases the intermediate radical can be trapped before it

Page 146: Photochemical Crosslinking Reactions in Polymers

125

disappears. A spin trap is a compound that, upon reaction with a short-lived radical, generates

a stable radical on itself. The structure of the initial radical can then be inferred from the EPR

spectrum of the spin adducts. The spin trap n-tert-butyl-∝-phenyl-nitrone (PBN) (Figure 5.2)

was chosen for this study. The reaction of PBN with a radical can be seen in Figure 5.1. The

highly polar nitrone, upon reaction with a transient radical, become a nitroxide with a stable

radical on the oxygen. The hyperfine interaction of this radical with both the ∝-hydrogen and

the atoms or groups bonded to the ∝-carbon as the C-N bond rotates probes the identity of the

attached groups and from there the identity of the initial radical is determined.

Figure 5.1: Reaction of the spin trap n-tert-butyl-∝-phenyl-nitrone (PBN) with a radical. A transient radical becomes essentially permanent when reacted with a spin trap.

In the previous chapter we found that hydrogen abstracting additives such as

benzophenone (BP) increase the density of macroradicals in polymer thin films when irradiated

with ultraviolet light. Subsequent radical reactions lead to macroradical combination and

macroradical scission reactions that change the molecular weight distributions (MWD) of the

polymer. We characterized and discussed the MWD changes in the previous chapter. In this

chapter we use EPR to investigate the radicals and macroradicals in the system.

In this investigation we used the unique ability of EPR to probe the radicals that form in

our systems, with an eye towards determining which hydrogen is abstracted from PS. We use

combinations of the four species in Figure 5.2 to identify the radicals in the systems before,

during, and after UV irradiations, which we perform in the EPR. While we were unable to

Page 147: Photochemical Crosslinking Reactions in Polymers

126

unequivocally prove the hydrogen abstraction location, we proved that PS acts to facilitate

hydrogen abstraction in the system even with toluene as a potential donor. We also found an

interesting reaction that appears to strip the ketone from benzophenone, though we did not

conclusively establish the mechanism, as that was beyond the scope of this work.

Figure 5.2: Molecular structure of the chemicals used in these investigations. A) Toluene B) Polystyrene C) Benzophenone D) n-tert-butyl-∝-phenyl-nitrone.

5.2. Materials

Polystyrene (PS) (Mn=46000, PDI=1.03, Waters Associates), HPLC-grade Toluene (Pharmco-

Aaper), n-tert-butyl-∝-phenyl-nitrone (PBN) (Sigma Aldrich) and Benzophenone (BP) (Fischer

Scientific) were used as received. The structures of these molecules are shown in Figure 5.2.

5.2.1. Sample Preparation

Solutions were prepared with various combinations of PS, BP, and PBN following the

ratios of (13). Solutions were as close to 1wt% PS, with a 90:1 molar ratio of spin trap to PS and

32:1 molar ratio of BP to PS as possible. If all three components were not present in a sample,

similar amounts of the individual material were used. Sample makeup is shown in Table 5.1.

Page 148: Photochemical Crosslinking Reactions in Polymers

127

Table 5.1: Samples used in EPR studies

Sample Toluene (mL) BP (mg) PS (mg) PBN (mg)

PS and Tol 0.5 N/A 4.100 N/A

PBN and Tol 0.5 N/A N/A 1.571

PS, PBN and Tol 0.5 N/A 4.362 1.980

BP, PBN and Tol 0.5 0.579 N/A 1.569

PS, BP, PBN and Tol 0.5 0.572 4.338 1.569

Solutions were prepared in 20mL scintillation vials and shaken for 20 minutes. They were

placed in quartz EPR liquid cells and gasses were removed using 4 freeze-pump-thawed cycles

in liquid nitrogen. Sample cells were wrapped in tin foil and placed in a dark drawer for 48

hours to allow any non-photochemical reactions to take place.

Used sample cells were washed with HPLC toluene, then acetone, then HPLC ethanol

and dried in a lab oven at 80oC for one hour. They were then filled with fresh piranha solution

(3:1 H2SO4:H2O2 DANGER: Piranha is extremely dangerous) and left until the next use. Prior

to the introduction of the sample, the piranha was removed and the sample cells were washed

consecutively with DI H20, acetone, HPLC toluene, acetone, and HPLC ethanol, covered with

aluminum foil and placed in a lab oven at 80oC for 1 hour to evaporate residual wash and

remove any residual organics.

5.2.2. EPR Measurements

Sample cells were placed in the EPR (EMX EPR Spectrometer, Bruker) and

measurements were taken both before, during, and after UV irradiation with a 350nm

ultraviolet source (UVP, Inc, 250 �W/cm2) as the EPR allows concurrent irradiation and sample

measurement. Initial EPR settings (Center field: 3480, Sweep Width: 250, Modulation

Page 149: Photochemical Crosslinking Reactions in Polymers

128

Amplitude: 1, Attenuation: 10, Frequency: 9.75, RC filter time constant: 20.48, and Conversion

time: 81.92) were tuned to produce the best calibration resonance.

EPR data was analyzed using the instrument software (Bruker WINEPR System v2.1 and

Bruker WINEPR SimFonia v1.25) and the EasySpin(14) MATLAB© toolbox. These software

suites also provided all simulations.

5.3. Results and Discussion

Solutions with various combinations of polystyrene, benzophenone, and n-tert-butyl-∝-

phenyl-nitrone in HPLC toluene were examined before, during, and after UV irradiation in an

EPR to explicate the radical and macroradical locations and reactions.

5.3.1. Controls and the Intrinsic PBN Peak

Initial EPR measurements of pure PS under irradiation proved that there were no PS

macroradicals visible at the temperature and time frame under investigation. Figure 5.3 plots

the EPR signal before, during, and after the concurrent UV irradiation. The flat signal indicates

that there are no detectable radicals. The plot is a representative of wider sweeps of the

magnetic spectrum and is centered where peaks appear in later investigations. A lack of radical

peaks does not prove that there are no radicals; it proves that if there are radicals present their

lifetime or density is not sufficient to be detected. There were no visible macroradicals in the

pure PS solution before, during or after irradiation. Other investigations of pure PS radicals are

usually performed at low temperatures using specially designed EPR cavities (4, 5) to freeze or

increase the lifetime of the radicals. A modified setup was beyond the scope of this work, so a

Page 150: Photochemical Crosslinking Reactions in Polymers

129

spin trap was chosen to allow room temperature investigations of the radicals in the system.

PBN was chosen because it is widely used in similar polymer investigations (10, 15, 16), it is

available from commercial chemical sources, and its hyperfine interactions are well studied

(17).

Figure 5.3: EPR spectra of a sample consisting solely of PS in toluene before, during, and after concurrent UV irradiation. There are no detectable radicals.

Pure PS had no detectable radicals. The spin trap was therefore required to make

radicals in the system permanent. PBN was first characterized in the EPR as a control. Despite

being both newly from Sigma-Aldrich and being kept in the dark, it has the background signal in

Figure 5.4, indicating a radical is already present on some of the PBN molecules. Subtracting

any two of the three signals in Figure 5.4 removes the signal from the noise; that the signal

stays unchanged throughout UV irradiation proves that it is intrinsic to the spin trap and not the

result of a UV generated radical in the PBN-toluene system.

0

Be

fore

0

Durin

g

342 343 344 345 346 347 348 349 350 351

0

B (mT)

After

Page 151: Photochemical Crosslinking Reactions in Polymers

130

Figure 5.4: PBN background signal before, during, and after irradiation. When any two of the three signals are subtracted, the bottom plot is produced.

The PBN intrinsic peak in Figure 5.4 appears as a triplet of peaks all of the same height.

Three equal peaks are characteristic of the radical interacting with one spin 1 atom (Table 2.2),

as two spin ½ atoms produce three peak with a height ratio of 1:2:1. The only spin 1 atom in

PBN is 14N (Table 2.1). We confirm this in Figure 5.5 which compares PBN in toluene data

during irradiation to a simulation of a radical interacting with a single 14N atom with an

�� = 1.54 �� hyperfine interaction. The fit is compelling, and a triplet hyperfine interaction of

1.54mT is associated with Nitrogen (17). When the fit is subtracted from the data the resulting

plot shows no peaks, though there is the hint of another peak we will discuss later. The PBN

intrinsic peak is due to the interaction between the radical and the 14N in the PBN.

0

Befo

re

0

During

0

After

341 342 343 344 345 346 347 348 349 350 351

0

B (mT)

Sub

tract

ed

Page 152: Photochemical Crosslinking Reactions in Polymers

131

Figure 5.5: EPR of pure PBN in toluene (dotted) and a simulation of the expected peaks formed by a single spin 1 atom (solid). The hyperfine interaction of the peaks is noted. The subtraction below shows nothing but noise, though there are hints of another set of peaks discussed later.

The identity of the PBN intrinsic peak allows us to look at the interaction between PS

and PBN in toluene. This system includes a polymer and spin trap, but no BP to photogenerate

radicals. We would therefore expect to only see radicals if PS reacts with either PBN or

toluene; there was no reaction in the PBN in toluene sample. Figure 5.6 compares the EPR

spectrum of PBN in toluene to the spectrum of PS and PBN in toluene. There is no difference;

only the PBN intrinsic peak is present. Pure PS does not generate macroradicals under UV

irradiation at experimental wavelengths and therefore does not change the signal. With a 90:1

molar ratio of PBN to PS in the liquid sample, any macroradicals that form should be trapped in

sufficient quantity to form a peak. There is no change in the signal after 40 minutes of

irradiation. This proves that the PS used in our experiments does not have impurities, initiator

fragments, or end groups that are photoactive at experimental wavelengths. Any macroradical

-0.25

-0.2

-0.15

-0.1

-0.05

0

0.05

0.1

0.15

0.2

0.25

Inte

nsi

ty

343 344 345 346 347 348 349-0.1

0

0.1

B (mT)

Fit Subtracted from Data

Inte

nsi

ty

N

g = 2.00991a = 1.5373

Page 153: Photochemical Crosslinking Reactions in Polymers

132

combination or scission seen is therefore due to macroradicals generated by BP, as BP is the

only radical source in the thin film systems discussed in Chapter 4.

Figure 5.6: Comparison of the EPR spectra of PBN in toluene and PS and PBN in toluene samples during irradiation. The plot of the subtraction shows no reaction has taken place.

Preliminary EPR investigations of pure PS in toluene were inconclusive; the lack of signal

indicated either no radicals or radicals too short lived to be detected. The inclusion of a spin

trap, which turned out to contain an intrinsic peak commensurate with a 14N interaction,

proved that few if any radicals were present in PS under UV irradiation at experimental

wavelengths. No change was seen in the EPR spectrum between irradiated samples of PBN in

toluene and irradiated samples of PS and PBN in toluene, demonstrating few if any radicals

were generated in PS, which confirms what we found in the GPC studies in Chapter 4.

5.3.2. Benzophenone and PBN: A Potential New Reaction

UV irradiation had no effect on PS in toluene, PBN in toluene, and PS and PBN in

toluene. The radicals trapped in those systems were brought in by an intrinsic PBN peak that

did not change throughout those experiments. Changes due to UV irradiation first appeared

0

Sp

in T

rap

0

PS

and S

pin

Tra

p

3410 3420 3430 3440 3450 3460 3470 3480 3490 3500 3510

0

B (mT)

Su

btr

act

ed

Page 154: Photochemical Crosslinking Reactions in Polymers

133

with the investigation of BP and PBN in toluene. This was not expected, but produced a very

interesting result that, while not proven in this study, is a wonderful candidate for further

research.

The spectra of the BP and PBN system before and during irradiation are shown in Figure

5.7. The before spectrum shows the same intrinsic PBN peak we would expect; after thirty

minutes of irradiation the spectrum has changed significantly. The pure before spectrum,

shown in the top plot, shows more pronounced secondary peaks on top of the 14N already

discussed. These secondary peaks, which were hinted at in Figure 5.5, are much more

prominent here and become obvious as a triplet of doubles. The two sample cells, post freeze-

pump-thaw, were wrapped in tin foil and placed in a dark drawer for 48 hours. This allowed

non-photochemical reactions to take place. This triplet of doubles is therefore the product of a

non-photochemical reaction that is sped up in the presence of but does not require BP, as both

samples spent about the same amount of time between preparation and the EPR. Thirty

minutes of UV irradiation in the EPR, as seen in the main plot of Figure 5.7, increases the

magnitude of the triplet of doubles while eliminating the intrinsic peak discussed in Figure 5.5.

Page 155: Photochemical Crosslinking Reactions in Polymers

134

Figure 5.7: Comparison of BP and PBN in Toluene before (solid) and after (dark solid) thirty minutes of irradiation within the EPR. Top plot is BP and PBN before irradiation.

The increase of the triplet of doubles under irradiation is incredibly rapid compared to

the 48 hours it took the peak to form in the first place; it is fully formed and does not

subsequently change in size after only 4 minutes of irradiation. Figure 5.8 shows the system

after 30 minutes of irradiation (the secondary intrinsic peak does not change in size after 4

minutes of irradiation, but all the peaks in the system are best visible and isolated after 30

minutes). It includes a simulation of the second peak. Capture of a radical causes the

formation of a nitroxide according to Figure 5.1. The hyperfine interaction between the

unpaired electron, the spin 1 nitrogen (���� = 1.37 ��) and the spin ½ alpha hydrogen

(���� = 0.19 ��) gives rise to the triplet of doubles with equal amplitudes. The nature of the

trapped radical is inferred indirectly by the value of ���� . According to literature, the ����

value decreases as the trapped radical progresses through methyl, primary, secondary, tertiary,

phenyl, diphenyl and triphenyl species (12). A combination of ���� = 1.37 �� and ���� =

0

Inte

nsi

ty

342 344 346 348 350 352 354 342 344 346 348 350

0

B (mT)

Inte

nsi

ty

Page 156: Photochemical Crosslinking Reactions in Polymers

135

0.19 �� is consistent with a phenyl or diphenyl radical (12, 17, 18). To be consistent with both

the suggestion of this second peak in PBN in toluene, its obvious presence in BP and PBN in

toluene, and its sudden increase in rate with irradiation, its structure must be the left one in

Figure 5.9 in PBD-toluene systems and a mixture of both structures in BP and PBD in toluene

systems. Over time in pure toluene, without irradiation, some toluene radicals form or are

present and react with PBN to generate the left nitroxide in Figure 5.9. The photogenerated

triplet BP accelerates this process by hydrogen abstracting the toluene to generate the phenyl

radical and generates its own diphenylmethyl radical through a process that explains the third

set of peaks.

Figure 5.8: EPR of BP and PBN in toluene (dotted) after 30 minutes of concurrent irradiation and a simulation of the interaction between the radical and a single spin 1 atom and a single spin ½ atom (solid). Hyperfine interactions are labeled.

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

Inte

nsi

ty

344 345 346 347 348 349 350-0.5

0

0.5

B (mT)

Inte

nsity

Fit Subtracted from Data

NHh

NHn

a = 0.19

g = 2.00829

a = 1.37

Page 157: Photochemical Crosslinking Reactions in Polymers

136

Figure 5.9: The two possible structures of PBN after trapping the radical that forms the triplet of doublets. The phenyl radical (left) comes from toluene while the diphenylmethyl radical (right) comes from BP.

The third set of peaks in the BP, PBN, and toluene system is a triplet of lines which are

not present in any previous system. They appear in Figure 5.7 between the triplets of doubles

only after irradiation has begun. This links them directly to the photogenerated triplet BP.

Simulations (Figure 5.10) of the third component show it is a single triplet with ��� = 0.79 ��.

No doublet split indicates there is no alpha hydrogen; there must be a double bond on the

alpha carbon of the nitroxide. ��� = 0.79 �� is an unusually small hyperfine interaction for a

nitroxide; the only possible species with that small of a pure 14N hyperfine interaction is a

carbonyl in the alpha position in relation to the nitroxide (17, 19). Given the absence of oxygen

due to the four cycles of freeze-pump-thaw at liquid nitrogen temperatures, the only O present

in the system is that of the benzophenone.

Page 158: Photochemical Crosslinking Reactions in Polymers

137

Figure 5.10: EPR of BP and PBN in toluene (dotted) after 30 minutes of concurrent irradiation and a simulation of the interaction between the radical and a single spin 1 (solid). Hyperfine interactions are labeled.

A plausible mechanism for this reaction is shown in Figure 5.11. During irradiation the

triplet benzophenone is trapped by the PBN. The unreacted BP radical hydrogen abstracts from

the alpha carbon of PBN, and the electrons rearrange to produce a stable carbonyl alpha to the

nitroxide on PBD and a stable diphenylmethyl radical. While this reaction is plausible, it is not

conclusively demonstrated here. There is no signal from the pure diphenylmethyl radical (20,

21), but the ���� = 0.19 �� hyperfine interaction is consistent with a diphenylmethyl radical

attached to PBN. Low temperature studies of the reaction in toluene, d8-toluene, and benzene

as solvents would be conclusive.

344 345 346 347 348 349 350-0.5

0

0.5Fit Subtracted from Data

B (mT)

Inte

nsi

ty

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Inte

nsi

ty

NOa = 0.79

g = 2.00889

Page 159: Photochemical Crosslinking Reactions in Polymers

138

Figure 5.11: Plausible mechanism for the creation of a carbonyl on the alpha carbon of the Nitroxide.

Figure 5.12: EPR of BP and PBN in toluene (dotted) after 30 minutes of concurrent irradiation and the sum of the previously discussed simulations (solid).

We have determined the major radicals present in the BP and PBN in toluene system.

Figure 5.12 compares the sum of the above simulations with the data. The fit is quite good and

we have accounted for the majority of the radicals in the system. The subtraction indicates

there may be an additional one, but its intensity needs to be increased before it can be

conclusively analyzed. We can say with confidence that the triplet BP preferentially hydrogen

abstracts over quenching and it is, in solution, involved in a very interesting side reaction that

transfers its ketone to PBN.

-0.5

-0.4

-0.3

-0.2

-0.1

0

0.1

0.2

0.3

0.4

0.5

Inte

nsi

ty

344 345 346 347 348 349 350-0.2

0

0.2

B (mT)

Fit Subtracted from Data

Inte

nsi

ty

Page 160: Photochemical Crosslinking Reactions in Polymers

139

5.3.3. Polystyrene, Benzophenone and PBN: Preferential Hydrogen

Abstraction

The previous sections demonstrated that toluene and PBN react on their own without

irradiation, and that the addition of BP and light speeds up this reaction and introduces a new

reaction that generates a ketone alpha to the nitroxide in PBN. In this section we introduce PS

into the system and observe that the EPR ketone peak increases in intensity significantly, which

is evidence that the BP preferentially hydrogen abstracts from the PS.

We have already determined the constituents of the EPR signal in solutions of BP and

PBN in toluene. The introduction of PS to the system introduces the opportunity for more

reactions but the already discussed reactions should still be present. We therefore begin with

the peaks already discussed previously and find that they fit the PS data quite well without the

introduction of new reactions. Figure 5.13 compares the previous fits in A) and B) and shows

that a weighted sum of the two fits the data in part C). From this we conclude that the PS does

not introduce new reactions into the system.

Page 161: Photochemical Crosslinking Reactions in Polymers

140

Figure 5.13: EPR of PS, BP, and PBN in toluene (dotted) and the simulation fits. A) and B) are the constitutive simulations from the BP and PBN in toluene system. The weighted sum of the fits in A) and B) is fitted to the new system in C). Subplots show the residual data after subtraction.

The introduction of PS does not introduce new reactions. But the triplet associated with

the ketone is five times larger than the same triplet without PS after the same amount of

irradiation (Figure 5.14). There are therefore five times more ketones in the PS system

compared to the pure BP system. The introduction of PS introduced a new source of hydrogen

-1.5

-1

-0.5

0

0.5

1

1.5

Inte

nsity

343 343.5 344 344.5 345 345.5 346 346.5 347 347.5-2

0

2Fit Subtracted from Data

B (mT)

Inte

nsity

g = 2.00852

NHhNHn

a = 0.19a = 1.37

342.5 343 343.5 344 344.5 345 345.5 346 346.5 347 347.5 348-0.5

0

0.5

B (mT)

Inte

nsity

-1.5

-1

-0.5

0

0.5

1

1.5

Inte

nsity

g = 2.00908

NOa = 0.79

-1.5

-1

-0.5

0

0.5

1

1.5

Inte

nsi

ty

343 343.5 344 344.5 345 345.5 346 346.5 347 347.5-0.5

0

0.5

B (mT)

Fit Subtracted from Data

Inte

nsi

ty

Page 162: Photochemical Crosslinking Reactions in Polymers

141

abstraction; the acceleration of the BP-PBN reaction is the result. This is most likely due to the

lower dissociation energy of hydrogen in polystyrene than in toluene (348.1 kJ/mol for the

tertiary carbon in PS (22, 23); 375.5 kJ/mol for toluene (23)) as well as the superior stability of a

tertiary radical over a primary radical.

Figure 5.14: Direct overlay of the raw data with (thin line) and without (thick line) PS in solution with the BP and PBN in toluene after thirty minutes of concurrent irradiation.

The involvement of the PS in the reaction suggests but does not prove that BP prefers

hydrogen abstraction from the tertiary carbon on the PS backbone. The abstraction results in a

radical on the PS and on BP. The PS radical is indistinguishable from toluene radical; the lack of

increase in the toluene radical signal implies that PBN is unable to bond the PS radical for steric

reasons. The radical on the BP leads to an increase in the ketone peak, which proves PS plays a

role in the reaction preferentially to toluene, and that the result of the reaction is a radical on

the BP, as that is the only source of ketones for PBN. The only PS bond whose dissociation

energy is less than that of the primary carbon on toluene is the tertiary carbon. The difference

is 27.4 kJ/mol, which is low enough to allow the continued generation of toluene radicals,

344 345 346 347 348 349 350

0

B (mT)

Page 163: Photochemical Crosslinking Reactions in Polymers

142

though at a lower rate as there is a more energetically favored competing reaction. The

dissociation energy required for the secondary PS hydrogen is an additional 34.7 kJ/mol above

that of toluene (22, 23). It does demonstrate that PS participates in the reaction in the same

way as toluene, and that they both generate radicals on the BP. There are no additional peaks

from the spin trapping of new species. Further investigation into the reaction is required, but

EPR data proves that there is a reaction between BP and PS that generates BP radicals, and that

it is preferred over available reactions with toluene. For the binary thin films in Chapter 4, we

have proven that there is a reaction between BP and PS and that it generates phenyl or

diphenyl radicals, which is only feasible for a hydrogen abstraction mechanism.

5.4. Conclusion

In this chapter we use EPR to study the radicals generated by combinations of BP, PS,

and PBN in toluene in hopes of determining the location of the radicals on the PS. We proved

that radicals are generated and a reaction occurs. The reaction is accelerated by the presence

of BP and UV irradiation and the hyperfine interactions correspond to the formation of a phenyl

or diphenyl radical. Radicals form on toluene, BP, and PS through hydrogen abstraction, and a

mechanism that an unsuccessful literature search suggests is novel strips the ketone from BP

and attaches it to PBN. The introduction of PS increases the ketone peak without increasing the

phenyl peak, which proves BP reacts preferentially with PS over toluene but PBN cannot trap

the PS radical.

Page 164: Photochemical Crosslinking Reactions in Polymers

143

5.5. Further Directions

Low temperature studies of the BP and toluene reaction to fully understand the

mechanism and allow the diphenylmethyl radical to be seen alone.

Further investigations of the PS and BP reaction using d8-tolune and benzene to isolate

the reaction pathway that leads to macromolecular crosslinks.

Page 165: Photochemical Crosslinking Reactions in Polymers

144

5.6. References

1. A. V. Kulikov et al., Russian Chemical Bulletin 51, 2216 (Dec, 2002). 2. W. G. Wamer, J. J. Yin, R. R. Wei, Free Radical Biology and Medicine 23, 851 (1997). 3. K. Vanheusden et al., J. Appl. Phys. 79, 7983 (May, 1996). 4. B. Yamada, D. G. Westmoreland, S. Kobatake, O. Konosu, Progress in Polymer Science 24, 565 (1999, 1999). 5. M. Kamachi, Advances in Polymer Science 82, 207 (1987). 6. B. J. Qu, Y. H. Xu, W. F. Shi, B. Ranby, Macromolecules 25, 5215 (Sep 28, 1992). 7. D. R. Bauer, J. L. Gerlock, in Advanced ESR Methods in Polymer Research, S. Schlick, Ed. (John Wiley & Sons, Inc., Hoboken, NJ, 2006). 8. J. L. Gerlock, D. F. Mielewski, D. R. Bauer, Polym. Degrad. Stabil. 20, 123 (1988, 1988). 9. J. Pilar, Macromolecules 33, (2000). 10. J. Pilar, J. Labsky, Macromolecules 36, 913 (Feb 11, 2003). 11. R. F. Boyer, P. L. Kumler, Macromolecules 10, 461 (1977). 12. E. G. Janzen, Accounts Chem. Res. 4, 31 (1971). 13. J. Pilar, Macromolecules 36, (2003). 14. S. Stoll, A. Schweiger, J. Magnetic Resonance 178, 13 (2006). 15. V. Sciannamea, J. Poly. Sci. A 47, 1085. 16. D. Cunliffe, J. E. Lockley, J. R. Ebdon, S. Rimmer, B. J. Tabner, Macromolecules 34, 3882 (Jun 5, 2001). 17. G. R. Buettner, Free Radical Biology and Medicine 3, 259 (1987, 1987). 18. E. G. Janzen, B. J. Blackburn, J. Am. Chem. Soc. 91, 4481 (1969/07/01, 1969). 19. A. Halpern, J. Chem. Soc., Faraday Trans. 1 83, 219 (1987). 20. D. R. Dalton, S. A. Liebman, H. Waldman, R. S. Sheinson, Tetrahedron Lett. 9, 145 (1968). 21. A. R. Bassindale et al., Tetrahedron Lett. 14, 3185 (1973). 22. N. P. Cheremisinoff, Handbook of Polymer Science and Technology: Synthesis and properties. (CRC Press, ed. 1, 1989), vol. 1, pp. 310. 23. in Comprehensive Handbook of Chemical Bond Energies, Y. Luo, Ed. (CRC Press, 2007), pp. 19-145.

Page 166: Photochemical Crosslinking Reactions in Polymers

145

6. Chapter 6: Preliminary Investigations of other Polymer and Crosslinker Systems

Previous chapters have delved deeply in to the photochemistry of benzophenone and its

interactions with polystyrene and poly(normal-butylacrylate). In this chapter we examine other

polymer-crosslinker systems, as discussed in Chapter 3. The following is a survey of the results.

Each polymer and crosslinker combination is examined for changes in the molecular weight

distribution (MWD). Changes are linked back to the discussion in Chapter 3 and further

directions are noted.

These investigations are incomplete. They were intended as a survey of the interactions

between the polymers and additives, but they instead revealed issues in both the experimental

apparatus and the materials. A new, more intense UV lamp was employed in these

investigations. This new lamp increased the photon dose per unit time in order to speed up the

reactions. It will be shown that the wavelength range of the lamp was too broad and it most

likely included a short wavelength UV tail that induced transesterification and homolytic bond

cleavage. Most of the additives used contain ester groups, mostly in the bis-functional species.

The new lamp therefore introduced additional reactions that make the interpretation of the

MWD difficult. The homolytic bond cleavage introduces additional end radicals into the

systems which cause macroradical recombination and a broadening of the MWD.

While the intent of the investigation was unachievable, there are some conclusions that

can be drawn from the results. In this chapter we discuss the conclusions that can be reached

from the data that is available. These experiments were meant as preliminary investigations;

Page 167: Photochemical Crosslinking Reactions in Polymers

146

they do not provide the information they were designed for, but some validation of additive

choice can be achieved. Polyisobutylene is not discussed because it was found to have a UV-

active polymeric contaminant. Poly(∝-methylstyrene) is not discussed because all its results

were inconclusive due to a complicated reaction in the control.

6.1. Materials

Polystyrene (PS) (Mn=46000, PDI=1.03, Waters Associates), poly(normal-butyl acrylate)

(PnBA) (Mn=60000 PDI=1.08, Polymer Source, Inc.), poly(1,4 butadiene) (PBD) (Mn=37500 PDI

1.03, Polymer Source, Inc), polyisobutylene (PIB) (Mn=47000 PDI=1.3, Polymer Source, Inc),

poly(∝-methylstyrene) (PamS) (Mn=32300 PDI=1.02, Polymer Source, Inc), and

polymethylmethacrylate (PMMA) (Mn=34000 PDI=1.06, Waters Associates), HPLC-grade

Toluene (Pharmco-Aaper) , Benzophenone (BP) (Fischer Scientific), Xanthone (XAN) (Fischer

Scientific), Phthalimide (PTH) (Fischer Scientific), and Phenylazide (FEN) (Sigma-Aldrich) were

used as received after purification.

The bi-functional bis-benzophenone adduct (BP-BP), bi-functional bis-xanthone adduct

(XAN-XAN), bi-functional bis-Phthalimide adduct (PTH-PTH), and bi-functional bis-phenylazide

(FEN-FEN) were synthesized following similar methods to those reported previously (1) by our

collaborator Jeffrey Lancaster(2).

6.1.1. Substrate Preparation

Silicon Wafers (Wafer World, Inc) were cut into 1cm2 squares and placed in Piranha

solution (3:1 H2SO4:H2O2 DANGER: Piranha is extremely dangerous) overnight. Wafers were

Page 168: Photochemical Crosslinking Reactions in Polymers

147

washed first in DI H2O and then in ethanol, and were subsequently dried with N2 in a clean

hood. After treatment in a UV-Ozone Cleaning System (UVOCS, ,Inc) for 20 minutes, the wafers

were again washed in DI H2O and Ethanol and dried with nitrogen in a clean hood. The cleaning

procedure removes all organic contaminants and produces a fresh oxide layer on the silicon.

6.1.2. Solution Preparation

Solutions of 20mg/ml polymer in Toluene were agitated with a vortexer for 30 minutes

and split into scintillation vials containing approximately 0.5mL of polymer solution each. Each

vial contained a single polymer. One vial was set aside as a control. Mono-functional additives

were added in a molar ratio of 8:1, while bis-functional additives were added in a molar ratio of

4:1. The molar ratios for the bis-functional additives were kept at half those of the mono-

functional to keep the overall content of the additive moieties equal.

6.1.3. Sample Preparation

Polymer solutions were spin coated (Laurell Technologies Co. WS-400A-6NPP/LITE) onto

the silicon substrates at 2000 rpm for 1 minute. The film thicknesses measured by ellipsometry

(alpha-SE, J.A. Woolam, Co.) were 130±20nm for all samples. Samples were irradiated with a

350nm ultraviolet source (UVP, Inc, 550 �W/cm2) under an Argon blanket in quartz-topped

chambers for 0min (control), 1 hr, 3 hr, 6 hr. The source was run for 30 min before the start of

each experiment to eliminate startup transients. Specimens were irradiated on a rotating stage

to optimize the homogeneity of irradiation.

Page 169: Photochemical Crosslinking Reactions in Polymers

148

6.1.4. GPC analysis

After irradiation, specimens (i.e. films on cut silicon wafers) were stored in scintillation

vials in a dark drawer to eliminate the possibility of further reactions. No differences in GPC

response were observed between a fresh specimen and a specimen stored in this fashion for 5

days confirming the appropriateness of this storage method.

Solutions for GPC analysis were prepared by placing 0.100 mL of HPLC toluene and 0.100 mL

of HPLC THF in each vial and followed by shaking on a vortexer for 30 min. 0.05mL of a 1mg/mL

solution of 2350 Dalton PS in THF was added as an internal standard subsequent to the shaking.

Solutions were then filtered through a 0.2 micrometer PTFE syringe filter and injected into the

GPC (2 Polymer Laboratories’ ResiPore™ Columns and 1 Polymer Laboratories’ MesoPore™ in a

Shimadzu HPLC system: LC-10ATvp pump, CTO-10ACvp column oven, SPD-10a UV-Vis detector,

RID-10A RI detector, SCL-10Avp system controller and a Wyatt miniDAWN TREOS static light

scattering detector). Analysis was performed using custom MATLAB™ programs and Origin

8.0™. Molecular weights were confirmed using a Wyatt miniDAWN light scattering detector.

dn/dc was measured using Wyatt miniDAWN in batch mode at 657nm as well as Brookhaven

Instruments differential refractometer (BI-DNDCW) at 535nm to determine wavelength

dependence (Section 3.4). GPC data were calibrated using linear PS standards (EasiVial PS-

M(2mL), Varian). Absolute polymer molecular weights were determined from the PS calibrated

values by correcting for hydrodynamic volumes using tabulated Mark-Houwink constants

(Section 3.5).

Page 170: Photochemical Crosslinking Reactions in Polymers

149

6.2. Polystyrene and Other Crosslinker Chemistries

The polystyrene and benzophenone interactions have been well characterized in previous

chapters. Here we expand the investigation to the reactions between polystyrene and the

crosslinkers based upon phenylazide and phthalimide. Polystyrene shows no reaction in the

control.

6.2.1. Polystyrene and Phenylazide

Phenylazide, as discussed in section 3.3.4 reacts through nitrene insertion (Figure 3.13).

It does not generate radicals on the polymer; instead it inserts into an C-H bond. Polystyrene

has seven C-H bonds per monomer. The C-H bond on the tertiary carbon has the lowest bond

dissociation energy (BDE). There are no mechanisms for scission, and only FEN-FEN can form

crosslinks. The mono-functional FEN can insert into the chain, but it neither generates radicals

nor participates in other reactions.

Figure 6.1: PS and FEN (left) and PS and FEN-FEN (right) after 0 (dotted), 3 (dashed), and 6 (solid) hours of irradiation. Inset graphs are the full peak of the respective plot.

The results of six hours of irradiation are shown in Figure 6.1. The MWD of FEN and

FEN-FEN both increase and decrease; scission and combination are occurring. FEN-FEN

produces a much larger number of crosslinks than FEN. Additional reactions are occurring

100 200 300 400 500 600 700 800 900 1000 1100 12000

Number of Monomers per Chain (N)

Norm

aliz

ed N

um

ber

of C

hain

s

100 200 300 400 500 600 700 800 900 1000 1100 12000

Number of Monomers in Chain (N)

Norm

aliz

ed N

um

ber

of C

hain

s

100 200 300 400 500 600 700 800 900 1000 1100 12000

Number of Monomers per Chain (N)

Norm

alize

d N

um

ber

of C

hain

s

100 200 300 400 500 600 700 800 900 1000 1100 12000

Number of Monomers in Chain (N)

Norm

alize

d N

um

ber

of C

hain

s

Page 171: Photochemical Crosslinking Reactions in Polymers

150

within the system. The nitrene insertion reaction (Figure 3.13) does not generate radicals and

FEN is not a bivalent molecule. Additional reactions that have not been previously accounted

for are the source of the reaction in the FEN system. If the additional reactions are responsible

for the crosslinking and scission in the FEN system, then FEN-FEN crosslinking may be partially

responsible for the differences between the FEN and FEN-FEN systems. Further investigations

into the FEN system are required to identify the source of the MWD changes.

6.2.2. Polystyrene and Phthalimide

Phthalimide, as discussed in section 3.3.3, self-catalyzes radical generating reactions

that generate polymer radicals in the deactivation of PTH (Figure 3.11). The PTH-PTH does not

react in polymer systems; the only radical mechanism lies with the PTH. Proton acceptors are

required to begin the photochemical reaction. No proton acceptors are present in this system,

so we do not expect a reaction in PS unless there is contamination.

Figure 6.2: PS and PTH and PS and PTH-PTH after 0 (dotted), 3 (dashed) and 6 (solid) hours of irradiation.

The results of six hours of irradiation are above in Figure 6.2. There is no reaction. Both

PTH and PTH-PTH are inert in PS binary systems. PTH can react with toluene derivatives (3);

100 200 300 400 500 600 700 800 900 1000 1100 12000

Number of Monomers per Chain (N)

Norm

aliz

ed N

um

ber

of C

hain

s

Page 172: Photochemical Crosslinking Reactions in Polymers

151

none of that occurs here. This provides more evidence that there are no impurities or

contaminates in our systems. The need for proton acceptors to begin the reaction allows for

intriguing design possibilities. Self-healing polymer films may be feasible with PTH-PTH and

provide one avenue for further investigation.

6.3. Poly(normal-butylacrylate) and Other Crosslinker

Chemistries

The poly(normal-butylacrylate) and benzophenone interactions have been well

characterized in previous chapters. Here we expand the investigation to the reactions between

poly(normal-butylacrylate) and the crosslinkers based upon xanthone, phenylazide and

phthalimide.

PnBA has a reaction in the control in this new system. There were no reactions after 9

hours in the systems of Chapter 4, but with the more intense light used in this study there are

crosslinking reactions. The reaction after 6 hours is visible in Figure 6.3. After 3 hours, there is

scission but little crosslinking. The intensity of the UV lamp used in this study, which was

intended to speed up the reaction time without affecting the reactions, is most likely unzipping

the PnBA by generating radicals on the chain due to a short UV wavelength tail.

Page 173: Photochemical Crosslinking Reactions in Polymers

152

Figure 6.3: PnBA after 0 (dotted), 3 (dashed) and 6 (solid) hours of irradiation

The lowest BDE is on the side chain (section 3.2.5). Forming a radical at that location would

allow crosslinking without scission, which is visible in the 6 hour data. This was also a new

batch of PnBA, so there may have been contaminants or impurities that were not present in the

previous batch.

The control reaction was more extensive than most of the reactions with other

crosslinkers. Figure 6.4 is a representative graph. The PnBA control irradiated for 6 hours

shows more reaction than the PnBA and FEN system after 6 hours of irradiation. The same

trend is visible in the systems with FEN-FEN, XAN, and XAN-XAN.

200 400 600 800 1000 1200 1400 1600 1800 20000

Number of Monomers in Chain (N)

No

rmaliz

ed

Nu

mber

of C

ha

ins

Page 174: Photochemical Crosslinking Reactions in Polymers

153

Figure 6.4: PnBA control (dotted), PnBA and FEN after 6 hours (dashed) and PnBA control after 6 hours (solid).

The decrease in the control reaction extent due to additive inclusion occurs because the

additives are inhibiting the reaction in some way. Both the XAN and FEN additives inhibit the

reaction to the same extent, as do their bis-functional adducts, despite different crosslinking

chemistries. Figure 6.5 demonstrates that the PTH additives do not affect the reaction.

Figure 6.5: PnBA control (dotted), PnBA and PTH (dashed), and PnBA and PTH-PTH (solid) after 6 hours irradiation.

200 400 600 800 1000 1200 1400 1600 1800 20000

Number of Monomers in Chain (N)

Norm

aliz

ed N

um

ber

of C

hain

s

200 400 600 800 1000 1200 1400 1600 1800 2000

0

Number of Monomers in Chain (N)

Norm

aliz

ed N

um

ber

of C

hain

s

Page 175: Photochemical Crosslinking Reactions in Polymers

154

The additives that react to inhibit the reaction all contain an ester group. PTH and PTH-PTH do

not. Transesterification, Figure 6.6, is most likely the competing reaction. The new lamp must

have a wavelength tail that includes short UV wavelengths that induce transesterification with

trace water or trace oxygen.

Figure 6.6: Transesterification.

The PnBA systems do not allow us to draw conclusions about the additive-polymer

reactions. Instead we discover that the higher intensity lamp that was intended to explore

higher photon doses in shorter times has a short UV tail that causes transesterification with

trace water or oxygen. This mechanism may also account for the additional reactions seen in

the PS and FEN systems (section 6.2.1).

6.4. Poly(1,4-butadiene) and Crosslinkers

Here we investigate the reactions between poly(1,4-butadiene) and the crosslinkers

based upon benzophenone, xanthone, phenylazide and phthalimide. Poly(1,4-butadiene) has a

reaction in the control system. This reaction is again most likely due to a lamp wavelength tail.

The plot of the MWD after the irradiation is included in Figure 6.7.

Benzophenone generates radicals through hydrogen abstraction. We predicted (section

3.2.6) that the PBD double bond would not allow hydrogen abstraction because the BDEs were

R1O R

OR2 OH+

R2O R

OR1 OH+

Page 176: Photochemical Crosslinking Reactions in Polymers

155

too high. Figure 6.7 proves this true, as the addition of BP and BP-BP to the system has no

effect on the PBD control reaction.

Figure 6.7: PBD control (dotted) and after 6 hours (solid), PBD and BP (dashed) and PBD and BP-BP (dashed) after 6 hours.

This should also be true of XAN and XAN-XAN, as they also crosslink using hydrogen abstraction.

We therefore expect all hydrogen abstracting mechanisms to be inert in PBD for our purposes.

Other mechanism may have better results. The results of the investigations with other

crosslinkers were inconclusive. No other conclusions could be drawn.

6.5. Polymethylmethacrylate and Crosslinkers

Here we investigate the reactions between polymethylmethacrylate and the crosslinkers

based upon benzophenone, xanthone, phenylazide and phthalimide. PMMA has a reaction in

the control (Figure 6.8) that is purely a combination reaction.

200 400 600 800 1000 1200 1400 1600 1800 20000

Number of Monomers in Chain (N)

Norm

aliz

ed N

um

ber

of C

hain

s

Page 177: Photochemical Crosslinking Reactions in Polymers

156

Figure 6.8: PMMA control before (dotted) and after (solid) 6 hours of irradiation

The lack of scission in the control reaction indicates that the reaction occurs on the

methacrylate pendant group. If the reaction occurred in the backbone, the polymer chains

would split and scission would occur. The crosslinking reaction is most likely a

transesterification involving the methacrylate pendant group. This would explain the lack of

scission in the control reaction. PMMA does not show any additional reactions in XAN, XAN-

XAN, FEN, or FEN-FEN. These additives all have ester groups on them.

6.5.1. Polymethylmethacrylate and Benzophenone

Benzophenone reacts through hydrogen abstraction. In section 3.2.4 we predicted that

the tertiary carbon on the methacrylate pendant group was the most likely location for

hydrogen abstraction because it had the lowest BDE of the PMMA C-H bonds. The plot of the

PMMA MWD with both BP and BP-BP (Figure 6.9) shows that there is reaction beyond the

control reaction in benzophenone systems. BP shows an increase in MW above that of the

control, and BP-BP forms a new peak. In both cases there is very little if any scission,

supporting the idea that all the reactions in the PMMA systems occur on the pendant group.

200 400 600 800 1000 1200 1400 1600 1800 20000

Number of Monomers in Chain (N)

Norm

aliz

ed N

um

ber

of C

hain

s

Page 178: Photochemical Crosslinking Reactions in Polymers

157

Figure 6.9: PMMA and BP-BP (dashed), PMMA and BP (bold line), and PMMA (line) after 6 hours irradiation. Dotted line is control.

BP and BP-BP, as sources of radicals in the system, are the most likely of all the additives to

induce β-scission of the alkyl backbone; this does not occur. BP and BP-BP in PMMA is an

excellent candidate for further investigations of purely crosslinking reactions in polymer thin

films. The other hydrogen abstractor, Xanthone and its derivative, showed no reactions in

PMMA.

6.5.2. Polymethylmethacrylate and Phthalimide

Phthalimide, as discussed in section 3.3.3, reacts with proton acceptors and then self-

catalyses a SET reaction that generates a phthalimide radical that can hydrogen abstract. These

reactions are not possible in the PTH-PTH because the alkyl chain linker is connected to the

secondary amine required for the SET reaction. Crosslinks in PTH systems are therefore the

result of macroradical recombination and should only produce four arm stars unless backbone

scission occurs. Scission of the backbone is unlikely in PMMA, as radicals are generated on the

200 400 600 800 1000 1200 1400 1600 1800 20000

Number of Monomers in Chain (N)

Norm

aliz

ed N

um

ber

of C

hain

s

Page 179: Photochemical Crosslinking Reactions in Polymers

158

pendant groups. Figure 6.10 shows the results of the reactions between PMMA and PTH, and

PMMA and PTH-PTH. There is no significant scission.

Figure 6.10: PMMA control (dotted) and control, PMMA and PTH (bold line), and PMMA and PTH-PTH (dashed) after 6 hours irradiation.

The mono-functional PTH generates more combination reactions than the control

reaction, and the bis-functional PTH-PTH shows little to no reaction as would be expected by

the predicted PTH reaction scheme. The presence of the PTH reaction indicates that there are

proton acceptors in the PMMA system. Transesterification is not a good candidate for

generating proton acceptors unless it generates a radical, which is unlikely. Short UV

interaction with trace oxygen or water is a more likely source for the generation of proton

acceptors.

6.6. Conclusions

The preliminary investigations were hampered by a change in the UV lamp. The higher

intensity introduced a short wavelength tail that caused reactions within the control samples.

100 200 300 400 500 600 700 800 900 1000 1100 12000

Number of Monomers in Chain (N)

No

rma

lized N

um

be

r of C

hain

s

Page 180: Photochemical Crosslinking Reactions in Polymers

159

Despite this, the data provided evidence for a number of conclusions that further our

understanding of the polymer and additive systems under investigation.

The polystyrene controls did not react and we were able to examine its reactions with

phenylazide and phthalimide. Polystyrene reacts in both the FEN and FEN-FEN systems, which

contradicts expectations and is only possible if another reaction is present. The larger amount

of crosslinking in the FEN-FEN system indicates that the FEN-FEN is forming covalent bridges

between macromolecules as it was designed to do and bis-functional phenylazide is a good

crosslinker for PS. The additional reaction was exposed by the PnBA systems.

The PnBA control reacted, while it did not in the previous investigations with the other

UV lamp. The lamp is therefore the cause of these additional reactions and the incompleteness

of these investigations. The reactions between PnBA and both the xanthones and the

phenylazides did not reach the same extent as the control reactions, indicating that the

additives compete with the additional reaction in the control. Neither PS nor PnBA reacted

with the phthalimides. The competing additive reactions reached the same extent after the

same amount of time, indicating the reaction is the same even though the xanthones and

phenylazides are designed to crosslink using different chemistries. Transesterification is

therefore the hypothesized additional reaction in the control, as PnBA, the xanthones and the

phenylazides have ester groups and the phthalimides do not. Transesterification requires

alcohols; trace oxygen and water in combination with short UV irradiation from the new lamp

both have mechanisms that allow transesterification.

Page 181: Photochemical Crosslinking Reactions in Polymers

160

The PBD control also reacted, and the addition of benzophenone derivatives did not

affect the reaction. The reaction is therefore not a radical reaction. PBD is a vinyl polymer, and

the addition of short UV has the potential to produce many reactions. The results in other

additive systems were therefore inconclusive.

PMMA reacted without additives; this further supports the transesterification

hypothesis as PMMA also includes an ester. The fact that no reaction in PMMA, whether with

or without additive, produced scission is further confirmation. Transesterification of the

pendant ester would not affect the polymer backbone. PMMA reacted with BP and BP-BP, with

BP-BP generating more macroradical combination than BP. Benzophenone can therefore be

used to crosslink PMMA. The BP and BP-BP radical reactions also did not generate scission

product. Hydrogen abstraction is therefore also occurring on the pendant methacrylate group

as was predicted by the group contribution calculations, and not the backbone methyl. A

radical on the backbone methyl could migrate to the secondary backbone carbon and cause

scission of the chain. PMMA’s reaction with phthalimide further supported the

transesterification hypothesis and proved, as predicted, that only the bis-functional phthalimide

species reacts to produce crosslinks.

These studies did not achieve what they set out to, but they provided important insights

into both the experimental setup and the polymer and additive reactions. The fact that none of

these reactions were seen in the thin films of Chapter 4 further validates those results.

Page 182: Photochemical Crosslinking Reactions in Polymers

161

6.7. Further Directions

Repeat these experiments with only low intensity, long wavelength (>300nm) UV to

confirm the results that suggest polymer-additive reactions.

Study macromolecule combination-only reactions and gain further insight into the early

stages of polymer network formation using the bis-functional phthalimide or

phenylazide additives. PMMA, is a prime candidate as there are no scission reactions

with the additive.

Study combinations of proton acceptors and bis-functional phthalimide to create self-

healing polymers.

Page 183: Photochemical Crosslinking Reactions in Polymers

162

6.8. References

1. G. T. Carroll, M. E. Sojka, X. Lei, N. J. Turro, J. T. Koberstein, Langmuir 22, 7748 (2006). 2. J. R. Lancaster, Columbia University (2011). 3. O. l. Michael, G. Axel, in CRC Handbookof Organic Photochemistry and Photobiology,

Volumes 1 & 2, Second Edition. (CRC Press, 2003).

Page 184: Photochemical Crosslinking Reactions in Polymers

163

7. Chapter 7: Conclusions

This thesis explored the photochemical crosslinking of polymers using photoactive

additives. Additives are used extensively in the post-synthesis modification of polymer

properties; this work focused on photocrosslinkers active in the ultraviolet. A survey of

common polymers was performed and the bond dissociation energies for the carbon-hydrogen

bonds on the monomers calculated. The calculation of bond dissociation energies allowed the

location of hydrogen abstraction and other chemical reactions to be predicted. A survey of

both mono- and bi-functional crosslinkers that utilize various reactive chemistries was

performed. Hydrogen abstraction, using benzophenone, was of primary interest as it has been

used extensively previously. The bond dissociation energies predicted that benzophenone

would preferentially hydrogen abstract from the backbone in polystyrene and from the

pendant group in poly(normal-butylacrylate). These two polymers were chosen for further, in-

depth investigations because they were the only two polymers that had clearly preferred

abstraction locations. Polystyrene is a glass and poly(normal-butylacrylate) is a rubber at

experimental temperatures. This provided further incentive, as the experiments could

investigate multiple diffusion regimes.

Polystyrene and poly(normal-butylacrylate) thin films were irradiated for multiple

lengths of time, with multiple ratios of additive to polymer, and with two additive

functionalities. The data, in conjunction with modeling of the scission and combination

reactions, indicated that the functionality of the additive is only important at high molar ratios.

Page 185: Photochemical Crosslinking Reactions in Polymers

164

Below the threshold, the functionality does not play a role in the reaction because diffusion and

kinetic considerations outweigh probabilistic ones. Instead, the higher diffusion rate in mono-

functional additive leads to more self-quenching of the triplet state due to aggregation and a

lower density of radicals in the system. Above the threshold, probabilistic considerations

dominate and a covalent macroradical bridging mechanism appears in the bi-functional

additive. The density of radicals above the threshold is so high that the normally second-order

macroradical recombination reactions become pseudo first-order.

The radical reactions in benzophenone-polystyrene systems were studied using EPR to

identify the location of the reaction and confirm the prediction of the bond dissociation

energies. The EPR studies confirmed that polystyrene is the preferred hydrogen abstraction

location when competing with toluene, but identification of the exact bond was prevented by a

previously unseen reaction between the spin trap and benzophenone. The reaction was shown

to strip the ketone from benzophenone and transfer it to the spin trap, generating a diphenyl

radical.

A broad survey of polymers and crosslinkers was then performed. The survey indicated

the bi-functional phenylazide was a good candidate for investigating combination mechanisms

without scission mechanisms present. Polymethylmethacrylate was also shown to ideal for

those investigations, as it preferentially reacted through the pendant group, as did poly(normal-

butylacrylate). Transesterification was also shown to be a viable alternative to additives when

designing systems to crosslink.

Page 186: Photochemical Crosslinking Reactions in Polymers

165

8. Chapter 8: Implications for Future Work

Throughout this thesis we have explored crosslinking reactions in a number of polymers

using a number of different additive chemistries. In Chapter 3, we used rational design

methodologies to predict the crosslinking location in a number of polymers for a number of

different chemistries. It was found that the bond dissociation energies and sterics allow the

prediction of reaction location; the EPR experiments in Chapter 5 confirmed some of the

predictions.

The location of the reaction plays a large role in the resulting network of reactions.

When the hydrogen is abstracted from the polymer backbone, significant polymer scission

occurs. But systems in which the abstraction occurs on the pendant group show no scission.

Non-radical based chemistries also showed a lack of scission. The smart design of polymer and

additive systems therefore allows one to design the exact network of reactions available to the

systems. Polystyrene, for example, is used extensively as a biocompatible material. We found

that radical-based crosslinking reactions generate significant scission products. But if the

polystyrene were modified by, for example, a para-methyl group, then the crosslinking

reactions would occur solely on the pendant group and scission would not be a viable reaction,

making the material safer for biological applications. Other chemistries, such as nitrene

insertion, are also available to generate crosslinks without the addition of radicals or other,

undesired reactions in the system.

In Chapter 4 we found that the functionality of the additive is only important at high

molar ratio of additive to polymer because that corresponds to high radical densities. This

Page 187: Photochemical Crosslinking Reactions in Polymers

166

applies to all radical-generating crosslinkers including hydrogen abstractors. The covalent

bridging mechanism that is often hypothesized to cause the property changes in polymer-

additive systems does not play a large role; instead it is the density of radicals that is important.

We would therefore expect high densities of mono-functional additives to produce the same

results as bi-functional additives. At low molar ratios, the functionality is even less important,

and the same results can be achieved irrespective of functionality. Our probability calculations

show that increasing the functionality of radical generating additives is a self-defeating process,

as the probability of even a single covalent bridge is low, and increasing the functionality makes

the probability even less. Mono- and bi-functional additives are therefore all that are required

for crosslinking reactions.

In summary, this thesis shows how best to design photoactive crosslinker systems. To

maximize the crosslinking and minimize scission, design the polymer and additive chemistry

such that the reaction occurs on the pendant group. The additive functionality is then not

important unless the chemistry requires two ends to form bonds between chains.

Page 188: Photochemical Crosslinking Reactions in Polymers

167

9. Appendix

The appendix begins with an explanation of the code used to calculate the Hamielec

Scission and the recombination models followed by the code itself. Then the appendix

continues with an explanation of the simulation principles used in the design of the simulation

code that is included later in the appendix. The files were used to simulate scission and

combination. In the end, their results were not used in the thesis but they were used in model

development, testing, and comprehension and served as confirmation of Hamielec results.

Aspects of the simulation were written with the advice of John Zhang.

The appendix then includes the computer code used in the data analysis and simulation

of the HPLC-SEC data. The code was written by Nicholas Carbone over the course of this thesis.

Code consists of custom MATLAB® m-files and was run in MATLAB® version 7.10. MATLAB® and

all related patents, copyrights and trademarks are the property of The Mathworks, Inc. and

were used under license.

EPR data was analyzed using the instrument software (Bruker WINEPR System v2.1 and

Bruker WINEPR SimFonia v1.25) and the EasySpin MATLAB© toolbox. These software suites

also provided all simulations. For more details on the commercial software, see their

respective websites.

M-files are listed individually, with a page break between them. M-files with

subfunctions have the subfunctions below with the names underlined. In MATLAB, code

preceded by a % is a comment.

Page 189: Photochemical Crosslinking Reactions in Polymers

168

Appendix Table of Contents

A. HAMIELEC MODELING ................................................................................................................................ 170

a. Modeling random macroradical �-scission (from Chapter 4) ............................................................. 170

b. Modeling random macroradical recombination ................................................................................ 170

c. Linking the model to HPLC-SEC data ................................................................................................. 171

B. HAMIELEC M-FILES.................................................................................................................................... 171

Hamielec.m ........................................................................................................................................................... 171

FitShape.m ....................................................................................................................................................... 177

FitPeaks.m ........................................................................................................................................................ 179

MacroRadicalCombination.m ............................................................................................................................ 181

AreaofPeaks.m.................................................................................................................................................. 182

Probabilities.m .................................................................................................................................................. 183

C. SIMULATION PRINCIPLES ............................................................................................................................. 184

D. SIMULATION M-FILES ................................................................................................................................. 189

CombinationScissionwithRg.m................................................................................................................................ 189

RadiusOfGyration.m .......................................................................................................................................... 198

E. DATA IMPORTATION M-FILES ....................................................................................................................... 206

GetGPCData.m ....................................................................................................................................................... 206

EPR.m .................................................................................................................................................................... 208

F. DATA PREPARATION M-FILES ....................................................................................................................... 210

TransformX.m ........................................................................................................................................................ 210

TransformPStoOther.m .......................................................................................................................................... 213

BaselineChoice.m ................................................................................................................................................... 214

FitLine.m .......................................................................................................................................................... 216

SubtractTail.m .................................................................................................................................................. 217

TheFit.m ........................................................................................................................................................... 219

G. DATA ANALYSIS M-FILES ............................................................................................................................. 221

Page 190: Photochemical Crosslinking Reactions in Polymers

169

Histogram.m .......................................................................................................................................................... 221

AreaNormalize.m ................................................................................................................................................... 223

HeightNorm.m ....................................................................................................................................................... 224

MolWDist.m .......................................................................................................................................................... 225

SameX.m ............................................................................................................................................................... 226

List of Appendix Figures

FIGURE 9.1: SIMULATION FLOWCHART ........................................................................................................................ 187

FIGURE 9.2: COMPARISON OF DATA AND SIMULATION RESULTS FOR PURE LINEAR CHAINS (RIGHT BLACK DOTS, G=1) AND BRANCHED

CHAINS (LEFT BLACK DOTS, G=VARIABLE) ASSUMING 10% OF ORIGINAL CHAINS COMBINED. DATA (LINE) IS 16:1 BIS:PS IRRADIATED

FOR 9 HOURS. ............................................................................................................................................... 188

Page 191: Photochemical Crosslinking Reactions in Polymers

170

A. Hamielec Modeling

a. Modeling random macroradical �-scission (from Chapter 4)

Chapter 4 describes the derivation and assumptions behind the model. Here we discuss

the implementation. Saito’s equation in Hamielec’s formulation adjusted for our use is

(7.1)(2.18):

�(�,�)−�(�,0)exp (−��)= exp (−��)���2 + �(� − �)

��(�,0)��

(7.1)

The second term on the left is the original weight fraction distribution. In these equations, the

degree of scission � is defined as the number of chain scission events per number of monomers

and is a time and kinetics independent radiation dose analog. � is derived by fitting the LHS of

equation (7.1). The data, �(�, �), is compared to the control, �(�, 0), and the exponential

term is fitted using a sum of least squares algorithm. FitShape.m is the sub program used.

As we’re only interested in fitting the differences, we ignore the first term of equation

(7.1) and use a for loop to calculate the integral (assuming we can transition back to a discrete

function without loss of solution quality). The calculated distribution is a weight fraction; we

convert the data to a weight fraction to compare it to the model. We arbitrarily scale it on the

yaxis for best fit using FitPeaks.m.

b. Modeling random macroradical recombination

For the combination modeling, the approach is simpler. Equation (7.2) is a simple

nested for loop over � and �. We arbitrarily scale it on the y axis for best fit.

Page 192: Photochemical Crosslinking Reactions in Polymers

171

�(�)∝ �(�)= ��(� − �)�(�)�(� − �)

���

= ��(�)�(� − �)

���

(7.2)

The entire recombination distribution is therefore the sum of equation (7.2) over all �.

c. Linking the model to HPLC-SEC data

The model uses number and weight fraction distributions of polymer chains of length �.

The experimental distributions from the HPLC-SEC must be converted into these distributions.

As discussed in the section on the HPLC-SEC detectors (section 2.2.2), RIDs measure the number

of chains of length � and UVDs measures the number of chromophores. All data used in this

paper was first converted to the number of chains of length �. If a different distribution was

needed for modeling, the data was converted to the needed distribution before calculations. If

� is the detector signal and � is the number fraction of chains of degree of polymerization, �,

then the following conversions were used before graphing and analysis:

����(�)=����(�) �⁄

∑(����(�) �⁄ )

����(�)=����(�)

∑ ����(�)

(7.3)

B. Hamielec M-Files

These files was used to simulate scission using the equations in (5) as well as the

combination reactions as described in the thesis.

Hamielec.m

This file is the heart of the Hamielec scission and combination model. It changes the

input data to number and weight fractions, then approximates the degree of scission u and

Page 193: Photochemical Crosslinking Reactions in Polymers

172

from there models the scission. It then uses the modeled scission to calculate end-to-end

combinations, the formation of 3-arm stars, and the formation of 4-arm stars. Once the

scission and combination have been modeled, it fits the results to the data.

function [NumberFractions WeightFractions uwt unum ufromsubtraction]= Hamielec(xaxis,DataDist,ControlDist,range,g1,g2,g3) %This function models random chain scission from an input distribution. % %The code is based upon equations 10 and 13 from %Triacca, V.J. et al; Polymer Engineering and Science; 33(8) 1993 pg445. % %xaxis: the number of monomers per chain %DataDist: the distribution of polymer after degradation %ControlDist: the distribution of polymer before degradation %range: the full width of the distribution for normalization %g1 to g3: the g-factors for end-to-end, 3-arm, and 4-arm branches. make %them all 1 to have no effect on anything. %we don't want to incorporate species too small or too large into the %calculations. We want the distributions to reflect the GPC range. We %therefore zero all the values of the inputs outside of the range DataDist([1:range(1), range(end):size(DataDist,1)]',:)=0; ControlDist([1:range(1), range(end):size(ControlDist,1)]',:)=0; %we assume the input distributions are the number fraction. %to make sure it is normalized correctly, we renormalize the number %fraction distributions within the range of the GPC columns. If it's %already normalized correctly this won't do anything; if not this will %correct it. This will also normalize just the range we're looking at DataDistnum=DataDist./repmat(sum(DataDist,1),size(DataDist,1),1); ControlDistnum=ControlDist./repmat(sum(ControlDist,1),size(ControlDist,1),1); %after normalization the sums should be 1 disp('The number fraction normalization; data and control') [sum(DataDistnum,1) sum(ControlDistnum,1)] %we then convert to weight fraction %numerator DataDistwt=DataDistnum.*xaxis.*repmat(sum(DataDistnum,1),size(DataDistnum,1),

1); ControlDistwt=ControlDistnum.*xaxis.*repmat(sum(ControlDistnum,1),size(Contro

lDistnum,1),1); %Area normalize to produce fraction on the same interval. %This calculates and scales by the denominator DataDistwt=DataDistwt./repmat(sum(DataDistwt,1),size(DataDistwt,1),1); ControlDistwt=ControlDistwt./repmat(sum(ControlDistwt,1),size(ControlDistwt,1

),1); %after normalization the sums should be 1 disp('The weight fraction normalization; data and control') [sum(DataDistwt,1) sum(ControlDistwt,1)]

Page 194: Photochemical Crosslinking Reactions in Polymers

173

%we now have the scaled number and weight fraction of the input %distributions. We need to determine u. %first we calculate the u based upon the weight fraction, with a range that %is chosen to be a 100 points around the peak of the distributions. This %guarantees a good fit around the peak, as there fewest changes are there %due to the large number of original MW species remaining. %find the index of the max value of the control within the range already provided [~,indexwt]=max(ControlDistwt(range,:)); [~,indexnum]=max(ControlDistnum(range,:)); %use the index of the max as the center and calculate the fit to get u and %the percent depletion. [fittedwt uwt depwt]=FitShape(xaxis,DataDistwt, ControlDistwt, (range(1)

+indexwt-50):(range(1)+indexwt+50)); [fittednum unum depnum]= FitShape(xaxis,DataDistnum, ControlDistnum,

(range(1)+indexnum-50):(range(1)+indexnum+50)); %the depletion is the constant we multiply the control by to fit the max %peak of the data; we subtract from 1 to get the fractional percent that %actually reacted. depwt=1-depwt; depnum=1-depnum; %estimates is the u; fitted the fitted control peaks. %we subtract the fitted controls from the data to get the subtraction dist Subtractionwt=DataDistwt-fittedwt; Subtractionnum=DataDistnum-fittednum; %We assume everything bigger than the control peak location is combination %and everything smaller is scission. We put them in their own variables. %NOTE: The data files are stored from the largest MW (first) to the smallest MW (last). SubtractionwtComb=zeros(size(Subtractionwt,1),1); SubtractionwtScis=zeros(size(Subtractionwt,1),1); SubtractionnumComb=zeros(size(Subtractionnum,1),1); SubtractionnumScis=zeros(size(Subtractionnum,1),1); SubtractionwtComb(range(1):range(1)+indexwt,:)=Subtractionwt(range(1):range(1

)+indexwt,:); SubtractionwtScis(range(1)+indexwt:range(end),:)=Subtractionwt(range(1)+index

wt:range(end),:); SubtractionnumComb(range(1):range(1)+indexnum,:)=Subtractionnum(range(1):rang

e(1)+indexnum,:); SubtractionnumScis(range(1)+indexnum:range(end),:)=Subtractionnum(range(1)+in

dexnum:range(end),:); SubtractionwtComb=SubtractionwtComb./repmat(sum(SubtractionwtComb,1),size(Sub

tractionwtComb,1),1); SubtractionwtScis=SubtractionwtScis./repmat(sum(SubtractionwtScis,1),size(Sub

tractionwtScis,1),1); SubtractionnumComb=SubtractionnumComb./repmat(sum(SubtractionnumComb,1),size(

SubtractionnumComb,1),1); SubtractionnumScis=SubtractionnumScis./repmat(sum(SubtractionnumScis,1),size(

Page 195: Photochemical Crosslinking Reactions in Polymers

174

SubtractionnumScis,1),1); %we now have the weight average and number average subtraction plots. %As a check, we calculate u from the number average scission %in control [Mncontrol,~,~]=MolWDist(xaxis,ControlDistnum,range); %in scission [Mnscis,~,~]=MolWDist(xaxis,SubtractionnumScis,range); ufromsubtraction=(1./Mnscis)-(1./Mncontrol); %Now that we have both the weight and number fraction, and we have also %isolated the scission area of the distribution, we use the equations of %Hamielec and Saito to predict the scission from the control weight %fraction %the data is usually input as MW high-low (the first value in the array is %the highest MW, the last the lowest. We check if this is true, and if so, %flip the arrays to make it increase for loop work. if xaxis(end,1)==1 xaxisflip=flipud(xaxis); ControlDistwtflip=flipud(ControlDistwt); ControlDistnumflip=flipud(ControlDistnum); end %--------------------------------------------------------- %Hamielec Scission %determine the largest chain to go to. maxMonomerNum=xaxis(min(range)); %create variables resultDist=zeros(size(ControlDistwtflip,1),1); %use a for loop to cycle through each w(r,u) of equation 13 on the r axis for r=1:1:maxMonomerNum % another for loop to cycle through the integral from r to inf in eq13 count=1; integral=0; for s=r:1:maxMonomerNum integral=integral+(2+uwt.*(s-r))./s.*ControlDistwtflip(s,1); count=count+1; end resultDist(r,1)=uwt.*r.*(integral./count).*exp(-uwt.*r); end %zero the resulting scission distribution everywhere except where the data %exists. %flip it back over so it again has the highest MWs first resultDist=flipud(resultDist); resultDist([1:range(1), range(end):size(resultDist,1)]',:)=0; %then calculate the weight and number fraction resultDistwt=resultDist./repmat(sum(resultDist,1),size(resultDist,1),1); resultDistnum=resultDistwt./xaxisflip.*repmat(sum(resultDistwt,1),size(result

Page 196: Photochemical Crosslinking Reactions in Polymers

175

Distwt,1),1); %make the number distribution a number fraction resultDistnum=resultDistnum./repmat(sum(resultDistnum,1),size(resultDistnum,1

),1); %------------------------------------------------------------------ %Now that we have all the distributions and have calculated the Hamielec %scission, we can calculate the recombinations based on our model %we employ another m-file, as there will be multiple combination %calculations %calculate the recombination of the control with itself %(4-arm star) ContContCombnum=MacroRadicalCombination(xaxis,ControlDistnumflip,ControlDistn

umflip,range); %calculate the recombination of the Hamielec scission with itself %(end-linking) HamHamCombnum=MacroRadicalCombination(xaxis,flipud(resultDistwt),flipud(resul

tDistwt),range); %calculate the recombination of the control with the Hamielec scission %(3-arm star) HamContCombnum=MacroRadicalCombination(xaxis,ControlDistnumflip,flipud(result

Distwt),range); %convert the combination number fractions to weight fractions ContContCombwt=ContContCombnum.*xaxis.*repmat(sum(ContContCombnum,1),size(Con

tContCombnum,1),1); ContContCombwt=ContContCombwt./repmat(sum(ContContCombwt,1),size(ContContComb

wt,1),1); HamHamCombwt=HamHamCombnum.*xaxis.*repmat(sum(HamHamCombnum,1),size(HamHamCom

bnum,1),1); HamHamCombwt=HamHamCombwt./repmat(sum(HamHamCombwt,1),size(HamHamCombwt,1),1)

; HamContCombwt=HamContCombnum.*xaxis.*repmat(sum(HamContCombnum,1),size(HamCon

tCombnum,1),1); HamContCombwt=HamContCombwt./repmat(sum(HamContCombwt,1),size(HamContCombwt,1

),1); %----------------------------------------------------- %we have the scission and the combination distributions. We here fit them %to the data in amplitude. This destroys the weight or number fraction, %but allows us to see how the data overlays %to set no end-linking, set noendlink=1. With end-linking set noendlink=0 noendlink=0; [peakfits, peakestimates]=FitPeaks(xaxis,Subtractionwt, Subtractionwt,

resultDistwt,HamHamCombwt,HamContCombwt,ContContCombwt,g1,g2,g3,range, noendlink);

Page 197: Photochemical Crosslinking Reactions in Polymers

176

%the peaks are fitted to the subtraction. We calculate the area of the %best fit combination and scission peaks. We use the data subtraction, %only the >0 parts, as we will be using this to calculate probabilities and %the data subtraction is obviously the data fit. This reduces the error. %The recombination areas are from the best fit peaks. Every input must be %large MW first, as the area calc uses a trapezoidal riemann sum using the %xaxis. [ScissionArea EndLinkingArea ThreeArmArea FourArmArea]= AreaofPeaks(xaxis(:,1),Subtractionwt(range(1)+indexwt:range(end),:),peakestim

ates(1).*HamHamCombwt,peakestimates(2).*HamContCombwt,peakestimates(3).*ContContCombwt);

%now that we have the areas, we use the probability arguments to determine %f, the fraction scission. We assume that depwt, the wt percent fractional %depletion that is the equivalent of exp(-x*u), is all the reacted species %and is equivalent to m. [fscissionm, fendlinkm, fthreearmm, ffourarmm, fscissionb, fendlinkb, fthreearmb, ffourarmb, f, Pmono, Pbis]= Probabilities(depwt,ScissionArea, EndLinkingArea, ThreeArmArea, FourArmArea); %--------------------------------------------------- %plot the data figure plot(xaxis,Subtractionwt,'.b') hold plot(xaxis,peakfits,'ob') plot(xaxis,peakestimates(4).*ControlDistwt,':r') plot(xaxis,peakestimates(4).*resultDistwt,':b') plot(xaxis.*g1,peakestimates(1).*HamHamCombwt,'r') plot(xaxis.*g2,peakestimates(2).*HamContCombwt,'g') plot(xaxis.*g3,peakestimates(3).*ContContCombwt,'k') legend({'Data';'Fit';'Scaled Control';'ScaledScission';'End linked';'3-arm';'4-arm'}) text(100,2E-4,{'End-link g-factor= ',num2str(g1),' 3-arm g-factor= ',num2str(g2),'4-arm g-factor= ',num2str(g3)}) axis([100 2000 -1E-5 3.5E-4]) %----------------------------------------------------- %Prepare data for output NumberFractions=[ControlDistnum DataDistnum fittednum Subtractionnum

SubtractionnumComb SubtractionnumScis resultDistnum ContContCombnum HamHamCombnum HamContCombnum];

WeightFractions=[ControlDistwt DataDistwt fittedwt Subtractionwt SubtractionwtComb SubtractionwtScis resultDistwt ContContCombwt HamHamCombwt HamContCombwt];

end

Page 198: Photochemical Crosslinking Reactions in Polymers

177

FitShape.m

This file scales the control peak by exp(-ru) to the data (the left side of the Hamielec scission

equation) and fits u to determine the degree of scission.

function [fitted uestimates depestimates]= FitShape(xData,yData,referenceData,range) %this function fits the control distribution to the data by multiplying by %a function of the scission extent. It also calculates the depletion, %defined as the fraction of the reference that best matches the data. This %is a single number equivalent of the exp(-x*u) used to fit u, the scission %extent. The depletion, in other words, has no x-dependence. if size(xData,2)==1 xData=repmat(xData,1,size(yData,2)); end if size(referenceData,2)==1 referenceData=repmat(referenceData,1,size(yData,2)); end %this for loop cycles through all data and fits to both the exponential and %to a pure multiple. The pure multiple is the %depletion: what percent of %the original chains have reacted for i=1:size(yData,2) r=referenceData(range,i); y=yData(range,i); x=xData(range,i); %exponential fit start_point = 0.0001; options=optimset('Display','none','MaxIter',10000,'MaxFunEvals',

10000,'TolX',1E-40,'TolFun',1E-40); uestimates(i,:) = fminsearch(@expfun, start_point,options); fitted(:,i)=referenceData(:,i).*exp(-xData(:,i).*uestimates(i,1)); %depletion fit start_point = rand(1,1); options=optimset('Display','none','MaxIter',10000,'MaxFunEvals',

10000,'TolX',1E-40,'TolFun',1E-40); depestimates(i,:) = fminsearch(@depletionfun, start_point,options); fitteddep(:,i)=referenceData(:,i).*depestimates(i,1); end

%function that calculates the sum of squared errors for the exponential function sse = expfun(params) m = params(1); line = r.*exp(-x.*m); FittedCurve=line; ErrorVector = y - FittedCurve; sse = sum(ErrorVector .^ 2); end

Page 199: Photochemical Crosslinking Reactions in Polymers

178

%function that calculates sum of squared errors for the depletion function sse = depletionfun(params) m = params(1); line = r.*m; FittedCurve=line; ErrorVector = y - FittedCurve; sse = sum(ErrorVector .^ 2); end end

Page 200: Photochemical Crosslinking Reactions in Polymers

179

FitPeaks.m

This file finds the fit of the combination peaks that best fit the data, with or without g-factor

incorporation. It minimizes the sum of squared errors between the sum of the 3 fitted peaks

and the data.

function [fitted estimates]= FitPeaks(xData,controlData,yData,Scission,convolution1,convolution2,

convolution3,g1,g2,g3,range,noendlink) if size(xData,2)==1 xData=repmat(xData,1,size(yData,2)); end %first we change the convolution in the xaxis using the g-factors %find the index of the max [~,index1]=max(convolution1); %we find the MW at g1 mw1=xData(index1,1); %we scale the MW by g1 and round to the nearest whole number newmw1=round(mw1.*g1); %we find the index of the new mw newindex1=find(xData(:,1)==newmw1,1); %calculate the shift shift1=newindex1-index1; %add that shift in zeros to the beginning of the convolution convolution1=[zeros(shift1,1);convolution1(1:end-shift1,:)]; %we do it again for the second g-factor %find the index of the max [~,index2]=max(convolution2); %we find the MW at g1 mw2=xData(index2,1); %we scale the MW by g1 and round to the nearest whole number newmw2=round(mw2.*g2); %we find the index of the new mw newindex2=find(xData(:,1)==newmw2,1); %calculate the shift shift2=newindex2-index2; %add that shift in zeros to the beginning of the convolution convolution2=[zeros(shift2,1);convolution2(1:end-shift2,:)]; %we do it again for the second g-factor %find the index of the max [~,index3]=max(convolution3); %we find the MW at g1 mw3=xData(index3,1); %we scale the MW by g1 and round to the nearest whole number newmw3=round(mw3.*g3); %we find the index of the new mw

Page 201: Photochemical Crosslinking Reactions in Polymers

180

newindex3=find(xData(:,1)==newmw3,1); %calculate the shift shift3=newindex3-index3; %add that shift in zeros to the beginning of the convolution convolution3=[zeros(shift3,1);convolution3(1:end-shift3,:)]; for i=1:size(yData,2) y=yData(range(1):6076,i); x=xData(range(1):6076,i); start_point = rand(1, 3); options=optimset('Display','none','MaxIter',10000,'MaxFunEvals',

10000,'TolX',1E-40,'TolFun',1E-40); estimates(i,:) = fminsearch(@expfun,start_point,options); estimates=abs(estimates); fitted(:,i)=(1-noendlink).*estimates(i,1).*convolution1+estimates(i,2).*

convolution2+estimates(i,3).*convolution3; estimates(i,1)=estimates(i,1).*(1-noendlink); end %function that calculates the sum of squared errors function sse = expfun(params) m = abs(params(1)); b = abs(params(2)); q = abs(params(3)); line = m.*convolution1.*(1-noendlink)+ b.*convolution2+q.*convolution3; FittedCurve=line(range(1):6076,1); ErrorVector = FittedCurve - y; sse = sum(ErrorVector .^ 2); end end

Page 202: Photochemical Crosslinking Reactions in Polymers

181

MacroRadicalCombination.m

This file calculates the macroradical recombination of any two input distributions and outputs it

for further use.

function resultDist=MacroRadicalCombination(xaxis,signal1,signal2,range) %this is a subfunction of Hamielec that calculates the macroradical %recombination of any 2 input distributions. It assumes the distributions %are correct (number fraction)). %we again employ nested for loops; the outer to cycle through r and the %inner to cycle through s. %signal1 and signal2 should be flipped so the small MWs are first; xaxis %should not be flipped and should have the large MWs first. %determine the largest chain to go to. maxMonomerNum=xaxis(min(range)); %create variables resultDist=zeros(size(signal1,1),1); %cycle through n(r) for r=2:1:maxMonomerNum %cycle through all smaller molecules count=1; thesum=0; for s=1:1:r-1 thesum=thesum + s.*(r-s).*signal1(s,:).*signal2(r-s,:); count=count+1; end %we divide by count to prevent from unintentionally multiplying by r, %as count=r, and the number of contributions to thesum will increase %with time. (at r=4, there will be 4 terms of thesum; at r=10 there %will be 10 terms of thesum; we divided by count to make a single term %and therefore remove the extra r) resultDist(r,1)=thesum./count; end %make sure it is a number fraction resultDist=resultDist./repmat(sum(resultDist,1),size(resultDist),1); %flip it so the large MWs are first resultDist=flipud(resultDist); end

Page 203: Photochemical Crosslinking Reactions in Polymers

182

AreaofPeaks.m

This file calculates the area of the input peaks using trapezoidal Riemann sums. It is used in the

determination of peak height/area and relative reactions.

function [ScissionArea EndLinkingArea ThreeArmArea FourArmArea]= AreaofPeaks(xaxis,ScissionPeak,EndLinkingPeak,ThreeArmPeak,FourArmPeak) %this function calculates the areas of the input peaks. It calculates the %ENTIRE AREA input; this is feasible because the combination peaks are zero %everywhere except at the peaks. We artificially impose this on the %ScissionPeak, which is data, by making all data<0=0. We also only use the %scission section, but that is set through the input. %This calculates area by doing a trapezoidal reiman sum of the %form: (1/2)*Q*[f(a)+2f(a+Q)+2f(a+2Q)+2f(a+3Q)+...+f(b)] %where a is the beginning of the xaxis, b the end, and Q the step %size. So it's .5*rate*[y(1)+2y(2)+2y(3)+...+2y(b-1)+y(b)] %Pad the scission data with zeros at the high MW so it's the same size %vector as the xaxis and then set the data scission peak to just the %positive scission for the area calc ScissionPeak=[zeros(size(xaxis,1)-size(ScissionPeak,1),1); ScissionPeak]; ScissionPeak(ScissionPeak<0,:)=0; %all xaxis are negative to offset the highMW to lowMW ScissionArea=trapz(-xaxis,ScissionPeak,1); EndLinkingArea=trapz(-xaxis,EndLinkingPeak,1); ThreeArmArea=trapz(-xaxis,ThreeArmPeak,1); FourArmArea=trapz(-xaxis,FourArmPeak,1); %if any of the areas are near zero, set the area to zero. That way the %floating point calculation errors don't propagate if ScissionArea<1E-10 ScissionArea=0; end if EndLinkingArea<1E-10 EndLinkingArea=0; end if ThreeArmArea<1E-10 ThreeArmArea=0; end if FourArmArea<1E-10 FourArmArea=0; end end

Page 204: Photochemical Crosslinking Reactions in Polymers

183

Probabilities.m

This uses the probabilities of each radical species to attempt to calculate the fraction reacted.

function [fscissionm, fendlinkm, fthreearmm, ffourarmm, fscissionb, fendlinkb, fthreearmb, ffourarmb, f, Pmono, Pbis]= Probabilities(m,ScissionArea, EndLinkingArea, ThreeArmArea, FourArmArea) %This function uses the probabilities and the depletion percent (in %decimal form) to calculate, f, the balance of chains that underwent %scission vs those that didn't. %we assume that the scission area reflects the dead chains that exist %after scission events %we are assuming that the depletion, exp(-ru), is equal to the sum of all %the processes. IE mass conservation. So: %depletion=dead polymer (scission) + endlinking + 3arm stars + 4arm stars %we calculate this so that each area is a fraction of the depletion totalarea=EndLinkingArea+ThreeArmArea+FourArmArea; %ScissionArea=ScissionArea./totalarea.*m; EndLinkingArea=EndLinkingArea./totalarea.*m; ThreeArmArea=ThreeArmArea./totalarea.*m; FourArmArea=FourArmArea./totalarea.*m; %for mono-functional fscissionm=ScissionArea./(m+m.^2); fendlinkm=sqrt(EndLinkingArea./m.^2); fthreearmm=roots([m.^2;-m.^2;ThreeArmArea]); ffourarmm=roots([m.^2;-2.*m.^2;m.^2-FourArmArea]); %for bi-functional fscissionb=ScissionArea./(m+m.^2+m.^4); fendlinkb=sqrt(EndLinkingArea./(m.^2+m.^4)); fthreearmb=roots([m.^4+m.^2;-1.*(m.^4+m.^2);ThreeArmArea]); ffourarmb=roots([m.^4+m.^2;-2.*(m.^4+m.^2);m.^4+m.^2-FourArmArea]); %create vectors of the values of the probabilities at all f for display %purposes f=(0:.01:1)'; %for mono-functional Pmono(:,1)=f.*(m+m.^2); Pmono(:,2)=f.^2.*m.^2; Pmono(:,3)=polyval([m.^2;-m.^2;0],f); Pmono(:,4)=polyval([m.^2;-2.*m.^2;m.^2],f); %for bi-functional Pbis(:,1)=f.*(m+m.^2+m.^4); Pbis(:,2)=(m.^2+m.^4).*f.^2; Pbis(:,3)=-polyval([m.^4+m.^2;-1.*(m.^4+m.^2);0],f); Pbis(:,4)=polyval([m.^4+m.^2;-2.*(m.^4+m.^2);m.^4+m.^2],f); end

Page 205: Photochemical Crosslinking Reactions in Polymers

184

C. Simulation Principles

The combination and scission of the chains in our system can be simulated simply to

quantitatively predict the distribution changes. We begin with a theoretical polymer length

distribution, �(�). �(�) is the number of polymer chains of length N monomers. In the

simulation, chains in the original distribution �(�) will randomly undergo combination or β-

scission reactions as radicals are generated on the backbone by BP, bi-functional BP, or any

other photosensitizer. For the model we are only interested in reactions that change the

molecular weight distribution; we assume only interpolymer combination and polymer scission

reactions, and that the radicals are generated directly on the backbone. We can therefore

neglect the photochemistry and the additive.

The principle of equal reactivity states that all monomers in a polymerization reaction

react with equal probability. Equivalently, we assert that all bonds in the (homopolymer)

polymer distribution can form radicals with equal probability. The probable location of a radical

in a polymer distribution �(�) is therefore (� − 1)�(�), as there is one less bond than

monomer in a polymer chain. �(�) can either be a theoretical distribution or an experimental

HPLC-SEC curve. To determing �(�) from experiment we begin with the distributions as

reported by a UV detector, ���(�), and/or a RI detector, ���(�), with N the number of

monomers per chain calculated from the calibrated retention-time to molecular-weight axis or

using a LS detector. UV detectors detect the concentration of UV-active monomer in the eluent

stream; monomer connectivity is ignored. Thus the distribution reported by a UV detector is

related to the N-mer distribution by ���(�)= ��(�). Above 10N , a RI detector detects

the concentration of polymer independent of molecular weight (8), so ���(�)= �(�). For the

simulations the non-irradiated control distributions were used as the input �(�)s.

The simulation models both combination and scission of the input distribution. The

relative probability of these two reactions is input as a constant as defined in equation (7.4)

below.

Page 206: Photochemical Crosslinking Reactions in Polymers

185

� =���������

��������� + ������������ (7.4)

Low values of � correspond to a high relative probability of combination while low values

correspond to a high relative probability of scission. A pseudorandom number generator

generates a value between zero and one; values above � switch to a combination calculation

while values below � switch to a scission calculation. A pseudorandom integer generator

randomly identifies one of the bonds in the distribution. In the simulation, each bond in the

distribution is labeled individually and consecutively. This facilitates the locating of the chain

to which it belongs; the cumulative number of bonds, when compared with the cumulative

��(�) easily locates the length, �, of the chain the bond comes from. Modular arithmetic

determines where on the chain the bond is located for use in calculation of branching

coefficients and non-ideality of the chain. After the first bond/chain is identified, the simulation

diverges based on reaction. If a scission reaction was predicted, then the chain is split at the

identified bond and the pieces are added to the distribution while the original chain is removed.

If a combination reaction was predicted, then the first chain is removed from the distribution to

prevent intrapolymer reactions and a second random bond is identified. The first chain is

reacted to the second chain at the bond locations and this combined chain is added to the

distribution. The same procedure is performed repeatedly on the constantly updating

distribution until a set percentage of the original chains have reacted. One can either proscribe

a percentage and �, or the variables can be fitted to the data using a least squares minimization

algorithm. The overall procedure is repeated multiple times and the results averaged together

in order to minimize the effects of outliers on the results. The program flowchart is shown in

Figure 9.1 below.

The program simulation predicts the chain lengths after reaction. We incorporate

branching using the equations of Zimm and Stockmayer(9). This allows the prediction of the

polymer independent radius of gyration, ���

��� =

��6� , where � is the branching coefficient.

Zimm and Stockmayer (9) do not incorporate excluded volume effects; we can therefore predict

Page 207: Photochemical Crosslinking Reactions in Polymers

186

that the simulation branching coefficients will be smaller than those of the real chains. Figure

9.2 compares the 16:1 bis:PS 9 hour data with the simulation results for pure linear chains

(�=1) as well as those with accurate branching coefficients as calculated following the methods

of Zimm and Stockmayer. The simulation in Figure 9.2 does not incorporate scission. The pure

linear combination peak has a radius of gyration larger than the data peak; the combination

peak from the calculated branching coefficients has a radius of gyration smaller than the data

peak as expected due to the lack of specific interactions. The experimental data therefore does

not correspond to the pure linear, end-to-end, combination reaction expected if the reaction

was end linking due to initiator fragments or other residual synthesis contaminants. We can

say with confidence that the combination reactions are along the backbone of the polymers.

The swelling of the chains in the eluent (a good solvent) due to excluded volume, straining, and

other specific interactions not included in Zimm and Stockmayer’s treatmeant increase the

polymer independent radii of gyration.

Page 208: Photochemical Crosslinking Reactions in Polymers

187

Figure 9.1: Simulation Flowchart

Page 209: Photochemical Crosslinking Reactions in Polymers

188

Figure 9.2: Comparison of Data and Simulation results for pure linear chains (right black dots, g=1) and branched chains (left black dots, g=variable) assuming 10% of original chains combined. Data (line) is 16:1 bis:PS irradiated for 9 hours.

Page 210: Photochemical Crosslinking Reactions in Polymers

189

D. Simulation M-Files

These files were used to simulate scission and combination. In the end, their results

were not used in the thesis but they were used in model development, testing, and

comprehension and served as confirmation of other results. They keep track of every chain in

the system.

CombinationScissionwithRg.m

This is the heart of the simulation. It does the calculation either on user-input distributions or

on user-defined ones by removing the commenting to the Gaussian definition.

function [CombScisPureN] = CombinationScissionWithRg(NIn,FNIn,FNcompare,fraction) %This function models the CombScis distribution. %designed in conjunction with John Zhang %define the distribution % NIn=(1:1:10)'; % FNIn=[1;2;3;3;2;2;1;0;0;0]; % NIn=(1:1:100)'; % FNIn=round(10000./(10.*sqrt(2.*pi)).*exp(-(NIn-50).^2./(2.*10^2))); %input is a fraction distribution; this converts to # of chains, and can be %adjusted for detail (more chains=more fine detail) as well as processor load %(fewer chains=less crashing and smaller memory footprint) FNIn=round(10000.*FNIn); %Define the axes Naxis=(1:1:max(NIn))'; Combaxis1=(1:1:max(Naxis))'; Combaxis=(1:1:max(Naxis))'; FNaxis=zeros(size(Naxis,1),1); %populate the axes with F(N) distribution. This allows distribution d for i=1:1:size(NIn,1) FNaxis(Naxis==NIn(i,1))=FNIn(i); end %what percent of the bonds do we want to react? %fraction=.10; %fractional percent of all initial bonds that react

Page 211: Photochemical Crosslinking Reactions in Polymers

190

balance=0; %distribution between pure combination (0) and pure scission %(1) %create a vector of every chain, by length. The lengths are repeated once %for every chain of that length. So if there are 3 chains of length 1 and %4 chains of length 2, the resulting vector will be % [1;1;1;2;2;2;2]. chains_static = cumsum(FNaxis); %bond_static = cumsum(FNaxis.*Combaxis); experiment_num=10; %do the "experiment" experiment_num times and then we can average to get a %better picture of the distribution. CombScis_repeat=zeros(max(Combaxis),experiment_num); CombScis_evolution=zeros(size(Combaxis,1),round(fraction.*max(chains_static))); for j=1:1:experiment_num %the first column is the distribution; the second column counts the %number of single branched polymers in each part of the distribution; %third column counts double branched; four counts triple branched FNaxisUse=FNaxis; toobig=zeros(1,experiment_num);

%keep track of chains that get too big to be tracked toomanybranches=zeros(1,experiment_num);

%keep track of chains that have too many branches to be tracked, just % %in case %variables that track the combination locations of the branches. These %are stored with each column a branched molecule. branchingarraychains=zeros(5,round(fraction.*max(chains_static))); branchingarraybonds=zeros(4,3.*round(fraction.*max(chains_static))); chain_counter=1; evolution_counter=1; while chain_counter <= round(fraction.*max(chains_static)) random_reaction=rand(1); if random_reaction>balance bond=cumsum(FNaxisUse(:,1).*(Combaxis-1)); %Where_It_Broke=[NaN,NaN,NaN]; %choose a random bond bond_rand(1,1) = randi([1,max(bond)],1,1); %locate the length of chains selected by the bonds and remove a %chain index=find(bond<bond_rand(1,1),1,'last')+1; FNaxisUse(index,1)=FNaxisUse(index,1)-1; %locate the exact bond within that chain bond_number1=mod(bond(index,1)-bond_rand(1,1),

Combaxis(index,1)-1); if bond_number1==0 bond_number1=Combaxis(index,1)-1; end

Page 212: Photochemical Crosslinking Reactions in Polymers

191

%we determine if this chain is branched. the branching array %consists of 4 columns per branch, with 5 rows. The %first column is the chain lengths in a branch, the second column %the first two bonds to react, the third the second two bonds to %react, and the fourth the third two bonds to react. The bottom %row is the total length of the chain. We assume the branched %chains are at the beginnning of the size.

%first we make sure the chain lengths of the previously %branched molecules are correct branchingarraychains(5,:)=sum(branchingarraychains(1:4,:),1);

%we know from above the length of the chosen chain; we search %the branching array for all chains of that length and compute %the number of bonds in them [row,column]=find(branchingarraychains(5,:)==index); branchbonds=sum(row,2).*(Combaxis(index,1)-1);

%calculate the number of bonds in branched molecules of that size chosenindex1=[]; %if the randomly selected bond is higher than the number of bonds in the %previous size plus the bonds in the branch for the current size, then it's %not branched if branchbonds+bond(index-1,1)<bond_rand(1,1) chain1is=0; %no branching %find the first empty column in the branching matrix to %make a new branch chosenindex1=find(branchingarraychains(5,:)==0,1); %add the chain length branchingarraychains(1,chosenindex1)=index; %add the bond number branchingarraybonds(1,3.*chosenindex1-2)=bond_number1; %if otherwise, it is branched elseif branchbonds+bond(index-1,1)>=bond_rand(1,1) %determine which of the branched molecules has been chosen, and the bond %within that molecule if rem(bond_rand(1,1)-bond(index-1,1),Combaxis(index,1)-1)==0 chosenchain=floor((bond_rand(1,1)-bond(index-

1,1))/(Combaxis(index,1)-1)); else chosenchain=floor((bond_rand(1,1)-bond(index-

1,1))/(Combaxis(index,1)-1))+1; end %get the chain itself from the column chosenindex1=column(:,chosenchain); %get info on chain-how many branches already? chain1is=find(branchingarraychains(1:4,chosenindex1)==0,1)-2; %if 1, one branch. if 2, two branch. if three,we must eliminate it %allows infinite branching off the same bond

if ~isempty(chain1is) %if it has less than 3 branches, add the new reacting bond. first determine %what chain within the already branched chain reacted if bond_number1<=branchingarraychains(1,chosenindex1); %first chain

%add the bond that reacted to the bon array

Page 213: Photochemical Crosslinking Reactions in Polymers

192

branchingarraybonds(1,3.*chosenindex1- 2+chain1is)=bond_number1;

elseif bond_number1>branchingarraychains(1,chosenindex1) && bond_number1<=sum(branchingarraychains(1:2, chosenindex1),1)

%second chain branchingarraybonds(2,3.*chosenindex1-2+chain1is) =bond_number1-branchingarraychains(1,chosenindex1);

elseif %third chain

bond_number1>sum(branchingarraychains(1:2,chosenindex1),1) && bond_number1<=sum(branchingarraychains(1:3,chosenindex1),1)

branchingarraybonds(3,3.*chosenindex1-2+chain1is) =bond_number1-sum(branchingarraychains(1:2,

chosenindex1),1); %error catching else disp('This should never print Combination Scission

WithRg:1') end elseif isempty(chain1is) chain1is=3; end else disp('This should never print CombinationScissionWithRg:2') end %recount the bonds to account for the selected chain %(This sets interpolymer reactions only) bond=cumsum(FNaxisUse(:,1).*(Combaxis-1)); %choose second random bond and remove the chain from the %distribution bond_rand(2,1) = randi([1,max(bond)],1,1); index2=find(bond<bond_rand(2,1),1,'last')+1; FNaxisUse(index2,1)=FNaxisUse(index2,1)-1; %locate the exact bond within that chain bond_number2=mod(bond(index2,1)-bond_rand(2,1),

Combaxis(index2,1)-1); if bond_number2==0 bond_number2=Combaxis(index2,1)-1; end %we determine again if this chain is branched %we know from above the length of the chosen chain; we search %the branching array for all chains of that length and compute %the number of bonds in them [row,column]=find(branchingarraychains(5,:)==index2); %calculate the number of bonds in branched molecules of that size branchbonds=sum(row,2).*(Combaxis(index2,1)-1);

Page 214: Photochemical Crosslinking Reactions in Polymers

193

%this insures there's not an error with the ~isempty(chosenindex2) below when %checking for a second branched molecule. Otherwise its value is held over %from the previous iteration. chosenindex2=[]; %if the randomly selected bond is higher than the number of bonds in the %previous size plus the bonds in the branch for the current size, then it's %not branched if branchbonds+bond(index2-1,1)<bond_rand(2,1) %no branching chain2is=0; %if otherwise, it is branched elseif branchbonds+bond(index2-1,1)>=bond_rand(2,1) if rem(bond_rand(2,1)-bond(index2-1,1), Combaxis(index2,1)-

1)==0 chosenchain=floor((bond_rand(2,1)-bond(index2-

1,1))/(Combaxis(index2,1)-1)); else chosenchain=floor((bond_rand(2,1)-bond(index2-

1,1))/(Combaxis(index2,1)-1))+1; End %get the chain itself from the column chosenindex2=column(:,chosenchain); %get info on chain-how many branches already? chain2is=find(branchingarraychains(1:4,chosenindex2)==0,1)-2; %if 1, one branch. if 2, two branch. if three, we must eliminate it end if isempty(chain2is) chain2is=3; end

%we only allow up to 3 branches. Check if there are more than 3 branches in %the combination, and make the chain(s) dissapear %if there are 3 branches if chain1is+chain2is+1>3 branchingarraychains=[branchingarraychains(:,1:chosenindex1-

1) branchingarraychains(:,chosenindex1+1:end)]; branchingarraybonds=[branchingarraybonds(:,1:3.*(chosenindex1-1)) branchingarraybonds(:,3.*chosenindex1+1:end)];

if ~isempty(chosenindex2) branchingarraychains=[branchingarraychains(:,1:chosenindex2-1) branchingarraychains(:,chosenindex2+1:end)]; branchingarraybonds=[branchingarraybonds(:,1:3.*(chosenindex2-1)) branchingarraybonds(:,3.*chosenindex2+1:end)];

end

toomanybranches(1,experiment_num)=toomanybranches(1,experiment_num)+1; %skip the rest of this loop continue end

switch chain2is

Page 215: Photochemical Crosslinking Reactions in Polymers

194

%if chain 2 is not branched; add it to the branching array case 0 emptyrow=find(branchingarraychains(1:4,chosenindex1)==0,1); %add the chain length

branchingarraychains(emptyrow,chosenindex1)=index2; %add the bond number

branchingarraybonds(emptyrow,3.*chosenindex1-2+chain1is)= bond_number2;

%if it has 1 branch case 1 %transplant the lengths directly to the first chain vector.

emptyrow=find(branchingarraychains(1:3,chosenindex1)==0,1); %should always be 2 for the same reason emptycolumn=find(sum(branchingarraybonds(:,3.*chosenindex1-2:3.*chosenindex1),1)==0,1); branchingarraychains(emptyrow:emptyrow+1,chosenindex1)=branchingarraychains(1:2,chosenindex2);

branchingarraybonds(:,3.*chosenindex1-3+emptycolumn: 3.*chosenindex1-2+emptycolumn)= circshift(branchingarraybonds(:,3.*chosenindex2-2:3.*chosenindex2-1),emptyrow-1);

%first determine what chain within the already branched chain reacted if bond_number2<=branchingarraychains(1,chosenindex2); %first chain %add the bond that reacted to the bond array

branchingarraybonds(emptyrow,3.*chosenindex1-2+chain1is)= bond_number2; %second chain elseif bond_number2>branchingarraychains(1,chosenindex2) && bond_number2<=sum(branchingarraychains(1:2,chosenindex2),1)

branchingarraybonds(emptyrow+1,3.*chosenindex1-2+chain1is)= bond_number2-branchingarraychains(1,chosenindex2);

else

disp('This should never print CombinationScissionWithRg:3') end %remove chain 2 from the branching arrays

branchingarraychains=[branchingarraychains(:,1:chosenindex2-1) branchingarraychains(:,chosenindex2+1:end)]; branchingarraybonds=[branchingarraybonds(:,1:3.*(chosenindex2-1)) branchingarraybonds(:,3.*chosenindex2+1:end)];

%if it has 2 branches

case 2 %transplant the lengths directly to the first chain vector.

Page 216: Photochemical Crosslinking Reactions in Polymers

195

%should always be 2 as we just created this branch in the first chain section emptyrow=find(branchingarraychains(1:2,chosenindex1)==0,1); %should always be 2 for the same reason emptycolumn=find(sum(branchingarraybonds(:,3.*chosenindex1-2:3.*chosenindex1),1)==0,1); branchingarraychains(emptyrow:emptyrow+2,chosenindex1)=branchingarraychains(1:3,chosenindex2);

branchingarraybonds(:,3.*chosenindex1-3+emptycolumn:3.*chosenindex1-2+emptycolumn)=circshift(branchingarraybonds(:,3.*chosenindex2-2:3.*chosenindex2-1),1);

%first determine what chain within the already branched chain reacted if bond_number2<=branchingarraychains(1,chosenindex2); %first chain %add the bond that reacted to the bon array

branchingarraybonds(emptyrow,3.*chosenindex1-4+emptycolumn)=bond_number2;

elseif bond_number2>branchingarraychains(1,chosenindex2) && bond_number2<=sum(branchingarraychains(1:2,chosenindex2),1) %second chain branchingarraybonds(emptyrow+1,3.*chosenindex1-4+emptycolumn)=bond_number2-branchingarraychains(1,chosenindex2);

elseif bond_number2>sum(branchingarraychains(1:2,chosenindex2),1) && bond_number2<=sum(branchingarraychains(1:3,chosenindex2),1) %third chain branchingarraybonds(emptyrow+2,3.*chosenindex1-4+emptycolumn)=bond_number2-sum(branchingarraychains(1:2,chosenindex2),1);

else

disp('This should never print CombinationScissionWithRg:4') end %remove chain 2 from the branching arrays

branchingarraychains=[branchingarraychains(:,1:chosenindex2-1) branchingarraychains(:,chosenindex2+1:end)]; branchingarraybonds=[branchingarraybonds(:,1:3.*(chosenindex2-1)) branchingarraybonds(:,3.*chosenindex2+1:end)];

otherwise

disp('This should never print CombinationScissionWithRg:5') end

Page 217: Photochemical Crosslinking Reactions in Polymers

196

%if the size of the combined chain is greater than 5 times the maximum input array size (irrespective of if there are any chains at that size; this is about just the array size) these chains "dissapear" from the distribution and are counted as reacted,though too big to be part of the output distribution. if index+index2>max(Combaxis) toobig(1,experiment_num)=toobig(1,experiment_num)+1; continue end %add the new, combined chain to the distribution. This variable is the pure %linear length; radius of gyration changes in output location are not %considered. FNaxisUse(index+index2,1)=FNaxisUse(index+index2,1)+1; %now the distribution is adjusted. %if it’s the first experiment, the changes are just the distribution if evolution_counter==1 CombScis_evolution(:,evolution_counter)=FNaxisUse(:,1); %account for multiple experiments by averaging the new experiment with %previous else

CombScis_evolution(:,evolution_counter)=mean([FNaxisUse(:,1), CombScis_evolution(:,evolution_counter)],2);

end chain_counter=chain_counter+2; evolution_counter=evolution_counter+1; elseif random_reaction<balance bond=cumsum(FNaxisUse(:,1).*(Combaxis-1)); %choose a random bond bond_rand = randi([1,max(bond)],1,1); %locate the index (length of chain) of that bond index=find(bond<bond_rand,1,'last')+1; %locate the exact bond with the chain that broke bond_number=mod(bond(index,1)-bond_rand,Combaxis(index,1)-1); %The size of one piece is the bond_number (since a break of bond 2 produces a %piece of size 2); the size of the other is the chain length minus the bond %number+1. Remove 1 chain from the original length and add 1 chain each to %each pieces' length. A bond_number of zero indicates the last bond in a %chain; this is equivalent to Naxis(index,1)-1 if bond_number==0 bond_number=Combaxis(index,1)-1; end FNaxisUse(index,1)=FNaxisUse(index,1)-1; FNaxisUse(bond_number,1)=FNaxisUse(bond_number,1)+1; FNaxisUse(Combaxis(index,1)-

bond_number,1)=FNaxisUse(Combaxis(index,1)-bond_number,1)+1;

Page 218: Photochemical Crosslinking Reactions in Polymers

197

%now the distribution is adjusted. %if it’s the first experiment, it's just the distribution if evolution_counter==1 CombScis_evolution(:,evolution_counter)=FNaxisUse(:,1); %account for multiple experiments by averaging the new experiment with previous else

CombScis_evolution(:,evolution_counter)=mean([FNaxisUse(:,1), CombScis_evolution(:,evolution_counter)],2);

end chain_counter=chain_counter+1; evolution_counter=evolution_counter+1; elseif random_reaction == balance continue else disp('This should never print error CombinationScissionWithRg:6') keyboard end end CombScis_repeat(:,j)=FNaxisUse; branchingarraychains(5,:)=sum(branchingarraychains(1:4,:),1); [linRg2b2(:,j) FNaxisRg(:,j) singlebranchchains twobranchchains

threebranchchains onebranchRg twobranchRg threebranchRg]= RadiusOfGyration(Combaxis,FNaxisUse,branchingarraychains,branchingarraybonds);

end %---------------------------------------------------------------------- %output, theory, and display code figure CombScisPureN=mean(CombScis_repeat,2); chains_static2 = sum(CombScisPureN,1); plot(Combaxis,CombScisPureN,'og') CombScisPureN=CombScisPureN./repmat(sum(CombScisPureN,1),size(CombScisPureN,1),1); %scale to 1 title('Direct Chain Lengths') %Pass results to g-factor code linRg2b2Ave=mean(linRg2b2,2); FNaxisRgAve=mean(FNaxisRg,2); BranchRgs=[onebranchRg twobranchRg threebranchRg]; [FittedFNaxisRg(:,j) gestimates(:,j)]=RadiusOfGyrationCompare(linRg2b2Ave,BranchRgs,CombScisPureN,FNcompare); figure plot(flipud(Combaxis./6),FNcompare./repmat(sum(FNcompare,1),size(FNcompare,1),1)) hold plot(linRg2b2Ave,FNaxisRgAve./repmat(sum(FNaxisRgAve,1),size(FNaxisRgAve,1),1),':r') plot(Combaxis./6,CombScisPureN,':k') title('Radius of Gyration/b^2')

Page 219: Photochemical Crosslinking Reactions in Polymers

198

RadiusOfGyration.m

This subfunction implements Zimm and Stockmayer’s calculations on the arrays from the

simulation and produces the radii of gyration of all species in the system.

function [linRg2b2 FNaxisRg singlebranchchains twobranchchains threebranchchains onebranchRg twobranchRg threebranchRg]=RadiusOfGyration(Naxis,FNaxisUse,Chains,Bonds) %This function uses the equations from %Zimm,Stockmeyer, J.Chem.Physics,v17#12,1949,pg 1301 %To calculate from an input distribution and branching locations the %relative radius of gyration for polymers with 0,1,2,3 branch units %first calculate linear dimention. These equations are in R^2/b^2 space to %eliminate the rotational, bond angle, and solvent effects within a %polymer. In other words, all below equations are R^2/b^2=... %check to insure the sums are correct Chains(5,:)=sum(Chains(1:4,:),1); %initial definitions in case the system does not have and branches of that %size onebranchRg=[]; twobranchRg=[]; threebranchRg=[]; %linear Rg axis: this is merely R^2=b^2*N/6 changed to R^2/b^2=N/6 linRg2b2=Naxis./6; %one branch R^2/b^2=1/N*sumv(Nv %find the chains in Chains that have only 1 branch [~,column1br]=find(Chains(1,:)~=0 & Chains(2,:)~=0 & Chains(3,:)==0 & Chains(4,:)==0 & Chains(5,:)~=0); %because the third row of chains is only if there are 2 or more branches if ~isempty(column1br) %if there are any branches %we get the chains and bonds of just the single chain species. To %calculate Rg, we need to count the number of monomers in each arm of the %single branch species. %this is an easy calculation. The chains variable has the length of the %chain; the bonds variable the bond that's reacted %for each branched chain, create a 4 row array with the branch lengths. singlebranchchains=Chains(:,column1br); %we have to do a bit of manipulation to get the bond columns from the %chain one %we only need these, as one branch only uses the first of the 3 available branch columns singlebranchbonds=Bonds(:,3.*column1br-2);

Page 220: Photochemical Crosslinking Reactions in Polymers

199

for i=1:1:size(singlebranchchains,2) branchesofsingle(:,i)=[singlebranchbonds(1:2,i);singlebranchchains(1:2,i)-singlebranchbonds(1:2,i)];

onebranchRg(1,i)=singlebranchchains(5,i); onebranchRg(2,i)=(1./singlebranchchains(5,i)).*sum((branchesofsingle(:,i).^2./2-branchesofsingle(:,i).^3./(3.*singlebranchchains(5,i))),1); gfactor(:,i)=(6./singlebranchchains(5,i).^2).*sum((branchesofsingle(:,i).^2./2-branchesofsingle(:,i).^3./(3.*singlebranchchains(5,i))),1);

end end %two branch %find the chains in Chains that have 2 branches [~,column2br]=find(Chains(3,:)~=0 & Chains(4,:)==0 & Chains(5,:)~=0); %because the third row of chains is only if there are 2 or more branches if ~isempty(column2br) %if there are any 2 branch chains twobranchchains=Chains(:,column2br); %we have to do a bit of manipulation to get the bond columns from the chain %one bonds1=3.*column2br-2; bonds2=3.*column2br-1; twobranchcolumns=sort([bonds1 bonds2]); %we need the first two of the 3 available branch columns twobranchbonds=Bonds(:,twobranchcolumns); for i=1:1:size(twobranchchains,2) %we eliminate a bug, sofar of unknown source, whereby there is %sometimes a column of all zeros counted as a two branch species. if sum(twobranchbonds(:,2.*i-1),1)==0 ||

sum(twobranchbonds(:,2.*i),1)==0 break end %we find the row in the bonds that has no zeros; this row is the %cross-chain that contains the interbranch ab chain. We make it a %boolean vector, with 1 the row and 0 the others row=twobranchbonds(:,2.*i-1)~=0 & twobranchbonds(:,2.*i)~=0; %we find the column of the row with the larger #; this helps determine how to %match up the ends. If inequality is 0 (false), then the first column is %greater; if it's 1 (true) then the second is greater. column=twobranchbonds(row,2.*i-1) < twobranchbonds(row,2.*i); %this allows the following definition: % column index of the greater=2.*i-1+column; % column index of the lesser2.*i-column;

Page 221: Photochemical Crosslinking Reactions in Polymers

200

%NOTE: if the two values are equal, then the inequality is 0 and everything %is correct except the interbranch ab chain is length 0 and the equation %reduces to the one branch equation, which is valid for arbitrary %functionality and therefore still valid in this calculation (on branch of %functionality 6). %we define branch a as the branch connected to the branch point with the %lesser of the two bonds connected to the interbranch ab chain %xor finds the row that contains both a number, and not a number in the %second column; the column is the lesser one as defined above branchesoftwoa(:,i)= [ twobranchbonds(xor(row,twobranchbonds(:,2.*i-column)),2.*i-

column); twobranchchains(xor(row,twobranchbonds(:,2.*i-column)),i) - twobranchbonds(xor(row,twobranchbonds(:,2.*i-column)),2.*i-column);

%and finds the row that has reacted with the other column, and is %of the lesser value. twobranchbonds(and(row,twobranchbonds(:,2.*i-column)),2.*i-column)

]; %branch b is the branch of the greater of the 2 connected bonds. %The only changes are the final row calculation, and the %substitution of 2.*i-1+column for the lesser code. %xor finds the row that contains both a number, and not a number in the second column; the column is the greater one as defined above branchesoftwob(:,i)=[ twobranchbonds(xor(row,twobranchbonds(:,2.*i-1+column)),2.*i-

1+column);

twobranchchains(xor(row,twobranchbonds(:,2.*i-1+column)),i) - twobranchbonds(xor(row,twobranchbonds(:,2.*i-1+column)),2.*i-1+column);

%and finds the row that has reacted with the other column, and is of the lesser value. twobranchchains(and(row,twobranchbonds(:,2.*i-1+column)),i)-t

wobranchbonds(and(row,twobranchbonds(:,2.*i-column)),2.*i- 1+column];

%branch ab is the branch that connects branch a to b. It's the %only branch with no ends. It's length is the difference between %the greater and lesser value in the row that connects the columns. branchesoftwoab(:,i)=[ twobranchbonds(and(row,twobranchbonds(:,2.*i-1+column)),2.*i-

1+column)-twobranchbonds(and(row,twobranchbonds(:,2.*i-column)),2.*i-column)];

%to make the radius of gyration calculation easier, we put all the %branches into a single array for part of it.

Page 222: Photochemical Crosslinking Reactions in Polymers

201

alltwobranches=[branchesoftwoa(:,i);branchesoftwob(:,i);branchesoftwoab(:,i)];

twobranchRg(1,i)=twobranchchains(5,i); %first calculate the single branch equation using the single array

parta=sum(alltwobranches.^2./2- alltwobranches.^3./(3.*twobranchchains(5,i)),1);

%then the modification for two branches

partb=branchesoftwoab(:,i)./twobranchchains(5,i).*sum(branchesoftwoa(:,i),1).*sum(branchesoftwob(:,i),1);

%and finally put the two parts of the equation together twobranchRg(2,i)=(1./twobranchchains(5,i)).*(parta+partb); gfactor2(:,i)=(6./twobranchchains(5,i).^2).*(parta+partb); end end

%three branch %find the chains in Chains that have 3 branches [~,column3br]=find(Chains(4,:)~=0 & Chains(5,:)~=0); %because the fourth row of chains is only if there are 3 or more branches if ~isempty(column3br) %if there are any three branch chains threebranchchains=Chains(:,column3br); %we have to do a bit of manipulation to get the bond columns from the %chain %one. bonds1=3.*column3br-2; bonds2=3.*column3br-1; bonds3=3.*column3br; threebranchcolumns=sort([bonds1 bonds2 bonds3]); threebranchbonds=Bonds(:,threebranchcolumns); %we need all 3 of the available branch columns for i=1:1:size(threebranchchains,2) %we eliminate a bug, sofar of unknown source, whereby there is %sometimes a column of all zeros counted as a three branch species.

if sum(threebranchbonds(:,3.*i-2),1)==0 || sum(threebranchbonds(:,3.*i-1),1)==0 || sum(threebranchbonds(:,3.*i),1)==0

break end bonds=threebranchbonds(:,3.*i-2:3.*i); %this calculation is relatively simple because they treat the 3 %branches as one "reference branch" which connects to both other %branches. So we split this into 2 2-branch problems. The

Page 223: Photochemical Crosslinking Reactions in Polymers

202

%reference branch and a, and the reference branch and b. This will %overcount the reference branch, but we can then remove the excess %branches. %first find the reference branch, which is the only one that has %both its reacted bonds connected to other branches

%we first boolean AND columns 1,2 and 2,3 and 1,3. This gives us the %matching row in each. A boolean OR gives us the boolean signature of the %reference column (the rows that aren't zero). We repeat the signature into %a 3 column matrix, then AND it with the bonds and sum that in the row %direction. The reference column will be the only one with sum==2

refcolumn=sum((bonds>0 & repmat((bonds(:,1)~=0 & bonds(:,2)~=0) | (bonds(:,1)~=0 & bonds(:,3)~=0) | (bonds(:,2)~=0 & bonds(:,3)~=0),1,3)),1)>=2;

%we have to account for the situation where one chain has all the %crosslinks (acts more like a comb polymer than a branch...but in %this code we still treat it like as a branch). We know this has %happened when the sum of refcolumn is not 1 (more than 1 %"refcolumn") if sum(refcolumn,2)~=1 %get the row row=sum(bonds~=0,2)==3; %find the columns that correspond to the max and min [~,maxcol]=max(bonds(row,:)); [~,mincol]=min(bonds(row,:)); %we can directly get the b branches from this branchb(:,1)=[ bonds(bonds(:,mincol)~=0,mincol); threebranchchains(xor(row,bonds(:,mincol)),i)-

bonds(xor(row,bonds(:,mincol)),mincol)]; branchb(:,2)=[ bonds(xor(row,bonds(:,maxcol)),maxcol); threebranchchains(xor(row,bonds(:,maxcol)),i)-

bonds(xor(row,bonds(:,maxcol)),maxcol); threebranchchains(and(row,bonds(:,maxcol)),i)-bonds(and(row,bonds(:,maxcol)),maxcol)];

%we identify the middle column; this determines the reference %crosslink. We do it by removing the max and min columns from %the bonds array if mincol<maxcol midcol=bonds(:,[1:mincol-1 mincol+1:maxcol-1 maxcol+1:end]); elseif mincol>maxcol midcol=bonds(:,[1:maxcol-1 maxcol+1:mincol-1 mincol+1:end]); end %we have the midcolumn thus brancha=[midcol(xor(row,midcol>0)); threebranchchains(xor(row,midcol>0),i)-midcol(xor(row,midcol>0)) ]';

Page 224: Photochemical Crosslinking Reactions in Polymers

203

%last we calculate the connecting branches from the single chain. Use an and %to extract the three values and sort them lowest to highest. thechain=sort(bonds(and(repmat(row,1,3),bonds))); branchab=[thechain(2)-thechain(1) thechain(3)-thechain(2)]; elseif sum(refcolumn,2)==1 %the negation of the refcolumn array gives us the nonreference bond %columns, while it itself gives us the reference bond column

refbonds=bonds(:,refcolumn); nonrefbonds=bonds(:,~refcolumn); %we can then use the code for 2 branches on each pair, nonreference1 and %reference, and nonreference2 and reference. We know for a fact which of the %2 contains the reference branch; we can therefore use that instead of the %greater/lesser method for part of it. What we do is change the variable %names and change the indexing from k to i and from 2.*i-1+column to 1+column %and 2.*i-column to 2-column for k=1:1:2 twobonds=[refbonds nonrefbonds(:,k)]; %we find the row in the bonds that has no zeros; this row is the cross-chain %that contains the interbranch ab chain. We make it a boolean vector, with 1 the row and 0 the others row=twobonds(:,1)~=0 & twobonds(:,2)~=0; %we find the column of the row with the larger #; this helps determine how to %match up the ends. If inequality is 0 (false), then the first column is %greater; if it's 1 (true) then the second is greater. column=twobonds(row,1) < twobonds(row,2);

%this allows the following definition: % column index of the greater=1+column; % column index of the lesser=2-column; %NOTE: if the two values are equal, then the inequality is 0 and everything %is correct except the interbranch ab chain is length 0 and the equation %reduces to the one branch equation, which is valid for arbitrary %functionality and therefore still valid in this calculation (on branch of %functionality 6). %brancha is the 2 ends off the reference, a branch %branchb are both the other branchs %branchab is the distance between the branches and branch a %we know that the reference chain is in column 1; %xor finds the row that contains a number in the nonreference column but no number in the reference; this is one of the linear branch ends. The number and the lenght of chain minus that number are 2 of the 3 arms. branchb(1:2,k)=[ twobonds(xor(row,twobonds(:,2)),2); threebranchchains(xor(row,twobonds(:,2)),i) –

twobonds(xor(row,twobonds(:,2)),2) ];

Page 225: Photochemical Crosslinking Reactions in Polymers

204

%if the second column of the branch is the lesser, then that is the value for %the third arm. If it's the greater, the value is the chain length minus the %value. The same is true for branchb; we can do both in one if loop if 1+column == 1 branchb(3,k)=twobonds(and(row,twobonds(:,2)),2); brancha(1,k)=threebranchchains(and(row,twobonds(:,1)),i)-

twobonds(and(row,twobonds(:,1)),1); elseif 1+column == 2 branchb(3,k)=threebranchchains(and(row,twobonds(:,2)),i)-

twobonds(and(row,twobonds(:,2)),2); brancha(1,k)=twobonds(and(row,twobonds(:,1)),1); else disp('This should not print RadiusOfGyration:1') keyboard end %branch ab is the branch that connects branch a to b. It's the only branch %with no ends. It's length is the difference between the greater and lesser %value in the row that connects the columns. branchab(:,k)=[ twobonds(and(row,twobonds(:,1+column)),1+column)-

twobonds(and(row,twobonds(:,2-column)),2-column) ]; end else disp('This should not print RadiusOfGyration:2') keyboard end

%to make the radius of gyration calculation easier, we put all the %branches into a single array for part of it. allthreebranches=[brancha;branchb;branchab]; allthreebranches=reshape(allthreebranches,numel(allthreebranches),1); threebranchRg(1,i)=threebranchchains(5,i); %first the all chain part of the equation parta=sum(allthreebranches.^2./2-

allthreebranches.^3/(3.*threebranchchains(5,i)),1); %then we deal with the branch a and branch b chain

partb=fliplr(branchab)./repmat(threebranchchains(5,i),1,2).*(repmat(sum(brancha,2),1,2)+sum(branchb,1)+branchab).*sum(fliplr(branchb),1);

%put it all together threebranchRg(2,i)=(1./threebranchchains(5,i)).*(parta+sum(partb,2)); gfactor3(:,i)=6./(threebranchchains(5,i).^2).*(parta+sum(partb,2)); end end %now we've determined the Rg of all the branched chains (up to 3 branches) %we are going to combine this data with the input original distribution %into a single Rg vs height

Page 226: Photochemical Crosslinking Reactions in Polymers

205

%one-,two-,and three-branchRg have the N in the top row and the Rg in the %bottom. We recombine them into one larger array; one branch then 2 branch %then 3 branch. BranchRgs=[onebranchRg twobranchRg threebranchRg]; %go through each column, remove a chain the appropriate N from the %FNaxisUse, then locate that %chain on the linear Rg axis and add a chain there. for i=1:1:size(BranchRgs,2) %we will "round" to nearest Rg on the linear Rg axis. %first, remove the chain from the linear N axis FNaxisUse(BranchRgs(1,i),1)=FNaxisUse(BranchRgs(1,i),1)-1;

%If the Rg of the current chain is closer to the higher linear Rg, add %the chain to that population. if it is closer to the lower linear Rg, %add the chain there. This is the "rounding". if ~isempty(find(BranchRgs(2,i)==linRg2b2,1)) index=find(BranchRgs(2,i)==linRg2b2); FNaxisUse(index,1)=FNaxisUse(index,1)+1; elseif abs(BranchRgs(2,i)-

linRg2b2(find(linRg2b2<BranchRgs(2,i),1,'last'),:)) > abs(BranchRgs(2,i)-linRg2b2(find(linRg2b2>BranchRgs(2,i),1,'first'),:))

index=find(linRg2b2>BranchRgs(2,i),1,'first'); FNaxisUse(index,1)=FNaxisUse(index,1)+1;

elseif abs(BranchRgs(2,i)-

linRg2b2(find(linRg2b2<BranchRgs(2,i),1,'last'),:)) < abs(BranchRgs(2,i)-linRg2b2(find(linRg2b2>BranchRgs(2,i),1,'first'),:))

index=find(linRg2b2<BranchRgs(2,i),1,'last'); FNaxisUse(index,1)=FNaxisUse(index,1)+1;

elseif abs(BranchRgs(2,i)-

linRg2b2(find(linRg2b2<BranchRgs(2,i),1,'last'),:)) == abs(BranchRgs(2,i)-linRg2b2(find(linRg2b2>BranchRgs(2,i),1,'first'),:))

%if it is exactly equidistant between the two (unlikely, but %possible), add .5 to each index=find(linRg2b2<BranchRgs(2,i),1,'last'); index2=find(linRg2b2>BranchRgs(2,i),1,'first'); FNaxisUse(index,1)=FNaxisUse(index,1)+0.5; FNaxisUse(index2,1)=FNaxisUse(index2,1)+0.5; else disp('Radius large then linear axis') end end FNaxisRg=FNaxisUse; End

Page 227: Photochemical Crosslinking Reactions in Polymers

206

E. Data Importation M-Files

These two m-files converted the instrumental data files into MATLAB® readable

formats.

GetGPCData.m

This function reads all Shimadzu™ HPLC software export files in the MATLAB® directory and

creates two large arrays, one of the RID data and the other of the UV data. Filenames and time

axes are also imported.

function [SampleNames TimeAxis UVData RIDData]=GetGPCData %get the list of exported GPC data in current directory (.dat.asc) a=dir('*.dat.asc'); %for each file get the filename and then extract the data for i=1:size(a,1)

Name=a(i).name; %get the numeric data and the header GPC=importdata(Name,'\r',13); SampleNames{1,i=sscanf(GPC.textdata{3},'Sample ID:%s%s%s%s'); SamplingRate=sscanf(GPC.textdata{8},'Sampling Rate:%f%f'); DataPoints=sscanf(GPC.textdata{9},'Total Data Points:%f%f'); Xmult=sscanf(GPC.textdata{12},'X Axis Multiplier:%f%f'); Ymult=sscanf(GPC.textdata{13},'Y Axis Multiplier:%f%f'); %pad UV and RID data with zeros if there are more data points in the %next set of data (this accounts for different collection times by zero %padding at THE END of the collection so as to not change the elution %time of the samples if DataPoints(1)==DataPoints(2) k=1; elseif DataPoints(1)>DataPoints(2) k=1; elseif DataPoints(1)<DataPoints(2) k=2 end %UV if i==1 || size(UVData,1)==DataPoints(1)

Page 228: Photochemical Crosslinking Reactions in Polymers

207

UVData(:,i)=GPC.data(1:DataPoints(1),1).*Ymult(1); TimeAxis(:,i)=0:Xmult(1)./SamplingRate(1):DataPoints(k)./(SamplingRate(1).*60);

elseif size(UVData,1)<DataPoints(1) UVData=[UVData;zeros(DataPoints(1)-size(UVData,1),size(UVData,2))]; TimeAxis=[TimeAxis;zeros(DataPoints(k)-size(TimeAxis,1),

size(TimeAxis,2))];

UVData(:,i)=GPC.data(1:DataPoints(1),1).*Ymult(1); TimeAxis(:,i)=(0:1:DataPoints(k)-1)./(10.*60) elseif size(UVData,1)>DataPoints(1) intermediateUV=GPC.data(1:DataPoints(1),1).*Ymult(1); UVData(:,i)=[intermediateUV;zeros(size(UVData,1)-DataPoints(1),1)]; TimeAxis(:,i)=(0:1:size(UVData,1)-1)./(10.*60); end %RID if i==1 || size(RIDData,1)==DataPoints(2)

RIDData(:,i)=GPC.data(DataPoints(1)+1:DataPoints(1)+ DataPoints(2).*Ymult(2);

elseif size(RIDData,1)<DataPoints(2)

RIDData=[RIDData;zeros(DataPoints(2)-size(RIDData,1), size(RIDData,2))];

RIDData(:,i)=GPC.data(DataPoints(1)+1:DataPoints(1)+

DataPoints(2)).*Ymult(2); elseif size(RIDData,1)>DataPoints(2)

intermediateRID=GPC.data(DataPoints(1)+1:DataPoints(1)+ DataPoints(2)).*Ymult();

RIDData(:,i)=[intermediateRID;zeros(size(RIDData,1)-DataPoints(2),1)];

end end end

Page 229: Photochemical Crosslinking Reactions in Polymers

208

EPR.m

This file, like GetGPCData.m, processes the raw EPR data files into MATLAB® readable

format. This is important because EASYSPIN is a MATLAB® toolbox and can only be used in

MATLAB®.

%processes epr signals into something understandable function [Gvalues,Avalues,Minutes,TextNames]=EPR %get the list of txt files a=dir('*.txt'); %first, extract the time in minutes from the .par file for i=1:size(a,1) %get the .txt filename for i Name=a(i).name; %first,extract just the name, not the extension, then add par to the %end justname='[a-z_A-Z0-9~]*\.'; filenamepart=regexp(Name,justname,'match'); TimeFileName=strcat(filenamepart,'par'); TextNames{i}=Name; %error catching to make sure the .par file exists b=dir(TimeFileName{1,1}); if isempty(b) continue else filetext=fileread(TimeFileName{1,1}); end %extract the time from the .par file, convert to number then minutes. expr='[0-9]*:[0-9]*'; time=regexp(filetext,expr,'match'); hours=str2num(time{1,1}(1,1:2)); minutes=str2num(time{1,1}(1,4:5)); Minutes(i)=60.*hours+minutes; end %next, extract the data. It's in a different for loop just for easy of %viewing. for i=1:size(a,1) Name=a(i).name; fid=fopen(Name); frewind(fid); data=textscan(fid,'%f%f%s','HeaderLines',6); indices=data{1,1}; values=data{1,2};

Page 230: Photochemical Crosslinking Reactions in Polymers

209

cell=data{1,3}; Gvalues(:,i)=values; for j=1:size(indices,1) Avalues(j,i)=str2num(cell{j,1}); end fclose(fid); end end

Page 231: Photochemical Crosslinking Reactions in Polymers

210

F. Data Preparation M-Files

These m-files were used to prepare the raw data for analysis.

TransformX.m

This file used the HPLC calibration curves to convert the x-axis from elution time to PS

molecular weight. It gives the user the option of choosing the calibration curve appropriate to

when the data was taken.

function MWXaxis=TransformX(xTime,oldnew) %This function transforms the x axis from time(min) to MW based on %the GPC calibration curve. %This changes the values in the time vector so that only the %Calibrated range is changed to MW. All other values are set to %the MW at 12.5minutes, which is outside the calibrated range %The letters are the opposite of those in the export files check=find(xTime<10.5); xTime(check)=10.5; xTime=xTime.*60; %convert to seconds if oldnew==0 %this is 2 column a=1.93165E7; b=2.81645E6; c=-1.75077E6; d=236534.44232; e=-14585.34916; f=435.78173; g=-5.13227; elseif oldnew==1 %this is 2column/2column (Resi-Resi/Mesi-Resi) a=-4.900882244E-11; b=2.058938486E-7; c=0.0003153731941; d=0.2025771423; e=-40.44872522; f=0; g=0; elseif oldnew==2 %this is 2/3 column switchable 2 column UV detector a=-2.866117528; b=0.04002332988; c=-5.382586768E-5; d=2.026961038E-8; e=0;

Page 232: Photochemical Crosslinking Reactions in Polymers

211

f=0; g=0; elseif oldnew==3 %this is 2/3 column switchable 2 column RID detector a=-3.177184773; b=.04084312483; c=-5.428356495E-5; d=2.026913156E-8; e=0; f=0; g=0; elseif oldnew==4 %this is 2/3 column switchable 2 column UV detector 12/3 with EasiVial a=-39.982; b=0.2059; c=-0.00033031; d=2.2402E-7; e=-5.5887E-11; f=0; g=0; elseif oldnew==5 %this is 2/3 column switchable 2 column RID detector 12/3 with EasiVial a=-45.63992636; b=0.2183448071; c=-0.0003307921861; d=2.116468266E-7; e=-4.962786325E-11; f=0; g=0; elseif oldnew==6 %this is 2/3 column switchable 2 column UV from 1/2010 a=-44.734588; b=0.2175808327; c=-0.0003343674405; d=2.169212995E-7; e=-5.156317144E-11; f=0; g=0; elseif oldnew==7 %this is 2/3 column switchable 2 column RID from 1/2010 a=-45.63992636; b=0.2183448071; c=-0.0003307921861; d=2.116468266E-7; e=-4.962786325E-11; f=0; g=0; elseif oldnew==8 %this is 2/3 column switchable 2 column UV from 3/2010 a=-43.58527992; b=0.2130353005; c=-.0003277485637; d=2.126715507E-7; e=-5.055083433E-11; f=0; g=0; elseif oldnew==9 %this is 2/3 column switchable 2 column RID from 3/2010 a=-44.70487894; b=0.214725017; c=-0.0003256570381; d=2.08442878E-7;

Page 233: Photochemical Crosslinking Reactions in Polymers

212

e=-4.888727017E-11; f=0; g=0; elseif oldnew==10 %this is 2/4 column switchable 4 column UV from 6/2010 a=11.23177528; b=-0.003088471478; c=-2.485245588E-6; d=1.647091749E-9; e=-2.915523588E-13; f=0; g=0; elseif oldnew==11 %this is 2/4 column switchable 4 column RID from 6/2010 a=16.61365117; b=-0.01482312072; c=7.087251289E-6; d=-1.781351636E-9; e=1.622308918E-13; f=0; g=0; elseif oldnew==12 %this is 2/4 column switchable 2 column UV from 8/2010 a=12.83444155; b=-0.01174746014; c=2.847842538E-6; d=-1.364719786E-10; e=0; f=0; g=0; elseif oldnew==13 %this is 2/4 column switchable 2 column RID from 8/2010 a=12.92780856; b=-0.01164914189; c=2.699787596E-6; d=-9.135393069E-11; e=0; f=0; g=0; else print('Error in second argument.') end lnXaxis=a+b.*xTime+c.*xTime.^2+d.*xTime.^3+e.*xTime.^4+f.*xTime.^5+g.*xTime.^6; MWXaxis=10.^lnXaxis; end

Page 234: Photochemical Crosslinking Reactions in Polymers

213

TransformPStoOther.m

This program converts a PS molecular weight axis to the molecular weight axis of any other

polymer in this thesis based upon the Mark-Houwink-Sakurada coefficients.

function NewMWs=TransformPStoOther(MWxaxisPS,choice) %THis converts from PS to another polymer using Mark Houwink Coefficients Kps=0.00863; aps=.736; %there are multiple polymers it can be converted to. We switch between %them switch choice case 1 %PnBA %need to redefine Kps and aps because this data is from a different %source. Therefore the relative values have to be maintained. Kps=.000114; aps=.716; K=0.000122; a=0.700; case 2 %Poly alpha methyl styrene (mL/g) K=0.0111; a=0.690; case 3 %polyisobutylene (mL/g) K=0.0266; a=.654; case 4 %polymethylmethacrylate (mL/g) K=0.00897; a=0.710; case 5 %1,4 polybutadiene K=.0252; a=0.727; end check=find(MWxaxisPS<1); MWxaxisPS(check)=1; logMWnew=(1./(1+a)).*log10(Kps./K)+((1+aps)./(1+a)).*log10(MWxaxisPS); NewMWs=10.^(logMWnew); end

Page 235: Photochemical Crosslinking Reactions in Polymers

214

BaselineChoice.m

This program allows the subtraction of baseline drift or noise from the data. Options are a

linear drift or a curve that is fit to the tail of the solvent peak, as some data included that tail.

The solvent peak was run 10 times, averaged together, and then a fit was produced. This code

fits the data baseline curve to that fit. Linear drift is also available.

function [Output FittedData]=BaselineChoice(xData,yData,FittedData,startover) %this allows one to compare the two baseline subtractions and adjust them %to best produce results. i=1; if startover==1 start=1; else load BaselineChoicetemp start=i; end clear i %for bulk processing, it saves outside of matlab every 5 spectra for i=start:1:size(yData,2) i if i==5 || i==10 ||i==15 || i==20 ||i==25 || i==30 ||i==35 || i==40 ||i==45 || i==50 ||i==55 || i==60 ||i==65 || i==70 ||i==75 || i==80 ||i==85 || i==90 ||i==95 || i==100 ||i==105 || i==110 ||i==115 || i==120 ||i==125 || i==130 ||i==135 || i==140 save BaselineChoicetemp end %this while loop is the heart; it cycles until the user says the fit is acceptable. choice=0; while choice==0 %plot the fits that may already be present close all LineFit(:,i)=FitLine(xData(:,i),yData(:,i),[100:9000]); plot(yData(:,i),'k') hold text(1000,1,num2str(i)); plot(FittedData(:,i),'r') plot(LineFit(:,i),'b') legend(['Original Data ';'Decay Fit Data';'Line Fit Data ']) xlabel('Index') ylabel('GPC') choice2=input('Redo any fit? 1 is Linear, 2 is Decay: '); if choice2==1 range=input('What is the range for a linear fit? Form [x:y]: '); LineFit(:,i)=FitLine(xData(:,i),yData(:,i),range);

Page 236: Photochemical Crosslinking Reactions in Polymers

215

FittedDecay=FittedData(:,i); elseif choice2==2 range=input('What is the range for a Decay fit?

Form [x:y]: '); LineFit(:,i)=ones(size(yData,1),1); FittedDecay=SubtractTail(yData(:,i),range); else FittedDecay=FittedData(:,i); end close all plot(yData(:,i),'k') hold text(1000,1,num2str(i)); plot(FittedDecay,'r') plot(LineFit(:,i),'b') if choice2==1 plot(yData(:,i)-LineFit(:,i),'g') elseif choice2==2 plot(yData(:,i)-FittedDecay,'g') else plot(yData(:,i)-FittedDecay,'g') plot(yData(:,i)-LineFit(:,i),'c') end legend(['Original Data ';'Decay Fit Data';'Line Fit Data

';'SubtractedData';'SubtractedLine']) xlabel('Index') ylabel('GPC') which=input('What fit do you prefer? 1 is line, 2 is Decay,

anything else reasks the question: '); switch which case 1 Output(:,i)=LineFit(:,i); choice=1; case 2 FittedData(:,i)=FittedDecay; Output(:,i)=FittedData(:,i); choice=1; otherwise choice=0; end end end end

Page 237: Photochemical Crosslinking Reactions in Polymers

216

FitLine.m

This is a linked function of BaselineChoice.m that fits a line to the baseline.

function [fitted estimates]=FitLine(xData,yData,range) if size(xData,2)==1 xData=repmat(xData,1,size(yData,2)); end %find the range of the data to fit, then pass data to fit function for i=1:size(yData,2) y=yData(range,i); x=xData(range,i); start_point = rand(1, 2); options=optimset('Display','none','MaxIter',10000,'MaxFunEvals',10000,

'TolX',1E-40,'TolFun',1E-40); estimates(i,:) = fminsearch(@expfun, start_point,options); fitted(:,i)=estimates(i,1).*xData(:,i)+estimates(i,2); end figure plot(xData,yData,'o') hold plot(xData,fitted,'-') function sse = expfun(params) m = params(1); b = params(2); line = m.*x+b; FittedCurve=line; ErrorVector = FittedCurve - y; sse = sum(ErrorVector .^ 2); end end

Page 238: Photochemical Crosslinking Reactions in Polymers

217

SubtractTail.m

This is another linked function to BaselineChoice.m. It fits the solvent tail discussed previously.

function [Fitted Subtracted estimates FVAL]=SubtractTail(yData,range) %this fits the experimental decay of the solvent peak to the beginning of %the data so as to remove whatever decay is present, if any. close all %Get the full fitting curve. Dynamically generating it wastes lots of %time. We will instead dynamically change the subset of it we investigate %for fit xData=((1:1:size(yData,1))./(10.*60))'; %do the fit for every data set in the array (column) for i=1:size(yData,2) y=yData(range,i); %make sure the error is smaller than 100. If not, redo fit with %new start points. 100 is arbitrary, a magic number. We also stop the %retesting if we can't find a smaller solution after 100 iterations of %the search (again, a magic number) skip = 0; counter = 1; start_point_lin=rand(26,1);

while skip==0;

options=optimset('Display','iter','MaxIter',10000,'MaxFunEvals',10000,'

TolX',1E-20,'TolFun',1E-20); [estimates(:,i) FVAL(:,i)]= fminsearch(@fitfun, start_point_lin,options);

coefficients=estimates(:,i); a = coefficients(1); b = coefficients(2); c = coefficients(3); d = coefficients(4); e = coefficients(5); f = coefficients(6); g = coefficients(7); h = coefficients(8); ii = coefficients(9); j = coefficients(10); k = coefficients(11); l = coefficients(12); m = coefficients(13); n = coefficients(14); o = coefficients(15); p = coefficients(16);

Page 239: Photochemical Crosslinking Reactions in Polymers

218

q = coefficients(17); r = coefficients(18); s = coefficients(19); t = coefficients(20); u = coefficients(21); v = coefficients(22); w = coefficients(23); xx = coefficients(24); yy = coefficients(25); z = coefficients(26);

Fitted(:,i)=f./(a.*(xData.^e+c))+d+g./(h.*(xData.^ii+j))+k./(l.*(xData.

^m+n))+o./(p.*(xData.^q+r))+s./(t.*(xData.^u+v))+w./(xx.*(xData.^yy+z));

%here you set the quality of the fit; it will recalculate until the termination parameters here are satisfied. if FVAL(:,i)<1000 || counter == 20 skip=1; elseif FVAL(:,i)>3000 start_point_lin=rand(26,1); skip=0; counter=counter+1 elseif FVAL(:,i)>1000 && FVAL(:,i)<3000 start_point_lin=coefficients; skip=0; end end end function sse=fitfun(coefficients) fitted=TheFit(coefficients,range(end),range(1)); %remove the importance of the first 5% of pts in the fit;

%this is where the largest differences are and it is the least %important part of the fit since the data comes out at the end.

thesize=max(size(range)); begin=round(0.05.*thesize); sse=sum((y-fitted).^2); end Subtracted=yData-Fitted; beep beep beep end

Page 240: Photochemical Crosslinking Reactions in Polymers

219

TheFit.m

A subfunction of SubtractTail, this contains the coefficients and assignments.

function FitOut=TheFit(coefficients,howmany,start) x=(start:1:howmany)'; x=x./(10.*60); %arrived at through iterative solver coefficients = [ 0.0244; -0.9764; 0.5183; 1.0842; 1.1943; 0.5741; 0.3423; 0.2008; 0.0031; -0.2028; -1.3604; -0.0181; 1.4508; 2.6194; 0.0079; -0.0015; 0.1100; 0.4387; -0.0482; 0.0011; 0.8196; 0.2022; 0.1109; 0.0032; 0.8422; 0.1687; ]; % in case a new fit needs to be made, just comment above and uncomment below % a = coefficients(1); % b = coefficients(2); % c = coefficients(3); % d = coefficients(4); % e = coefficients(5); % f = coefficients(6); % g = coefficients(7); % h = coefficients(8); % i = coefficients(9); % j = coefficients(10); % k = coefficients(11); % l = coefficients(12); % m = coefficients(13); % n = coefficients(14); % o = coefficients(15);

Page 241: Photochemical Crosslinking Reactions in Polymers

220

% p = coefficients(16); %q = coefficients(17); % r = coefficients(18); % s = coefficients(19); % t = coefficients(20); % u = coefficients(21); % v = coefficients(22); % w = coefficients(23); % xx = coefficients(24); % yy = coefficients(25); % z = coefficients(26); FitOut=f./(a.*(x.^e+c))+d+g./(h.*(x.^i+j))+k./(l.*(x.^m+n))+o./(p.*(x.^q+r))+s./(t.*(x.^u+v))+w./(xx.*(x.^yy+z)); end

Page 242: Photochemical Crosslinking Reactions in Polymers

221

G. Data Analysis M-Files

After the data has the baseline removed, it was analyzed. The following m-files were

used in the analysis.

Histogram.m

Histogram linearized the distribution. It takes the molecular weight axis in number of

monomers which, due to the nonlinearity of separation, are often fractional according to the

calibration curve and converts the x-axis into degree of polymerization, N. It also bins the data

such that it is also only for whole number molecules. It does this by averaging all the fractional

polymer lengths and placing the result in the nearest whole number.

function [linaxis,Histograms]=Histogram(Naxis,Data) %this takes an N axis and a set of Data, and fills the gaps in the data %using a basic average, or reduces the range of a single N to one data %point %round Naxis roundN=round(Naxis); roundN=repmat(roundN,1,size(Data,2)); %create a new axis that goes by 1 from 1 to the max of Naxis (up to 10000 %monomers) if max(roundN)<=10000 linaxis=(max(roundN):-1:1)'; elseif max(roundN)>10000 linaxis=(10000:-1:1)'; end linaxis=repmat(linaxis,1,size(Data,2)); %now we map the Data to the new axis Histograms=zeros(size(linaxis,1),size(linaxis,2)); for k=1:1:size(linaxis,2) %cycle through lengths and directly match sizes for i=1:1:size(linaxis,1) %find the indices of the data that matches the current length x=find(roundN(:,k)==linaxis(i,k)); if isempty(x)

Page 243: Photochemical Crosslinking Reactions in Polymers

222

%if no data is of the current length, amplitude is zero Histograms(i,k)=0; else %if some data is of current length, average that data for %amplitude Histograms(i,k)=sum(Data(x,k),1)./size(x,1); end end

%fill in the holes in the distribution by averaging the two nearest %values

for i=2:1:size(linaxis)-1 %if the amplitude is zero, find the next nonzero amplitude if Histograms(i,k)==0 for j=i+1:size(linaxis,1) if Histograms(j,k)~=0 %finds where the next number is and sets j to that index break end end %the amplitude is the average of the previous amplitude and the %next Histograms(i,k)=mean([Histograms(i-1,k);Histograms(j,k)]); end end end end

Page 244: Photochemical Crosslinking Reactions in Polymers

223

AreaNormalize.m

This m-file uses the trapezoidal reiman sum to calculate the area under a user-specified

curve.

function [Norm,Area,AreaAfter]=AreaNormalize(xaxis,yaxis,range) %This normalizes by area by doing a trapezoidal reiman sum of the %form: (1/2)*Q*[f(a)+2f(a+Q)+2f(a+2Q)+2f(a+3Q)+...+f(b)] %where a is the beginning of the xaxis, b the end, and Q the step %size. So it's .5*rate*[y(1)+2y(2)+2y(3)+...+2y(b-1)+y(b)] step=xaxis(range,1); ycalc=yaxis(range,:); Area=trapz(-step,ycalc,1); Norm=yaxis./abs(repmat(Area,size(yaxis,1),1)); AreaAfter=trapz(-step(:,1),Norm(range,:),1) end

Page 245: Photochemical Crosslinking Reactions in Polymers

224

HeightNorm.m

This m-file takes user-specified data and normalizes its maximum height in a specified

range to the same range in a control. It was extended to also subtract off the control peak after

normalization, determine the percent depletion, scale the subtracted peak by that, and

calculate the area of the residual change in distribution. All the advanced options are included

in varargin.

function [Normed Depletion Subtracted SubtractedScaled AreaComb AreaScis]=HeightNorm(Data,control,range,varargin) if size(control,2)==1 control=repmat(control,1,size(Data,2)); end %find the max value in the control and data within the specified range. Then %calculate depletion and scale the data an appropriate amount. [index location]=max(Data(range,:),[],1); cont=max(control(range,:),[],1); for i=1:size(Data,2) Normed(:,i)=cont(i).*Data(:,i)./index(i); Depletion(:,i)=index(i)./cont(i); if Depletion(:,i)>1; Depletion(:,i)=1./Depletion(:,i); end Depletion(:,i)=100-Depletion(:,i).*100; end Depletion(:,1)=0; %subtract the scaled data from the control to highlight changes Subtracted=Normed-repmat(Normed(:,1),1,size(Normed,2)); SubtractedScaled=Subtracted./repmat(abs(Depletion),size(Subtracted,1),1); %if calculations of the area of the changes are needed, do them if nargin==5 %varargin{1} is the xaxis while varargin{2} is the outer bounds of the area calculation for i=1:size(Subtracted,2) [~,AreaComb(i,:),~]=AreaNormalize(varargin{:,1}(:,1),Subtracted(:,i),

varargin{:,2} (1,1):location(:,i)+range(1)); [~,AreaScis(i,:),~]=AreaNormalize(varargin{:,1}(:,1),Subtracted(:,i),

location(:,i)+range(1):varargin{:,2}(end,end)); end AreaComb=AreaComb'; AreaScis=AreaScis'; end end

Page 246: Photochemical Crosslinking Reactions in Polymers

225

MolWDist.m

Calculates the number average, weight average, and PDI of the input distribution.

function [Mn Mw PDI]=MolWDist(MWaxis,Amplitude,range) %Mn calc %insure all positive amplitudes Amplitude=abs(Amplitude(range,:)); %calculate the numerator (Number of chains*MW of chains) num=sum(Amplitude.*repmat(MWaxis(range,:),1,size(Amplitude,2)),1); %calculate the denominator (Number of chains) den=sum(Amplitude,1); Mn=(num./den)'; %Mw calc %calculate the numerator (Number of chains*MW of chains^2) num2=sum(Amplitude.*(repmat(MWaxis(range,:),1,size(Amplitude,2))).^2,1); %calculate the denominator (Number of chains*MW of chains) den2=sum(Amplitude.*repmat(MWaxis(range,:),1,size(Amplitude,2)),1); Mw=(num2./den2)'; %PDI Calc PDI=Mw./Mn; end

Page 247: Photochemical Crosslinking Reactions in Polymers

226

SameX.m

Shifts the distribution on the x-axis to match the location of a user specified peak. Used

in adjustment to internal standards.

function Shifted=SameX(yData,AveData,range) %find the x-axis location of the int standard peak and that of the control [~,loc]=max(yData(range,:),[],1) [~,loc2]=max(AveData(range,:),[],1) %add zeros to the beginning or end of the data array to push the peak to the %same x location diff=loc2-loc absdiff=abs(diff) for i=1:size(yData,2) if diff(i)<0 Shifted(:,i)=[yData(absdiff(i)+1:end,i); zeros(absdiff(i),1)]; elseif diff(i)>0 Shifted(:,i)=[zeros(absdiff(i),1); yData(1:end-absdiff(i),i)]; elseif diff(i)==0 Shifted(:,i)=yData(:,i); else disp('Shift Error') end end %check for error and output the shift for user information [~,locE]=max(Shifted(range,:),[],1); [~,loc2E]=max(AveData(range,:),[],1); err=locE-loc2E plot(Shifted) end

Page 248: Photochemical Crosslinking Reactions in Polymers

227

9.1. Appendix References

1. O. Saito, J. Phys. Soc. Jpn. 13, 1451 (1958). 2. O. Saito, J. Phys. Soc. Jpn. 13, 198 (1958). 3. O. Saito, J. Phys. Soc. Jpn. 13, 1465 (1958). 4. O. Saito, Ed., Statistical Theories of Crosslinking, vol. 1 (Academic Press, New York, ed. 1,

1972), vol. 1, 1, pp. 369. 5. V. J. Triacca, P. E. Gloor, S. Zhu, A. N. Hrymak, A. E. Hamielec, Polym. Eng. Sci. 33, 445

(Apr, 1993). 6. Y. Kodera, B. J. McCoy, AIChE J. 43, 3205 (Dec, 1997). 7. M. Wang, J. M. Smith, B. J. McCoy, AIChE J. 41, 1521 (Jun, 1995). 8. R. P. W. Scott, Liquid Chromatography. E. S. P. Company, Ed., Journal of

Chromatography Library (Elsevier, New York, ed. 2, 1986), vol. 33, pp. 271. 9. B. H. Zimm, W. H. Stockmayer, J. Chem. Phys. 17, (1949).


Recommended