polyisobutylene chain end transformations: block copolymer

268
The University of Southern Mississippi The University of Southern Mississippi The Aquila Digital Community The Aquila Digital Community Dissertations Spring 5-2010 Polyisobutylene Chain End Transformations: Block Copolymer Polyisobutylene Chain End Transformations: Block Copolymer Synthesis and Click Chemistry Functionalizations Synthesis and Click Chemistry Functionalizations Andrew Jackson David Magenau University of Southern Mississippi Follow this and additional works at: https://aquila.usm.edu/dissertations Part of the Polymer Chemistry Commons Recommended Citation Recommended Citation Magenau, Andrew Jackson David, "Polyisobutylene Chain End Transformations: Block Copolymer Synthesis and Click Chemistry Functionalizations" (2010). Dissertations. 872. https://aquila.usm.edu/dissertations/872 This Dissertation is brought to you for free and open access by The Aquila Digital Community. It has been accepted for inclusion in Dissertations by an authorized administrator of The Aquila Digital Community. For more information, please contact [email protected].

Upload: others

Post on 16-Oct-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Polyisobutylene Chain End Transformations: Block Copolymer

The University of Southern Mississippi The University of Southern Mississippi

The Aquila Digital Community The Aquila Digital Community

Dissertations

Spring 5-2010

Polyisobutylene Chain End Transformations: Block Copolymer Polyisobutylene Chain End Transformations: Block Copolymer

Synthesis and Click Chemistry Functionalizations Synthesis and Click Chemistry Functionalizations

Andrew Jackson David Magenau University of Southern Mississippi

Follow this and additional works at: https://aquila.usm.edu/dissertations

Part of the Polymer Chemistry Commons

Recommended Citation Recommended Citation Magenau, Andrew Jackson David, "Polyisobutylene Chain End Transformations: Block Copolymer Synthesis and Click Chemistry Functionalizations" (2010). Dissertations. 872. https://aquila.usm.edu/dissertations/872

This Dissertation is brought to you for free and open access by The Aquila Digital Community. It has been accepted for inclusion in Dissertations by an authorized administrator of The Aquila Digital Community. For more information, please contact [email protected].

Page 2: Polyisobutylene Chain End Transformations: Block Copolymer

The University of Southern Mississippi

POLYISOBUTYLENE CHAIN END TRANSFORMATIONS:

BLOCK COPOLYMER SYNTHESIS AND CLICK CHEMISTRY

FUNCTIONALIZATIONS

by

Andrew Jackson David Magenau

Abstract of a Dissertation Submitted to the Graduate School

of The University of Southern Mississippi in Partial Fulfillment of the Requirements

for the Degree of Doctor of Philosophy

May 2010

Page 3: Polyisobutylene Chain End Transformations: Block Copolymer

ii

ABSTRACT POLYISOBUTYLENE CHAIN END TRANSFORMATIONS: BLOCK COPOLYMER

SYNTHESIS AND CLICK CHEMISTRY FUNCTIONALIZATIONS

by Andrew Jackson David Magenau

May 2010

The primary objectives of this research were twofold: (1) development of

synthetic procedures for combining quasiliving carbocationic polymerization (QLCCP)

of isobutylene (IB) and reversible addition fragmentation chain transfer (RAFT)

polymerization for block copolymer synthesis; (2) utilization of efficient, robust, and

modular chemistries for facile functionalization of polyisobutylene (PIB). Two site

transformation strategies were employed to create block copolymers effectively linking

PIB with either poly(methylmethacrylate) (PMMA), polystyrene (PS), and poly(N-

isopropylacrylamide) (PNIPAM) block segments. Functionalization of PIB was

accomplished by utilizing two click chemistries, the azide-alkyne 1,3-dipolar cyclo

addition and the thiol-ene hydrothiolation reaction, and by efficient transformation of the

thiol functional group.

In the first study block copolymers consisting of PIB, and either PMMA or PS

block segments, were synthesized by a site transformation approach combining living

cationic and reversible addition-fragmentation chain transfer (RAFT) polymerizations.

The initial PIB block was synthesized via quasiliving cationic polymerization using the

TMPCl/TiCl4 initiation system and was subsequently converted into a hydroxyl-

terminated PIB. Site transformation of the hydroxyl-terminated PIB into a macro chain

Page 4: Polyisobutylene Chain End Transformations: Block Copolymer

iii

transfer agent (PIB-CTA) was accomplished by N,N’-

dicyclohexylcarbodiimide/dimethylaminopyridine-catalyzed esterification with 4-cyano-

4-(dodecylsulfanylthiocarbonylsulfanyl)pentanoic acid. Structure of the PIB-CTA was

confirmed by both 1H and 13C NMR spectroscopy. The PIB-CTA was then employed in

a RAFT polymerization of either methyl methacrylate or styrene resulting in PIB block

copolymers with narrow polydispersity indexes and predetermined molecular weights,

confirmed by both 1H NMR and GPC.

In the second study another site transformation approach was developed to

synthesize a novel block copolymer, composed of PIB and PNIPAM segments. The PIB

block was prepared via quasiliving cationic polymerization and end functionalized by in-

situ quenching to yield telechelic halogen-terminated PIB. Azido functionality was

obtained by displacement of the terminal halogen through nucleophilic substitution,

which was confirmed by both 1H and 13C NMR. Coupling of an alkyne-functional chain

transfer agent (CTA) to azido PIB was successfully accomplished through a copper

catalyzed click reaction. Structure of the resulting PIB-based macro-CTA was verified

with 1H NMR, FTIR, and GPC; whereas coupling reaction kinetics were monitored by

real time variable temperature (VT) 1H NMR. Subsequently, the function of this macro-

CTA was demonstrated by RAFT polymerization of NIPAM for synthesis of the second

block. RAFT kinetics was investigated under a variety of reaction conditions using VT

NMR, and the resulting block copolymers were characterized by 1H NMR and GPC.

Aqueous solution properties were probed by dynamic light scattering confirming the

presence of self assembled aggregates with reversible temperature sensitive

responsiveness.

Page 5: Polyisobutylene Chain End Transformations: Block Copolymer

iv

In a third study, a click chemistry functionalization procedure was developed

based upon the azide-alkyne 1,3-dipolar cycloaddition reaction. 1-(ω-

Azidoalkyl)pyrrolyl-terminated PIB was successfully synthesized both by substitution of

the terminal halide of 1-(ω-haloalkyl)pyrrolyl-terminated PIB with sodium azide and by

in situ quenching of quasiliving PIB with a 1-(ω-azidoalkyl)pyrrole. Azide substitution

of the terminal halide was carried out in 50/50 heptane/DMF at 90°C for 24 h using

excess azide. The 1-(ω-haloalkyl)pyrrolyl-PIB precursors included 1-(2-

chloroethyl)pyrrolyl-PIB, 1-(2-bromoethyl)pyrrolyl-PIB, and 1-(3-bromopropyl)pyrrolyl-

PIB. In-situ quenching involved direct addition of 1-(2-azidoethyl)pyrrole to quasiliving

PIB initiated from 5-tert-butyl-1,3-di(1-chloro-1-methylethyl)benzene (t-Bu-m-

DCC)/TiCl4 at -70 °C in hexane/CH3Cl (60/40, v/v). 1H NMR analysis of the quenched

product revealed mixed isomeric end groups in which PIB was attached at either C2 or C3

of the pyrrole ring (C2/C3 = 0.40/0.60). GPC indicated the absence of coupled PIB under

optimized conditions, confirming exclusive mono-substitution on each pyrrole ring. 1-(3-

Azidopropyl)pyrrolyl-PIB was reacted in modular fashion with various functional

alkynes, propargyl alcohol, propargyl acrylate, glycidyl propargyl ether, and 3-

dimethylamino-1-propyne, via the Huisgen 1,3-dipolar cycloaddition “Click” reaction,

using Cu(I)Br/N,N,N′,N′′,N′′-pentamethyldiethylnetriamine or bromtris(triphenyl-

phosphine)Cu(I) as catalyst. The reactions were quantitative and produced PIBs bearing

terminal hydroxyl, acrylate, glycidyl, or dimethylaminomethyl groups attached via

exclusively 4-substituted triazole linkages.

In a fourth study, radical thiol-ene hydrothiolation “Click” chemistry was

explored and adapted to easily and rapidly modify exo-olefin PIB with an array of thiol

Page 6: Polyisobutylene Chain End Transformations: Block Copolymer

v

compounds bearing useful functionalities, including primary halogen, primary amine,

primary hydroxyl, and carboxylic acid. The thiol-ene “click” procedure was shown to be

applicable to both mono and difunctional exo-olefin polyisobutylene. Telechelic mono-

and difunctional exo-olefin PIBs were synthesized via quasiliving cationic

polymerization followed by quenching with the hindered amine, 1,2,2,6,6-

pentamethylpiperidine. Lower reaction temperatures were found to increase exo-olefin

conversion to near quantitative amounts. Thiol-ene reactions with cysteamine and

cysteamine hydrochloride resulted in no thioether formation. Primary amine-terminated

PIB was successfully obtained via a two–step, one-pot procedure, by using a tert-

butoxycarbonyl (BOC) protected amine during the hydrothiolation step and subsequent

deblocking with triflouroacetic acid (TFA). A sunlight-activated hydrothiolation reaction

was also demonstrated; although detectable byproducts resulted.

In the fifth study, thiol-terminated polyisobutylene (PIB-SH) was synthesized by

reaction of thiourea with α,ω-bromine-terminated PIB in a three step one-pot procedure.

First the alkylisothiouronium salt was produced using a 1:1 (v:v) DMF:heptane cosolvent

mixture at 90°C. Hydrolysis of the salt by aqueous base produced thiolate chain ends,

which were then acidified to form the desired thiol functional group. Structural evidence

of the thiol functionality was provided by 1H and 13C NMR, indicating complete

conversion of the terminal halogen. Competing sulfide formation was effectively

suppressed during the base-hydrolysis step, as verified by GPC, by increased reaction

temperature. Utility of PIB-SH was demonstrated through a series of thiol-based “click”

reactions. Alkyne-terminated PIB was synthesized by a phosphine-catalyzed thiol-ene

Michael addition reaction with propargyl acrylate. An extension of this reaction was

Page 7: Polyisobutylene Chain End Transformations: Block Copolymer

vi

performed by a sequential thiol-ene/thiol-yne procedure to produce tetra-hydroxy

functionalized PIB. 1H NMR was used to confirm formation of both alkyne and tetra-

hydroxyl functional species. Further utility of PIB-SH was demonstrated by base

catalyzed thiol-isocyanate reactions. A model reaction was conducted with phenyl

isocyanate in THF using triethylamine as the catalyst. Similar conditions were then used

to successfully synthesize PIB-based polythiolurethanes, with and without a small-

molecule dithiol chain extender. Increased molecular weights and conversion of thiol

functional groups were observed with GPC and 1H NMR, respectively. Last, conversion

of PIB-SH directly into a RAFT macro-CTA was accomplished, as shown by 1H NMR,

by treatment of PIB-SH with triethylamine in carbon disulfide and subsequent alkylation

with 2-bromopropionic acid.

Page 8: Polyisobutylene Chain End Transformations: Block Copolymer

COPYRIGHT BY

ANDREW JACKSON DAVID MAGENAU

2010

Page 9: Polyisobutylene Chain End Transformations: Block Copolymer

viii

DEDICATION

This dissertation is dedicated to my family for their love and support.

To my grandfather and grandmother, John Martin Jr. and Carol Magenau, for

encouragement of an open mind, creativity, and temperance.

To my grandmother, Mary Patrone, for her fortitude and discipline.

To my brother, John Martin Magenau IV, and his wife, Heather Simpkins, for their

inspiration.

To my parents, John Martin III and Ann Marie Rose Magenau, for their wisdom,

guidance, and dedication.

To Audrey Magenau for her kindness.

For everyone, I am eternally grateful.

Page 10: Polyisobutylene Chain End Transformations: Block Copolymer

ix

ACKNOWLEDGEMENTS

I would like to acknowledge my advisor, Professor Robson F. Storey, for being a

great friend, mountaineer, and mentor. His scientific expertise, creativity, and guidance

have had a profound impact on my graduate career and life. I would also like to

acknowledge my graduate committee members for their time and guidance: Dr. William

Jarrett, Dr. Sarah Morgan, Dr. Daniel Savin, and Dr. Daniel Savin. I would like to thank

the faculty of the School of Polymers as well as the entire College of Science and

Technology for offering resources, expertise, and a wonderful graduate education.

I would like to thank all of my fellow graduate students, my first year class, and

the Storey Research Group, for they have been a source of knowledge, inspiration, and

camaraderie. In particular, I would like to thank Dr. Nemesio Martinez Castro, Laura

Fosselman, Irene Gorman, Todd Hartlage, Dr. Andrew Jackson, Dr. Lisa Kemp, David

Morgan, Dr. Scott Morovek, Megan Powell, and Yaling Zhu. Special gratitude is

extended to the past Dr. Charles E. Hoyle for his insight, fruitful discussion, and

encouragement in my graduate career.

I would like to thank all of my close friends for their endless support.

Page 11: Polyisobutylene Chain End Transformations: Block Copolymer

x

TABLE OF CONTENTS

ABSTRACT ........................................................................................................................ ii

DEDICATION ................................................................................................................. viii

ACKNOWLEDGMENTS ................................................................................................. ix

LIST OF ILLUSTRATIONS ............................................................................................ xii

LIST OF TABLES ......................................................................................................... xviii

CHAPTER

I. BACKGROUND ........................................................................................ 1

Quasiliving Carbocationic Polymerization ..................................................2 Reversible Addition Fragementation Chain Transfer (RAFT) Polymerization .............................................................................................7

Azide-alkyne Click Chemistry ...................................................................11 Thiol-ene Click Chemistry .........................................................................14 References ..................................................................................................20

II. OBJECTIVES OF RESEARCH ................................................................48

III. SITE TRANSFORMATION OF POLYISOBUTYLENE CHAIN ENDS INTO FUNCTIONAL RAFT AGENTS FOR BLOCK COPOLYMER SYNTHESIS ..............................................................................................50

Introduction ................................................................................................50 Experimental ..............................................................................................52 Results and Discussion ..............................................................................58 Conclusion .................................................................................................64 References ..................................................................................................66

IV. POLYISOBUTYLENE RAFT CTA BY A CLICK CHEMISTRY SITE TRANSFORMATION APPROACH: SYNTHESIS OF POLY(ISOBUTYLENE-b-N-ISOPROPYLACRYLAMIDE)..................93

Introduction ................................................................................................93 Experimental ..............................................................................................96 Results and Discussion ............................................................................102 Conclusion ...............................................................................................109 References ................................................................................................111

Page 12: Polyisobutylene Chain End Transformations: Block Copolymer

xi

V. FUNCTIONAL POLYISOBUTYLENES VIA A CLICK CHEMISTRY APPROACH ............................................................................................132

Introduction ..............................................................................................132 Experimental ............................................................................................134 Results and Discussion ............................................................................142 Conclusion ...............................................................................................154 Acknowledgements ..................................................................................155 References ................................................................................................156

VI. FACILE POLYISOBUTYLENE FUNCTIONALIZATION VIA THIOL-ENE CLICK CHEMISTRY .......................................................183 Introduction ..............................................................................................183 Experimental ............................................................................................185 Results and Discussion ............................................................................189 Conclusion ...............................................................................................193 References ................................................................................................194

VII. TRANSFORMATIONS OF α,ω-THIOLPOLYISOBUTYLENE: EFFICIENT ROUTES TO A CHAIN TRANSFER AGENT, TELECHELICS, AND PIB BASED THIOURETHANES .....................213 Introduction ..............................................................................................213 Experimental ............................................................................................216 Results and Discussion ............................................................................220 Conclusion ...............................................................................................229 References ................................................................................................230

Page 13: Polyisobutylene Chain End Transformations: Block Copolymer

xii

LISTOF ILLUSTRATIONS

Figure 1.1 Polymeric architectures via controlled living polymerization ...................28

1.2 Chain breaking events in the cationic polymerization of IB ......................29

1.3 Chain equilibrium between (a) dormant, (b) ion pair, (c) free-ion pair chains. P is the propagating chain end, X is the counterion, Keq is the ionization equilibrium constant, and Kdis is the ion pair dissociation equilibrium constant...................................................................................30

1.4 GPC (RI) traces of PIB prepared by the TMPCl/TiCl4 system in the absence and the presence of ED.................................................................31

1.5 Proposed mechanism for the formation of common ions through the scavenging of protic impurities by 2,4-dimethylpyridine. .........................32

1.6 CRP equilibrium between active and dormant chains with (a) SFRP, (b) ATRP, and (c) RAFT .................................................................................33

1.7 The RAFT polymerization mechanism ......................................................34

1.8 Thermal initiators employed in RAFT polymerizations ............................35

1.9 Generic structures of RAFT chain transfer agents .....................................36

1.10 Dipolar cycloadditions between alkynes and azides ..................................37

1.11 Proposed catalytic cycle for the Cu(I)-catalyzed azide alkyne reaction ....38

1.12 Preparation of chain-end functionalized PS via a combination of ATRP and click chemistry ....................................................................................39

1.13 Incorporation of azide end groups in polystyrene via an end group modification procedure ..............................................................................40

1.14 Introduction of terminal alkyne functionality in polymers utilizing functionalized ATRP initiator ....................................................................41

1.15 Anti-markovnikov hydrothiolation of C=C bond ......................................42

1.16 Radical mediated thiol-ene photoinitiated reaction ...................................43

1.17 Base/Nucleophile catalyzed hydrothiolation .............................................44

1.18 Activated double bonds applicable to the base/nucleophile-mediated hydrothiolation ...........................................................................................45

1.19 Base catalyzed hydrothiolation mechanism ...............................................46

1.20 Nucleophilic phosphine catalyzed hydrothiolation mechanism ................47

3.1 1H NMR spectrum of exo-olefin PIB .........................................................71

3.2 13C NMR spectrum of exo-olefin PIB ........................................................72

Page 14: Polyisobutylene Chain End Transformations: Block Copolymer

xiii

3.3 GPC traces of pre-quenched and quenched exo-olefin PIB .......................73

3.4 1H NMR spectrum of hydroxyl PIB...........................................................74

3.5 13C NMR spectrum of hydroxy PIB...........................................................75

3.6 GPC traces of hydroxy PIB and exo-olefin PIB ........................................76

3.7 Site transformation from living carbocationic polymerization to RAFT polymerization ...........................................................................................77

3.8 1H NMR spectrum of CTA ........................................................................78

3.9 13C NMR spectrum of CTA .......................................................................79

3.10 gHSQC of CTA..........................................................................................80

3.11 13C NMR spectrum of CTA: attached proton test ......................................81

3.12 1H NMR spectrum of PIB-CTA .................................................................82

3.13 13C NMR spectrum of PIB-CTA................................................................83

3.14 13C NMR spectra of CTA(top), PIB-OH(middle), and PIB-CTA(bottom) ......................................................................................84

3.15 Partial 13C NMR spectra of CTA(top), PIB-OH(middle), and PIB-CTA(bottom) ......................................................................................85

3.16 Expanded 13C NMR of CTA(top), PIB-OH(middle), and PIB-CTA (bottom) .....................................................................................86

3.17 Partial 13C NMR spectra comparing PIB-CTA to CTA and PIB-OH precursors ...................................................................................................87

3.18 1H NMR spectrum of PIB-b-PMMA .........................................................88

3.19 GPC traces of PIB-CTA and PIB-b-PMMA after purification ..................89

3.20 Effect of system concentration on rate of RAFT polymerization ..............90

3.21 GPC traces of reaction aliquots removed from reaction R-3 at various conversions, along with the final purified block copolymer......................91

3.22 Dependency of blocking efficiency on monomer selection: MMA vs. S .................................................................................................92

4.1 Synthesis of PIB-CTA for RAFT polymerization ...................................118

4.2 Real-time 1H NMR analysis of click reaction between CTA and PIB44-N3 ...................................................................................................119

4.3 1H NMR spectra of (top)PIB-N3, (middle)CTA, and (bottom)PIB-CTA ....................................................................................120

4.4 FTIR spectra of (top) PIB-N3 and (bottom) PIB-CTA ............................121

4.5 GPC trace before and after click reaction ................................................122

4.6 RAFT polymerization of NIPAM with PIB-CTA ...................................123

Page 15: Polyisobutylene Chain End Transformations: Block Copolymer

xiv

4.7 1H NMR spectrum of PIB-b-PNIPAM ....................................................124

4.8 GPC traces before and after RAFT polymerization of PNIPAM ............125

4.9 Conversion vs. time and first-order kinetic plot (inset) for RAFT polymerization from PIB44-CTA as a function of temperature.

[NIPAM] = 1.2 M; [NIPAM]:[CTA]:[I] = 250:1:0.25 ............................126

4.10 Ln kapp for RAFT polymerization vs. reciprocal temperature to determine activation energy and prefactor (min-1) ...................................127

4.11 Conversion vs. time and first-order kinetic plot (inset) for RAFT polymerization from PIB44-CTA as a function of thermal initiator concentration.

[NIPAM] = 1.2 M; [NIPAM]:[CTA] = 250:1; T = 90 °C .......................128

4.12 Conversion vs. time and first-order kinetic plot (inset) for RAFT polymerization as a function of PIB-CTA molecular weight.

[NIPAM] = 1.2 M; [NIPAM]:[CTA]:[I] = 250:1:0.25; T = 85 °C ..........129

4.13 Conversion vs. time and first-order kinetic plot (inset) for RAFT polymerization as a function of PIB-CTA concentration.

[NIPAM] = 1.2 M; [NIPAM]:[I]=1:0.001; T = 85 °C .............................130

4.14 Diameter of Micelles with Temperature (Θ = 90°) .................................131

5.1 Clickable polyisobutylene: synthetic approaches and utilization ............168

5.2 1H NMR spectrum of 1-(2-azidoethyl)pyrrole-PIB prepared by reaction of 1-(2-bromoethyl)pyrrole-PIB with sodium azide (Table 1, Entry 5).

The product is a mixture of isomers with PIB substituted at the 3- (major) and 2- (minor) positions of the pyrrole ring. Inset shows the tether

proton resonances of the bromide precursor for comparison ...................169

5.3 1H NMR spectrum of difunctional 1-(3-azidopropyl)pyrrole-PIB, with peak integrations, prepared by reaction of difunctional 1-(3-bromopropyl)pyrrole-PIB with sodium azide (Table 1, Entry 8). The product is a mixture of isomers with PIB substituted at the 3- (major)

and 2- (minor) positions of the pyrrole ring .............................................170

5.4 1H NMR spectrum of difunctional 1-(2-azidoethyl)pyrrole-PIB prepared by post-polymerization reaction of the corresponding difunctional 1-(2-bromoethyl)pyrrole-PIB with sodium azide. The product is a mixture of isomers with PIB substituted at the 3- (major) and 2- (minor) positions

of the pyrrole ring (Table 1, Entry 10) .....................................................171

5.5 Reaction conversion vs. time for reaction of difunctional 1-(2-bromoethyl)pyrrole-PIB with sodium azide: 50/50 (v/v) heptanes/DMF, 90°C, [CE] = 0.10 mol/L, [NaN3] = 0.30 mol/L; C3 isomer (triangles),

C2 isomer (squares), overall (circles) .......................................................172

5.6 Reaction conversion vs. time for reaction of monofunctional 1-(2-bromoethyl)pyrrole-PIB with sodium azide: 50/50 (v/v) heptanes/DMF,

Page 16: Polyisobutylene Chain End Transformations: Block Copolymer

xv

90°C, [CE] = 0.10 mol/L, [NaN3] = 0.30 mol/L; C3 isomer (circles), C2 isomer (squares), overall (triangles) ........................................................173

5.7 1H NMR spectrum of difunctional 1-(2-azidoethyl)pyrrole-PIB, with peak integrations, prepared by end-quenching of difunctional tert-

chloride-terminated PIB with N-(2-Azidoethyl)pyrrole in 60/40 (v/v) hexane/MeCl at -70°C (Table 5.2, Entry 1) .............................................174

5.8 13C NMR spectrum of difunctional 1-(2-azidoethyl)pyrrole-PIB prepared by end-quenching of difunctional tert-chloride-terminated PIB with N-(2-Azidoethyl)pyrrole in 60/40 (v/v) hexane/MeCl at -70°C

(Table 5.2, Entry 1) ..................................................................................175

5.9 GPC traces of difunctional tert-chloride-terminated PIB before (dotted line) and after (solid line) end-quenching with 1-(2-azidoethyl)pyrrole (Table 5.4, Entry 1) ..................................................................................176

5.10 1H NMR spectrum of 1-[2-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)ethyl] pyrrole from reaction of 1-(2-azidoethyl)pyrrole with propargyl alcohol catalyzed with Cu(I)Br/PMDETA (Table 5.4, Entry 2) ..........................177

5.11 Reaction kinetics of azide-terminated PIB (2- or 3-carbon tether) with propargyl alcohol in THF at 25°C with initial [alkyne]/[azide] = 2.0: A) Cu(I)Br/PMDETA catalyst with 3-carbon tether; B) bromotris(triphenylphosphine)copper(I) with 2-carbon tether (solid circles) or 3-carbon tether (open circles) .................................................178

5.12 1H NMR spectrum of 1-[3-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl) propyl] pyrrole-PIB prepared by Click chemistry

(Table 5.5, Entry 1) ..................................................................................179

5.13 1H NMR spectrum of 1-[3-(4-vinylcarbonyloxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole-PIB prepared by Click chemistry

(Table 5.5, Entry 2) ..................................................................................180

5.14 1H NMR spectrum of 1-[3-(4-glycidyloxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole-PIB prepared by Click chemistry

(Table 5.5, Entry 3) ..................................................................................181

5.15 1H NMR spectrum of 1-[3-(4-dimethylaminomethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole-PIB prepared by Click chemistry

(Table 5.5, Entry 4) ..................................................................................182

6.1 Photo-initiated radical addition of functional thiols (halo, NH-Boc, carboxylic acid, hydroxyl) to exo-olefin-terminated PIB and subsequent Boc deprotection ......................................................................................199

6.2 GPC traces of mono-functional PIB before (solid line, tert-chloride-terminated PIB) and after quenching with PMP (dashed line, exo-olefin-terminated PIB) ........................................................................................200

Page 17: Polyisobutylene Chain End Transformations: Block Copolymer

xvi

6.3 GPC traces of difunctional PIB before (solid line, tert-chloride-terminated PIB) and after quenching with PMP (dashed line, exo-olefin-terminated PIB) ..........................................................................................................201

6.4 1H NMR spectra of exo-olefin PIB (top) and PIB-Cl (bottom) ...............202

6.5 PIB exo-olefin conversion vs. time for reaction with 2-(tert-butoxycarbonylamino)ethanethiol, monitored by observing diminution of the C=C stretch at 1645 cm-1 using real-time FTIR .................................203

6.6 GPC traces of monofunctional PIB before (dashed line, exo-olefin-terminated PIB) and after thiol-ene reaction (solid line, Boc-protected

and halogen-terminated PIB) ...................................................................204

6.7 1H NMR spectra of PIB-Boc (top) and PIB-NH2 (bottom) .....................205

6.8 FTIR spectra of exo-olefin-terminated PIB (top), Boc-protected PIB (middle), and amine-terminated PIB (bottom).........................................206

6.9 1H NMR spectra of exo-olefin-terminated PIB (top) and carboxylic acid-terminated PIB (bottom) ..........................................................................207

6.10 FTIR spectra of exo-olefin-terminated PIB (top) and carboxylic acid-terminated PIB (bottom) ..........................................................................208

6.11 GPC traces of monofunctional PIB before (dashed line, exo-olefin-terminated PIB) and after thiol-ene reaction (solid line, acid-terminated PIB) ..........................................................................................................209

6.12 1H NMR spectra of exo-olefin-terminated PIB (top) and hydroxyl-terminated PIB (bottom) ..........................................................................210

6.13 FTIR spectra of exo-olefin-terminated PIB (top) and hydroxyl- terminated PIB (bottom) ..........................................................................211

6.14 GPC traces of monofunctional PIB before (dashed line, exo-olefin-terminated PIB) and after thiol-ene reaction (solid line, hydroxyl-terminated PIB) ........................................................................................212

7.1 1H NMR Spectrum of α,ω-bis[4-(3-bromopropoxy)phenyl] PIB............233

7.2 13C NMR Spectrum of α,ω-bis[4-(3-bromopropoxy)phenyl] PIB ..........234

7.3 GPC traces of prequenched (dotted line) and quenched (solid line) PIB ...........................................................................................................235

7.4 1H NMR spectra of α,ω-bis[4-(3-thiopropoxy)phenyl] PIB with comparison of the endgroup regions to the precursor bromine-terminated PIB...........................................................................236

7.5 13C NMR of α,ω-bis[4-(3-thiopropoxy)phenyl]PIB with comparison of the endgroup chemical shifts to those of the bromine-terminated PIB ...........................................................................................................237

7.6 gHSQC of α,ω-bis[4-(3-bromopropoxy)phenyl] PIB .............................238

Page 18: Polyisobutylene Chain End Transformations: Block Copolymer

xvii

7.7 gHSQC of α,ω-bis[4-(3-thiopropoxy)phenyl]polyisobutylene ................239

7.8 Temperature effects on sulfide formation during base hydrolysis step ...240

7.9 Conversion of bromine terminated PIB to isothiouronium salt terminated PIB verses time ......................................................................241

7.10 1H NMR of α,ω PIB-CTA ........................................................................242

7.11 1H NMR of thiol terminated PIB (top) and alkyne terminated PIB (bottom)....................................................................................................243

7.12 13C NMR of alkyne terminated PIB with expansion of alkyne PIB (bottom) and expansion of thiol terminated PIB (top) .............................244

7.13 gHSQC of alkyne terminated PIB............................................................245

7.14 1H NMR of tetra hydroxyl functional PIB ...............................................246

7.15 1H NMR of phenyl thiourethane terminated PIB .....................................247

7.16 GPC traces of (a) chain extended thiolurethane with hexane dithiol (b) chain extended thiourethane without hexane dithiol (c) thiol terminated PIB .........................................................................................248

Page 19: Polyisobutylene Chain End Transformations: Block Copolymer

xviii

LIST OF TABLES

Table 1.1 Boiling points of alcohols and thiols .........................................................27

3.1 Characterization of exo-Olefin PIB Precursors ..........................................69

3.2 RAFT Polymerizations of PMMA and PS ................................................70

4.1 PIB-Br Precursors ....................................................................................116

4.2 RAFT Polymerization Data .....................................................................117

5.1 Conversion of N-(ω-Haloalkyl)pyrrole-PIB to N-(ω-Azidoalkyl)pyrrole-PIB via Nucleophilic Displacement of Halide by NaN3 in N,N-Dimethylformamide/heptane (50/50, v/v) ...............................................159

5.2 End-Quenching of Difunctional tert-Chloride-Terminated Masterbatch PIB with N-(2-Azidoethyl)pyrrole in 60/40 (v/v) Hexane/MeCl at

-70°C ........................................................................................................160

5.3 End-Quenching of Quasiliving IB Polymerizations with N-(2-azidoethyl)pyrrole in 60/40 (v/v) Hexane/CH3Cl at -70°C .............161

5.4 Synthesis of Model Compounds 1-[2-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)ethyl]pyrrole and 1-[3-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole ......................................................................................162

5.5 Experimental conditions and results of functionalization of 1-(3-azidopropyl)pyrrole -PIB by Click chemistry in tetrahydrofuran....163

5.6 Chemical shift assignments for 2-and 3-PIB-1-(2-azidoethyl)pyrrole .......................................................164

5.7 Chemical shift assignments for 2-and 3-PIB-1-(2-azidoethyl)pyrrole .......................................................165

5.8 Chemical shift assignments for 2-and 3-PIB-1-(3-azidopropyl)pyrrole .....................................................166

5.9 Chemical shift assignments for 2-and 3-PIB-1-(3-azidopropyl)pyrrole .....................................................167

6.1 Mono- and difunctional exo-olefin PIB precursors .................................197

6.2 Thiol-ene reactions with exo-olefin PIB ..................................................198

Page 20: Polyisobutylene Chain End Transformations: Block Copolymer

1

CHAPTER I

BACKGROUND

Polyisobutylene (PIB) is a fully saturated hydrocarbon elastomer possessing

superior gas barrier and mechanical damping characteristics, oxidative and chemical

stability, and biocompatibility.1-3 PIB is synthesized from isobutylene exclusively

through carbocationic polymerization. When combined with other materials, the inherent

properties of PIB help produce commercially viable materials such as butyl rubbers,

thermoplastic elastomers, and biomaterials. PIB-based materials have found applications

ranging from inner liners, tubeless tires, electrical insulators, adhesives, motor oil

dispersants, sealants, to coronary stents. Over the last 25 years, methods for controlled or

quasiliving carbocationic polymerization (QLCCP) have been developed. These

advancements have enabled new synthetic strategies to produce precisely tailored block

copolymers from PIB and to efficiently functionalize PIB with a variety of useful

functional groups.

The work presented in this dissertation focuses on transformation of PIB chain

ends. Depending on the transformation used, chain ends can be converted into macro-

initiators for block copolymer synthesis, or a multitude of other useful functional end

groups through the use of robust, efficient, and modular chemistries. Novel approaches

were developed to combine QLCCP and reversible addition fragmentation chain transfer

(RAFT) polymerizations for block copolymers synthesis. In addition, two click

chemistry approaches and efficient thiol group modifications were utilized as versatile

functionalization tools for PIB. This introduction will give a brief background of

controlled polymerizations, focusing on QLCCP and RAFT polymerization, and two

Page 21: Polyisobutylene Chain End Transformations: Block Copolymer

2

click chemistries, namely, 1) the 1,3-dipolar Huisgen cycloaddition reaction of azides

with terminal alkynes, and 2) the thiol-ene family of reactions including radical

hydrothiolation and nucleophilic Michael addition to electrophile enes.

Quasiliving Carbocationic Polymerization

Since discovery of the first living polymerizations, by Szwarc in 1956,

preparation of precise polymers and complex architectures has been possible.4-6 Living

polymerizations, unlike conventional chain polymerizations, proceed in the absence of

termination and chain transfer events. Szwarc’s original definition stated, that

“propagation proceeds with the exclusion of termination and chain transfer, and yields

polymers retaining, virtually indefinitely, their ability to add further monomer whenever

supplied to the system.” Suppression of unwanted termination and chain transfer has

allowed for the synthesis of well-defined polymers and block copolymers of

predetermined molecular weight and narrow molecular weight distributions, and other

advanced architectures. A few examples of the architectures obtainable through

controlled/living polymerizations include statistical (a), alternating (b), AB diblock (c),

and ABA triblock (d) copolymers, ABC block terpolymers (e), tapered block (f) and graft

copolymers(g), and star (h) structures, as shown in Figure 1.1.

Living anionic polymerization was the first technique to enable production of

polymers and block copolymers with controlled molecular weights and narrow PDIs in

the absence of any detectable chain transfer or termination. Extending this idea of

“ideal” living polymerizations to carbocationic systems was initially thought to be

impossible.7 Because carbenium ions are highly unstable (i.e. possess high reactivity), in

comparison to carbanions, they readily participate in undesired side reactions. These

Page 22: Polyisobutylene Chain End Transformations: Block Copolymer

3

reactions typically consist of β-proton elimination, chain transfer to monomer, or

unimolecular carbenium ion rearrangement (Figure 1.2). Highly reactive β-protons are

extremely susceptible to abstraction by basic species within polymerization systems (e.g.

monomers, anionic counterions, and other basic additives).8,9 Therefore, carbocation

lifetimes are typically short, on the order of a few milliseconds, resulting in uncontrolled

polymerizations plagued with termination/transfer events.

In order to realize prolonged lifetime of the growing chain ends (i.e QLCCP), and

create controlled high molecular weight polymers, three main factors had to be

recognized and implemented: (1) control of initiation, (2) realization of reversible

termination, (3) suppression of chain transfer.10 Chain transfer and termination were

circumvented by employing cryogenic reaction conditions, introducing common ions to

suppress ion-pair dissociation, depicted in Figure 1.3 as the equilibrium between species

(b) and (c), and implementation of reaction conditions whereby termination becomes

reversible, i.e., establishment of a dynamic equilibrium between “dormant” and “active”

chain ends, depicted in Figure 1.3 as the equilibrium between species (a) and (b). In the

latter equilibrium, the “dormant” polymer chains (a) cannot participate in chain transfer

or irreversible termination; whereas, further polymerizations can occur by ionization to

the paired ion or “active” state (b). Free ions (c) generally possess higher rates of

decomposition compared to paired ions and are thus suppressed using the common ion

effect.

Early developments toward living cationic polymerization began with the

recognition that certain known systems partially resembled those of a living system. For

example, polymerization of p-methoxystyrene initiated by I2 (i.e. nucleophilic

Page 23: Polyisobutylene Chain End Transformations: Block Copolymer

4

counterion) was observed to yield fairly low PDIs, increased molecular weight upon

sequential monomer addition, and a linear correlation between Mn and monomer

conversion; however, chain end concentrations were observed to increase with monomer

conversion indicating slow initiation.11,12 Another important early cationic

polymerization system was the so-called inifer method, useful for the synthesis of

telechelic PIB, which utilized a molecule functioning as both initiator and chain transfer

agent.13,14 The inifer system was recognized to provide controlled initiation and chain-

end functionality, but it was thought to be non-living. The authors believed that collapse

of ion pairs in these cationic polymerizations was an irreversible termination event.

Later, however, Faust et al. demonstrated that in monomer starved environments, with a

cumyl chloride/BCl3/α-methylstyrene system, a rapid and reversible chain end

equilibrium existed which exhibited “quasiliving” characteristics.15 The basic premise of

quasiliving polymerization is that if termination or chain transfer reactions are present,

but reversible, the system will yield a polymerization that kinetically behaves as a living

system.

The first QLCCP was reported in 1984 by Higashimura et al.16 This system

involved a binary initiating system of HI/I2, in equimolar amounts, for the “living”

cationic polymerization of isobutyl vinyl ether (IBVE). Solvent effects were substantial

due to their profound influence on the dormant-active equilibrium (Keq). It was found

that n-hexane at -15 °C provided the best results by reducing the amount of ionized

species within the system. As previously mentioned, I2 produces the nucleophilic

counteranion, I3-, which provides control by stabilization of the carbocation, i.e.

diminishing chain transfer and spontaneous termination.17 Selection of a “suitably

Page 24: Polyisobutylene Chain End Transformations: Block Copolymer

5

nucleophilic counterion” was of paramount importance; the correct nucleophilicity is

required to create a reversible ionization equilibrium that causes the majority of chain

ends to exist in their dormant state. It was proposed that both initiation and propagation

occurred through a monomer insertion mechanism. “Living” behavior in this system was

attributed to fast initiation, a result of rapid addition of HI to monomer, thereby

producing an initiating species in situ, and the simultaneous absence of chain transfer

reactions due to carbocation stabilization. These systems achieved predetermined (Xn =

[IBVE]/[HI]) and narrow (Mw/Mn ≈ 1.1) molecular weights, proportional increases in Mn

with monomer conversion, constant chain end concentrations, and continued

polymerizations with addition of further monomer. Expansion of this dual initiator

system to other reactive monomers, including p-methoxystyrene18 and N-

vinylcarbazole19, was also demonstrated. Further variations were explored using other

dual initiator systems (e.g. HI/ZnI2) which allowed controlled polymerizations over larger

temperature ranges (-40 °C to 40 °C) and with higher polymerization rates.17 Near the

same time, Faust and Kennedy were using a similar “suitably nucleophilic counterion”

approach with IB in alkyl and aromatic ester or ether/BCl3 initiation systems.20,21 When

targeting low molecular weight PIB, fairly monodisperse (PDI ≈ 1.2-1.3) polymers were

successfully synthesized.

The next significant advancement to further suppress chain transfer and

termination, and improve PDIs in cationic polymerization was the addition of external

electron donors (EDs) to create the common ion effect. Higashimura and coworkers

demonstrated that IBVE polymerizations with strong lewis acids (e.g. SnCl2, TiCl4)

resulted in an uncontrolled system, but could be easily transformed to have quasiliving

Page 25: Polyisobutylene Chain End Transformations: Block Copolymer

6

behavior by charging the reactor with chloride salts.22,23 Improved control was also

observed with less reactive styrene and IB monomers by including salts within the

polymerization.24,25,26 Styrene polymerizations were conducted with 1-phenylethyl

chloride/SnCl4 coinitiation system in CH2Cl2 at -15 °C. To achieve controlled behavior it

was necessary to add nBu4N+Cl-, which resulted in narrow PDIs and targeted molecular

weights. Analogous results were found for IB/TiCl4 polymerizations conducted in the

presence of EDs, including DMSO and DMA.27 Kaszas et al. demonstrated the profound

effect of EDs and their ability to impart control by monitoring molecular weights (Figure

1.4). Interestingly, IB polymerizations conducted using TiCl4 or BCl3 were found to

proceed optimally under conditions of excess of lewis acid relative to ED; whereas IBVE

systems appeared to require excess ED relative to lewis acid.

The exact function of externally added EDs in quasiliving carbocationic

polymerization has been a subject of much debate. Initial theories proposed that EDs

provide “carbocation stabilization.”22,27,28 Later, Faust and coworkers declared their

function was solely to serve as protic scavenger.29 The latter work demonstrated that IB

polymerization rates, using a TiCl4/cumyl-type coinitiation system, were independent of

ED concentration. If indeed the ED served to provide carbocation stability the

polymerization rate should have slowed with higher ED concentrations. The authors

believed that protic initiation was extremely fast,30 and if not suppressed would pre-empt

the much slower controlled QLCCP. Storey and coworkers supported the proposal by

Faust et al. that the ED served as a protic scavenger, but they also postulated that protic

scavenging produced common ion salts, as shown in Figure 1.5, which reduced the rate

of polymerization and conferred livingness by eliminating propagation by unpaired

Page 26: Polyisobutylene Chain End Transformations: Block Copolymer

7

ions.31 The ED function was therefore two-fold: prevention of protic initiation and in situ

creation of common ions. The latter suppress ion pair dissociation via the “common ion

effect.”32 In terms of the Winstein spectrum of ionicities33 shown in Figure 1,3, due to

mass action, addition of the common ion X- shifts the population of chains dramatically

away from dissociated ions (c) in the direction of paired ions (b).

Living carbocationic processes are referred to as “quasiliving” to indicate the

existence of the dormant-active equilibrium, i.e., all of the chains are not active all of the

time. QLCCPs display characteristics of living behavior (i.e. targeted molecular weights,

narrow PDIs, ability to form block copolymers); however, they all exhibit termination.

Since these termination reactions are completely reversible, as illustrated by the dormant-

active chain equilibrium, the polymerizations appear to be truly living.

Reversible Addition-Fragmentation Chain Transfer (RAFT) Polymerization

Although ionic polymerizations give precise control of polymeric architecture,

these techniques require stringent reaction conditions and have a limited selection of

monomers. Radical polymerizations, however, offer many advantages including a large

selection of monomers, tolerance to a broad scope of functionalities and reaction

conditions, and are fairly simple and inexpensive in comparison to other methods.34 In

spite of these advantages, free radicals terminate at nearly diffusion controlled rates

leading to limited control of molecular weights, PDIs, and polymer composition and

architecture. Controlled/living polymerizations (CRP) were developed in an effort to

maintain the versatile monomer selection and robust reaction conditions of conventional

free radical polymerization and advanced architectures achievable with living

polymerizations.

Page 27: Polyisobutylene Chain End Transformations: Block Copolymer

8

CRP relies on the same principles characteristic of living polymerizations by

reducing termination through an equilibrium strongly favoring dormant chains over

propagating chains. This equilibrium effectively lowers radical concentrations, which

reduces the overall rate of polymerization, but in a greater magnitude suppresses the rate

of termination due to its second order dependence on radical concentration. The most

commonly used CRP techniques can be categorized by the mechanism with which

activated/deactivated chains are created. The majority of CRPs rely on a reversible

termination mechanism (Figure 1.6 (a,b)) to impart control; these include iniferters,35,36

stable free radical polymerization,37,38 and atom transfer radical polymerization.39,40 The

second mechanism used to impart control in radical polymerizations is a degenerative

chain transfer process involving a rapid exchange between polymer species (Figure 1.6

(c)) and can be illustrated by RAFT.

RAFT polymerization was first reported in literature by Rizzardo and Moad in

1988.41,42 It is considered to be one of the most versatile controlled/free radical

polymerizations, because of its wide range of applicable monomers and reaction

conditions.43 Controlled/living RAFT polymerizations have been reported with acrylates

and methacrylates, acrylamides and methacrylamides, acrylonitrile, styrene and styrenic

derivatives, butadiene, vinyl acetate, and N-vinylpyrrolidone. In addition, RAFT is

tolerant to a vast range of solvent systems (e.g. organic, heterogeneous, and aqueous),

initiators (e.g. azo, peroxide, photo, γ – radiation, and redox initiating systems), and

functionalities (e.g. OH, NR2, CO2H, SO3H, CONR2). RAFT polymerizations are similar

to conventional radical polymerizations (i.e. same solvents, temperatures, monomers, and

initiator) with the exception of one additional component, a chain transfer agent (CTA).

Page 28: Polyisobutylene Chain End Transformations: Block Copolymer

9

In common with living polymerizations, RAFT polymerizations operate with constant

concentration of growing chain ends, linear increases in molecular weight with

conversion, and preservation of active chains at the end of polymerization.

A series of reversible chain transfer reactions (i.e. addition fragmentation

equilibrium) is responsible for imparting control in RAFT polymerization; the accepted

mechanism is shown in Figure 1.7.44,45 Initiation and termination occur in the same

manner as conventional radical polymerization. Commonly, azo initiators (Figure 1.8)

are used to generate radials (I ) by thermal decomposition in the initiation stage (Figure

1.7 (a)). It is generally believed that these initiator-derived radicals first add to monomer

to produce an oligomeric propagating radical species (Pn ) (Figure 1.7 (b)), due to large

relative ratio of monomer to initiator.46 In the early stages of polymerization, Pn adds to

the thiocarbonyl compound (Z-(C=S)-SR) to produce an intermediate carbon centered

radical, followed by fragmentation into an oligomeric CTA (Pn-(C=S)-SR) species and a

new CTA derived radical (R ) (Figure 1.7 (c)).44 The pre-equilibrium is defined as the

time required for all R fragments to form propagating chains Pm , and is governed by

the four rate constants kadd, k-add, kb and k-b. This new CTA derived radical (R ) then

proceeds through re-initiation, forming another oligomeric propagating radical species

(Pm ). Eventually a rapid equilibrium is reached between the actively propagating

radicals (Pn and Pm ) and the dormant polymeric thiocarbonylthio compounds (Figure

1.7 (d)). The main reversible degenerative chain transfer equilibrium provides an equal

probability for all propagating chains to grow resulting in narrow polydispersity

polymers. Most monomer consumption occurs during the main equilibrium and the

frequency of monomer additions can vary depending on reaction conditions; however, it

Page 29: Polyisobutylene Chain End Transformations: Block Copolymer

10

has been shown that in most RAFT polymerizations, less than one monomer units adds to

a propagating chain per transfer step.47

RAFT polymerizations are properly formulated so that the CTA concentration

relative to initiator concentration is high, therefore ensuring that most chains are initiated

by CTA derived radical (R ) instead of initiator radicals (I ). Initiator-derived chains are

thought to have a negative effect on the control of the molecular weight. With free

radical processes, termination inevitably occurs through radical coupling and

disproportionation, and the rate of termination is inherently related to the radical

concentration. When termination is a result of bimolecular combination, the number of

dead chains is equal to half the number of initiator derived chains; whereas when

disproportionation is dominant the number of dead chains is equal to the total number of

initiator derived chains.41 The RAFT process effectively suppresses termination events

by limiting the instantaneous concentration of radicals and typically preventing the

number of terminated chains from exceeding 5%.45

The key component of RAFT polymerization is the CTA, which consists of a

thiocarbonylthio moiety of the general structure Z-(C=S)-S-R.48,49,44 CTAs can be

categorized into four main classes depending on their Z group, including dithioesters,

xanthates, dithiocarbamates, and trithiocarbonates (Figure 1.9). To design a successful

RAFT polymerization the appropriate CTA must be selected in order to establish the

appropriate balance between reversible addition and fragmentation reactions (Figure 1.7);

otherwise loss of control, retardation, and long induction periods can occur. RAFT

agents are chosen based on their Z and R groups. Generally, the Z group controls the

level of activation/deactivation of the thiocarbonyl double bond. Activating Z groups

Page 30: Polyisobutylene Chain End Transformations: Block Copolymer

11

efficiently promote rapid radical addition thereby inhibiting prolonged propagation by

initiator and CTA derived radicals.48 If the Z group is not chosen correctly it can lead to

an overly stable intermediate radical, increasing the probably of polymerization

retardation50 and intermediate radical termination.51,52 The R group is directly related to

the pre equilibrium and must have the necessary stability to function as an efficient

leaving group with the ability to reinitiate polymerization of monomer.49

According to the RAFT mechanism, two potential sources of polymer chains are

possible, either initiator-derived or CTA-derived (i.e. R group). Therefore the theoretical

number-average molecular weight (Mn) is defined by equation 1,

[ ] [ ] ( ) MWtk

oo

MWothn CTA

eIfCTA

MMM

d+

−+=

−12][

,

ρ (1)

where [M]0 is the initial monomer concentration, MMW is the molecular weight of the

monomer, ρ is the monomer conversion, [CTA]0 is the initial CTA concentration, f is

the initiator efficiency, [I]0 is the starting initiator concentration, kd is the initiator

decomposition rate constant, and CTAMW is the molecular weight of the CTA.43,45 In

RAFT polymerizations with high CTA to initiator ratios, the effective amount of

initiator-derived chains should be less than 5 % allowing these terms to be neglected.45

Thus Equation 1 can be approximated as the following Equation 2.

[ ] MW

o

MWothn CTA

CTA

MMM +=

ρ][,

(2)

Azide-alkyne Click Chemistry

With the advent of living polymerizations and click chemistry, complex and

precise polymeric architectures are now possible.53 Sharpless originally defined click

chemistry as a reaction that, “must be modular, wide in scope, give very high yields,

Page 31: Polyisobutylene Chain End Transformations: Block Copolymer

12

generate only inoffensive byproducts that can be removed by nonchromatographic

methods, and be stereospecific (but not necessarily enantioselective). The required

process characteristics include simple reaction conditions (ideally, the process should be

insensitive to oxygen and water), readily available starting materials and reagents, the

use of no solvent or a solvent that is benign (such as water) or easily removed, and

simple product isolation. Purification – if required – must be by nonchromatographic

methods, such as recrystallization or distillation, and the product must be stable under

physiological conditions.”54

The most popular click chemistry reaction is the copper (I)-catalyzed azide-alkyne

1-3 dipolar cycloaddition.53 Until recently, this reaction was unable to meet the stringent

click chemistry criteria, failing to satisfy the requirements of simple reaction conditions

and stereospecificity. Initially, the azide-alkyne (Huisgen) cycloaddition was performed

at high temperature resulting in two regioisomers (i.e. 1,4- and 1,5-substituted-1,2,3-

triazoles) as shown in Figure 1.10. Later, research groups directed by Sharpless55 and

Meldal56 reported the synthesis of strictly 1,4-substituted-1,2,3-triazoles at room

temperature (i.e. accelerated reaction rates) through copper (I) catalysis. The proposed

mechanism offered by Sharpless (Figure 1.11)55 briefly states that first a copper(I)

acetylide complex is generated. Then through a “stepwise, ligation process” a six

member intermediate is formed, and lastly the triazole is formed releasing the copper(1)

catalyst. In 2005, Jia et al. also discovered that ruthenium (II) catalysts, instead of copper

(I), would reverse the stereospecificity producing only the 1,5-substituted-1,2,3 triazole.57

Because of the many advantages of click type reactions, a profound impact and

wide scale usage of these chemistries has been seen in many areas of research. Various

Page 32: Polyisobutylene Chain End Transformations: Block Copolymer

13

literature examples have been reported in biochemistry,58,59 surface functionalization,60,61

organic synthesis,62 drug discovery,63 and in polymer chemistry.53,64 -66 The first reported

usage of click chemistry in the field of polymer science was by Hawker, Sharpless, and

coworkers,67 and since that time, click chemistry has been demonstrated to have a broad

range of utility in the field. A few examples include step growth polymerizations of

dialkyne and diazide monomers/polymers,68,69 end and side chain functionalizations,70-74

block copolymers, 75,76 cyclic polymers,77 hyperbranched/dendritic structures, stars, and

cross-linked networks.

Of particular interest in this dissertation is end and pendent functionalization of

well defined polymers. Lutz et al. performed a unique combination of ATRP and click

chemistry to prepare functional polymers (Figure 1.12).70 Narrow molecular weight

telechelic polystyrene was synthesized, and the terminal bromide was substituted with

azide. Afterward, a 1,3-dipolar cycloaddition was conducted with copper bromide/4,4′-

di-(5-nonyl)-2,2′-bipyridine (dNbipy) to achieve quantitative conversion of bromide

chain ends into ω-hydroxyl, ω-carboxyl, and ω-methyl vinylidene end groups. In another

example involving ATRP, Opsteen and van Hest synthesized PMMA, PS, and PEG

polymers with terminal alkynes and azides.75 Subsequently, using click chemistry, the

authors successfully coupled these polymers to create tri and diblock copolymers.

Alkyne-terminal polymers were quantitatively synthesized using two approaches: use of a

unique blocked-alkyne ATRP initiator (Figure 1.13 (a)), which was deprotected after

polymerization, and use of carbodiimide coupling chemistry (Figure 1.13 (b)). Azide

terminal polymers were synthesized with a variety of substitution chemistries (Figure

Page 33: Polyisobutylene Chain End Transformations: Block Copolymer

14

1.14) to produce both mono and difunctional polymers. The resulting block copolymers

were easily purified and had monomodal, narrow PDIs.

Functionalization of polymer pendent groups has been demonstrated in

combination with numerous controlled polymerization techniques including ring opening

metathesis polymerization (ROMP),72 ATRP,73 and nitroxide mediated polymerization

(NMP).74 Binder and Kluger prepared functionalized poly(oxynorbornene)s by two

synthetic methods: (a) incorporation of functionality through 1,3-dipolar cycloaddition to

azide and alkyne functional 7-oxynorborene monomers prior to ROMP; (b) performing

ROMP and subsequent attachment of the functional groups to the prefabricated

polymer.72 Both synthetic strategies proved successful in attaching several functional

moieties both before and after ROMP polymerization while maintaining a well defined

poly(oxynorborne). Matyjaszewski and coworkers,73 attempted direct ATRP

polymerization of functionalized monomers containing both azido and acetylene pendent

groups. Although limited success was achieved with propargyl methacrylate (i.e. Mn/Mw

> 3), theorized to be a result of catalyst interaction, radical addition to the acetylene,

and/or chain transfer, well controlled polymers were synthesized with 3-azidopropyl

methacrylate. Later these polymers were utilized as highly functional scaffolds for click

reactions; producing alcohol, acid, halogen, and triphenylphosphine functionalized

macromolecules.

Thiol-ene Click Chemistry

The thiol and hydroxyl groups share some common features, but fundamentally

they, are different, owing to their disparate chemical and physical properties. To begin,

sulfur has less electronegative character and a larger atomic radius, attributing to its

Page 34: Polyisobutylene Chain End Transformations: Block Copolymer

15

longer and weaker covalent bonds.78 Geometrically, when comparing thiol and alcohol

functional groups, the bonds angles in alcohols are larger. For example, methanethiol has

a bond angle of 100.3°, whereas methanol has a larger bond angle of 109.5°. Other

characteristic differences include the inability of thiols to participate in hydrogen

bonding, unlike their alcohol counterpart, resulting in typically lower boiling points. A

few examples of this can be viewed in Table 1.1. One crucial attribute of the thiol group

is the labile nature of the hydrogen sulfur bond in comparison to a hydrogen oxygen

bond. Thiols are readily deprotonated (heterolytic bond cleavage), which is exemplified

by their more acidic nature in contrast to alcohols (CH3S-H pKa = 10.3; CH3O-H pka =

15.5).78,79 As a result, thiols easily converted by bases into highly nucleophilic thiolate

anions. Thiols also readily undergo homolytic cleavage of the S-H bond to produce thiyl

radicals when exposed to photo or thermal initiating processes. Various types of thiol

chemistries exist and have been studied extensively;78 although more recently, a series of

thiol-based reactions including thiol-ene,80-83,100 thiol-yne,84,93 and thio-isocyanate85,86

have been recognized for their “click” characteristics.87

Hydrothiolation, more commonly known as the “thiol-ene” reaction, is the

addition of a thiol across any unsaturated carbon-carbon double bond regardless of the

reaction mechanism (Figure 1.15).87 In the past, much work was conducted by Hoyle et

al.88 and Bowman et al.89,90,91 in the materials/polymer fields by utilizing the thiol-ene

reaction as a means to prepare near-perfect networks. Recently though, the thiol-ene

reaction has been recognized as a powerful synthetic tool for preparation of complex

functional architectures because of its many attractive features and “click”

characteristics.92,81 This process has proven to be versatile and can proceed through a

Page 35: Polyisobutylene Chain End Transformations: Block Copolymer

16

broad range of reaction conditions including a radical-mediated pathway (i.e. homolytic

S-H lysis),88 base/nucleophile-mediated addition with activated enes (i.e. heterolytic S-H

lysis),93,94 through solvent-mediated conditions,95 and by supramolecular catalysis.96 This

being the case, an extensive assortment of tools are available to the synthetic chemist,

including a broad range of mono and multivalent ene substrates with varying reactivities,

and a vast array of functional thiols of variable S-H bond strength. Numerous click

characteristics are exemplified by the thiol ene reaction, including its extremely rapid

reaction rate, tolerance to oxygen and water, and (near) quantitative conversions to a

regioselective thioether product.

Of the many thiol-ene reactions, the photochemically induced radical-mediated

addition process is most often used.87,88,97,98 Extensive work has been conducted with this

technique for network formation,88,89 polymer modification,99,83 and synthesis of complex

polymeric architectures.93,100 In general, the radical-mediated process proceeds through a

chain mechanism consisting of initiation, propagation, chain transfer, and termination

steps (Figure 1.16). Initiation begins by irradiation of thiol groups in the presence of

photoinitiator, which rapidly produces thiyl radicals and other photoinitiator fragments.

Radical generation can also be accomplished thermally.101 After thiyl radical generation,

propagation begins by radical addition across a double bond to produce a carbon-centered

radical species. Through homolysis of another sulfur hydrogen bond, chain transfer

occurs by hydrogen abstraction from the thiol to the carbon centered radical and

simultaneous regeneration of a new thiyl radical. Lastly, termination processes involve

typical radical-radical coupling reactions.

Page 36: Polyisobutylene Chain End Transformations: Block Copolymer

17

Reactivities of the thiol-ene radical mediated process can vary considerably based

upon the thiol and double bond structures.87 Double bond reactivity is typically a

function of electron density; where electron rich species have increased reactivity in

comparison to their electron deficient counterparts.87,88,97 In general, double bond

reactivity is observed in the following order: norbornene > vinyl ether > propenyl >

alkene ≈ vinyl ester > N-vinylamide > allyl ether ≈ allyl triazine ≈ alkyl isocyanurate >

acrylate > N-substituted maleimide > acrylonitrile ≈ methacrylate > styrene > conjugated

diene. With the exception of norbornene and the last three species, reactivity decreases

with diminished electron density. The exceptionally high reactivity of norbornene can be

explained by relief of ring strain after thiyl radical addition; whereas, the stability of the

carbon centered radicals, in the last three molecules, may impede facile hydrogen

abstraction, retarding their reactivity. While considerable analysis of double bond

reactivity has been conducted, limited literature is available on thiol reactivity. Some

general trends exist, showing that mercaptopropionates and thiolglycolates have

significantly higher reactivities relative to alkyl thiols.102 Some authors explain this as a

hydrogen bonding phenomenon, while other researchers claim this may be a result of

polar effects.87

In addition to the radical-mediated thiol-ene reaction, hydrothiolation can be

performed with activated double bonds with mildly basic or nucleophilic catalysts

(Figure 1.17). Activated double bonds are electron deficient; containing carbonyl, cyano,

or similar electron withdrawing groups adjacent to the α-carbon. This type of thiol-ene

reaction, commonly referred to as the thiol-Micheal addition reaction, can be performed

with a vast array of available substrates. A few examples of activated double bond

Page 37: Polyisobutylene Chain End Transformations: Block Copolymer

18

substrates are shown in Figure 1.18. In comparison to typical Michael additions, only

mildly basic catalysts are required for the thiol-Michael reaction.87 Because of the

slightly more acidic nature of thiols, in comparison to their corresponding alcohols,

thiolate anions are readily formed.78

During the base-catalyzed thiolation mechanism (Figure 1.19), first deprotonation

by a basic species (e.g.triethylamine) occurs to form the thiolate anion and

triethylammonium cation. The thiolate anion, a strong nucleophile, then attacks the

electron deficient double at the β-carbon producing an even stronger intermediate carbon

centered anion (i.e. enolate). Rapid proton transfer, from either the ammonium cation or

another thiol, to the enolate species results in the desired regioselective anti-Markovnikov

thioether product and concomitant regeneration of another thiolate anion.

In contrast to the base-catalyzed mechanism, recently an alternative nucleophile-

based mechanism was proposed for thiol-ene Michael reaction with phosphine catalysis

(Figure 1.20).93 With a similar reaction, but for the hydration and hydroxylation of

activated double bonds, Steward et al. declared that the catalyst served as a nucleophile

instead of as a base.103 Evidence to support this nucleophilic mechanism was provided

by another study involving a series of thiol-Michael reaction kinetics mediated with

amine catalysis.87 In one study, three catalysts were examined, all of which had similar

basicities (i.e. all pKa values between 10.56 and 11.0) but varying degrees of

nucleophilicity. It was found that reaction rates increased with more powerful

nucleophiles, e.g. hexylamine, di-n-propylamine, and triethylamine achieved

approximately 90, 60, and > 1% conversion in 500 s, respectively. However, when

catalyst nucleophilicity was kept fairly constant and instead basicity was varied, little

Page 38: Polyisobutylene Chain End Transformations: Block Copolymer

19

variation in reaction rates were observed for pyridine (pKa = 5.15), aniline (pKa = 9.34),

and 1,8-bis(dimethylamino)naphthalene (pKa = 12.1). These results imply that the

mechanism is not governed purely by catalyst basicity but instead by nucleophilicity. For

the phosphine-mediated thiol-Michael addition, the mechanism dictates that initial attack

on the activated carbon occurs at the β-carbon and a subsequent zwitteronic enolate

species is then formed. This intermediate species is a powerful base therefore promoting

rapid proton abstraction from another thiol to begin the extremely rapid thiol-ene chain

process. Thiol-Michael additions have been reported to proceed very rapidly to 100%

conversion without sensitivity to moisture.

Page 39: Polyisobutylene Chain End Transformations: Block Copolymer

20

References 1. Kennedy, J. P. In Polymer Chemistry of Synthetic Elastomers: Part I; Kennedy, J.P.;

Tornqvist, E. G. M., Eds.; Interscience: New York, 1968; Chapter 5.

2. Kennedy, J. P. TRIP 1993, 12, 381.

3. L. Pinchuk, G. J. Wilson, J. J. Barry, R. T Schoephoerster, Parcel, J-M.; Kennedy, J.

P., Biomaterials 2008, 29, 448.

4. Szwarc, M. Nature 1954, 178, 1168.

5. Szwarc, M.; Levy, M.; Milkovitch, R. J. Am. Chem. Soc. 1956, 78, 2656-2657.

6. Swarc, M. Carbanions, Living Polymers, and Electron Transfer Processes; Wiley,

New York, 1968.

7. Pepper, D. C. J. Poly. Sci., Polym. Chem. Ed. 1977, 15, 329.

8. Matyjaszewski, K. Cationic Polymerizations: Mechanisms, Synthesis, and

Applications; Marcel Dekker, Inc.: New York, 1996.

9. Storey, R. F.; Curry, C. L.; Brister, L. B. Macromolecules 1998, 31(4), 1058

10. Kennedy, J. P. J. Poly. Sci., Polym. Chem. Ed. 1999, 37, 2285.

11. Higashimura, T.; Nishii, H. J. Poly. Sci., Polym. Chem. Ed. 1977, 15, 329.

12. Higashimura, T.; Mitsuhashi, M.; Sawamoto, M. Macromolecules 1979, 12, 178.

13. Kennedy, J. P.; Smith, R. A. J. Polym. Sci., Polym. Chem. Ed. 1980, 18, 1523.

14. Kennedy, J. P.; Ross, L. R.; Lackey, J. E.; Nuyken, O. Polym. Bull. 1981, 4, 67.

15. Faust, R.; Fehervari, A.; Kennedy, J. P. J. Macromol. Sci-Chem. 1982, A18, 1209.

16. Miyamoto, M.; Sawamoto, M.; Higashimura, T. Macromolecules 1984, 17, 265.

17. Sawamoto, M.; Okatmoto, C.; Higashimura, T. Macromolecules 1987, 17, 2693.

Page 40: Polyisobutylene Chain End Transformations: Block Copolymer

21

18. Kojima, K.; Higashimura, T.; Sawamoto, M. Polym. Bull. 1988, 19, 7.

19. Sawamoto, M.; Okamoto, C.; Higashimura, T. Macromolecules 1987, 20, 2693.

20. Faust, R.; Kennedy, J. P. Polym. Bull. 1986, 15, 317.

21. Faust, R.; Kennedy, J. P. J. Polym. Sci. 1987, A25, 1847.

22. Aoshima, S.; Higashimura, T. Polym. Bull. 1986, 15, 417.

23. Kimigaito, M.; Maeda, Y.; Sawamoto, M.; Higashimura, T. Macromolecules 1993,

26, 1643.

24. Higashimura, T.; Ishihama, Y.; Sawamoto, M. Macromolecules 1993, 26, 744.

25. Pernecker, T.; Kennedy, J. P. Polym. Bull. 1991, 26, 305.

26. Pernecker, T.; Kelen, T.; Kennedy, J. P. J. Macromol. Sci., Pure Appl. Chem. 1993,

A30, 399.

27. Kaszas, G.; Puskas, J.; Kennedy, J. P.; Chen, C. C. J. Macromol. Sci. Chem. 1989,

A26, 1099.

28. Kaszas, G.; Puskas, J. E.; Chen, C. C,; Kennedy, J. P. Polym. Bull. 1988, 20, 413.

29. Gyor, M.; Wang, H. C.; Faust, R. J. Macromol. Sci., Pure Appl. Chem. 1992, A29,

639.

30. Balogh, L.; Faust, R. Poly. Bull. 1992, 281, 367.

31. Storey, R. F.; Curry, C. C.; Hendry, L. K. Macromolecules 2001, 34, 5416.

32. Storey, R. F.; Curry, C. C.; Hendry, L. K. Macromolecules 2001, 34, 5416.

33. Winstein, S.; Robinson, G. C. J. Am. Chem. Soc. 1958, 80, 169.

34. Moad, G.; Rizzardo, E.; Thang, S. H. Aust. J. Chem. 2005, 58, 379-410.

35. Otsu, T.; Yoshida, M. Makromol. Chem. Rapid Commun. 1982, 3, 127.

Page 41: Polyisobutylene Chain End Transformations: Block Copolymer

22

36. Otsu, T. J. Polym. Sci., Polym. Chem. 2000, 38, 44.

37. Hawker, C. J.; Bosman, A. W.; Harth, E. Chem. Rev. 2001, 101, 3661-3688.

38. Chung, T. C.; Janvikul, W.; Lu, H. L. J. Am. Chem. Soc. 1996, 118, 705-706.

39. Kamigaito, M.; Ando, T.; Sawamoto, M. Chem. Rev. 2001, 101, 3689-3745.

40. Matyjaszewski, K.; Xia, J. Chem. Rev. 2001, 101, 2921-2990.

41. Chiefari, J.; Chong, Y, K.; Ercole, F.; Krstina, J.; Jeffery, J.; Le, T. P. T.; Mayadunne,

R. T. A.; Meijs, G. F.; Moad, C. L.; Moad, G.; Rizzardo, E.; Thang, S. H.;

Macromolecules 1998, 31, 5559.

42. Le, T. P.; Moad, G.; Rizzardo, E.; Thang, S. H. Int. Pat. 9801478.

43. Moad, G.; Rizzardo, E.; Thang, S. H. Aust. J. Chem. 2006, 59, (10), 669-692.

44. Moad, G.; Solomon, D. H. The Chemistry of Radical Polymerization; Elsevier, Ltd.:

Oxford, UK, 2006.

45. Moad, G.; Rizzardo, E.; Thang, S. H. Aust. J. Chem. 2005, 58, (6), 379-410.

46. Matyjaszewski, K.; Davis, T. P., Handbook of Radical Polymerization. Wiley:

Hoboken, 2002.

47. Wang, A. R.; Zhu, S. P. J. Polym. Sci. Polym. Chem. 2003, 41, (11), 1553-1566.

48. Chiefari, J.; Mayadunne, R. T. A.; Moad, C. L.; Moad, G.; Rizzardo, E.; Postma, A.;

Skidmore, M. A.; Thang, S. H. Macromolecules 2003, 36, (7), 2273-2283.

49. Chong, Y. K.; Krstina, J.; Le, T. P. T.; Moad, G.; Postma, A.; Rizzardo, E.; Thang, S.

H. Macromolecules 2003, 36, (7), 2256-2272.

50. Mayadunne, R. T. A.; Rizzardo, E.; Chiefari, J.; Chong, Y. K.; Moad, G.; Thang, S.

H. Macromolecules 1999, 32, (21), 6977-6980.

Page 42: Polyisobutylene Chain End Transformations: Block Copolymer

23

51. Kwak, Y.; Goto, A.; Tsujii, Y.; Murata, Y.; Komatsu, K.; Fukuda, T. Macromolecules

2002, 35, (8), 3026-3029.

52. Barner-Kowollik, C.; Quinn, J. F.; Nguyen, T. L. U.; Heuts, J. P. A.; Davis, T. P.

Macromolecules 2001, 34, (22), 7849-7857.

53. Fournier, D.; Hoogenboom, R.; Schubert, U. S. Chem. Soc. Rev. 2007, 36, 1369-

1380.

54. Kolb, H. C.; Finn, M. G.; Sharpless, K. B. Angew. Chem. Int. Ed. 2001, 40, 2004.

55. Rostovtsev, V. V.; Green, L. G.; Fokin, V. V.; Sharpless, K. B. Angew. Chem. Int.

Ed. 2002, 41, 2595.

56. Tornoe, C. W.; Christensen, C.; Meldal, M. J. Org. Chem. 2002, 67, 3057.

57. Zhang, L.; Chen, X.; Xue, P.; Sun, H. H. Y.; Williams, I. D.; Sharpless, K. B.; Fokin,

V. V.; Jia, G. J. Am. Chem. Soc. 2005, 127, 15998.

58. Lee, L. V.; Mitchell, M. L.; Huang, S.-J.; Fokin, V. V.; Sharpless, K. B.; Wong, C.-H.

J. Am. Chem. Soc. 2003, 125, 9588.

59. Krasinski, A.; Radic, Z.; Manetsch, R.; Raushel, J.; Taylor, P.; Sharpless, K. B.;

Kolb, H. C. J. Am. Chem. Soc. 2005, 127, 6686.

60. Collman, J. P.; Devaraj, N. K., Chidsey, C. E. D. Langmuir 2004, 20, 1051.

61. Lee, J. K.; Chi, Y. S.; Choi, I. S. Langmuir 2004, 20, 3844.

62. Bock, V. D.; van Maarseveen, J. H.; Hiemstra, H. Eur. J. Org. Chem. 2006, 51.

63. Kolb, H. C.; Sharpless, K. B. Drug Discovery Today 2003, 8, 1128.

64. Hawker, C. J.; Wooley, K. L. Science 2005, 309, 1200.

65. Lutz, J.-F. Angew. Chem. Int. Ed. 2007, 46, 1018.

Page 43: Polyisobutylene Chain End Transformations: Block Copolymer

24

66. Sumerlin, B. S.; Tsarevsky, N. V.; Gao, H.; Golas, P.; Louche, G.; Lee, R. Y.;

Matyjaszewski, K. ACS Symp. Ser. 2006, 944, 140.

67. Wu, A.; Feldman, K.; Nugent, A. K.; Hawker, C. J.; Scheel, A.; Voit, B.; Pyun, J.;

Fre´chet, J. M. J.; Sharpless, K. B.; Fokin, V. V. Angew. Chem., Int. Ed. 2004, 43, 3928.

68. van Steenis, D. J. V. C.; David, O. R. P.; van Strijdonck, G. P. F.; van Maarseveen, J.

H.; Reek, J. N. H. Chem. Commun. 2005, 4333.

69. Tsarevsky, N. V.; Sumerlin, B. S.; Matyjaszewski, K. Macromolecules 2005, 38,

3558.

70. Lutz, J.-F.; Börner, H. G.; Weichenhan, K. Macromol. Rapid Commun. 2005, 26,

514.

71. Lutz, J.-F.; Börner, H. G.; Weichenhan, K., Macromolecules 2006, 39, 6376.

72. Binder, W. H.; Kluger, C. Macromolecules 2004, 37, 9321.

73. Sumerlin, B. S.; Tsarevsky, N. V.; Louche, G.; Lee, R. Y.; Matyjaszewski, K.

Macromolecules 2005, 38, 7540.

74. Malkoch, M.; Thibault , R. J.; Drockenmuller, E.; Messerschmidt, M.; Voit, B.;

Russell, T. P.; Hawker, C. J. J. Am. Chem. Soc. 2005, 127, 14942.

75. Opsteen, J. A.; van Hest, J. C. M. Chem. Commun. 2005, 57.

76. Durmaz, H.; Dag, A.; Altintas, O.; Erdogan, T.; Hizal, G.; Tunca, U. Macromolecules

2007, 40, 191.

77. Laurent, B. A.; Grayson, S. M. J. Am. Chem. Soc. 2006, 128, 4238.

78. Patai, S. The Chemistry of the Thiol Group, Parts 1-2, John Wiley and Sons, Inc.,

New York, 1974.

Page 44: Polyisobutylene Chain End Transformations: Block Copolymer

25

79. Serjeant, E. P.; Dempsey, B. Ionisation Constants of Organic Acids in Aqueous

Solution, Pergamon Press, New York, 1979.

80. Li, M.;De, P.; Gondi, S. R.; Sumerlin, B. S. J. Polym. Sci., Part A: Polym. Chem.

2008, 46, 5093-5100.

81. Becer, C. R.; Hoogenboom, R.; Schubert, U. S. Angew. Chem. Int. Ed. 2009, 48,

4900.

82. Dondoni, A. Angew. Chem. Int. Ed. 2008, 47, 8995-8997.

83. Hordyjewicz-Baran, Z.; You, L.; Smarsly, B.; Sigel, R.; Schlaad, H. Macromolecules

2007, 40, 3901

84 Hensarling, R.; Doughty, V.; Chan, J.; Patton, D. J. Am. Chem. Soc. 2009, 131, 41,

14673.

85 Li, H.; Yu, B.; Matsushima, H.; Hoyle, C. E.; Lowe, A. B. Macromolecules 2009, 42,

6537.

86 Shin, J.; Matsushima, H.; Chan, J. W.; Hoyle, C. E. Macromolecules 2009, 42, 3294.

87. Lowe, A. Polym. Chem., 2010, DOI: 10.1039/b9py00216b

88. C. E. Hoyle, T. Y. Lee, and T. Roper. J. Polym. Sci., Part A: Polym. Chem. 2004, 42,

5301-5338.

89. Carioscia, J. A.; Stansbury, J. W.; Bowman, C. N. Polymer, 2007, 48, 1526-1532.

90. Cramer, N. B.; Reddy, S. K.; O’Brian, A. K.; Bowman, C. N. Macromolecules 2003,

36, 7964-7969.

91. Khire, V. S.; Benoit, D. S. W.; Anseth, K. S.; Bowman, C. N. J. Polym. Sci., Part A:

Polym. Chem. 2006, 1006, 44, 7027-7039.

Page 45: Polyisobutylene Chain End Transformations: Block Copolymer

26

92. Lodge, T. P. Macromolecules 2009, 42, 3827.

93. Chan, J. W.; Hoyle, C. E.; Lowe, A. B. J. Am. Chem. Soc. 2009, 131, 5751-5753.

94. Chan, J. W.; Yu, B.; Hoyle, C. E.; Lowe, A. B. Chem. Commun. 2008, 4959-4961.

95. Kakwere, H.; Perrier, S.; J. Am. Chem. Soc. 2009, 131, 1889-1895.

96. Krishnaveni, N. S.; Surendra, K.; Rao, K. R. Chem. Commun. 2005, 669-671.

97. Morgan, C. R.; Magnotta, F.; Ketley, A. D. J. Polym. Sci., Polym. Chem. Ed. 1977,

15, 627-645.

98. Griesbaum, K. Angew. Chem., Int. Ed. Engle. 1970, 9, 273-287.

99. Campos, L. M.; Killops, K. L.; Sakai, R.; Paulusse, J. M. J.; Damiron, D.;

Drockenmuller, E.; Messomore, B. W.; Hawker, C. J. Macromolecules 2008, 41, 7063-

7070.

100. Killops, K. L.; Campos, L. M.; Hawker, C. J. J. Am. Chem. Soc. 2008, 130, 5062

101. Voronkov, M. G.; Deryagina, E, N. Russ. Chem. Rev. 1990, 59, 778-791.

102. Stock, L. M. J. Chem. Educ. 1972, 49, 400-404.

103. Steward, I. C.; Bergman, R. G.; Toste, F. D. J. Am. Chem. Soc. 2003, 125, 8686

Page 46: Polyisobutylene Chain End Transformations: Block Copolymer

27

Table 1.1. Boiling points of alcohols and thiols.

Thiols Alcohols Methanethiol 6 °C Methanol 65 °C Ethanethiol 35 °C Ethanol 78 °C

1-Butanethiol 98 °C 1-Butanethiol 117 °C

Page 47: Polyisobutylene Chain End Transformations: Block Copolymer

Figure 1.1. Polymeric architectures

Polymeric architectures via controlled living polymerization.

28

controlled living polymerization.

Page 48: Polyisobutylene Chain End Transformations: Block Copolymer

29

Figure 1.2. Chain breaking events in the cationic polymerization of IB.

Ti2Cl9n n

Ti2Cl9

Ti2Cl9n n

HCl 2 TiCl4

Ti2Cl9n

Ti2Cl9n

H

1. Bimolecular Chain Transfer

2. Spontaneous β-Proton Elimination (Unimolecular Chain Transfer)

2. Carbenium Ion Rearrangement (Unimolecular Chain Termination)

Page 49: Polyisobutylene Chain End Transformations: Block Copolymer

30

Figure 1.3. Chain equilibrium between (a) dormant, (b) ion pair, (c) free-ion pair chains.

P is the propagating chain end, X is the counterion, Keq is the ionization equilibrium

constant, and Kdis is the ion pair dissociation equilibrium constant.

PX

(a)

P X P X

(b) (c)

Page 50: Polyisobutylene Chain End Transformations: Block Copolymer

31

Figure 1.4. GPC (RI) traces of PIB prepared by the TMPCl/TiCl4 system in the absence

and the presence of ED.

Page 51: Polyisobutylene Chain End Transformations: Block Copolymer

32

Figure 1.5. Proposed mechanism for the formation of common ions through the

scavenging of protic impurities by 2,4-dimethylpyridine.

H2O Ti2Cl8N

Ti2Cl8

Ti2Cl8

N

H Ti2Cl8OH

N

H

Ti2Cl7OHTi2Cl9

Page 52: Polyisobutylene Chain End Transformations: Block Copolymer

33

Figure 1.6. CRP equilibrium between active and dormant chains with (a) SFRP, (b)

ATRP, and (c) RAFT.

Pm Xkact

kdeact

Pm X+

Pm X Y+kact

kdeact

Pm X+

kp

M

kp

M

Pm + Pn X

Y

kexch

k-exch

PnPm X +

kp

M

kp

M

(a)

(b)

(c)

Page 53: Polyisobutylene Chain End Transformations: Block Copolymer

34

Figure 1.7. The RAFT polymerization mechanism.

(e) I, R, Pn, Pmkt

Dead Polymer

PmS

CZ S Pn+kaddk-add

C

Z

SS PnPmPn

S

CZ S Pnk-addkadd

+

Monomer

Pn

S

CZ S R+kaddk-add

C

Z

SS RPm

S

CZ S Pn + Rkbk-b

Initiatorkd

I Pn(a)

(b)

(c)

(d)

initiation

reversible chain transfer/propagation

reinitiation

PmR

reversible (degenerate) chain transfer/propagation

M

M

M M

termination

+

Page 54: Polyisobutylene Chain End Transformations: Block Copolymer

35

Figure 1.8. Thermal initiators employed in RAFT polymerizations.

N N

NN

N N

NN

OON N

NN

OHHO

O O

AIBN V-70 V-501

Page 55: Polyisobutylene Chain End Transformations: Block Copolymer

36

Figure 1.9. Generic structures of RAFT chain transfer agents.

Z SR

S

O SR

S

R'N S

R

S

R'

R'S S

R

S

R'

Dithioester Xanthate DithiocarbamateTrithiocarbonate

Page 56: Polyisobutylene Chain End Transformations: Block Copolymer

37

Figure 1.10. Dipolar cycloadditions between alkynes and azides.

R1 N N N

+

R2

NN

N

R1

R2 NN

N

R1

R2

heat

Cu(I) catalyst

Ru(II) catalyst

+

NN

N

R1

R2

NN

N

R1

R2

Page 57: Polyisobutylene Chain End Transformations: Block Copolymer

38

Figure 1.11. Proposed catalytic cycle for the Cu(I)-catalyzed azide alkyne reaction.

Page 58: Polyisobutylene Chain End Transformations: Block Copolymer

39

Figure 1.12. Preparation of chain-end functionalized PS via a combination of ATRP and click chemistry.

Page 59: Polyisobutylene Chain End Transformations: Block Copolymer

40

Figure 1.13. Incorporation of azide end groups in polystyrene via an end group modification procedure.

(a)

(b)

Page 60: Polyisobutylene Chain End Transformations: Block Copolymer

41

Figure 1.14. Introduction of terminal alkyne functionality in polymers utilizing

functionalized ATRP initiator.

Page 61: Polyisobutylene Chain End Transformations: Block Copolymer

42

Figure 1.15. Anti-markovnikov hydrothiolation of C=C bond.

R'+RS-H

S H

R'

R

Page 62: Polyisobutylene Chain End Transformations: Block Copolymer

43

Figure 1.16. Radical mediated thiol-ene photoinitiated reaction.

Initiation

R-SH PIhv

R-S

R-S

Propagation

R'

S

R'

R

Chain Transfer

S

R'

R R-SHS

R'

R HR-S

Page 63: Polyisobutylene Chain End Transformations: Block Copolymer

44

Figure 1.17. Base/Nucleophile catalyzed hydrothiolation.

EWG+RS-H

S H

EWG

Rcatalyst

EWG = ester, amide, cyano

Page 64: Polyisobutylene Chain End Transformations: Block Copolymer

45

Figure 1.18. Activated double bonds applicable to the base/nucleophile-mediated

hydrothiolation.

O

OR

O

OR

O

NR2

O

NR2

O

ORNC

CN

CN

NR

O

O

O

OR

O

RO

O

OO

OO

O

O

OO

O

O

O

Page 65: Polyisobutylene Chain End Transformations: Block Copolymer

46

Figure 1.19. Base catalyzed hydrothiolation mechanism.

+RS-H NEt3

EWG

SEWG

R

H+SEWG

R

RS

thiol-ene product

Page 66: Polyisobutylene Chain End Transformations: Block Copolymer

47

Figure 1.20. Nucleophilic phosphine catalyzed hydrothiolation mechanism.

thiol-ene product

R''S

O

OR'

R'' S OR'

O

R''S HR'' S OR'

O

O

OR'

R3P

O

OR'

R''S H

R''S

R3P

O

OR'+R3P

Page 67: Polyisobutylene Chain End Transformations: Block Copolymer

48

CHAPER II

OBJECTIVES OF RESEARCH

As stated earlier, the primary objectives of this research were twofold: (1)

development of synthetic procedures for combining quasiliving carbocationic

polymerization (QLCCP) of isobutylene (IB) and reversible addition fragmentation

transfer chain (RAFT) polymerization for block copolymer synthesis; (2) utilization of

efficient, robust, and modular chemistries for facile functionalization of polyisobutylene

(PIB). Two site transformation strategies were employed to create block copolymers

effectively linking PIB with either poly(methylmethacrylate) (PMMA), polystyrene (PS),

and poly(N-isopropylacrylamide) (PNIPAM) block segments. Functionalization of PIB

was accomplished by utilizing two click chemistries, the azide-alkyne 1,3-dipolar cyclo

addition and thiol-ene hydrothiolation reactions, and by efficient transformation of the

thiol functional group. Specifically, the objectives of this work are briefly stated below:

1. Development of a synthetic strategy to combine quasiliving carbocationic IB

polymerization and RAFT polymerization.

2. Development of a synthetic strategy which combines quasiliving carbocationic IB

polymerization and RAFT polymerization to synthesize amphiphilic block

copolymers.

3. Investigate the utility of the azide-alkyne 1,3-dipolar cyclo addition to create

functional polyisobutylenes from a common azide functional precursor.

4. Investigate the utility of the radical mediated thiol-ene click reaction to create

functional polyisobutylenes from a common exo-olefin precursor.

Page 68: Polyisobutylene Chain End Transformations: Block Copolymer

49

5. Synthesis and transformation of α,ω-thiol-functional isobutylene into multi-

functional telechelics, macro-CTAs, and PIB based thiourethanes.

Page 69: Polyisobutylene Chain End Transformations: Block Copolymer

50

CHAPTER III

SITE TRANSFORMATION OF POLYISOBUTYLENE CHAIN ENDS INTO

FUNCTIONAL RAFT AGENTS FOR BLOCK COPOLYMER SYNTHESIS

Introduction

Polyisobutylene (PIB) is a fully saturated hydrocarbon elastomer with outstanding

oxidative and chemical resistance, superior gas-barrier and mechanical damping

characteristics, and excellent biocompatibility. Due to these characteristics, block

copolymers based on PIB elastomeric segments are currently of great interest as self

assembling materials with unique properties. For example, PIB-based triblock

copolymers have become an attractive candidate as a biomaterial and have found a niche

market in this field.1 Poly(styrene-b-isobutylene-b-styrene) (SIBS) is currently being

used as a drug-eluting coating for coronary stents because of its effectiveness as a drug

delivery matrix, vascular biocompatibility, and advantageous mechanical properties.2,3

Faust et al. have studied drug release characteristics of similar materials possessing outer

blocks derived from methyl methacrylate (MMA), 2-hydroxyethyl methacrylate

(HEMA), and hydroxyl and acetylated styrene (S) derivatives.4,5 Recently, much

progress has been made in the synthesis of self assembling PIB-based polymersomes and

micelles to create new delivery/encapsulation systems and interpolyelectrolyte complexes

for biotechnology and medicine.6,7 In other application areas, Binder and Machl have

synthesized poly(ether ketone-b-PIB–b-ether ketone) triblock copolymers with potential

uses as high-temperature thermoplastic elastomers (TPEs) with outer block Tg’s above

150°C.8 ABA rod-coil-rod triblock copolymers containing a PIB center block and

Page 70: Polyisobutylene Chain End Transformations: Block Copolymer

51

mesogen-jacketed liquid crystalline polymer outer blocks exhibit liquid crystalline

properties and have been suggested to have potential electrochemical applications.9

The examples above demonstrate that a number of methods have been devised for

the creation of novel block copolymers, in addition to traditional sequential monomer

addition. Specifically, the technique of site transformation can be used to greatly expand

the library of polymer segments that can be mated to PIB to form new and interesting

block copolymers. In this method, PIB block segments derived through cationic

polymerization are converted into functional macroinitiators for a chain polymerization

process other than cationic. For PIB-based systems this has typically involved

combinations of living cationic polymerization with atom transfer radical polymerization

(ATRP),10,11,12,13 condensation polymerization,14 or anionic polymerization.15,16

The relatively new controlled/living radical polymerization technique, reversible

addition-fragmentation chain transfer (RAFT) polymerization, has proven to be very

versatile regarding monomers, solvents, and reaction conditions to yield macromolecules

with predetermined molecular weights and narrow polydispersities.17 In addition, chain

transfer agents (CTAs) used in the RAFT process can be synthesized to carry many

useful reactive end-groups. These end groups can later be used for coupling reactions to

various macromolecules allowing for subsequent block copolymer synthesis.18 This type

of approach has been demonstrated with poly(ethylene glycol) (PEO),19 commercially

available Kraton polymers,20 and solid polymer supports derived from cotton,21 and was

successfully used to convert these macromolecules into functional macro–CTA’s for

block and graft copolymer synthesis. Because of the versatility of RAFT polymerizations

and its capability to polymerize many monomers which are inherently troublesome for

Page 71: Polyisobutylene Chain End Transformations: Block Copolymer

52

other polymerization techniques, RAFT is potentially an ideal polymerization technique

to combine with cationic polymerization by site transformation. This unique combination

would allow the synthesis of a variety of PIB based block copolymers with potentially

new or greatly improved properties compared to what is now currently available. Herein,

we report an initial example of combining living cationic polymerization of isobutylene

with the subsequent RAFT polymerization of methyl methacrylate (MMA) and styrene

(S) through site transformation of PIB chain ends into functional macro-CTAs. With

these model monomers as a template, further expansion of this technique can be used for

the synthesis of PIB based copolymers with monomers traditionally inaccessible to PIB

and cationic polymerization.

Experimental

Materials

Hexane (anhydrous, 95 %), TiCl4 (99.9 %, packaged under N2 in sure-seal

bottles), 2,6-lutidine (redistilled, 99.5%), chloroform-d (99.8 atom% D), N,N′-

dicyclohexyl carbodiimide (99%) (DCC), 4-(dimethylamino)pyridine (99%) (DMAP),

anhydrous dichloromethane (99.8%) (DCM), 1,2,2,6,6-pentamethylpiperidine (97%)

(PMP), and 1,3,5-trioxane (≥ 99%) were purchased from Sigma-Aldrich and used as

received. Methyl methacrylate (MMA) (99%), benzene (≥ 99.0%), and styrene (S)

(99.5%) were distilled from calcium hydride under a N2 atmosphere. 2,2′-Azobis(2-

methylpropionitrile) (98%) (AIBN) was recrystallized from ethanol. Isobutylene (IB)

(BOC, 99.5 %) and CH3Cl (Alexander Chemical Corp.) were dried through columns

packed with CaSO4 and CaSO4/4 Å molecular sieves, respectively.

Page 72: Polyisobutylene Chain End Transformations: Block Copolymer

53

Synthesis of Exo-olefin and Hydroxyl Terminated PIB

Two monofunctional exo-olefin-terminated PIB precursors were synthesized

using quasiliving polymerization of isobutylene followed by in-situ quenching with

1,2,2,6,6-pentamethylpiperidine according to a previously reported method.22 A

representative procedure is as follows: quasiliving polymerization of IB with TMPCl as

an initiator was carried out within a N2 atmosphere glovebox, equipped with an integral

cryostated hexane/heptanes bath. Into a round bottom flask with a mechanical stirrer,

infrared probe, and thermocouple were added 572 mL CH3Cl, 860 mL hexane, 2.50 mL

(2.19 g, 0.0147 mol) TMPCl, and 0.86 mL (0.79 g, 0.0074 mol) of 2,6-lutidine. The

mixture was allowed to equilibrate to -60 °C, and then 29.0 mL (19.9 g, 0.354 mol) of IB

was added to the reactor and allowed to reach thermal equilibrium. To begin the

polymerization, 4.8 mL (8.3 g, 0.044 mol) of TiCl4 was charged to the reactor. Full

monomer conversion (≥ 98 %) was achieved in 90 min, after which time 8.0 mL (6.9 g,

0.044 mol) PMP and an additional 4.8 mL TiCl4 (8.3 g, 0.044 mol) were added to the

polymerization. PMP was allowed to react with the living chain ends for 90 min.

Finally, the reaction was terminated by addition of excess prechilled methanol. The

contents of the reaction flask were allowed to warm to room temperature, and the

polymer in hexane was immediately washed with methanol and then precipitated into

methanol from hexane. The precipitate was collected by dissolution in hexane; the

solvent was washed with water, dried over MgSO4, and concentrated on a rotary

evaporator. Residual solvent was removed under vacuum at 40 °C. A representative 1H

and 13C NMR spectra of exo-olefin PIB is shown in Figure 3.1 and 3.2 respectively,

whereas GPC traces before and after quenching is shown in Figure 3.3. Their

Page 73: Polyisobutylene Chain End Transformations: Block Copolymer

54

characterization result is summarized in Table 3.1. The exo-olefin PIB precursors were

converted into hydroxyl PIBs through hydroboration-oxidation as reported by Ivan et

al.23 Representative 1H NMR, 13C NMR, and GPC are in Figures 3.4, 3.5, and 3.6

respectively.

Synthesis of 4-Cyano-4-(dodecylsulfanylthiocarbonylsulfanyl)-pentanoic Acid (CTA)

The chain transfer agent (CTA) was synthesized according to previously reported

literature methods.18 n-Dodecylthiol (15.4 g, 76 mmol) was added over 10 min to a

stirred suspension of sodium hydride (60% in oil) (3.15 g, 79 mmol) in diethyl ether (150

mL) at a temperature between 5 and 10 °C. A vigorous evolution of hydrogen was

observed and the greyish sodium hydride was transformed to a thick white slurry of

sodium thiododecylate. The reaction mixture was cooled to 0 °C and carbondisulfide

(6.0 g, 79 mmol) added to provide a thick yellow precipitate of sodium S-dodecyl

trithiocarbonate which was collected by filtration and used in the next step without

purification.

A suspension of sodium S-dodecyl trithiocarbonate (14.6 g, 0.049 mol) in diethyl

ether (100 mL) was treated by portion-wise addition of solid iodine (6.3 g, 0.025 mol).

The reaction mixture was then stirred at room temperature for 1 h when the white sodium

iodide which settled was removed by filtration. The yellow–brown filtrate was washed

with an aqueous solution of sodium thiosulfate to remove excess iodine and water and

dried over sodium sulfate and evaporated to leave a residue of bis-

(dodecylsulfanylthiocarbonyl)disulfide. A solution of 4,4′-azobis(4-cyanopentanoic acid)

(2.10 g, 0.0075 mol) and the above bis-(dodecylsulfanylthiocarbonyl) disulfide (2.77 g,

0.005 mol) in ethylacetate (50 mL) was heated at reflux for 18 h. After removal of the

Page 74: Polyisobutylene Chain End Transformations: Block Copolymer

55

volatiles in vacuum, the crude product was extracted with water (5 X 100 mL) to afford

4-cyano-4-(dodecylsulfanylthiocarbonyl) sulfanyl pentanoic acid as a pale yellow solid

after recrystallization from hexane.

Synthesis of 4-Cyano-4-(dodecylsulfanylthiocarbonylsulfanyl)pentanoic Acid-

Functionalized PIB (PIB-CTA)

To a 25 mL one-neck round-bottom flask equipped with a magnetic stir bar were

added DCC (0.32 g, 1.54 mmol), DMAP (38 mg, 0.31 mmol), and CTA (0.39 g, 0.96

mmol) under a dry nitrogen atmosphere. In a separate vessel, hydroxyl-functional PIB-1

(1.25 g, 0.77 mmol) was dissolved in 13.5 mL DCM and the resulting solution was

charged to the reaction flask. After 12 h, the reaction was filtered, and the solvent was

removed under reduced pressure. The resulting product was dissolved in hexane, washed

with methanol, and then precipitated into methanol from hexane. The precipitate was

dissolved in hexane and washed first with a saturated NaCl solution and then with

deionized water. The solution was then dried over magnesium sulfate and filtered, and

the hexane was stripped under reduced pressure until a constant weight was reached.

Polymerization of MMA and S from PIB-CTA

A representative RAFT polymerization was conducted as follows. To a 25 mL

Schlenk-style, long-neck round-bottom flask were charged PIB-1-CTA (0.069 g, 0.031

mmol), 1,3,5-trioxane (0.048g, 0.533 mmol), MMA (0.394 g, 3.94 mmol), and AIBN

(0.0025 g, 0.015 mmol) in 0.18 mL of benzene. After dissolution of the reagents an

initial aliquot was taken to establish the initial monomer concentration relative to the

internal standard 1,3,5-trioxane, via 1H NMR spectroscopy. The solution was then

subjected to three freeze-pump-thaw cycles to remove oxygen, sealed under N2, and

Page 75: Polyisobutylene Chain End Transformations: Block Copolymer

56

submerged in an oil bath at 60°C. After approximately 16 h the reaction was exposed to

oxygen and quenched in liquid nitrogen. A final aliquot was taken for 1H NMR analysis,

and then the crude reaction product was precipitated into hexane and placed under

vacuum until a constant mass was reached. Conversion was calculated from the initial

and final monomer concentrations relative to 1,3,5-trioxane.

Instrumentation

NMR spectra were acquired using a Varian Mercuryplus 300 MHz NMR

spectrometer. Samples were dissolved in chloroform-d (3-7 %, w/v) and analyzed using

5 mm NMR tubes. 13C and 1H resonances were correlated with gradient enhanced

heteronuclear single-quantum coherence (gHSQC) spectroscopy, using the average of 16

transients for each of 2 x 512 increments and phase sensitive detection in the F1

dimension.

Variable-temperature NMR (VT NMR) data were acquired using a Varian

Mercuryplus 300 MHz NMR spectrometer fitted with a Bruker Eurotherm VT controller.

RAFT polymerizations were performed in benzene-d at 333 K, with a 250 s pre-

acquisition delay between each spectrum. The probe was allowed to equilibrate for 10

min prior to data acquisition. Temperatures reported in this study are within plus/minus 2

degrees, based on ethylene glycol calibration.24,25 RAFT polymerizations for VT NMR

were prepared by first charging a Schlenk-style round bottom flask with the crude

reaction mixture and performing three freeze-thaw-pump cycles. After degassing, and

just prior to analysis, the contents were transferred into an air-tight NMR tube within a

dry nitrogen atmosphere glovebox. Conversion was calculated by comparison of the

vinyl proton areas of the monomer to the internal standard, 1,3,5-trioxane.

Page 76: Polyisobutylene Chain End Transformations: Block Copolymer

57

Number average molecular weight (Mn) and polydispersity index (PDI) of the

polymeric materials were measured using a Gel Permeation Chromatography (GPC)

system consisting of a Waters Alliance 2695 separation module, an on-line multiangle

laser light scattering (MALLS) detector fitted with a gallium arsenide laser (power: 20

mW) operating at 690 nm (MiniDAWN, Wyatt Technology Inc.), an interferometric

refractometer (Optilab DSP, Wyatt Technology Inc.), and two Polymer Laboratories

mixed E columns (pore size range 50-103 Å, 3 µm bead size) connected in series. Freshly

distilled THF served as the mobile phase and was delivered at a flow rate of 1.0 ml/min.

Sample concentrations were ca. 6-7 mg of polymer/mL of THF, and the injection volume

was 100 µL. The detector signals were simultaneously recorded using ASTRA software

(Wyatt Technology Inc.), and absolute molecular weights were determined by MALLS.

The dn/dc values for PIB homopolymer, PIB-CTA, PIB-b-poly(methyl methacrylate)

(PIB-b-PMMA), and PIB-b-polystyrene (PIB-b-PS) were calculated from the response of

the Optilab DSP and assuming 100% mass recovery from the columns.

Diblock Copolymer Molecular Weight and Blocking Efficiency by 1H NMR

The number average molecular weights of the diblock copolymers were

calculated using 1H NMR and equation 1 for PIB-b-PMMA and equation 2 for PIB-b-PS.

Amethoxy and Aaromatic represent the area of the methoxy protons of PMMA and the

aromatic protons of PS; MMMA and MS are the molecular weights of the corresponding

monomer units, and Mn,PIB-CTA equals 2,200 g/mol for PIB-1 and 3,100 g/mol for PIB-2.

The methylene protons on carbon two of the CTA were used to normalize Amethoxy and

Aaromatic.

methoxyn,NMR MMA n,PIB-CTA

AM = ×M +M

3 (1)

Page 77: Polyisobutylene Chain End Transformations: Block Copolymer

58

aromaticn,NMR S n,PIB-CTA

AM = ×M +M

5 (2)

Blocking efficiency (Beff %) was calculated using GPC (equation 3) and 1H NMR

(equation 4).

n,theo n,PIB-CTAeff

n,GPC n,PIB-CTA

M -MB %=

M -M (3)

n,theo n,PIB-CTAeff

n,NMR n,PIB-CTA

M -MB %=

M -M (4)

Mn,GPC is the number average molecular weight of the diblock copolymer determined

using GPC-MALLS. Mn,theo is the theoretical number average molecular weight of the

diblock copolymer calculated according to equation 5 for PIB-b-PMMA (the calculation

is analogous for PIB-b-PS), where p is monomer conversion for the RAFT

polymerization and [MMA]o and [PIB-CTA]o refer to the initial MMA and PIB-CTA

concentrations, respectively.

o MMAn,PIB-CTA

o

p[MMA] M+M

[PIB-CTA] (5)

Results and Discussion

The method employed for site transformation from living carbocationic

polymerization to RAFT polymerization is shown in Figure 3.7. The initial PIB block

was synthesized by quasiliving cationic polymerization using the 2-chloro-2,4,4-

trimethylpentane/TiCl4 initiation system. After reaching full conversion of the IB

monomer, in-situ quenching with the hindered nucleophile, PMP, was used to convert the

quasiliving cationic chain ends into exo-olefin functionality.22 After quenching, 1H NMR

integration was used to characterize the end group composition of the product by

Page 78: Polyisobutylene Chain End Transformations: Block Copolymer

59

assuming that tert-chloride, endo-olefin, exo-olefin, and coupled PIB chain ends

represent 100 % of the chain ends. The results indicated that the PIB precursors were

functionalized to near quantitative amounts (Table 3.1) showing only trace quantities of

tert-chloride, endo-olefin, and coupled products. Terminal exo-olefin PIB functionality

was selected because of the convenience of in-situ quenching and the facile

transformation of this end group into a variety of other functional groups. Low molecular

weight PIB was targeted in this work to facilitate accurate NMR analysis. For discussion

purposes all NMR and GPC characterization results will correspond to PIB-1, Table 3.1,

unless otherwise stated. 1H and 13C NMR spectra of exo-olefin PIB are shown in Figures

3.1 and 3.2, respectively. GPC traces of both the pre-quench tert-chloride PIB and

quenched exo-olefin PIB are also shown in Figure 3.3.

Exo-olefin PIB was converted into hydroxyl PIB by hydroboration oxidation.23

Complete characterization of hydroxyl PIB using 1H/13C NMR along with the GPC traces

before and after functionalization demonstrated its successful synthesis (Figures 3.4-6).

From 1H NMR the disappearance of the methylene (2.00 ppm), methyl (1.78 ppm), and

olefinic protons (4.85 and 4.64 ppm) and formation of the new methylene protons (3.53-

3.41 ppm and 3.36-3.26 ppm) adjacent to the hydroxyl group indicated near quantitative

conversion from the exo-olefin PIB. 13C NMR revealed that the exo-olefin carbons at

144.53 and 114.03 ppm were absent in the hydroxyl PIB, and new shifts for these carbons

appeared at 31.99 and 69.86 ppm, respectively. Carbons immediately adjacent to the

exo-olefin functionality also shifted from 53.87 to 49.58 ppm and from 25.91 to 20.04

ppm. The GPC traces in Figure 3.6 did not show significant changes in Mn and indicated

the absence of any coupling.

Page 79: Polyisobutylene Chain End Transformations: Block Copolymer

60

The CTA, 4-cyano-4-(dodecylsulfanylthiocarbonylsulfanyl)pentanoic acid, was

synthesized according to previously reported literature methods and characterized using

1H and 13C NMR, as shown in Figure 3.8 and 3.9. This particular CTA was chosen due

to its solubility in low-polarity media, facile synthesis, reduced odor,17 and established

ability to polymerize MMA.18 1H NMR peak assignments were obtained from the

literature.18

13C NMR peak assignments were made using two dimensional gHSQC (Figures

3.10) and further bolstered by an attached proton test, which can be found in Figure 3.11.

In Figure 3.9, the carbon assignments were made by correlation with their respective

proton peaks; although some ambiguity exists in identification of carbon 12 (associated

with protons b′ in Figure 3.8) due to overlap of its signal with the numerous carbon peaks

of the dodecyl group. Carbons 11, 14, 16, and 17 had no correlations during the HSQC

experiment, confirming that they were indeed quaternary carbons.

Synthesis of a PIB-based macro-CTA was accomplished by esterification of the

CTA with hydroxyl PIB. The esterification reaction utilized DMAP as a catalyst and

DCC as a water scavenger. This reaction successfully yielded CTA-functionalized PIB

to near quantitative conversion. Structural evidence of the resulting PIB-CTA can be

seen in the 1H and 13C NMR spectra shown in Figures 3.12 and 3.13, respectively. In

Figure 3.12, the methylene protons (a′) adjacent to the hydroxyl group on the PIB shifted

from their original location at approximately 3.53-3.26 to 4.00-3.74 ppm. In addition, no

residual methylene protons of the hydroxyl-terminated PIB were visible, indicating

complete conversion was achieved. Also, the methylene protons (b′) adjacent to the acid

Page 80: Polyisobutylene Chain End Transformations: Block Copolymer

61

functionality on the CTA shifted slightly upfield from their original location after

coupling with the hydroxyl PIB.

In the 13C NMR spectrum of the PIB-CTA (Figure 3.13) peaks from both the

CTA and hydroxyl PIB are visible, indicating that both structures are present and coupled

together after purification. If the CTA were not covalently attached to the PIB, it would

have been removed during washing and precipitation of the polymer. Due to the number

of peaks present in Figure 3.13, direct comparisons of expanded, partial spectra of the

CTA, hydroxyl PIB, and PIB-CTA were performed for accurate identification of each

peak. These expanded spectral comparisons can be found in Figures 3.14-16. Strong

shifts were observed for the methylene carbon (C-10) adjacent to oxygen in the PIB

precursor and the carbonyl carbon of the CTA (C-11), as displayed in Figure 3.17. After

esterification, the methylene carbon shifted from 69.86 to 71.33 ppm; whereas the

carbonyl carbon shifted from 177.39 to 171.69 ppm. No residual carbonyl carbon or

methylene carbon peaks from the precursors were visible in the resulting product.

Carbon atoms two and three bonds away also shifted slightly including C-9 (20.04 to

20.37 ppm) and C-8 (31.99 to 29.97 ppm) of hydroxyl PIB and C-12 (29.64 to 28.70

ppm) and C-13 (33.38 to 34.11 ppm) of the CTA.

Upon addition of the CTA to hydroxyl PIB a molecular weight increase should

have been observed. GPC results confirmed that the PIB-CTA molecular weights were

2,200 and 3,100 g/mol for PIB-1 and PIB-2, respectively. These values are close to the

predicted molecular weights after the coupling reaction.

Once the CTA-functionalized PIB was obtained, RAFT polymerizations were

conducted for the synthesis of both PIB-b-PMMA and PIB-b-PS. Initial RAFT

Page 81: Polyisobutylene Chain End Transformations: Block Copolymer

62

polymerizations were conducted with MMA using a variety of reaction conditions to

optimize polymerization rate and blocking efficiency of the macro CTA. This

preliminary experimentation revealed that monomer concentrations below 3 M and

[CTA]:[AIBN] ratio above five resulted in unreasonably long reaction times, and that

lower temperatures (e.g. 60 – 70 °C) afforded increased blocking efficiency at the

expense of longer reaction times. Figures 3.17 and 3.18 show the 1H NMR spectrum and

GPC traces, respectively, of a representative PIB-b-PMMA copolymer polymerized at

optimal conditions, after purification. Due to the relatively long PMMA block, fractional

precipitation of the crude reaction product into hexanes was effective at removing

unreacted PIB-CTA. The 1H NMR spectrum shows characteristic peaks from both PIB

and PMMA. The CTA methylene proton peak (b) was used to normalize the methoxy

peak of the PMMA for NMR molecular weight calculations. In addition, the GPC traces

illustrate a substantial decrease in elution volume from the macro PIB-CTA starting

material to the PIB-b-PMMA indicating an increase in molecular weight.

Next, a series of RAFT polymerizations was performed in which monomer

concentration ([M]), [M]:[CTA]:[AIBN], and type of monomer were varied in order to

probe their influence on blocking efficiency. Table 3.2 summarizes the conditions of

each polymerization and its respective conversion, PDI, molecular weight, and blocking

efficiency. Blocking efficiency was determined using both GPC and 1H NMR as

described in the Experimental section.

Experiments R – 1,2,3,4 were all conducted with near identical [M]:[CTA]:[I]

concentration ratios but differ in overall system concentration, which was dictated by the

selected monomer concentration [M]. No clear trends can be observed concerning the

Page 82: Polyisobutylene Chain End Transformations: Block Copolymer

63

blocking efficiency of the macro CTA regarding the system concentration. It is possible

that an optimal system concentration may exist at or around [M] of 4.5 M (R-3) indicated

by blocking efficiency values greater than 95 %. Additionally, less concentrated

systems (R – 2, 3, 4) resulted in slightly improved blocking efficiencies compared to that

of the more concentrated system (R-1). The various polymerizations detailed in Table

3.2 demonstrate the capability of this system to yield targeted molecular weights over a

fairly broad range (12,000 to 18,000 g/mol) and low PDIs, even when the

polymerizations are taken to high monomer conversion.

Experiments R-3 and R-4 were divided into two portions prior to initiation. One

portion was transferred into a round bottom flask for a traditional RAFT polymerization

and the other into an NMR tube for VT NMR. These VT NMR experiments were

utilized to obtain real time conversion and polymerization rate verses time data, as a

function of system concentration, as shown in Figure 3.20. Except for an initial induction

period, the RAFT polymerizations exhibited linear first-order plots, indicating an

approximately constant number of growing species. Such an induction period is

commonly observed in RAFT polymerizations,26,27 and is generally attributed to the

establishment of the main RAFT equilibrium. As expected, the more concentrated

system, curve A, achieved larger conversion values in the same duration of time, without

causing any increase in PDI or decrease in the degree of molecular weight control (see

data in Table 3.2). This shows that shorter reactions times can be achieved with higher

system concentrations without deleterious effects to the resulting block copolymer.

In addition to monitoring the conversion of the system with 1H NMR, aliquots

were taken during the polymerization for GPC analysis. Figure 3.21 shows the GPC

Page 83: Polyisobutylene Chain End Transformations: Block Copolymer

64

traces of reaction aliquots removed from reaction R-3 at various conversions, along with

the final purified block copolymer. A gradual increase in molecular weight with

conversion can be seen. At full conversion a small fraction of residual PIB-CTA

remained in the crude reaction product. This residual fraction was isolated by fractional

precipitation and found to contain some PMMA in its structure. This suggests that the

failure of this fraction to form the desired second block may result from a combination of

poor initiation efficiency, i.e., all MMA is consumed before all PIB-CTA’s react, and

early termination, resulting in a very short PMMA block.

To test whether the less-than-quantitative initiation efficiency might be due

specifically to the selection of MMA, styrene was selected as a second monomer to

conduct RAFT polymerization from the PIB macroinitiator. Styrene polymerization was

conducted in the bulk thereby simplifying the system greatly by elimination of solvent

and initiator, and styrene polymerization has been shown to proceed in a controlled

fashion in the RAFT process.28 In Figure 3.22 the same PIB-CTA was used to

polymerize MMA (R-1) and S (R-8). It is apparent that styrene yielded substantially

improved blocking efficiency as evidenced by the significantly reduced amount of

residual PIB-CTA. The molecular weight data in Table 3.2 show that styrene also

yielded diblock copolymers with low PDI. However, the polymerization of styrene

yielded experimental Mn values (both GPC and NMR) slightly lower than that of the

theoretical value giving apparent blocking efficiency values exceeding 100 %.

Conclusion

Site transformation of PIB chain ends into functional macro-CTAs was

successfully accomplished. The macro CTA synthesized in this work represents the first

Page 84: Polyisobutylene Chain End Transformations: Block Copolymer

65

report of combining cationic IB polymerization and subsequent RAFT polymerization for

block copolymer synthesis. Structural of the macro CTA was confirmed by both 1H and

13C NMR after coupling the CTA to PIB. RAFT polymerizations using the PIB-CTA

were demonstrated with both MMA and S to yield block copolymers with predetermined

molecular weights and narrow PDIs, as characterized by GPC and 1H NMR. VT NMR

experiments verified that MMA polymerizations progressed in a controlled fashion and

that the rate was concentration dependant. Blocking efficiency values were found to be

slightly improved with S compared MMA.

Page 85: Polyisobutylene Chain End Transformations: Block Copolymer

66

References

1. Kennedy, J.P. J. Polym. Sci.: Part A: Polym. Chem. 2005, 43, 2951-2963.

2. Ranade, S.V.; Miller, K.M.; Richard, R.E.; Chan, A.K.; Allen, M.J.; Helmus, M.N. J.

Biomed. Mater. Res. 2004, 71A, 625-634.

3. Pinchuk, L.; Wilson, G.J.; Barry, J.J.; Schoephoerster, R.T.; Parel, J.-M.; Kennedy, J.P.

Biomaterials 2008, 29(4), 448-460.

4. Cho, J.C.; Cheng, G.; Feng, D.; Faust, R.; Richard, R.; Schwarz, M.; Chan, K.; Boden,

M. Biomacromolecules 2006, 7(11), 2997-3007.

5. Sipos, L.; Som, A.; Faust, R.; Richard, R.; Schwarz, M.; Ranade, S.; Boden, M.; Chan,

K. Biomacromolecules 2005, 6(5), 2570-2582.

6. Binder, W.H.; Sachsenhofer, R. Macromol. Rapid Commun. 2008, 29(12-13), 1097-

1103.

7. Burkhardt, M; Ruppel, M.; Tea, S.; Drechsler, M.; Schweins, R.; Pergushov, D.V.;

Gradzielski, M.; Zezin, A.B.; Müller, A.H.E. Langmuir 2008, 24(5), 1769-1777.

8. Binder, W.H.; Machl, D. J. Polym. Sci.: Part A: Polym. Chem. 2005, 43(1), 188-202.

9. Gao, L.-C.; Zhang, C.-L.; Liu, X.; Fan, X.-H.; Wu, Y.-X.; Chen, X.-F.; Shen, Z.; Zhou,

Q.-F. Soft Matter 2008, 4(6), 1230-1236.

10. Storey, R. F.; Scheuer, A. D.; Achord, B. C. Polymer 2005, 46 (7) 2141-2152.

11. Breland, L.K.; Storey, R.F. Polymer 2008, 49(5), 1154-1163.

12. Kennedy, J.P.; Fang, Z. J. Polym Sci.: Part A: Polym. Chem. 2002, 40(21), 3662-

3678.

Page 86: Polyisobutylene Chain End Transformations: Block Copolymer

67

13. Jakubowski, W.; Tsarevsky, N.V.; Higashihara, T.; Faust, R.; Matyjaszewski, K.

Macromolecules 2008, 41(7), 2318-2323.

14. Zaschke, B.; Kennedy, J. P. Macromolecules 1995, 28(13), 4426-4432.

15. Martinez-Castro, N.; Lanzendörfer, M.G.; Müller, A.H.E.; Cho, J.C.; Acar, M.H.;

Faust, R. Macromolecules 2003, 36(19), 6985-6994.

16. Kwon, Y.; Faust, R.; Chen, C.X.; Thomas, E.L. Macromolecules 2002, 35(9), 3348-

3357.

17. Moad, G.; Rizzardo, E; Thang, S.H. Polymer 2008, 49(5), 1079-1131.

18. Moad, G.; Chong, Y.K.; Postma, A.; Rizzardo, E.; Thang, S.H. Polymer 2005,

46(19), 8458-8468.

19. Shi, L.; Chapman, T.M.; Beckman, E.J. Macromolecules 2003, 36(7), 2563-2567.

20. De Brouwer, H.; Schellekens, M.A.J.; Klumperman, B.; Monteiro, M.J.; German,

A.L. J. Polym. Sci.: Part A: Polym. Chem. 2000, 38(19), 3596-3603

21. Perrier, S.; Takolpuckdee, P.; Westwood, J.; Lewis, D.M. Macromolecules 2004,

37(8), 2709-2717

22. Simison, K.L.; Stokes, C.D.; Harrison, J.J.; Storey, R.F. Macromolecules 2006, 39(7),

2481-2487.

23. Iván, B.; Kennedy, J.; Chang, V.S.C. J. Polym. Sci., Polym. Chem. 1980, 18(11),

3177-3191.

24. English, A. D. J. Magnetic Res. 1984, 57(3), 491-493.

25. Van Geet, A. L. Anal. Chem. 1968, 40(14), 2227-2229.

26. De, P.; Gondi, S.R.; Sumerlin, B. S. Biomacromolecules 2008, 9(3), 1064-1070.

Page 87: Polyisobutylene Chain End Transformations: Block Copolymer

68

27. Barner-Kowollik, C.; Buback, M.; Charleux, B.; Coote, M.L.; Drache, M.; Fukuda,

T.; Goto, A.; Klumperman, B.; Lowe, A.B.; Mcleary, J.B.; Moad, G.; Monteiro, M. J.;

Sanderson, R. D.; Tonge, M. P.; Vana, P. J. Polym. Sci.: Part A: Polym. Chem. 2006,

44(2), 5809-5831.

28. Mayadunne, R.T.A.; Rizzardo, E.; Chiefari, J.; Krstina, J.; Moad, G.; Postma, A.;

Thang, S.H. Macromolecules 2000, 33(2), 243-245.

Page 88: Polyisobutylene Chain End Transformations: Block Copolymer

69

Table 3.1. Characterization of exo-Olefin PIB Precursors

Sample Mna

(g/mol) PDI

(Mn/Mw) exo-olefinb

(%)

PIB – 1 1,600 1.04 98 PIB – 2 2,500 1.03 97

a determined by GPC b determined by 1H NMR

Page 89: Polyisobutylene Chain End Transformations: Block Copolymer

Table 3.2. RAFT Polymerizations of PMMA and PS

Reaction Conditions Mn (g/mol)

Blocking Efficiency

(%) Exp. PIB-

CTA

Monomer [M]

(mol/L)

[M]:[CTA]:[I] Time2

(h)

Conv. PDI GPC NMR Theo GPC NMR

R-1 PIB-2 MMA 6.50 130:1:0.52 10.8 94.6 1.04 17,000 17,840 15,410 88.6 83.5 R-2 PIB-2 MMA 5.52 142:1:0.53 11.8 93.0 1.04 17,100 16,810 16,300 94.3 96.3 R-3 PIB-2 MMA 4.48 146:1:0.53 19 100 1.04 18,200 17,820 17,480 95.2 97.7 R-4 PIB-2 MMA 3.51 143:1:0.51 14.8 80.9 1.06 15,900 14,200 14,670 90.4 104.2 R-5 PIB-1 MMA 6.30 125:1:0.54 15.3 70.4 1.06 12,000 12,500 11,020 90.0 85.6 R-6 PIB-1 MMA 6.48 85:1:0.51 13.3 100 1.03 14,000 11,370 10,720 72.2 92.9 R-7 PIB-1 MMA 6.51 139:1:0.26 14 100 1.04 18,300 16,290 16,130 86.5 98.9 R-81 PIB-2 S Bulk 301:1:0.0 20 82.2 1.02 26,700 26,160 28,910 109.4 111.9 R-91 PIB-2 S Bulk 200:1:0.0 20.8 52.8 1.04 13,000 14,100 14,080 110.9 99.8 1 Bulk polymerization of styrene with thermal initiation at 100°C 2 Elapsed time between initiation and quenching of RAFT polymerization

70

Page 90: Polyisobutylene Chain End Transformations: Block Copolymer

71

Figure 3.1. 1H NMR spectrum of exo-olefin PIB.

8 7 6 5 4 3 2 1 0

ppm

918_31a_1_1

n

q

p

a1 , a2

2.1 2.0 1.9 1.8 1.7

q

p

5.2 5.1 5.0 4.9 4.8 4.7 4.6 4.5 4.4 4.3

a1 a2

Page 91: Polyisobutylene Chain End Transformations: Block Copolymer

72

Figure 3.2. 13C NMR spectrum of exo-olefin PIB.

40 39 38 37 36 35 34 33 32 31 30 29 28

62 61 60 59 58 57 56 55 54 53 52

250 200 150 100 50 0

ppm

12

3 6

48

7

105

9

10

74

6

5

n

3 6

8 9

1

2

3

Page 92: Polyisobutylene Chain End Transformations: Block Copolymer

73

Figure 3.3. GPC traces of pre-quenched and quenched exo-olefin PIB.

0 5 10 15 20 25

time (min)

Page 93: Polyisobutylene Chain End Transformations: Block Copolymer

74

Figure 3.4. 1H NMR spectrum of hydroxyl PIB.

5 4 3 2 1 0

ppm

918_35b_1

n

OHa’

a’

Page 94: Polyisobutylene Chain End Transformations: Block Copolymer

75

Figure 3.5. 13C NMR spectrum of hydroxy PIB.

200 150 100 50 0

ppm

918_35b_1_C

61 60 59 58 57 56 55 54 53 52 51 50

ppm

40 39 38 37 36 35 34 33 32 31 30 29 28

ppm

n

OH1

2

3 6

48

7 105

93 6

10

8

97

56

4

1

2 3

Page 95: Polyisobutylene Chain End Transformations: Block Copolymer

76

Figure 3.6. GPC traces of hydroxy PIB and exo-olefin PIB.

0 5 10 15 20 25 30

time (min)

Page 96: Polyisobutylene Chain End Transformations: Block Copolymer

77

n m

N

O

O S

S

S C12H25

OO

n S

S

S C12H25

N

O

O

n OH

nCl

+TiCl4 Nuc

n9BBN H2O2

S

S

S C12H25

N

O

HO

DCC

MMAAIBN

DMAP

Figure 3.7 Site transformation from living carbocationic polymerization to RAFT polymerization.

Page 97: Polyisobutylene Chain End Transformations: Block Copolymer

78

Figure 3.8. 1H NMR spectrum of CTA.

7 6 5 4 3 2 1

ppm

3.5 3.0 2.5 2.0 1.5

HO

O

N

S

S

S

9

h

g

f

f

e

e

d

d

c’b’

c’

g

h

b’

Page 98: Polyisobutylene Chain End Transformations: Block Copolymer

79

Figure 3.9. 13C NMR spectrum of CTA

HO

O

N

S

S

S

8

250 200 150 100 50 0

ppm

886_27a_3

50 45 40 35 30 25 20 15 10

886_27a_3

17

11

12/20

16

1418 13

2022

211519

1112

1314

15

16

17 1819

20 2221

Page 99: Polyisobutylene Chain End Transformations: Block Copolymer

80

Figure 3.10. gHSQC of CTA

0.501.001.502.002.503.003.50

10.0

15.0

20.0

25.0

30.0

35.0

40.0

Page 100: Polyisobutylene Chain End Transformations: Block Copolymer

81

Figure 3.11. 13C NMR spectrum of CTA: attached proton test.

250 225 200 175 150 125 100 75 50 25 0 -25

ppm

Page 101: Polyisobutylene Chain End Transformations: Block Copolymer

82

Figure 3.12. 1H NMR spectrum of PIB-CTA

3.00 2.75 2.50 2.25 2.00

8 7 6 5 4 3 2 1 0

ppm

918_38b_1

beh

d

fc

4.25 4.00 3.75 3.50 3.25 3.00

n

O

O

N

S

S

S

9

ab

c

d e

f

g

h

a’

a

PIB-OH

PIB-CTA

gPIB-CTA

CTA

a

b’

b

c’

c

Page 102: Polyisobutylene Chain End Transformations: Block Copolymer

83

Figure 3.13. 13C NMR spectrum of PIB-CTA

O

O

N

S

S

S

8n

250 225 200 175 150 125 100 75 50 25 0

ppm

918_38b_1_C

62 60 58 56 54 52 50 48 46 44

ppm

4 147

33 32 31 30 29

40 38 36 34 32 30 28 26 24 22 20 18

ppm

92115

18

5

13

17 11 16 10 22

20

19

8 12

6

20

1

2

3

33 66 9

1 2 4 78

1315

1114

10

16

1819

2021

175

12

22

Page 103: Polyisobutylene Chain End Transformations: Block Copolymer

84

Figure 3.14. 13C NMR spectra of CTA (top), PIB-OH (middle), and PIB-CTA (bottom).

225 200 175 150 125 100 75 50 25 0

PIB-CTA

PIB-OH

ppm

CTA

17

17

11

11

16

16

Page 104: Polyisobutylene Chain End Transformations: Block Copolymer

85

Figure 3.15. Partial 13C NMR spectra of CTA (top), PIB-OH (middle), and PIB-CTA (bottom).

75 70 65 60 55 50 45 40 35 30 25 20 15 10 5

PIB-CTA

PIB-OH

ppm

CTA

21

219

9

20

20

15

15

19

19

10

10

14

147

74

4

Page 105: Polyisobutylene Chain End Transformations: Block Copolymer

86

Figure 3.16. Expanded 13C NMR of CTA (top), PIB-OH (middle), and PIB-CTA (bottom).

40 39 38 37 36 35 34 33 32 31 30 29 28 27

PIB-CTA

PIB-OH

ppm

CTA

18

18

13

13

20

85

5

19

1920

8

6

6

12

12

20

20

Page 106: Polyisobutylene Chain End Transformations: Block Copolymer

87

Figure 3.17. Partial 13C NMR spectra comparing PIB-CTA to CTA and PIB-OH precursors

180 179 178 177 176 175 174 173 172 171 170

PIB-CTA

ppm

CTA

74 73 72 71 70 69 68

PIB-CTA

PIB-OH

ppm

11

11

10

10

(a)

(b)

Page 107: Polyisobutylene Chain End Transformations: Block Copolymer

88

Figure 3.18. 1H NMR spectrum of PIB-b-PMMA

2.6 2.5 2.4

n

O

O

N

S

S

S

9O

O

n

m

4 3 2 1 0

ppm

s

r ij

k

s

r

k

j

i

b

b

Page 108: Polyisobutylene Chain End Transformations: Block Copolymer

89

Figure 3.19. GPC traces of PIB-CTA and PIB-b-PMMA after purification.

0 5 10 15 20 25 30

PIB-CTA

time (min)

PIB-b-PMMA

Page 109: Polyisobutylene Chain End Transformations: Block Copolymer

90

Figure 3.20. Effect of system concentration on rate of RAFT polymerization.

0

10

20

30

40

50

60

70

80

90

100

0 2 4 6 8 10 12 14

Co

nve

rsio

n (

%)

Polymerization time (h)

[M] [M]:[CTA] [CTA]:[I]

(A) 4.5 125 2.0

(B) 3.5 125 2.0

(A) (B)

0.00

0.20

0.40

0.60

0.80

1.00

1.20

1.40

1.60

0 2 4 6 8 10 12 14

Ln([

M0]/

[M])

Polymerization Time (h)

(A) (B)

Page 110: Polyisobutylene Chain End Transformations: Block Copolymer

91

Figure 3.21. GPC traces of reaction aliquots removed from reaction R-3 at various conversions, along with the final purified block copolymer.

0 5 10 15 20

time (min)

p ~ 20 %

p ~ 30 %

p ~ 50 %

p ~ 100 %

p ~ 0 %

PIB-b-PMMA (purified)

Page 111: Polyisobutylene Chain End Transformations: Block Copolymer

92

Figure 3.22. Dependency of blocking efficiency on monomer selection: MMA vs. S.

0 5 10 15 20 25 30

PIB-b-PMMA

time (min)

PIB-CTA

0 5 10 15 20 25 30

time (min)

PIB-b-PS PIB-CTA

Page 112: Polyisobutylene Chain End Transformations: Block Copolymer

93

CHAPTER IV

POLYISOBUTYLENE RAFT CTA BY A CLICK CHEMISTRY SITE

TRANSFORMATION APPROACH: SYNTHESIS OF POLY(ISOBUTYLENE-b-N-

ISOPROPYLACRYLAMIDE)

Introduction

Development of controlled/living polymerizations has enabled synthetic polymer

chemists to design complex polymeric architectures with great precision.1 Block

copolymers in particular have gained significant attention because of their unique

material properties resulting from compositionally different block segments, and ability

to self assemble into highly organized structures.2 Self assembling materials have many

potential applications, both as dispersions and in the solid state, due to their ability to

form versatile and functional morphologies.1 Thus, block copolymer systems are being

developed for and have found numerous applications in the areas of biomedical

devices,3,4,5,6 biomaterials,4,7,8 thermoplastic elastomers (TPEs),9 fuel cells,10,11 and

electronics.12,13

Polyisobutylene (PIB) is a completely saturated hydrocarbon rubber that can only

be produced through cationic polymerization. Inherent attractive properties of this

synthetic elastomer are its exceptional oxidative and chemical resistance, superior gas

barrier and mechanical damping characteristics, and excellent biocompatibility. Because

of these properties PIB-based block copolymers have continued to be a topic of great

interest. Poly(styrene-b-isobutylene-b-styrene) (SIBS) block copolymer has found a

niche market as a biomaterial and is used as the matrix polymer for the drug-containing

coating on a commercial drug-eluting coronary stent.14,15,16 Recently, prospective

Page 113: Polyisobutylene Chain End Transformations: Block Copolymer

94

materials for biotechnology and medicine were created from self-assembling PIB-based

polymersomes and micelles creating novel delivery/encapsulation systems17,18 and

interpolyelectrolyte complexes.19 In addition, Kennedy et al. have recently developed

PIB-based segmented thermoplastic polyurethane/ureas (TPUs) exhibiting unprecedented

resistance to oxidative/hydrolytic degradation for long term medical applications.20,21 In

other areas, novel PIB-based TPEs were synthesized, with poly(ether ketone) outer

blocks for increased thermal stability22 and dendritic PIB-core TPEs with carbon black

and silica fillers for increased mechanical properties.23

PIB-based block copolymers have been synthesized by four different strategies:

sequential monomer addition, macromolecular coupling, dual-site initiators,24 and site

transformation.9 Sequential monomer addition offers unparalleled simplicity, but in the

case of PIB-based block copolymers, allows combinations of cationically polymerizable

monomers only. This selection of monomers is confined essentially to isobutylene,

isoprene, vinyl ethers, styrenics, and N-vinylcarbazole, and the process is no longer

simple when the monomer pair exhibit large reactivity differences.25 Macromolecular

coupling17,26 of previously fabricated functional polymers requires quantitative

functionalization of the chain ends of both polymers, a high-conversion coupling

reaction, and a common solvent system; otherwise the yield of the desired block

copolymer is low and a tedious separation of homopolymers is often involved. Dual-site

initiators contain initiating sites for two mechanistically different polymerizations and

require no intermediate transformation steps to affect the second polymerization after the

first. Site transformation involves conversion of the head or tail group of the PIB block

segment into an initiator for a mechanistically different chain polymerization process.

Page 114: Polyisobutylene Chain End Transformations: Block Copolymer

95

This technique effectively expands the library of polymers segments that can be mated to

PIB. Block copolymer synthesis using site transformation has been successfully

demonstrated with atom transfer radical polymerization (ATRP),27,28,29 condensation

polymerization,30 anionic polymerization,31,32 and most recently by reversible addition

fragmentation transfer (RAFT) polymerization.33

RAFT has proven to be very versatile regarding monomers, solvents, and reaction

conditions, and yields macromolecules with predetermined molecular weights and narrow

polydispersities.34,35 Because of its versatility and capability to polymerize many

monomers that are inherently troublesome for other polymerization techniques, RAFT is

potentially an ideal polymerization technique to combine with cationic polymerization by

site transformation. In addition, chain transfer agents (CTAs) used in the RAFT process

can be synthesized to carry many useful reactive end groups including azide36 and

alkyne37 for copper(I)-catalyzed Huisgen [3 + 2] dipolar cycloadditions. This reaction,

categorized by Sharpless et al. as a “click” reaction, is known to be quantitative, devoid

of side reactions and byproducts, tolerant to a wide range of functional groups, and

requiring only mild reaction conditions.38,39 Due to these attributes, the azide/alkyne

reaction has received significant attention in macromolecular chemistry and has found

use in specific systems for postpolymerization functionalization, novel polymers

synthesis, and chain extension for block copolymer synthesis.40

Stimuli-responsive block copolymers are of immense importance and have

attracted significant attention as “smart” materials.41 Poly(N-isopropylacrylamide)

(PNIPAM) is known to display a sharp, thermally-reversible phase transition in aqueous

solution at approximately 32 °C.42 Investigation of this phase transition has revealed that

Page 115: Polyisobutylene Chain End Transformations: Block Copolymer

96

PNIPAM macromolecules experience dehydration and collapse from a hydrated,

extended coil to a hydrophobic globule, upon raising the temperature through the cloud

point, ultimately resulting in intermolecular aggregation.43 Various hydrophobically-

modified PNIPAM-based polymers have been reported in literature and are being studied

for many potential applications.44 Functional PNIPAM’s have been synthesized to be

joined with various hydrophobic polymer blocks including, for example, polystyrene45,46

and tetrahydrofuran-protected 2-hydroxyethyl methacrylate,47 as well as alkyl groups,48

alkyl chain transfer agents,49 and chromophores.50 Herein, we report the synthesis of a

novel amphiphilic diblock copolymer, composed of PIB and PNIPAM segments, through

the combination of quasiliving cationic polymerization and RAFT by a click chemistry

site transformation procedure.

Experimental

Materials

Hexanes (anhydrous, 95%), 2,6–lutidine (redistilled, 99.5%), TiCl4 (99.9%,

packaged under N2 in sure-seal bottles), chloroform-d (99.8 atom% D), DMF (anhydrous,

99.8%), bromotris(triphenylphosphine)copper(I) (Ph3PCuBr) (98%), and 1,3,5-trioxane

(≥99%) were purchased from Sigma-Aldrich and used as received. 1-(3-

Bromopropyl)pyrrole (PyBr) (>95%, TCI America), THF, heptanes, and dioxane were

distilled from calcium hydride under a N2 atmosphere. N-Isopropylacrylamide (NIPAM)

was recrystallized twice from a hexane/benzene mixture (3/2, v/v). 2,2′-Azobis(2-

methylpropionitrile) (98%) (AIBN) was recrystallized twice from ethanol. Isobutylene

(IB) (BOC, 99.5%) and CH3Cl (Alexander Chemical Corp.) were dried by passing the

gaseous reagent through columns packed with CaSO4 and CaSO4/4 Å molecular sieves,

Page 116: Polyisobutylene Chain End Transformations: Block Copolymer

97

respectively, and condensed within a N2-atmosphere glove box immediately prior to use.

2-Chloro-2,4,4-trimethylpentane (TMPCl) was prepared by bubbling HCl gas through

neat 2,4,4-trimethyl-1-pentene (Sigma-Aldrich) at 0°C. The HCl-saturated TMPCl was

stored at 0oC, and was neutralized with NaHCO3, dried over anhydrous MgSO4, and

filtered immediately prior to use. Alkyne-functional CTA, propargyl 2-(1-

dodecylsulfanylthiocarbonylsulfanyl)-2-methylpropionate, was synthesized as previously

reported.37

Synthesis of 1-(3-Azidopropyl)pyrrole-terminated PIB (PIB-N3)

Three 1-(3-bromopropyl)pyrrole-terminated PIB precursors (Table 4.1) were

synthesized by quenching TMPCl-initiated quasiliving IB polymerizations with 1-(3-

bromopropyl)pyrrole.51 A representative procedure to produce PIB44-Br was as follows.

Quasiliving polymerization of IB with TMPCl as an initiator was carried out within a N2

atmosphere glovebox, equipped with an integral, cryostated hexane/heptanes bath. Into a

round-bottom flask equipped with a mechanical stirrer, infrared probe, and thermocouple

were added 144 mL CH3Cl, 216 mL hexanes, and 0.15 mL (0.14 g, 1.3 mmol) of 2,6-

lutidine. The mixture was allowed to equilibrate to – 70 °C, and then 20.4 mL (14.2 g,

0.253 mol) of IB was charged to the reactor. After thermal equilibration, 1.113 mL

(0.9730 g, 6.545 mmol) of TMPCl was added to the reactor. To begin the

polymerization, 2.16 mL (3.74 g, 0.0197 mol) of TiCl4 was charged to the reactor. Full

monomer conversion ( >98 %) was achieved in 40 min, after which time a prechilled

solution of PyBr, prepared by dissolving 1.81 mL of PyBr (2.46 g, 13.1 mmol) into 25

mL of hexane/CH3Cl (60/40, v/v, – 70 °C), was added to the polymerization. PyBr was

allowed to react with the living chain ends for 40 min. Finally, the reaction was

Page 117: Polyisobutylene Chain End Transformations: Block Copolymer

98

quenched by addition of excess prechilled methanol. The contents of the reaction flask

were allowed to warm to room temperature, and the polymer in hexane was immediately

washed with methanol and then precipitated into methanol from hexane. The precipitate

was collected by dissolution in hexane; the solution was washed with water, dried over

MgSO4, and concentrated on a rotary evaporator. Residual solvent was removed under

vacuum at room temperature.

Each of three 1-(3-azidopropyl)pyrrole-terminated PIB’s was prepared by reaction

of the corresponding 1-(3-bromopropyl)pyrrole-terminated PIB with NaN3 in a mixture

of heptanes/DMF. A representative procedure to produce PIB44-N3 was as follows. To a

200 mL round bottom flask, under a dry nitrogen atmosphere, were added 1-(3-

azidopropyl)pyrrole-terminated PIB44 (11.56 g, 4.4 mmol) and 25.5 mL of dry heptanes.

The reaction vessel was agitated until complete dissolution of the PIB-N3 occurred

resulting in a clear solution. To the resulting solution was then added a separate solution

of sodium azide (0.853 g, 13.1 mmol) in 25.5 mL of DMF. After stirring for

approximately 20 min, the resulting biphasic mixture was submerged in an oil bath at 90

°C and soon became monophasic as the reaction reached equilibrium temperature. The

reaction was allowed to proceed with agitation for 17 h. Afterward, the monophasic

reaction mixture was removed from the oil bath and allowed to cool to room temperature

whereupon a biphasic mixture again formed. At this point, hexanes and DI water were

added and the DMF/water layer was separated. The organic phase was immediately

washed with DI water and then precipitated into methanol from hexane. The precipitate

was collected by dissolution in hexane; the solution was washed with water, dried over

Page 118: Polyisobutylene Chain End Transformations: Block Copolymer

99

MgSO4, filtered, and concentrated on a rotatary evaporator. Residual solvent was

removed under vacuum at room temperature.

Synthesis of PIB-CTA by Click Chemistry

To a 25 mL one-neck round-bottom flask equipped with a magnetic stir bar were

added PIB44-N3 (2.032 g, 0.79 mmol) and 2.5 mL THF. The reaction vessel was agitated

until complete dissolution of the PIB-N3 occurred resulting in a clear solution. In a

separate vessel, CTA (0.622 g, 1.55 mmol) and (Ph3P)3CuBr (0.0515g, 0.0554 mmol)

were dissolved in 2.5 mL of THF, and the resulting solution was charged to the one-neck

round-bottom flask. After 38 h, the solvent was removed under reduced pressure, and the

crude product was redissolved in hexane. The resulting solution was washed with

methanol three times, filtered, and precipitated twice into methanol from hexanes. The

precipitate was collected by dissolution in hexanes and then stripped of solvent under

reduced pressure until a constant weight was reached. A clear, yellow, viscous material

was obtained.

RAFT polymerization of NIPAM with PIB-CTA

A representative RAFT polymerization was conducted as follows. To a 25 mL

Schlenk-style, long-neck round-bottom flask were added three solutions: first a solution

of PIB44−CTA (0.0493 g, 0.0165 mmol) in 0.69 mL of heptanes, a second solution of

1,3,5-trioxane (0.0458 g, 0.509 mmol) and NIPAM (0.5467g, 5.0961mmol) in 2.82 mL

dioxane, and a finally 0.5 mL of a 0.0099 M AIBN stock solution in dioxane. After

mixing the three solutions an initial aliquot was taken to establish the initial monomer

concentration relative to the internal standard 1,3,5-trioxane, via 1H NMR spectroscopy.

The solution was then subjected to three freeze−pump−thaw cycles to remove oxygen,

Page 119: Polyisobutylene Chain End Transformations: Block Copolymer

100

sealed under N2, and submerged in an oil bath thermostatted at 85 °C. After 4 h the

reaction was exposed to oxygen and quenched in liquid nitrogen. A final aliquot was

taken for 1H NMR analysis, and then the crude reaction product was precipitated into

hexanes twice, dialyzed against water, and lyophilized. Conversion was calculated from

the initial and final monomer concentrations relative to 1,3,5-trioxane.

Instrumentation

NMR spectra were acquired using a Varian Mercuryplus 300 MHz NMR

spectrometer. Samples were dissolved in chloroform-d (3−7%, w/v) and analyzed using

5 mm NMR tubes.

Variable-temperature NMR (VT NMR) data were acquired using a UNITYInova

500 MHz NMR spectrometer equipped with a Highland VT controller. RAFT

polymerizations were performed in dioxane-d between 355.5 and 363 K, with a 45 s pre-

acquisition delay between each spectrum. The probe was allowed to equilibrate for 15

min prior to data acquisition. Temperatures reported in this study are within ±2 °C based

on ethylene glycol calibration.52,53 RAFT polymerizations for VT NMR were prepared

by charging a J Young NMR tube equipped with a Teflon seal with the crude reaction

mixture and performing three freeze−pump−thaw cycles. After degassing, the NMR tube

was backfilled with N2. Conversion was calculated by comparison of the vinyl proton

areas of the monomer to the internal standard, 1,3,5-trioxane.

Number-average molecular weight (Mn) and polydispersity index (PDI) of the

polymeric materials were measured using gel permeation chromatography (GPC). The

GPC system, operating at 35 °C for PIB-Br, PIB-N3, and PIB-CTA, and 25 °C for PIB-b-

PNIPAM, consisted of a Waters Alliance 2695 Separations Module, an on-line

Page 120: Polyisobutylene Chain End Transformations: Block Copolymer

101

multiangle laser light scattering (MALLS) detector (MiniDAWNTM, Wyatt Technology

Inc.), an interferometric refractometer (Optilab rEXTM, Wyatt Technology Inc.), an on-

line differential viscometer (ViscoStarTM, Wyatt Technology, Inc.), and either two mixed

E (3 µm bead size) or two mixed D (5 µm bead size) PL gel (Polymer Laboratories Inc.)

GPC columns connected in series. Freshly distilled THF for PIB-homopolymer (PIB-Br,

PIB-N3, and PIB-CTA) and 0.25 wt% tetrabutylammonium bromide in THF (25 °C)54 for

PIB-b-PNIPAM served as the mobile phase at a flow rate of 1.0 mL/min. Sample

concentrations were 10-12 mg/mL, with an injection volume of 100 µL. The detector

signals were recorded using ASTRATM software (Wyatt Technology Inc.) and molecular

weights were determined using dn/dc values calculated from a known dn/dc equation

reported elsewhere51 or by the response of the Optilab DSP assuming 100% mass

recovery from the columns.

Dynamic light scattering (DLS) was used to determine the hydrodynamic size of

particles in aqueous solution, as well as their distribution of sizes. Solutions of PIB-b-

PNIPAM block copolymers were prepared by dissolving the polymer into purified water

(Millipore) to a concentration of 0.01 wt%. Samples were agitated overnight to ensure

complete dissolution and then filtered through a 0.22 µm PVDF syringe-driven filter

(Millipore) directly into the scattering cell. Samples were then sonicated and allowed to

equilibrate to temperature for 20 min prior to analysis. Scattering was performed using

incident light at 633 nm from a Spectra Physics HeNe laser operating at 40 mW. For

DLS, the angular dependence of the autocorrelation functions was measured using a

Brookhaven BI-200SM goniometer with a TurboCorr correlator. Correlation functions

were analyzed according to the method of cumulants using the companion software, from

Page 121: Polyisobutylene Chain End Transformations: Block Copolymer

102

which a hydrodynamic diameter is extracted via the Stokes-Einstein relation. All data

reported correspond to the average decay rate and normalized variance (polydispersity)

obtained from the second-order cumulant fit.

Results and Discussion

Synthesis of PIB-CTA

Site transformation of PIB into a macro-CTA for RAFT polymerization was

accomplished using a click chemistry site transformation approach. This synthetic

strategy, shown in Figure 4.1, offers a number of advantages over our previously

published method, which involved quenching of quasiliving PIB with a hindered base to

form exo-olefin PIB, followed by hydroboration/oxidation to form hydroxyl-terminated

PIB, followed finally by esterification with a carboxylic acid-functional trithiocarbonate

CTA.55 Most notably, in situ quenching of quasiliving PIB with an N-substituted pyrrole

provides higher chain-end functionality compared to the exo-olefin route, and the

expensive and difficult hydroboration/oxidation reaction is eliminated.

As shown in Figure 4.1, PIB-N3 was synthesized in two steps. The first involved

in situ quenching of a TiCl4-activated quasiliving cationic polymerization of IB with 1-

(3-bromopropyl)pyrrole to obtain the primary bromide-terminated PIB.51 This reaction

was quantitative and yielded a mixture of two isomers, corresponding to attachment of

the PIB chain at either the C2 or C3 position of the pyrrole ring. The C3 isomer was the

major isomer and constituted about 60% of the chain ends. Table 4.1 lists molecular

weights (Mn) and PDIs (Mw/Mn) of three bromide-functional PIBs synthesized by this

method. This series of functional polymers was designed to probe the influence of PIB-

CTA molecular weight on the RAFT polymerization of NIPAM and the properties of the

Page 122: Polyisobutylene Chain End Transformations: Block Copolymer

103

resulting block copolymers. The results obtained from GPC and 1H NMR agree very

well, differing by about 6% in the worst case. In addition, the PDI’s were uniformly low.

The second step involved displacement of the terminal bromide by azide ion via

nucleophilic substitution. The reaction was biphasic at room temperature, with the

hydrophobic PIB located in the top heptanes layer, and the polar sodium azide located in

the bottom DMF layer. Upon heating, the reaction became homogeneous, and after

approximately 12 h at 90 °C, complete substitution of the halogen to the terminal azide

was accomplished without any noticeable coupling or loss of terminal functionality.

Upon cooling, the layers separated once again, and this greatly facilitated removal of

excess sodium azide and the byproduct sodium bromide.

The final step to create the PIB-CTA was the copper-catalyzed Huisgen

cycloaddition reaction between PIB-N3 and the alkyne-functionalized CTA (Figure 4.1).

The latter was synthesized by DCC/DMAP coupling as previously reported in literature.37

The azide-alkyne click reaction is attractive for the final coupling step since it can be

carried out quantitatively at room temperature and in the presence of both oxygen and

water. Conversion of PIB-N3 into PIB-CTA (Figure 4.2) was monitored using real-time

1H NMR analysis by observing the decrease in area of the propargyl methylene proton

signal of the CTA. From this graph nearly 80 % of the product was obtained after 8 h;

whereas the remaining 20 % required an additional 27 h to reach full conversion. Since

both the CTA and PIB-N3 had to be synthesized, it was impractical to use the former in

large excess and thereby reduce the overall order of the reaction. Thus, the initial molar

ratio (M) of alkyne to azide ([CTA]:[PIB-N3]) was set only to 2. As expected under these

conditions, the click reaction displayed second order kinetics, as shown in the inset to

Page 123: Polyisobutylene Chain End Transformations: Block Copolymer

104

Figure 4.2. Initial scouting experiments did reveal that the reaction rate can be

dramatically increased by small increases in temperature (30 – 35 °C), higher catalyst

concentrations, or by using a more reactive copper catalyst. Although full conversion

takes approximately 35 h under these conditions, the simplicity and forgiving nature of

the reaction makes it very appealing. The product was purified by vacuum-stripping of

THF, dissolution in hexanes, and washing the resulting solution with methanol. The

solution was then filtered to remove the copper catalyst, which has limited solubility in

hexane, and finally, the polymer was isolated by precipitation into methanol, which also

served to remove any residual catalyst and unreacted alkyne CTA. After removal of

methanol, the purified PIB-CTA was a transparent yellow color instead of its originally

colorless appearance, as shown in Figure 4.2.

The structure of PIB-CTA was confirmed by 1H NMR, FTIR, and GPC. Figure

4.3 shows 1H NMR spectra of the PIB-N3 precursor (top), the alkyne CTA (middle) , and

the final PIB-CTA (bottom). A clear downfield shift of the PIB-N3 tether protons

(a′,b′,c′) and disappearance of the alkyne (d′) and methylene (e′) protons of CTA were

observed upon triazole formation. No residual resonances from either PIB-N3 or alkyne

CTA were detectable in the resulting product indicating quantitative functionalization.

Formation of the triazole ring brings the alkyne proton into the general

aromatic/heteroaromatic proton region. This proton (d) was observed as two separate

peaks reflecting the mixture of 2-PIB and 3-PIB pyrrole isomers.

FTIR spectroscopy, Figure 4.4, clearly showed the disappearance of the azide at

2100 cm-1 and presence of three CTA frequencies representing the carbonyl, ester, and

thiocarbonyl at approximately 1735, 1150 and 1124, and 1067 cm-1 respectively. Further

Page 124: Polyisobutylene Chain End Transformations: Block Copolymer

105

characterization with GPC, Figure 4.5, confirmed a small increase in molecular weight

after coupling, due to addition of the 403 g/mol CTA, and absence of any chain extension

or other side reactions as evidenced by a uni-modal trace. Additionally, no detectable

residual lower molecular weight PIB-N3 was visible in the PIB-CTA trace.

RAFT Polymerization of NIPAM

After successful synthesis of PIB-CTA, it was employed in the RAFT

polymerization of NIPAM, as depicted in Scheme 4.6. A mixed solvent system of

dioxane/heptane was used to achieve a homogeneous reaction medium. For some

formulations, the reaction was slightly heterogeneous upon initial mixing at room

temperature, but the system became homogenous after submerging the reaction in an oil

bath during polymerization. It should also be noted that dissolution of the PIB-CTA in

heptanes prior to its addition to the other reagents added in its facile solvation

substantially.

Table 4.2 summarizes the results of RAFT polymerizations formulated using

various combinations of PIB-CTA molecular weight ( n,PIB-CTAM ), [NIPAM]:[CTA] ratio,

and NIPAM fractional conversion (pNIPAM), conducted to demonstrate control of both the

PIB and PNIPAM block lengths. Theoretical degree of polymerization of the PNIPAM

block ( n, theoX ) was targeted as pNIPAM×[NIPAM]/[CTA]. Reactions R1–R3 employed the

same PIB-CTA and targeted different n, theoX ranging from approximately 190 to 370.

Experiments R-3 through R-5 utilized three different PIB-CTA molecular weights

ranging from approximately 3,000 to 6,000 g/mol. In all cases, good to excellent

agreement was observed between theoretical and experimental degree of polymerization

Page 125: Polyisobutylene Chain End Transformations: Block Copolymer

106

of the PNIPAM block ( n,GPCX ), and the resulting diblock copolymers all possessed

narrow PDI.

The resulting block copolymer was characterized using 1H NMR and GPC. 1H

NMR (Figure 4.7) shows clear, strong resonances from both segments of the block

copolymer. PIB backbone resonances are identified “a” and “b” while the PNIPAM

resonances are labeled “c” through “g.” Purification of the block copolymer was

accomplished by precipitation into hexanes, for removal of residual PIB-CTA, and

dialysis against water for removal of unreacted NIPAM. A representative GPC trace,

Figure 4.8, shows lower elution time of the block copolymer in comparison to the PIB-

CTA homopolymer, indicative of increased molecular weight of the former. Moreover,

the crude GPC trace shows a small amount of residual PIB-CTA and its effective removal

by the hexanes precipitation.

RAFT Polymerization Kinetics

Kinetics of RAFT polymerizations from various PIB-CTA was monitored using

variable temperature real time 1H NMR. Figure 4.9 shows the temperature dependence

of the RAFT polymerization with a NIPAM concentration of 1.2 M and

[NIPAM]:[CTA]:[I] = 250:1:0.25. After a characteristic induction period, the

polymerizations were first-order up to pNIPAM ≈ 0.7, indicating a constant concentration of

actively propagating radicals; whereas at higher conversions the plots displayed a slight

downward curvature implying a decreasing concentration of active species. Eventually a

maximum conversion was reached beyond which no further polymerization occurred; for

the higher polymerization temperatures this was at pNIPAM ≈ 0.9. The apparent first-order

rate constants, kapp, extracted from the linear portion of the plots, were fitted to the

Page 126: Polyisobutylene Chain End Transformations: Block Copolymer

107

Arrhenius equation (Figure 4.10) to yield activation energy of 31.0 kcal/mol and

prefactor of 2.71 x 1017 min-1. This is very close to the reported activation energy for

AIBN dissociation and confirms the strong effects of reaction temperature on

polymerization rate. In addition, the duration of the induction period decreased with

increasing temperature and was practically eliminated at 90 °C.

The kinetics and energetics information obtained in the VT NMR experiments

were consistent with RAFT polymerization kinetics observed in the batch reactions of

Table 4.2. For the latter, the external bath temperature was 85°C; however, the internal

reactor temperature was typically only 78-79°C, after a thermal equilibration period of

about 15-20 min. A replicate of experiment R-2 was performed in which aliquots were

removed for conversion analysis using 1H NMR. The data revealed first-order kinetics

(kapp = 0.0127 s-1) and an induction period of 50 min (taken as the x-intercept from the

first-order plot). This kapp value agrees very well with that predicted from the Arrhenius

relationship (kapp = 0.0139 s-1) above for a reaction temperature of 78°C. The previously

mentioned Arrhenius relationship and experimentation help to verify the slower reaction

rate observed for the batch reaction, specifically R-2, which accounts for it achieving a

lower than expected conversion value (pNIPAM = 0.376). Using either of the above rate

constants and the empirically determined induction period, theoretical conversion values

between 0.32 and 0.34 can be calculated which fall in close proximity to the experimental

conversion value of R-2.

Figure 4.11 shows the dependence of RAFT polymerization kinetics on the

concentration of the thermal initiator, AIBN. The maximum conversion achieved in the

polymerization decreased with lower concentrations of initiator, from just over 0.9 at

Page 127: Polyisobutylene Chain End Transformations: Block Copolymer

108

[CTA]:[I] = 0.25 to approximately 0.73 at [CTA]:[I] = 0.125. In addition, increasing

initiator concentration reduced the duration of the induction period and increased the

apparent rate constant of the polymerization, as expected.

Conversion and rate of RAFT polymerizations were observed to be independent

of both PIB-CTA molecular weight and concentration (Figures 4.12 and 4.13).

Aqueous Self-assembly of PIB-b-PNIPAM

Below the critical point of PNIPAM (ca. 32 oC), amphiphilic PIB-PNIPAM block

copolymers are expected to self-assemble into higher-ordered structures. In addition,

these structures are expected to show temperature-dependent dimensions near the critical

point for PNIPAM. The self-assembly and temperature responsiveness for PIB44-

PNIPAM450 (R-6 from Table 4.2) was studied using DLS. Based on the large weight

fraction of hydrophilic PNIPAM, it is expected that spherical micelles will form.56 At

room temperature, direct solvation of PIB44-PNIPAM450 resulted in a clear solution

(Figure 4.14). The correlation functions fit well to a second-order cumulant relation,

indicative of a relatively monodisperse, unimodal distribution of scattering species. A

plot of the average decay rate of the correlation function (Γ) vs. the square of the

scattering vector (q2) yields a linear relation (Figure 4.14 inset), indicating that the

scattering comes solely from Brownian motion of spherical aggregates.

The hydrodynamic diameter (Dh) was determined as a function of temperature at a

90o scattering angle. At 15 oC, the Dh was 107 nm. As the temperature increases, the

hydrodynamic size remains relatively constant until 26 oC where it begins to decrease.

When the temperature is further increased towards the nominal critical point of PNIPAM,

the aggregate diameter continues decreasing to 76 nm. The decrease in size with

Page 128: Polyisobutylene Chain End Transformations: Block Copolymer

109

increased temperature is explained by the PNIPAM chains transforming from a coil to a

globular conformation.57 This conformational change results in the expulsion of water

due to the increased hydrophobicity of the PNIPAM corona chains. It is important to

note that the overall structure of the aggregates does not change over the temperature

range studied, i.e.: the Γ vs. q2 plot retains linearity over the temperature range,

suggesting that spherical aggregates are still present at 32 oC (Figure 4.14 inset). In

addition, the polydispersity (normalized variance) in the size distribution systematically

decreased from 0.187 to 0.133 as temperature increased from 15 to 33 oC. Finally, a

fairly broad transition ranging from 26 to 33 °C occurs with PIB44-b-PNIPAM450 block

copolymer, in contrast to sharp transitions which occur in PNIPAM homopolymer

systems.42 Broadened PNIPAM transitions are theoretically expected with

hydrophobically modified PNIPAM58 and have been observed with PS-b-PNIPAM

systems.59 Once the temperature of the system exceeded 33 °C intermolecular

aggregation occurred between assemblies, which was accompanied by a visual change in

the aqueous solution from transparent to turbid (Figure 4.14).

Conclusion

A novel block copolymer composed of PIB and PNIPAM block segments was

synthesized by a unique site transformation process combining quasiliving cationic and

RAFT polymerizations. PIB with a terminal halogen was functionalized by in situ

quenching and successfully convert into a “clickable” PIB by substitution with an azide

ion. Afterward, the azido functionalized PIB was transformed into a macro-CTA by

copper catalyzed click chemistry utilizing an alkyne functionalized CTA. The utility of

the macro-CTA was demonstrated by polymerization of NIPAM to produce a novel block

Page 129: Polyisobutylene Chain End Transformations: Block Copolymer

110

copolymer with predetermined molecular weights and low PDIs. The kinetic behavior of

this RAFT polymerization was studied and found to be strongly dependant on

temperature and initiator concentration. Conversely, the PIB-CTA molecular weight and

concentration had little effect on the kinetic behavior. Light scattering experiments

confirmed the presence of polymeric aggregates with stimuli responsive behavior display

a broad temperature induced transition.

Page 130: Polyisobutylene Chain End Transformations: Block Copolymer

111

References 1. Matyjaszewski, K. Prog. Polym. Sci. 2005, 30, 858-875.

2. Ruzette, A.; Leibler, L. Nature Materials 2005, 4, 19-31.

3. Lindman, B.; Alexandridis, P. In Amphiphilic block copolymers: Self Assembly and

Applications, 1st Ed.; Elsevier Science B. V., Netherlands 2000; Ch.15, p 347 – 376.

4. Kopecek, J. Nature 2002, 417, 388 – 391.

5. Jeong, B.; Bae, Y. H.; Lee, D. S.; Kim, S. W. Nature 1997, 388, 860 – 862.

6. McIlroy, D.; Barteau, B.; Cany, J.; Richard, P.; Gourden, C.; Conchon, S.; Pitard, B.

Molecular Therapy 2009, doi:10.1038/mt.2009.84.

7. Pinchuk, L.; Wilson, G. J.; Barry, J. J.; Schoephoerster, R. T.; Parel, J. M.; Kennedy, J.

P. Biomaterials 2008, 29, 448–460.

8. Place, E. S.; Evans, N. D.; Stevens, M. M. Nature Materials 2009, 8, 457-470.

9. Hadjichristidis, N.; Pispas, S.; Floudas, G. A. In Block Copolymers: Synthetic

Strategies, Physical Properities, and Applications, 1st Ed.; John Wiley & Sons, New

York 2002; Ch.21, p 383-406.

10. Diat, O.; Gebel, G. Nature Materials 2008, 7, 13-14.

11. Lazzari, M.; Liu, G.; Lecommandoux, S. In Block Copolymers in Nanoscience, 1st

Ed.; Wiley-VCH 2006, Weinheim; Ch.15, p 337 – 366.

12. Tang, C.; Lennon, E. M.; Fredrickson, G. H.; Kramer E. J.; Hawker, C. J. Science

2008, 332, 429-432.

13. Ryan, A. Nature 2008, 456, 334-336.

Page 131: Polyisobutylene Chain End Transformations: Block Copolymer

112

14. Ranade, S. V.; Miller, K. M.; Richard, R. E.; Chan, A. K.; Allen, M. J.; Helmus, M.

N. J. Biomed. Mater. Res. 2004, 71A, 625–634.

15. Pinchuk, L.; Wilson, G. J.; Barry, J. J.; Schoephoerster, R. T.; Parel, J.-M.; Kennedy,

J. P. Biomaterials 2008, 29, 448–460.

16. Fray, M. E.; Prowans, P.; Puskas, J. E.; Altstadt, V. Biomacromolecules 2006, 7, 844-

850.

17. Binder, W. H.; Sachsenhofer, R. Macromol. Rapid Commun. 2008, 29, 1097–1103.

18. Nagy, M.; Szollosi, L.; Keki, S.; Faust, R.; Zsuga, M. JMS-Part A: Pure Appl. Chem.

2009, 46, 331-338.

19. Burkhardt, M.; Ruppel, M.; Tea, S.; Drechsler, M.; Schweins, R.; Pergushov, D. V.;

Gradzielski, M.; Zezin, A. B.; Mueller, A. H. E. Langmuir 2008, 24, 1769-1777.

20. Jewrajka, S. K.; Yilgor, E.; Yilgor, I.; Kennedy, J. P. J. Polym. Sci,: Part A: Polym.

Chem. 2008, 47, 38-48.

21. Jewrajka, S.K.; Kang, J.; Erdodi, G.; Kennedy, J.P. J. Polym. Sci.: Part A: Polym.

Chem. 2009, 47, 2787-2797.

22. Binder, W. H.; Machl, D. J. Polym. Sci., Part A: Polym. Chem. 2005, 43, 188–202.

23. Puskas, J. E.; Dos Santos, L. M.; Kaszas, G.; Kulbara, K. J.Polym. Sci.: Part A:

Polym. Chem. 2009, 47, 1148-1158.

24. Zhu, Y.; Storey, R.F. ACS Div. Polym. Chem., Polym. Preprs. 2009, 50(2), 535-536.

25. Kwon, Y.; Faust, R. Adv. Polym. Sci. 2004, 167, 107-135.

26. Takacs A.; Faust, R. Macromolecules 1995, 28, 7266.

Page 132: Polyisobutylene Chain End Transformations: Block Copolymer

113

27. Jakubowski, W.; Tsarevsky, N. V.; Higashihara, T.; Faust, R.; Matyjaszewski, K.

Macromolecules 2008, 41, 2318-2323.

28. Storey, R. F.; Scheuer, A. D.; Achord, B. C. Polymer 2005, 46, 2141-2152.

29. Breland, L.K.; Murphy, J.C.; Storey, R.F. Polymer 2006, 47, 1852-1860.

30. Zaschke, B.; Kennedy, J. P.; Macromolecules 1995, 28, 4426-4432.

31. Martinez-Castro, N.; Lanzendorfer, M. G.; Muller, A. H. E.; Cho, J. C.; Acar, M. H.;

Faust, R. Macromolecules 2003, 36, 6985-6994.

32. Feng, D.; Higashihara, T.; Faust, R. Polymer 2008, 49, 386-393.

33. Magenau, A. J. D.; Martinez-Castro, N.; Storey, R. F. Macromolecules 2009, 42,

2353-2359.

34. Moad, G.; Chong, Y. K.; Postma, A.; Rizzardo, E.; Thang, S. H. Polymer 2005, 46,

8458–8468.

35. Moad, G.; Rizzardo, E.; Thang, S. H. Polymer 2008, 49, 1079–1131.

36. Gondi, S. R.; Vogt, A. P.; Sumerlin, B. S. Macromolecules 2007, 40, 474-481.

37. Ranjan, R.; Brittain, W. J. Macromolecules 2007, 40, 6217-6223.

38. Kolb, H. C.; Finn, M. G.; Sharpless, K. B. Angew. Chem., Int. Ed. 2001, 40, 2004-

2021.

39. Rostovtsev, V. V.; Green, L. G.; Fokin, V. V.; Sharpless, K. B. Angew. Chem., Int.

Ed. 2002, 114, 2708-2711.

40. Lodge, L. P. Macromolecules 2009, 42, 3827-3829.

41. Ballauff, M.; Lu, Y. Polymer 2007, 48, 1815-1823.

42. Schild, H. G. Prog. Polym. Sci. 1992, 17, 163-249.

Page 133: Polyisobutylene Chain End Transformations: Block Copolymer

114

43. Yamazaki, A.; Song, J. M.; Winnik, F. M.; Brash, J. L. Macromolecules 1998, 31,

109-115.

44. Wei, H.; Zhang, X.; Cheng, C.; Cheng, S. Zhou, R. Biomaterials 2007, 28, 99-107.

45. Zhou, X.; Ye, X.; Zhang, G.; J. Phys. Chem. .B 2007, 111, 5111-5115.

46. Nuopponen, M.; Ojala, J.; Tenhu, H. Polymer 2004, 45, 3643-3650.

47. Klaikherd, A.; Nagamani, C.; Thayumanavan, S. J. Am. Chem. Soc. 2009, 131, 4830-

4838.

48. Schild, H.G.; Tirrell, D. A. Langmuir 1991, 7, 1319-1324.

49. Kujawa, P.; Segui, F.; Shaban, S.; Diab, C.; Okada, Y.; Tanaka, F.; Winnik, F. M.

Macromolecules 2006, 39, 341-348.

50. Yamazaki, A.; Song, J.; Winnink, F.; Brash, J. L. Macromolecules 1998, 31, 109-115.

51. Martinez-Castro, N.; Morgan, D. L..; Storey, R. F. Macromolecules 2009, 42(14),

4963–4971.

52. English, A. D. J. Magn. Reson. 1984, 57 (3) 491–493

53. Van Geet, A. L. Anal. Chem. 1968, 40, 2227-2229.

54. Schilli, C.; Lanzendorfer, M. G.; Muller, A. H. E. Macromolecules 2002, 35, 6819-

6827.

55. Magenau, A.J.D.; Martinez-Castro, N.; Storey, R.F. Macromolecules 2009, 42, 2353–

2359.

56. Discher, D. E.; Ahmed, F. Annu. Rev. Biomed. Eng. 2006, 8, 323-341.

57. Wang, X.; Wu, C. Macromolecules 1999, 32, 4299-4301.

Page 134: Polyisobutylene Chain End Transformations: Block Copolymer

115

58. Zhulina, E. B.; Borisov, O. V.; Pryamitsyn, V. A.; Birshtein, T. M. Macromolecules

1991, 24, 140.

59. Zhang, W.;Zhou, X.; Li, H.; Fang, Y.; Zhang, G. Macromolecules 2005, 38, 909-914.

Page 135: Polyisobutylene Chain End Transformations: Block Copolymer

116

Table 4.1. PIB-Br Precursors

a Mn, GPC determined using 100 % mass recovery dn/dc b Mn, GPC determined using known dn/dc

Sample Mn, GPCa Mn, GPCb Mn, NMR Mw/Mn

(g/mol) (g/mol) (g/mol)PIB44-Br 2,600 2,600 2,748 1.03

PIB71-Br 4,200 4,000 4,178 1.02

PIB95-Br 5,600 5,500 5,758 1.06

Page 136: Polyisobutylene Chain End Transformations: Block Copolymer

Table 4.2. RAFT Polymerization Dataa RAFT Polymerization Conditions PNIPAM Block PIB-b-PNIPAM

Sample n,PIB-CTAM b

(g/mol)

[NIPAM]:[CTA]:[I] Time

(h)

pNIPAM n,TheoX c

n,GPCX d n,TheoM e

(g/mol)

n,GPCM f

(g/mol)

PDI

R-1 PIB44-b-PNIPAM500 2,990 487:1:0.5 4.00 0.766 373 328 45,199 40,100 1.08 R-2 PIB44-b-PNIPAM188 2,990 492:1:0.5 1.33 0.376 185 193 23,925 24,800 1.04 R-3 PIB44-b-PNIPAM300 2,990 309:1:0.3 4.00 0.760 235 210 29,583 26,700 1.02 R-4 PIB71-b-PNIPAM300 4,535 292:1:0.3 4.00 0.795 232 197 30,788 26,800 1.03 R-5 PIB95-b-PNIPAM300 5,959 288:1:0.3 4.00 0.813 234 236 32,438 32,700 1.05 R-6 PIB44-b-PNIPAM450 2,990 485:1:0.5 8.00 0.887 431 450 51,762 53,200 1.09

a Stirred glass reactors; temperature (external bath) = 85°C, internal reactor temperature = 78-79°C; [NIPAM] = 1.2 M b n,PIB-CTAM = n,PIB-BrM (100 % mass recovery dn/dc) + (MWN3 - MWBr) + MWCTA

c n,theoX = pNIPAM×[NIPAM]/[CTA] d n,GPCX = ( n,GPCM - n,PIB-CTAM )/MWNIPAM e n,TheoM = n,PIB-CTAM + n,theoX × MWNIPAM f n,GPCM determined using 100 % mass recovery dn/dc

117

Page 137: Polyisobutylene Chain End Transformations: Block Copolymer

118

+TiCl4

N BrnCl

nN Br

nN Br NaN3

S

S

S C12H25

O

O

S

S

S C12H25

O

O

NN

NNn

nN N3

Ph3PCuBrTHF

PIB-N3

Figure 4.1. Synthesis of PIB-CTA for RAFT polymerization.

Page 138: Polyisobutylene Chain End Transformations: Block Copolymer

119

Figure 4.2. Real-time 1H NMR analysis of click reaction between CTA and PIB44-N3.

0

20

40

60

80

100

0 10 20 30 40

conv

ersi

on (

%)

time (h)

0.00

0.50

1.00

1.50

2.00

2.50

3.00

3.50

4.00

4.50

5.00

0 2 4 6 8 10 12

Ln(

(M-p

)/(M

(1-p

)))

[PIB-N3]0(M-1)t

“CLICK”

Page 139: Polyisobutylene Chain End Transformations: Block Copolymer

120

Figure 4.3. 1H NMR spectra of (top) PIB-N3, (middle) CTA, and (bottom) PIB-CTA.

8 7 6 5 4 3 2 1 0ppm

O

O

S

S

S C12H25

PIB

NN

NN CTA

nN N3

a

b

d e

a bcd

e

d’ e’

d’

e’

c

a’ c’

a’ b’c’

b’

Page 140: Polyisobutylene Chain End Transformations: Block Copolymer

121

Figure 4.4. FTIR spectra of (top) PIB-N3 and (bottom) PIB-CTA.

PIB- N3

PIB- CTA

-N3

C=O

C=SC-O

3500 3000 2500 2000 1500 1000 500

wavenumber (cm-1)

Page 141: Polyisobutylene Chain End Transformations: Block Copolymer

122

Figure 4.5 GPC trace before and after click reaction.

6 8 10 12 14 16 18time (m)

PIB-CTA PIB-N3

time (min)

Page 142: Polyisobutylene Chain End Transformations: Block Copolymer

123

[NIPAM]:[CTA]:[I]250 : 1 : 0.25

83/17 (v/v)dioxane/heptane

C12H25N N

N NO

OS

OHN

S

S

mPIB

N N

N NO

O

S

S

S C12H25PIB

PIB-CTA

PIB-b-PNIPAM

Figure 4.6. RAFT polymerization of NIPAM with PIB-CTA.

Page 143: Polyisobutylene Chain End Transformations: Block Copolymer

124

Figure 4.7. 1H NMR spectrum of PIB-b-PNIPAM.

8 7 6 5 4 3 2 1 0

2.0 1.5 1.0 0.5

ppm

C12H25N N

N NO

OS

OHN

S

S

mn

a

b

a

c d

e

g f

b

g

e

f

c d

Page 144: Polyisobutylene Chain End Transformations: Block Copolymer

125

Figure 4.8. GPC traces before and after RAFT polymerization of PNIPAM.

11 12 13 14 15 16 17 18

time (m)

PIB-b-PNIPAM PIB-CTA

Crude

Purified

time (min)

Page 145: Polyisobutylene Chain End Transformations: Block Copolymer

126

Figure 4.9. Conversion vs. time and first-order kinetic plot (inset) for RAFT polymerization from PIB44-CTA as a function of temperature. [NIPAM] = 1.2 M; [NIPAM]:[CTA]:[I] = 250:1:0.25.

0

10

20

30

40

50

60

70

80

90

100

0 20 40 60 80 100 120 140 160 180

conv

ersi

on (%

)

time (min)

90.0 °C

85.0 °C

82.5 °C

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0 10 20 30 40 50 60 70

Ln(

[M0]

/[M

])

Page 146: Polyisobutylene Chain End Transformations: Block Copolymer

127

Figure 4.10. Ln kapp for RAFT polymerization vs. reciprocal temperature to determine activation energy and prefactor (min-1).

y = -15589x + 40.14

-5.00

-4.50

-4.00

-3.50

-3.00

-2.50

-2.00

-1.50

-1.00

2.75E-03 2.76E-03 2.77E-03 2.78E-03 2.79E-03 2.80E-03 2.81E-03 2.82E-03

Ln

(ka

pp)

1/T (K-1)

Page 147: Polyisobutylene Chain End Transformations: Block Copolymer

128

Figure 4.11. Conversion vs. time and first-order kinetic plot (inset) for RAFT polymerization from PIB44-CTA as a function of thermal initiator concentration. [NIPAM] = 1.2 M; [NIPAM]:[CTA] = 250:1; T = 90 °C

0

10

20

30

40

50

60

70

80

90

100

0 20 40 60 80 100 120 140 160 180

conv

ersi

on (%

)

time (min)

[CTA]:[I] = 0.250[CTA]:[I] = 0.167[CTA]:[I] = 0.125

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0 10 20 30 40 50 60 70

Ln(

[M0]

/[M

])

Page 148: Polyisobutylene Chain End Transformations: Block Copolymer

129

Figure 4.12. Conversion vs. time and first-order kinetic plot (inset) for RAFT polymerization as a function of PIB-CTA molecular weight. [NIPAM] = 1.2 M; [NIPAM]:[CTA]:[I] = 250:1:0.25; T = 85 °C.

0

10

20

30

40

50

60

70

80

90

100

0 20 40 60 80 100 120 140 160 180

con

ve

rsio

n (%

)

time (m)

Mn = 3,000

Mn = 4,500

Mn = 6,000

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0 10 20 30 40 50 60 70

Ln([

M0]/

[M])

Page 149: Polyisobutylene Chain End Transformations: Block Copolymer

130

Figure 4.13. Conversion vs. time and first-order kinetic plot (inset) for RAFT polymerization as a function of PIB-CTA concentration. [NIPAM] = 1.2 M; [NIPAM]:[I]=1:0.001; T = 85 °C.

0

10

20

30

40

50

60

70

80

90

100

0 20 40 60 80 100 120 140 160 180

con

ve

rsio

n (%

)

time (m)

[CTA] = 2.4 x 10-3 M

[CTA] = 3.2 x 10-3 M

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0 10 20 30 40 50 60 70

Ln([

M0]/

[M])

Page 150: Polyisobutylene Chain End Transformations: Block Copolymer

131

Figure 4.14. Diameter of Micelles with Temperature (Θ = 90°)

60

70

80

90

100

110

120

130

14 16 18 20 22 24 26 28 30 32 34

Dh

(nm

)

temperature (°C)

PIB44-b-PNIPAM450

T

T

0.0E+00

5.0E+02

1.0E+03

1.5E+03

2.0E+03

2.5E+03

3.0E+03

3.5E+03

4.0E+03

4.5E+03

0.0E+00 2.0E+14 4.0E+14 6.0E+14

Г(s

-1)

q2

T = 15 °C

T = 32 °C

Page 151: Polyisobutylene Chain End Transformations: Block Copolymer

132

CHAPTER V

FUNCTIONAL POLYISOBUTYLENES VIA A CLICK CHEMISTRY APPROACH

Introduction

Polyisobutylene (PIB) is recognized for its environmental stability, gas-barrier

and mechanical damping characteristics, and excellent biocompatibility, and thus PIB

chain segments carrying functional end groups are useful intermediates toward a variety

of products including fuel and lubricating oil additives,1 soluble catalyst supports,2

polyurethane/urea thermoplastic elastomers,3,4,5 and biomedical devices.6,7 Often though,

placement of the desired functionality onto the PIB chain end has required cumbersome

and expensive post-polymerization chemistries, ultimately making it a less feasible

option in some research and commercial applications.

With the recent developments of in-situ quenching (a.k.a. end-capping)

techniques8,9,10,11 and highly effective Sharpless-style “Click” chemistries,12 simpler

chain end reactions and new functionalities on PIB are now possible. In particular, the

copper(I)-catalyzed Huisgen 1,3-dipolar cycloaddition of azides and terminal alkynes has

proven to be a highly reliable, selective, and quantitative reaction.13,14 It is tolerant of a

wide variety of reactions conditions, proceeds at ambient temperature, is insensitive to

both H2O and O2, and can easily be performed in the presence of many types of

functional groups without side reactions.13 Click chemistry has shown great utility

toward the syntheses of triazole-based monomers,15,16 stimuli responsive polymers,17,18

step growth polymers,18 miktoarm star polymers,19 block copolymers,20,21 dendrimers,22

and polymer-modified substrates.23,24

Page 152: Polyisobutylene Chain End Transformations: Block Copolymer

133

Despite the above-mentioned efforts, limited research has been conducted on the

unique combination of PIB and Click chemistry. Binder et al. functionalized trivalent

PIB with hydrogen bonding moieties through Click chemistry to generate novel

supramolecular gels.25,26 Bergbreiter et al. reported the synthesis of PIB-supported Cu(I)

catalyst complexes by reacting azide-terminated PIBs with acetylene functionalized

ligands;27 this system exploited the inherent selective solubility of the PIB support for

facile catalyst removal. Storey et al.28 reacted azide-terminated PIBs with an alkyne-

functional chain transfer agent for reversible addition-fragmentation chain transfer

(RAFT) polymerization as a site transformation approach to novel PIB-based block

copolymers. With the exception of these reported examples, to our knowledge, no other

investigations into this area of research have been conducted.

Preparation of PIB chain ends for the azido/alkyne Click reaction has become

facile due to the ready availability of primary halide-terminated PIB via in situ quenching

techniques.8,9,10 In this report we show that 1-(ω-haloalkyl)pyrroles9 can be used to place

primary bromo- or chloroalkyl functionalities onto quasiliving PIB, followed by

displacement of halogen by azide ion, or alternatively, 1-(ω-azidoalkyl)pyrrole may be

used as quencher to directly yield the terminal azide (Figure 5.1). This process produces

a “Clickable” PIB intermediate that may be reacted with functional alkynes to tether a

variety of useful functional groups to the PIB chain ends through a common 1,4-

disubstituted-1,2,3-triazole linkage. We have demonstrated this modular process by

producing PIBs with quantitative end-group functionalities including hydroxyl, acrylate,

glycidyl, and tertiary amine.

Page 153: Polyisobutylene Chain End Transformations: Block Copolymer

134

Experimental

Materials

Hexane (anhydrous, 95%), TiCl4 (99.9%, packaged under N2 in sure-seal bottles),

2,6-lutidine (26Lut) (redistilled, 99.5%), N,N-dimethylformamide (anhydrous, 99.8%),

heptane (anhydrous, 99%), tetrahydrofuran (anhydrous, 99.9%), propargyl alcohol

(≥99%), sodium azide (≥99.5%), bromotris(triphenylphosphine)copper(l) (98%),

propargyl acrylate (99%), glycidyl propargyl ether (≥90%), 3-dimethylamino-1-propyne

(97%), N,N,N',N'',N''-pentamethyldiethylenetriamine (PMDETA, 99%), copper(I)

bromide, CuBr, (99.999%) and chloroform-d (0.01% H2O maximum) were purchased

from Sigma-Aldrich Co. and used as received. 1-(2-Bromoethyl)pyrrole (PyBr) and 1-(2-

chloroethyl)pyrrole (PyCl) were purchased from TCI and distilled from calcium hydride.

Isobutylene (IB) and CH3Cl (both BOC, 99.5%) were dried through columns packed with

CaSO4 and CaSO4/4 Å molecular sieves, respectively. 2-Chloro-2,4,4-trimethylpentane

(TMPCl)29 and 5-tert-butyl-1,3-di(1-chloro-1-methylethyl)benzene (t-Bu-m-DCC)30 were

synthesized according to the literature. 1-(3-Bromopropyl)pyrrole (PyBrP) was

synthesized by N-alkylation of pyrrolyl sodium salt with 1,3-dibromopropane in DMSO

according to the literature31 and purified by fractional distillation.

Polyisobutylene precursors. 1-(ω-Haloalkyl)pyrrolyl-terminated PIBs were

synthesized as previously described.9 tert-Chloride-terminated masterbatch PIBs,

difunctional ( nM =2,100 g/mol) and monofunctional ( nM =1,970 g/mol), were prepared

via the BCl3-catalyzed polymerization of isobutylene from t-Bu-m-DCC and TMPCl,

respectively, in methyl chloride32 at -60oC.

Page 154: Polyisobutylene Chain End Transformations: Block Copolymer

135

1-(ω-Azidoalkyl)pyrrole quenchers. Synthesis of 1-(2-azidoethyl)pyrrole (PyAz)

was carried out under a dry nitrogen atmosphere in a round bottom flask according to the

following representative procedure. To a solution of 1-(2-bromoethyl)pyrrole (3.57 mL,

5.0 g, 28.7 mmol) in N,N-dimethylformamide (3 mL) was added sodium azide (3.0 g, 46

mmol). The mixture was reacted for 12 h at 90 °C, with stirring. It was allowed to cool

to room temperature and excess sodium azide was removed by filtration. The filtrate was

dissolved in CH2Cl2, washed with water, dried over MgSO4, and concentrated on a rotary

evaporator. The crude product was finally purified by vacuum distillation. 1H NMR

(CDCl3): δ= 3.59 (t, 2H, CH2N3), 4.06 (t, 2H, NCH2), 6.24 (d, 2H, 3-H), 6.72 (d, 3H, 2-

H) ppm. 13C NMR (CDCl3): δ= 48.51 (NCH2), 52.16 (CH2N3), 108.88 (C3), 120.36 (C2)

ppm.

The synthesis of 1-(3-azidopropyl) pyrrole (PyPAz) was analogous to that of 1-(2-

azidoethyl)pyrrole. The crude product was purified by distillation. 1H NMR (CDCl3): δ=

2.02 (m, 2H, CH2), 3.28 (t, 2H, CH2N3), 4.01 (t, 2H, NCH2), 6.19 (d, 2H, 3-H), 6.67 (d,

2H, 2-H) ppm. 13C NMR (CDCl3): δ= 30.58 (CH2), 46.04 (NCH2), 48.09 (CH2N3),

107.35 (C3), 120.09 (C2) ppm.

Instrumentation

NMR spectra were acquired using a Varian Mercuryplus 300 MHz NMR

spectrometer. Samples were prepared by dissolving the sample in chloroform-d (5-7 %,

w/v) and charging this solution to a 5 mm NMR tube.

Molecular weights and polydispersities (PDI) of the polymeric materials were

measured using Size Exclusion Chromatography (SEC) (Waters Alliance 2695

separations module and two mixed E Polymer Laboratories Inc. columns) with on-line

Page 155: Polyisobutylene Chain End Transformations: Block Copolymer

136

multi-angle laser light scattering detection (Wyatt Technology Inc. MiniDAWN), as

described previously.9

ATR-FTIR spectroscopic monitoring, using a ReactIR 1000 reaction analysis

system (ASI Applied Systems, Millersville, MD), previously described,33,34 was

integrated with an inert atmosphere glovebox (MBraun Labmaster 30) to obtain real-time

FTIR spectra and temperature profiles of the isobutylene polymerizations. Reaction

conversion was determined by monitoring the area, above a two-point baseline, of the

absorbance centered at 887 cm-1, associated with the =CH2 wag of IB.

Synthesis of 1-(ω-Azidoalkyl)pyrrolyl-PIB by Nucleophilic Substitution

Functionalizations of monofunctional PIB with sodium azide were carried out

under a dry nitrogen atmosphere in a flask according to the following representative

procedure (Table 5.1, Entry 5). 1-(2-Bromoethyl)pyrrolyl-PIB (2.0 g, 1.00 mmol) was

dissolved in 5 mL of dry heptane in a flask, and then sodium azide (0.196 g, 3.02 mmol)

in 5 mL of DMF was added. The resulting biphasic mixture was stirred and heated to 90

°C, upon which it formed a single, homogeneous solution, and the reaction was

conducted at 90°C for 24 h. Finally, the monophasic reaction mixture was allowed to

cool, whereupon a biphasic mixture re-formed, and the heptane and DMF layers were

separated. The heptane phase was washed with methanol, and subsequently, the polymer

was precipitated two times into methanol. The precipitate was collected by dissolution in

hexane; the solution was concentrated on a rotary evaporator, and the polymer was finally

vacuum dried at room temperature.

Page 156: Polyisobutylene Chain End Transformations: Block Copolymer

137

Synthesis of Difunctional 1-(ω-Azidoalkyl)pyrrolyl-PIB using Masterbatch PIB

Synthesis of difunctional 1-(2-azidoethyl)pyrrolyl-PIB was carried out under a

dry nitrogen atmosphere in a glove box according to the following representative

procedure (Table 5.2, Entry 1). Large culture tubes equipped with Teflon-lined caps

were used as reactors. Into a test tube containing 0.53 g of masterbatch difunctional PIB

(Mn= 2,100 g/mol, 0.25 mmol) were added 10 mL of CH3Cl, 15 mL of hexane, and

0.009 mL (0.008 g, 0.07 mmol) of 2,6-lutidine. The polymer was allowed to dissolve,

and the solution was equilibrated to -70 °C with agitation. Then, 0.55 mL (0.95 g, 5.0

mmol) of TiCl4 was transferred to the reactor, followed by a pre-chilled solution

containing 0.124 g (0.911 mmol) 1-(2-azidoethyl)pyrrole in 25 mL of hexane/CH3Cl

(60/40, v/v, -70°C). The color of the reaction mixture changed from slightly yellow to

brown. PyAz was allowed to react with the living chain ends for 10 min. Finally, the

reaction was quenched by addition of prechilled methanol. The reactor contents were

allowed to warm to room temperature, and the polymer in hexane was washed with

methanol and then precipitated one time into methanol from hexane. The precipitated

polymer was collected by dissolution in hexane; the solution was concentrated on a rotary

evaporator, and the polymer was finally vacuum dried at room temperature.

Isobutylene Polymerization and N-(ω-Azidoalkyl)pyrrole Quenching

Quasiliving polymerizations of IB with either TMPCl or t-Bu-m-DCC as initiator

were carried out within a N2 atmosphere glovebox, equipped with an integral, cryostated

hexane/heptane bath according to the following representative procedure (Table 5.3,

Entry 3). Into a round-bottom flask equipped with a mechanical stirrer, infrared probe,

and thermocouple were added 72 mL of CH3Cl, 108 mL of hexane, 0.07 mL (0.06 g, 0.6

Page 157: Polyisobutylene Chain End Transformations: Block Copolymer

138

mmol) of 2,6-lutidine, 5.4 mL (3.8 g, 67 mmol) of IB, and 0.323 g (1.12 mmol) of t-Bu-

m-DCC. The mixture was allowed to equilibrate to -70°C, and then, 0.99 mL (1.7 g, 9.0

mmol) of TiCl4 was charged to the reactor. The reaction was allowed to proceed for 37

min, and then a pre-chilled solution of 1-(2-azidoethyl)pyrrole (PyAz) prepared by

dissolving 0.551 g PyAz (4.05 mmol) into 25 mL of hexane/CH3Cl (60/40, v/v, -70C),

was added, followed by an additional 1.48 mL (2.56 g, 13.5 mmol) TiCl4. The color of

the solution changed from slightly yellow to brown. PyAz was allowed to react with the

living chain ends for 60 min. Finally, the reaction was quenched by addition of excess

prechilled methanol. The reactor contents were allowed to warm to room temperature,

and the polymer in hexane was washed with methanol and then precipitated one time into

methanol from hexane. The precipitate was collected by dissolution in hexane; the

solution was washed with water, dried over MgSO4, and concentrated on a rotary

evaporator. The polymer was finally vacuum dried at room temperature.

Synthesis of 1-[2-(4-Hydroxymethyl-1H-1,2,3-triazol-1-yl)ethyl]pyrrole

(Table 5.4, Entry 2)

Under a dry nitrogen atmosphere, 1-(2-azidoethyl)pyrrole (1.23 g, 9.03 mmol)

and propargyl alcohol (0.76 mL, 0.73 g, 13 mmol) were mixed in a round-bottom flask.

In a separate vessel, a solution of copper(I) bromide (0.0057 g, 0.040 mmol) and

PMDETA (0.0083 mL, 0.0069 g, 0.040 mmol) in 1 mL THF was prepared. With stirring,

the catalyst solution in THF was added to the azide/alkyne mixture. A very exothermic

reaction ensued. After 5 min, the mixture was diluted with tetrahydrofuran and passed

through a column of aluminum oxide (neutral) using THF as eluent. The resulting

solution of 1-[2-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)ethyl]pyrrole was collected and

Page 158: Polyisobutylene Chain End Transformations: Block Copolymer

139

concentrated on a rotary evaporator, and the crude product was purified by distillation.

1H NMR (CDCl3): δ= 3.19 (broad, 1H, OH), 4.35 (t, 2H, 1-CH2-triazole), 4.64 (t, 2H,

CH2-pyrrole), 4.69 (s, 2H, 4-CH2-triazole), 6.13 (d, 2H, 3-H-pyrrole), 6.45 (d, 2H, 2-H-

pyrrole), 6.93 (s, 1H, 5-H-triazole) ppm. 13C NMR (CDCl3): δ= 49.24 (CH2-pyrrole),

51.47 (1-CH2-triazole), 56.18 (4-CH2-triazole), 109.33 (C3-pyrrole), 120.09 (C2-pyrrole),

123.47 (C5-triazole), 147.64 (C4-triazole) ppm.

Synthesis of 1-[3-(4-Hydroxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole

(Table 5.4, Entry 3)

Synthesis of 1-[3-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole was

analogous to the synthesis of 1-[2-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)ethyl]pyrrole.

1H NMR (CDCl3): δ= 2.37 (m, 2H, CH2), 3.36 (broad, 1H, OH), 3.92 (t, 2H, 1-CH2-

triazole), 4.25 (t, 2H, CH2-pyrrole), 4.77 (s, 2H, 4-CH2-triazole), 6.16 (d, 2H, 3-H-

pyrrole), 6.62 (d, 2H, 2-H-pyrrole), 7.47 (s, 1H, 5-H-triazole) ppm. 13C NMR (CDCl3):

δ= 31.56 (CH2), 46.04 (CH2-pyrrole), 47.20 (1-CH2-triazole), 56.27 (4-CH2-triazole),

108.71 (C3-pyrrole), 120.44 (C2-pyrrole), 121.87 (C5-triazole), 147.47 (C4-triazole)

ppm.

Synthesis of Monofunctional 1-[3-(4-Hydroxymethyl-1H-1,2,3-triazol-1-

yl)propyl]pyrrolyl-PIB by Click Chemistry

Synthesis of hydroxyl-PIB by Click chemistry was carried out under a dry

nitrogen atmosphere in a flask according to the following procedure (Table 5.5, Entry 1).

1-(3-Azidopropyl)pyrrolyl-PIB ( nM = 1,380 g/mol, 0.537 g, 0.389 mmol) was dissolved

in 5 mL of dry THF in a flask, and then 0.061 mL of propargyl alcohol (0.059 g, 1.0

mmol) and 0.074 g of bromotris(triphenylphosphine)copper (I) (0.080 mmol ) were

Page 159: Polyisobutylene Chain End Transformations: Block Copolymer

140

added. With stirring, the resulting mixture was heated to 55 °C, and the reaction was

allowed to proceed for 24 h. Finally, the reaction mixture was allowed to cool, and THF

and excess propargyl alcohol were removed at 40 °C using a rotary evaporator.

Subsequently, the polymer was dissolved in hexane and washed with methanol and then

precipitated one time into methanol from hexane. The precipitate was collected by

dissolution in hexane; the solution was concentrated on a rotary evaporator, and the

polymer was finally vacuum dried at room temperature.

Synthesis of Monofunctional 1-[3-(4-Vinylcarbonyloxymethyl-1H-1,2,3-triazol-1-

yl)propyl]pyrrolyl–PIB by Click Chemistry

Synthesis of acrylate-terminated PIB by Click chemistry was carried out under a

dry nitrogen atmosphere in a flask according to the following procedure (Table 5.5, Entry

2). 1-(3-Azidopropyl)pyrrolyl-PIB ( nM = 2,950 g/mol, 1.4 g, 0.47 mmol) was dissolved

in 3 mL of dry THF in a flask, and then 0.105 mL of propargyl acrylate (0.105 g, 0.950

mmol) and 0.074 g of bromotris(triphenylphosphine)copper (I) (0.080 mmol ) were

added. The resulting mixture was stirred and allowed to react for 24 h at 25 °C. The

reaction mixture was then allowed to cool, and THF and excess propargyl acrylate were

removed at 40 °C using a rotary evaporator. Subsequently, the polymer was dissolved in

hexane, filtrated and then precipitated one time into methanol from hexane. The

precipitate was collected by dissolution in hexane; the solution was concentrated on a

rotary evaporator, and the polymer was finally vacuum dried at room temperature.

Page 160: Polyisobutylene Chain End Transformations: Block Copolymer

141

Synthesis of Monofunctional 1-[3-(4-Glycidyloxymethyl-1H-1,2,3-triazol-1-

yl)propyl]pyrrolyl-PIB by Click Chemistry

Synthesis of glycidyl-terminated-PIB by Click chemistry was carried out under a

dry nitrogen atmosphere in a flask according to the following procedure (Table 5.5, Entry

3). 1-(3-Azidopropyl)pyrrolyl-PIB ( nM = 2,950 g/mol, 1.0 g, 0.34 mmol) was dissolved

in 2.5 mL of dry THF in a flask, and then 0.073 mL of glycidyl propargyl ether (0.076 g,

0.68 mmol) and 0.074 g of bromotris(triphenylphosphine)copper (I) (0.080 mmol ) were

added. The resulting mixture was stirred and allowed to react for 24 h at 25 °C. The

reaction mixture was then allowed to cool, and THF and excess glycidyl propargyl ether

were removed at 40 °C using a rotary evaporator. Subsequently, the polymer was

dissolved in hexane, filtrated and then precipitated one time into methanol from hexane.

The precipitate was collected by dissolution in hexane; the solution was concentrated on

a rotary evaporator, and the polymer was finally vacuum dried at room temperature.

Synthesis of Monofunctional 1-[3-(4-Dimethylaminomethyl-1H-1,2,3-triazol-1-

yl)propyl]pyrrolyl–PIB by Click Chemistry

Synthesis of dimethylamino-terminated PIB by Click chemistry was carried out

under a dry nitrogen atmosphere in a flask according to the following procedure (Table

5.5, Entry 4). 1-(3-Azidopropyl)pyrrole-PIB ( nM = 2,950 g/mol, 1.0 g, 0.34 mmol) was

dissolved in 2.5 mL of dry THF in a flask, and then 0.073 mL of 3-dimethylamino-1-

propyne (0.056 g, 0.68 mmol) and 0.074 g of bromotris(triphenylphosphine)copper (I)

(0.080 mmol ) were added. The resulting mixture was stirred and allowed to react for 24

h at 25 °C. The reaction mixture was then allowed to cool, and THF and excess 3-

dimethylamino-1-propyne were removed at 40 °C using a rotary evaporator.

Page 161: Polyisobutylene Chain End Transformations: Block Copolymer

142

Subsequently, the polymer was dissolved in hexane, filtrated and then precipitated one

time into methanol from hexane. The precipitate was collected by dissolution in hexane;

the solution was concentrated on a rotary evaporator, and the polymer was finally

vacuum dried at room temperature.

Results and Discussion

Synthesis of N-(ω-Azidoalkyl)pyrrolyl-PIB by Nucleophilic Substitution

We recently reported9 that quasiliving cationic PIB reacts quantitatively with 1-

(ω-haloalkyl)pyrroles to yield an isomeric mixture of 2- and 3-PIB-1-(ω-

haloalkyl)pyrroles, with no detectable di-substitution (coupled) products. The resulting

primary halide end groups are useful intermediates toward many alternative end groups

including azides.

To obtain the desired 1-(ω-azidoalkyl)pyrrolyl-PIB, two synthetic strategies were

attempted, as shown in Figure 5.1. The first approach involved substitution of the

terminal halogen of 1-(ω-haloalkyl)pyrrolyl-PIB with sodium azide (Figure 5.1, Route

A); whereas the second involved in-situ functionalization with 1-(2-azidoethyl)pyrrole

(Route B). The 1-(ω-haloalkyl)pyrrole-PIBs used for the first approach were synthesized

as previously reported.9

Direct conversion of the terminal halide to the corresponding azide was achieved

using sodium azide,35 which is readily available, inexpensive, and soluble in polar

solvents such as N,N-dimethylformamide. However, the hydrophobic nature of PIB

necessitated a cosolvent system consisting of a 1:1 (v/v) mixture of heptane and N,N-

dimethylformamide. This cosolvent mixture was optimal for dissolution of both PIB and

NaN3 at elevated temperatures, and it offered the useful process advantage of forming

Page 162: Polyisobutylene Chain End Transformations: Block Copolymer

143

two immiscible phases at room temperature. Upon cooling the reaction, this allowed

facile separation of the PIB-azide product, contained in the upper heptane-rich layer,

from excess NaN3 and Na halide byproduct, which resided in the lower DMF-rich layer.

Initial experimentation showed that for chain end concentrations [CE] in the range

0.06 to 0.12 mol/L, and for [azide]/[CE] = 3, reaction temperatures less than 90°C

resulted in unreasonably long reaction times (in excess of 24 h) for either chloride or

bromide substrates. For example, Table 5.1, Entries 1 and 2 show that for chloride at

70°C, no greater than 40% conversion was reached in 24 h depending upon [CE]. At

90°C, however, complete reaction was observed within 24 h, for either chloride or

bromide regardless of tether length (2 or 3 carbons) over a range of chain end

concentrations as demonstrated in Table 5.1, Entries 3-9.

Figure 5.2 shows the 1H NMR spectrum with peak assignments of 1-(2-

azidoethyl)pyrrolyl-PIB from 1-(2-bromoethyl)pyrrolyl-PIB (Table 1, Entry 5, which is

representative). Quantitative reaction was indicated by the complete absence of peaks

due to the tether unit of the bromide precursor, which would have been visible as triplets

centered at 3.53, 3.55, 4.18, and 4.31 ppm (see inset, Figure 5.2). To confirm

quantitative reaction, difunctional 1-(3-azidopropyl)pyrrolyl-PIB was synthesized (Table

5.1, Entry 8) and characterized by 1H NMR (Figure 5.3). Integration of the resonances

due to the terminal 1-(3-azidopropyl)pyrrole moieties in comparison with the aromatic

initiator resonance at 7.17 ppm indicated near-quantitative capping and production of

difunctional, telechelic primary azide PIB. Difunctional 1-(2-azidoethyl)pyrrolyl-PIB

was also synthesized and quantitative functionalization was confirmed by 1H NMR

analysis (Figure 5.4). Complete 1H and 13C NMR chemical shift assignments for

Page 163: Polyisobutylene Chain End Transformations: Block Copolymer

144

monofunctional 2- and 3-PIB-1-(2-azidoethyl)pyrrole are listed in Tables 5.6 and 5.7,

respectively. The same data for the 2- and 3-PIB-1-(3-azidopropyl)pyrroles are listed in

Tables 5.8 and 5.9, respectively.

Examination of the products obtained in low conversion reactions such as Entries

1 and 2 (Table 5.1) indicated that the isomeric 2- and 3-PIB-1-(ω-haloalkyl)pyrroles

might possess different reactivities toward the SN2 reaction with azide. Thus, the kinetics

of azide substitution at 90°C, with [azide]/[CE] = 3, were examined in detail using 1H

NMR spectroscopy. Figure 5.2 provides a comparison of the methylene tether

resonances of the bromide reactant (inset) and azide product, and thus indicates the

characteristic chemical shifts that were observed upon substitution. Reaction progress

was monitored by observing the resonance due to the methylene unit adjacent to the

pyrrole nitrogen. For the C3 isomer, appearance of the azide product was monitored

using the peak at 3.95 ppm; for the C2 isomer, disappearance of the halogen reactant was

observed at 4.27 (Cl) or 4.31 ppm (Br). Figure 5.5 shows a plot of reaction conversion

vs. time for 1-(2-bromoethyl)pyrrolyl-PIB (difunctional), with C3 and C2 isomer

conversions plotted individually. Complete reaction of both isomers was achieved after

about 10 h. As suspected, reaction at the C3 isomer was considerably faster than at the C2

isomer. This is probably a steric effect; since for SN2 reactions, steric effects are

generally more important than electronic effects for substituents located at the β position

or more remotely.36 Figure 5.6 shows a plot of reaction conversion vs. time for 1-(2-

bromopropyl)pyrrolyl-PIB, with C3 and C2 isomer conversions plotted individually.

Reactivites were observed to be opposite of those found with the two carbon tether,

although reactive isomer reactive was more similar.

Page 164: Polyisobutylene Chain End Transformations: Block Copolymer

145

Rate of the azide substitution reaction can be dramatically increased by using a

large excess of the inexpensive NaN3 (pseudo-first-order conditions) and maximizing the

temperature to the reflux temperature of the heptane/DMF cosolvents. In recent

preparative work, we have reduced reaction times for the bromide substrate to about 45

min using a 20-fold excess of azide at reflux temperature. Further increases in azide

concentration produce diminishing returns due to solubility limitations in the cosolvent

system.

Synthesis of 1-(ω-Azidoalkyl)pyrrolyl-PIB by Quenching with 1-(2-Azidoethyl)pyrrole

To avoid the post-polymerization substitution reaction, we next attempted to

quench quasiliving carbocationic PIB directly with a 1-(ω-azidoalkyl)pyrrole (Figure 5.1,

Route B). We expected that an aliphatic primary azide would not decompose TiCl4 and

therefore would not prevent electrophilic aromatic substitution at the pyrrole ring. For

this purpose, 1-(2-azidoethyl)pyrrole was synthesized from the corresponding bromide by

nucleophilic substitution with NaN3 in DMP at 90°C for 12 h. Reaction progress was

monitored using 1H NMR by observing the disappearance of the tether resonances of the

bromide reactant at 4.30 ppm (Py-CH2-CH2-Br) and 3.61 ppm (Py-CH2-CH2-Br) and the

appearance of the corresponding resonances of the product at 4.06 and 3.59 ppm.

To demonstrate its utility as a quencher, 1-(2-azidoethyl)pyrrole was reacted with

difunctional tert-chloride-terminated masterbatch PIB in the presence of TiCl4 under the

conditions listed in Table 5.2. Typically, alkylation of a 1-substituted pyrrole by PIB

cations is sufficiently rapid that ionization of tert-chloride PIB is rate limiting.9,11

However, the rate of ionization is dependent on the effective or free [TiCl4], and the latter

is difficult to specify because of uncertainty with regard to the extent of complexation of

Page 165: Polyisobutylene Chain End Transformations: Block Copolymer

146

TiCl4 by the quencher and, particularly, the alkylated product. For example, Cheradame

et al. reported the synthesis of azide-terminated PIB initiated from 1,4-di(1-azido-1-

methylethyl)benzene/TiCl4 at -70 °C in CH2Cl2 (the products actually possessed mixed

azide, tert-chloride, and olefin end groups).37,38 These authors reported that relatively

high [TiCl4] was necessary for polymerization due to a complexation between TiCl4 and

the azide groups.37 The data in Table 5.2, obtained using a uniform quencher

concentration of 0.018 mol/L, indicate that a nominal [TiCl4] = 0.10 mol/L was sufficient

to achieve quantitative alkylation in under 3 min (Exp. 1, 4, and 5); whereas nominal

[TiCl4] = 0.02 (Exp.2) or 0.03 mol/L (Exp. 4) was insufficient even for a reaction time of

15 min.

The 1H NMR spectrum of the product of Table 5.2, Entry 1, acquired after a

quenching time of 10 min is shown in Figure 5.7. Resonances at 1.68 and 1.96 ppm due

to residual tert-chloride end groups were not observed, indicating complete conversion.

A new set of peaks at 1.67, 3.52, 3.95, 6.07, 6.40 and 6.57 ppm (C3-isomer, major) and

1.72, 3.64, 4.13, 5.90, 6.10 and 6.60 ppm (C2-isomer, minor) appeared due to the

presence of N-(2-azidoethyl)pyrrolyl moieties at the chain ends. The integration data

indicated that the combined peak areas (C2 and C3 isomers) of the CH2N3 protons at 3.52

and 3.64, and the NCH2 protons at 3.95 and 4.13 ppm were 1.31 and 1.32 times that of

the aromatic proton resonance at 7.2 ppm, indicating that 98-99% of the chain ends were

functionalized with the desired azide groups; a similar analysis applied to the pyrrole ring

protons yielded a slightly lower estimate of about 97%. The balance of the chain ends

consisted of mixed olefins, visible in the spectrum at 4.64 and 4.85 (exo) and 5.05 ppm

(endo). A prequench control aliquot revealed that there were exactly two tert-chloride

Page 166: Polyisobutylene Chain End Transformations: Block Copolymer

147

end groups per aromatic initiator residue and no detectable olefin; therefore, this small

amount of olefin was created during the quenching reaction. Similar quenching reactions

with 1-(ω-haloalkyl)pyrroles did not produce olefin,9 and so we believe that elimination

is induced by competitive nucleophilic interaction of azide groups of the quencher with

the PIB carbocations.

The integration data in Figure 5.7 revealed a C2/C3 isomer ratio of 40/60. This is

consistent with our general experience with 1-substituted pyrrole quenching.9,11,39 Mixed

isomers are invariably observed, with the C3 isomer dominant, but the nature of the 1-

substituent attached to the pyrrole ring, and the tether length, effect the ratio. In previous

reports, we showed that 1-methylpyrrole yielded a nearly balanced C2/C3 isomer ratio of

48/52;39 1-(3-bromopropyl)pyrrole showed a greater tendency toward C3, yielding an

average C2/C3 = 39/61, and 1-(2-haloethyl)pyrroles showed the greatest tendency toward

C3, averaging C2/C3 = 28/72.9

Functionalization was also confirmed by 13C NMR spectroscopy, by observing

the disappearance of the resonances at 71.9 and 35.2 ppm, representing the quaternary

and geminal dimethyl carbons, respectively, adjacent to the terminal tert-chloride group,

and appearance of new peaks in both the aromatic and the aliphatic regions of the

spectrum. The 13C NMR spectrum of 1-(2-azidoethyl)pyrrolyl-PIB, along with peak

assignments, is shown in Figure 5.8.

Figure 5.9 compares GPC traces of 1-(2-azidoethyl)pyrrole-quenched PIB with an

aliquot removed from the reaction prior to quenching. The GPC traces prior to and after

quenching were indistinguishable, indicating the absence of any coupling reactions or

polymer degradation.

Page 167: Polyisobutylene Chain End Transformations: Block Copolymer

148

Entries 1 and 4 (difunctional) and 5 (monofunctional) of Table 5.2 demonstrate

the reproducibility of the quenching reaction. In all three cases, quantitative conversion

of tert-chloride chain ends was obtained, but the resulting polymers all contained 1 to 2

% olefinic chain ends (mixed exo and endo). No difference was observed between

difunctional and monofunctional PIBs; particularly, difunctional samples did not show a

greater tendency toward coupling via dialkylation of a pyrrole ring compared to the

monofunctional sample. Although quenching required less than 3 min under these

conditions, the product of Entry 1 was indistinguishable from the product of Entry 3; thus

no apparent harm was caused by prolonged reaction time after complete conversion.

However, such was not the case for prolonged reaction times at incomplete conversion.

For the two experiments run at lower [TiCl4] (Entries 2 and 4), GPC analysis of the PIB

after quenching showed that coupling products were present.

The experiments of Table 5.2, which involved pre-formed, tert-chloride-

terminated PIB, showed that quenching with a 1-(ω-azidoalkyl)pyrrole requires relatively

high [TiCl4]. If used for IB quasiliving polymerization, such high [TiCl4] would produce

inconveniently high rates; therefore, for the case of IB polymerization followed

immediately by quenching, a lower [TiCl4] was used during polymerization, with an

additional charge of TiCl4 added for quenching. Table 5.3 lists several in situ quenching

experiments that illustrate this strategy. Entry 1 was a control experiment in which the

additional charge of TiCl4 was omitted. With [TiCl4] = 0.053 M, this recipe yielded a

convenient polymerization time of about 30 min, but quenching was only 36% complete

after a 40 min quenching reaction, with the balance of the chain ends returned as tert-

chloride. Entry 2 used a polymerization recipe similar to Entry 1, but in this case an

Page 168: Polyisobutylene Chain End Transformations: Block Copolymer

149

additional increment of TiCl4 was added with the quencher to yield a total [TiCl4] = 0.060

during the quench. In this case the extent was quenching was 87% after a quenching time

of 60 min. However, an aliquot removed from the reaction after 2 min indicated that

65% alkylation had already occurred, suggesting that the initial rate of reaction was

sufficiently fast but that it soon stalled, presumably due to loss of catalyst. In Entry 3, the

additional TiCl4 increment was increased to yield a total [TiCl4] = 0.10 during the

quench. This resulted in quantitative functionalization within 2 min.

The structure of the polymer obtained from Table 5.3, Entry 3 above was

essentially identical in all respects to that of the polymers obtained from the pre-formed

PIB of Table 2. About 97-99% of the chains carried the desired 1-(2-azidoethyl)pyrrole

functions and the remainder of the chains were mixed olefin. It is noteworthy to mention

that Entries 1 and 2, Table 3, produced polymers with visible coupling in the GPC

chromatogram. This is consistent with the results for quenching of pre-formed PIB

(Table 5.2) and suggests that coupling is generally observed when functionalization is

incomplete.

Functionalization of 1-(ω-Azidoalkyl)pyrrolyl-PIB by Click Chemistry

An important advantage of 1-(ω-azidoalkyl)pyrrolyl-PIB is its ability to react with

alkynes via Click chemistry, i.e., the copper(I)-catalyzed Huisgen 1,3-dipolar

cycloaddition reaction. This reaction is a versatile tool for macromolecular engineering

and allows the fabrication of complex structures with various functionalities. In the

present context, it provides a modular approach to various telechelic PIBs.

To define appropriate conditions for Click chemistry at the PIB chain end,

including the preferred catalyst, 1-(2-azidoethyl)pyrrole and 1-(3-azidopropyl)pyrrole,

Page 169: Polyisobutylene Chain End Transformations: Block Copolymer

150

were used to mimic the end group of 1-(ω-azidoalkyl)pyrrolyl–PIB in model reactions

with a representative alkyne, propargyl alcohol (Table 5.4). Two copper(I)-based

catalysts were examined, bromotris(triphenylphosphine)copper(I) and

Cu(I)Br/PMDETA. Both catalysts promoted high regioselectivity to yield exclusively

the 4-hydroxymethyltriazole products. For example, Figure 5.10 shows the 1H NMR

spectrum of 1-[2-(4-hydroxymethyl-1H-1,2,3-triazolyl)ethyl]pyrrole obtained from

reaction of 1-(2-azidoethyl)pyrrole with propargyl alcohol using copper(I)

bromide/PMDETA as catalyst (Table 5.4, Entry 2). Conversion exclusively to the 4-

substituted regio isomer was indicated by a single resonance for the C5 triazole proton at

6.93 ppm. The methylene and hydroxyl protons of the propargyl alcohol residue were

observed at 4.69 and 3.19 ppm, respectively, and the protons of the ethylene tether were

observed at 4.35 and 4.64 ppm. The length of the alkylene tether appeared to make little

difference in the reaction; both azide model compounds behaved similarly with regard to

reactivity toward the alkyne and structure of the resulting triazole (Table 5.4, Entries 2

and 3).

Copper(I) bromide/PMDETA produced very rapid, highly exothermic Click

reactions. Complete conversion was obtained within 5 min for the two model reactions

in Table 5.4 (Entries 2 and 3). However, Cu(I)Br/PMDETA is very soluble in organic

solvents, including non-polar solvents such as hexane, which are good solvents for PIB.

Thus we found it necessary to pass the product through an alumina column to remove the

catalyst. Compared to Cu(I)Br/PMDETA, bromotris(triphenylphosphine)copper(I)

produced much slower reactions; full conversion of 1-(2-azidoethyl)pyrrole to the triazole

required 12 h (Table 5.4, Entry 1). However, although

Page 170: Polyisobutylene Chain End Transformations: Block Copolymer

151

bromotris(triphenylphosphine)copper(I) is soluble in most organic solvents, its solubility

in hexane is limited. This suggested an attractive strategy for Click reactions at the PIB

chain end, involving bromotris(triphenylphosphine)copper(I)-catalyzed reaction in THF,

followed by dissolution of the product in hexane and removal of the catalyst by simple

filtration and precipitation.

After obtaining positive Click chemistry results with the model compounds, we

next explored the reaction of propargyl alcohol with azide-terminated PIB. Kinetics were

measured using real-time 1H NMR analysis by observing the decrease in area of the

propargyl methylene protons signal of propargyl alcohol. Figure 5.11 shows plots of

reaction conversion vs. time, at 25°C in THF, for an initial molar ratio of alkyne/azide of

2.0. The inset plots show the data fitted to an integrated second-order rate equation. The

upper plot (A) shows reaction of 1-(3-azidopropyl)pyrrolyl-PIB (3-carbon tether) using

1/1 CuBr(I)/PMDETA; the lower plot (B) (solid circles) shows reaction of the same

polymer in the presence of bromotris(triphenylphosphine)copper(I). In plot (A), as

expected, reaction rate for CuBr(I)/PMDETA was very high, requiring less than 20 min

for complete reaction, and on the order of 30 times higher than for

bromotris(triphenylphosphine)copper(I). Poor temperature control due to exotherm of

the CuBr(I)/PMDETA-catalyzed reaction was suggested by the upward curvature in the

second-order plot. The second-order rate constant calculated from the initial, linear

portion of the curve (first 5 points) was 22 L/mol-s. In plot (B), an induction period was

observed for bromotris(triphenylphosphine)copper(I) catalysis, and complete reaction

required approximately 16-20 h. The second-order rate constant for the 3-carbon tether

Page 171: Polyisobutylene Chain End Transformations: Block Copolymer

152

(solid circles), using the entire data set, was 0.68 L/mol-s; the 2-carbon tether (open

circles) produced a slightly faster reaction, with a rate constant of 0.91 L/mol-s.

To demonstrate the modularity of the Click chemistry approach toward chain-end

functionalization, PIBs with various terminal groups were prepared as listed in Table 5.5.

Bromotris(triphenylphosphine)copper(I) was used as catalyst in all cases due to the ease

of its separation from the polymer (limited solubility in hexane). Reaction of propargyl

alcohol at the PIB chain end was identical to the model experiments in terms of structure

(Table 5.5, Entry 1). Figure 5.12 shows the 1H NMR spectrum of the resulting polymer

with peak assignments. Functionalization was indicated by the disappearance of the

peaks at 1.95/2.07 ppm (--PIB-Py-CH2-CH2-CH2-N3), 3.22/3.44 ppm (--PIB-Py-CH2-

CH2-CH2-N3), and 3.90/4.04 ppm (--PIB-Py-CH2-CH2-CH2-N3), and appearance of new

peaks at 2.35, 3.85, 4.24, 6.07, 6.40 and 6.57 ppm (3-isomer, major), 2.43, 4.04, 4.46,

5.90, 6.10 and 6.60 ppm (2-isomer, minor) and 4.81 ppm due to presence of the 1-[3-(4-

hydroxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole moieties at the chain ends.

Reaction of azide-terminated PIB with propargyl acrylate lead to the easy

isolation of a PIB acrylate macromer (Table 5.5, Entry 2), potentially useful for

introducing PIB grafts into acrylic copolymers. The reaction was carried out at room

temperature in THF solution. Isolation of the polymer consisted simply of dissolving the

product in hexanes, filtering to remove catalyst, and evaporation of the solvent and

excess propargyl acrylate. To remove all traces of unreacted propargyl acrylate,

optionally the polymer may be precipitated into methanol. The 1H NMR spectrum of the

resulting 1-[3-(4-vinylcarbonyloxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrolyl-PIB is

shown in Figure 5.13. Functionalization was indicated by the disappearance of

Page 172: Polyisobutylene Chain End Transformations: Block Copolymer

153

resonances associated with azide end groups and appearance of a new set of resonances

2.35, 3.85, 4.24, 6.07, 6.40 and 6.57 ppm (3-isomer, major), 2.43, 4.04, 4.46, 6.10 and

6.60 ppm (2-isomer, minor) and 5.31, 6.09, 6.12, 6.15, 6.18, 6.42, 6.47 ppm due to

presence of the 1-[3-(vinylcarbonyloxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole

moieties at the chain ends.

Propargyl glycidyl ether was “Clicked” onto azide-terminated PIB to obtain

epoxy-terminated PIB (Table 5.5, Entry 3). This chemistry would be useful for the

introduction of PIB graft segments (monofunctional) or network chain segments

(difunctional) into epoxy-based networks for toughening. The reaction was carried out at

room temperature in THF solution. Isolation of the polymer consisted of dissolving the

product in hexanes, filtering to remove catalyst, and precipitation into methanol. Figure

5.14 shows the 1H NMR spectrum of the resulting polymer with peak assignments. The

spectrum shows the characteristic 1,3-propylene tether peaks for both isomers and

resonances at 2.63, 2.81, 3.19, 3.45, 4.72 ppm due to the 1-[3-(4-glycidyloxymethyl-1H-

1,2,3-triazol-1-yl)propyl]pyrrole moieties at the chain ends.

Click reactions involving amine-functional alkynes are potentially challenging

due to complexation of copper by the amino group, which could hinder purification of the

functionalized polymer. To address this issue, we reacted 1-(3-azidopropyl)pyrrolyl-PIB

with 3-dimethylamino-1-propyne in THF at 25 °C in the presence of

bromotris(triphenylphosphine)copper(I) (Table 5.5, Entry 4). The progress of the Click

reaction itself did not appear to be effected by complexation. However, purification of

the product did prove to be more laborious in comparison to the other functionalizations

carried out by Click chemistry. Formation of 1-[3-(4-dimethylaminomethyl-1H-1,2,3-

Page 173: Polyisobutylene Chain End Transformations: Block Copolymer

154

triazol-1-yl)propyl]pyrrolyl-PIB was verified by 1H NMR (Figure 5.15) by the

appearance of characteristic tether peaks from both isomers and resonances at 2.28 and

3.62 ppm due to the dimethylamionomethyl moieties).

Conclusion

Synthesis of 1-(ω-azidoalkyl)pyrrolyl-PIB was successfully accomplished by

either reaction of 1-(ω-haloalkyl)pyrrolyl-PIB with sodium azide or by quenching of

quasiliving PIB with 1-(2-azidoethyl)pyrrole. The former method was found to yield

essentially quantitative azide functionality; however, quenching with 1-(2-

azidoethyl)pyrrole produced 1-2 % olefin (mixed exo and endo) chain ends in addition to

the desired azide. The nucleophilic substitution reaction with azide ion was found to

proceed substantially more rapidly with the 3-pyrrole isomer compared to the 2-pyrrole

isomer.

Structure of 1-(ω-azidoalkyl)pyrrolyl-PIB was investigated by both 1H and 13C

NMR. The product of end-quenching was found to be a mixture of 2- and 3-PIB-1-(2-

azido)pyrroles, consistent with our previous report concerning end-quenching with 1-(ω-

haloalkyl)pyrroles. However, compared to the haloalkylpyrroles, 1-(2-azidoethyl)pyrrole

quenching yielded a more balanced isomer ratio of C2/C3 = 0.40/0.60, similar to the

nearly perfectly balanced ratio (0.48/0.52) obtained with 1-methylpyrrole. To our

knowledge, the 1-(2-azidoethyl)pyrrolyl-PIB synthesized in this work represents the first

successful quenching of quasiliving PIB with a capping agent containing azide groups.

We have described here an easy way to functionalize PIB and provide fast access

to a large variety of interesting and potentially useful functional oligomers of uniform

molecular weight and high purity. 1-(ω-Azidoalkyl)pyrrolyl-PIB was shown to provide a

Page 174: Polyisobutylene Chain End Transformations: Block Copolymer

155

modular approach to various telechelic PIBs via the Huisgen 1,3-dipolar cycloaddition

(Click reaction) between azides and terminal alkynes. Click reactions, catalyzed by

bromotris(triphenylphosphine)copper(I), were conducted between 1-(ω-

azidoalkyl)pyrrolyl-PIB and propargyl alcohol, propargyl acrylate, propargyl glycidyl

ether, and 3-dimethylamino-1-propyne. PIBs containing hydroxyl, acrylate, glycidyl, and

dimethylamino groups were cleanly synthesized by this process. These materials have

potential commercial applications in many areas.

Acknowledgments

The research upon which this material is based was generously supported by

Chevron Oronite Co., LLC.

Page 175: Polyisobutylene Chain End Transformations: Block Copolymer

156

References 1. Mach, H.; Rath, P. Lubrication Science 1999, 11, 175.

2. Bergbreiter, D. E.; Tian, J. Tetrahedron Lett. 2007, 48, 4499.

3. Jewrajka, S. K.; Yilgor, E.; Yilgor, I.; Kennedy, J. P. J. Polym. Sci.: Part A: Polym.

Chem. 2009, 47, 38.

4. Erdodi, G.; Kang, J.; Kennedy, J.P.; Yilgor, E.; Yilgor, I. J. Polym. Sci.: Part A:

Polym. Chem. 2009, 47, 5278.

5. Ojha, U.; Kulkarni, P.; Faust, R. Polymer 2009, 50, 3448.

6. Puskas, J. E.; Chen, Y. Biomacromolecules 2004, 5, 1141.

7. Cho, J. C.; Cheng, G.; Feng, D.; Faust, R.; Richard, R; Schwarz, M.; Chan, K.; Boden,

M. Biomacromolecules 2006, 7, 2997.

8. De, P.; Faust, R. Macromolecules 2006, 39, 6861.

9. Martinez-Castro, N.; Morgan, D. L.; Storey, R. F. Macromolecules 2009, 42, 4963.

10. Morgan, D.L.; Storey, R.F. Macromolecules 2009, 42, 6844.

11. Morgan, D. L.; Storey, R. F. Macromolecules submitted.

12. Binder, W.H.; Sachsenhofer, R. Macromol. Rapid Comm. 2007, 28, 15.

13. Rostovtsev, V. V.; Green, L. G.; Fokin, V. V.; Sharpless, K. B. Angew. Chem. 2002,

114, 2708.

14. Kolb, H. C.; Finn, M. G.; Sharpless, K. B. Angew. Chem. Int. Ed. 2001, 40, 2004.

15. Thibault, R. J.; Takizawa, K.; Lowenheilm, P.; Helms, B.; Mynar, J. L.; Fréchet, J.

M. J.; Hawker, C. J. J. Am. Chem. Soc. 2006, 128, 12084.

Page 176: Polyisobutylene Chain End Transformations: Block Copolymer

157

16. Gacal, B.; Akat, H.; Balta, D. K.; Arsu, N.; Yagci, Y. Macromolecules 2008, 41,

2401.

17. Meudtner, R. M.; Hecht, S. Macromol. Rapid Commun. 2008, 29, 347.

18. Jiang, X.; Vogel, E. B.; Smith, M. R., III; Baker, G. L. Macromolecules 2008, 41,

1937.

19. Altintas, O.; Hizal, G.; Tunca, U. J. Polym. Sci.: Part A: Polym. Chem. 2008, 46,

1218.

20. Shi, G.-Y.; Tang, X.-Z.; Pan, C.-Y. J. Polym. Sci.: Part A: Polym. Chem. 2008, 46,

2390.

21. Mespouille, L.; Vachaudez, M.; Suriano, F.; Gerbaux, P.; Coulembier, O.; Degée, P.;

Flammang, R.; Dubois, P. Macromol. Rapid Commun. 2007, 28, 2151.

22. Urbani, C. N.; Bell, C. A.; Whittaker, M. R.; Monteiro, M. J. Macromolecules 2008,

41, 1057.

23. Ostaci, R.-V.; Damiron, D.; Capponi, S.; Vignaud, G.; Léger, L.; Grohens, Y.;

Drockenmuller, E. Langmuir 2008, 24, 2732.

24. Ranjan, R.; Brittain, W. J. Macromol. Rapid Commun. 2007, 28, 2084.

25. Petraru, L.; Binder, W. H. ACS Div. Polym. Chem., Polym. Preprs. 2006, 47, 440.

26. Roth, T.; Groh, P. W.; Pálfi, V.; Iván, B.; Binder, W. H. ACS Div. Polym. Chem.,

Polym. Preprs. 2005, 46, 1166.

27. Bergbreiter, D. E.; Hamilton, P. N.; Koshti, N. M. J. Am. Chem. Soc. 2007, 129,

10666.

Page 177: Polyisobutylene Chain End Transformations: Block Copolymer

158

28. Magenau, A.J.D.; Martinez-Castro, N.; Savin, D.A.; Storey, R.F. Macromolecules

2009, 42, 8044.

29. Storey, R. F; Lee, Y. J. Polym. Sci.: Part A: Polym. Chem. 1991, 29, 317.

30. Storey, R. F.; Choate, K. R., Jr. Macromolecules 1997, 30, 4799.

31. Bidan, G. Tetrahedron Lett. 1985, 26, 735.

32. Storey, R. F.; Maggio, T.L. Macromolecules 2000, 33, 681.

33. Storey, R. F.; Donnalley, A. B.; Maggio, T. L. Macromolecules 1998, 31, 1523.

34. Storey, R. F.; Thomas, Q. A. Macromolecules 2003, 36, 5065.

35. Li, J.; Sung, S.; Tian, J.; Bergbreiter, D. E. Tetrahedron 2005, 61, 12081.

36. Okamoto, K.; Kita, T.; Araki, K.; and Shingu, H. Bull. Chem. Soc. Japan 1967, 40,

1913.

37. Cheradame, H.; Habimana, J.; Rousset, E.; Chen, F. J. Macromolecules 1994, 27,

631.

38. Cheradame, H.; Habimana, J.; Rousset, E.; Chen, F. J. Makromol. Chem. 1991, 192,

2777.

39. Storey, R.F.; Stokes, C.D.; Harrison, J.J. Macromolecules 2005, 38, 4618.

Page 178: Polyisobutylene Chain End Transformations: Block Copolymer

159

Table 5.1. Conversion of N-(ω-Haloalkyl)pyrrole-PIB to N-(ω-Azidoalkyl)pyrrole-PIB via Nucleophilic Displacement of Halide by NaN3 in N,N-Dimethylformamide/heptane (50/50, v/v)a

Entry [PIB]

(mol/L)

Mn

(g/mol)

PDI End

Group

[NaN3]

(mol/L)

Temp.

(°°°°C)

Azide

Functionalitye

1 0.113 1,760 1.05 PyClb 0.339 70 0.39 2 0.067 1,980 1.04 PyCl 0.201 70 0.19 3 0.085 1,760 1.05 PyCl 0.255 90 ~1.0 4 0.085 1,980 1.04 PyCl 0.255 90 ~1.0 5 0.100 1,980 1.04 PyBrc 0.300 90 ~1.0 6 0.085 1,730 1.04 PyBr 0.255 90 ~1.0 7 0.144 1,380 1.05 PyBrPd 0.432 90 ~1.0 8f 0.070 2,120 1.02 PyBrP 0.424 90 ~2.0 9f 0.071 2,780 1.07 PyBr 0.424 90 ~2.0

a Reaction time 24 h b 1-(2-chloroethyl)pyrrole c 1-(2-bromoethyl)pyrrole d 1-(3-bromopropyl)pyrrole e Determined by 1H NMR integration of tether proton signals of reactant and product f Difunctional PIB

Page 179: Polyisobutylene Chain End Transformations: Block Copolymer

160

Table 5.2. End-Quenching of Difunctional tert-Chloride-Terminated Masterbatch PIBa with N-(2-Azidoethyl)pyrrole in 60/40 (v/v) Hexane/MeCl at -70°C

Entry [tert-Cl]

(mol/L)

[Lut]

(mol/L)

[TiCl4]

(mol/L)

[PyAz]

(mol/L)

Functionalityb

time (min)

3 10 15

1 0.010 0.0015 0.10 0.018 -- 0.98-0.99 -- 2 0.010 0.0015 0.020 0.018 -- 0.23 -- 3 0.010 0.0015 0.10 0.018 0.99 -- -- 4 0.010 0.0015 0.030 0.018 -- -- 0.32 5c 0.010 0.0014 0.10 0.018 0.99 -- --

a nM = 2,100 g/mol b Determined by 1H NMR integration of tether proton signals relative to aromatic proton signals of bDCC initiator residue c Monofunctional PIB: nM = 1,970 g/mol

Page 180: Polyisobutylene Chain End Transformations: Block Copolymer

Table 5.3. End-Quenching of Quasiliving IB Polymerizations with N-(2-azidoethyl)pyrrole in 60/40 (v/v) Hexane/CH3Cl at -70°Ca Entry [IB]

mol/L

[bDCC]

mol/L

[TiCl4]b

mol/L

Mn

g/mol

PDI [PyAz]

[CE]

[TiCl4]c

mol/L

Functionalityd

Time (min)

2 40 60

1 0.38 0.0059 0.053 3,350 1.01 1.8 -- - 0.36 - 2 0.36 0.0053 0.048 3,280 1.01 1.8 0.053 - - 0.87 3 0.36 0.0053 0.048 3,200 1.01 1.8 0.105 0.99 - -

a [2,6-Lutidine] = 0.0028 mol/L b [TiCl4] during polymerization c [TiCl4] during quenching d Determined by 1H NMR integration of tether proton signals relative to aromatic proton signals of bDCC initiator residue

161

Page 181: Polyisobutylene Chain End Transformations: Block Copolymer

Table 5.4. Synthesis of Model Compounds 1-[2-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)ethyl]pyrrole and 1-[3-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrolea

Entry [PyAz]b

mol/kg

[PyPAz]c

mol/kg

[PrgOH]d

mol/kg

[(PPh3)3CuBr]e

mol/kg

[Cu(I)Br]

mol/kg

[PMDETA]f

mol/kg

Time

h

1 2.17 2.9 0.023 12 2 3.16 - 4.6 - 0.014 0.014 0.083 3 - 2.84 3.4 - 0.017 0.017 0.083

a Initial reaction temperature = 25°C; Entries 2 and 3 were very exothermic b 1-(2-Azidoethyl)pyrrole. c 1-(3-Azidopropyl)pyrrole. d Propargyl alcohol. e Bromotris(triphenylphosphine)copper(I) f

N,N,N',N '',N ''-pentamethyldiethylenetriamine

162

Page 182: Polyisobutylene Chain End Transformations: Block Copolymer

163

Table 5.5. Experimental conditions and results of functionalization of 1-(3-azidopropyl)pyrrole -PIB by Click chemistry in tetrahydrofuran Entry [PIB-N3]

mol/L

Mn x10-3

g/mol

Mw/Mn [(PPh3)3CuBr]

mol/L

[PrgOH]a

mol/L

[PrgAcr]b

mol/L

[GPrgE]c

mol/L

[DAP]d

mol/L

1 0.069 1.38 1.05 0.014 0.19 --- --- --- 2 0.10 2.95 1.03 0.017 --- 0.21 --- --- 3 0.093 2.95 1.03 0.022 --- --- 0.19 --- 4 0.093 2.95 1.03 0.022 --- --- --- 0.19

aPropargyl alcohol bPropargyl acrylate cGlycidylpropargyl ether d3-Dimethylamino-1-propyne

Page 183: Polyisobutylene Chain End Transformations: Block Copolymer

164

Table 5.6. Chemical shift assignments for 2-and 3-PIB-1-(2-azidoethyl)pyrrole

Assignment 1H NMR chemical shift (ppm)

k 0.99 h 1.11 i 1.41 f 1.65, s f′ 1.73, s a 3.52, m a′ 3.64, m c 3.95, m c′ 4.13, m 3′ 5.90, m 4 6.05, m 4′ 6.07, m 2 6.40, m 5 6.56, m 5′ 6.59, m

CH2N CH2 N3ac

2

54

e

f

h

i

kn a'c'

5'4'

3'

e'

f' CH2 CH2N

N3

Page 184: Polyisobutylene Chain End Transformations: Block Copolymer

165

Table 5.7. Chemical shift assignments for 2-and 3-PIB-1-(2-azidoethyl)pyrrole

Assignment 13C NMR chemical shift (ppm)

e′ -- e -- h 31.25 k 32.46 j 32.58 d -- d′ -- g 38.14 c′ 47.21 c 48.64 a′ 52.08 a 52.54 f′ 55.90 f 58.52 i 59.51 3′ 107.13 4′ 107.59 4 107.69 2 116.50 5 119.85 5′ 121.19 3 135.60 2′ 138.85

CH2N CH2 N3ac

2

54

e

f

h

i

kn 3

dgj 2'd'a'c'

5'4'

3'

e'

f' CH2 CH2N

N3

Page 185: Polyisobutylene Chain End Transformations: Block Copolymer

166

Table 5.8. Chemical shift assignments for 2-and 3-PIB-1-(3-azidopropyl)pyrrole

Assignment 1H NMR chemical shift (ppm)

k 0.99 h 1.11 i 1.41 f 1.65, s f′ 1.73, s b 1.95, m b′ 2.07, m a 3.22, m a′ 3.44, m c 3.90, m c′ 4.04, m 3′ 5.88, m 4 6.02, m 4′ 6.05, m 2 6.38, m 5′ 6.59, m 5 6.55, m

CH2N CH2 CH2 N3ab

2c

4

e

f

h

i

kn

5

f'

e'

3'4'

c' b' a'

5'

CH2 CH2 CH2N

N3

Page 186: Polyisobutylene Chain End Transformations: Block Copolymer

167

Table 5.9. Chemical shift assignments for 2-and 3-PIB-1-(3-azidopropyl)pyrrole

Assignment 13C NMR chemical shift (ppm)

e′ -- a 29.54 a′ 30.32 e -- h 31.23 k 32.44 j 32.58 b 34.65 b′ 35.90 g 38.14 c 45.66 c′ 46.06 f′ 55.76 f 58.37 i 59.51 4′ 106.55 3′ 107.03 4 107.22 2 116.55 5 119.86 5′ 121.07 3 135.27 2′ 139.22

jCH2N CH2 CH2 N3

ab2

c

4

e

f

h

i

kn 3

dg

5

f'

e'

3'

4'

c' b' a'd' 2'

5'

CH2 CH2 CH2N

N3

Page 187: Polyisobutylene Chain End Transformations: Block Copolymer

168

CH2 Z

TiCl4

nCl

PIB-N3

n Ti2Cl9

A B

N NNN

CH2 Zn

N Br

N N3

N Brn

N N3

n

NaN3

(Ph3P)3CuBrTHF

Figure 5.1. Clickable polyisobutylene: synthetic approaches and utilization.

Page 188: Polyisobutylene Chain End Transformations: Block Copolymer

169

Figure 5.2. 1H NMR spectrum of 1-(2-azidoethyl)pyrrole-PIB prepared by reaction of 1-(2-bromoethyl)pyrrole-PIB with sodium azide (Table 1, Entry 5). The product is a mixture of isomers with PIB substituted at the 3- (major) and 2- (minor) positions of the pyrrole ring. Inset shows the tether proton resonances of the bromide precursor for comparison.

7.0 6.0 5.0 4.0 1.6 1.2 0.8

PIB-N3

f

f '

hi

e

e'

a'c'

5'4'

3'

e'

f' CH2 CH2N

N3

CH2N CH2 N3ac

2

54

e

f

h

i

kn

δ ppm

k

PIB-Br

c

a'

a

c'3'4'

425

5'

6.5 6.0 4.5 4.0 3.5

4.5 4.0 3.5

Page 189: Polyisobutylene Chain End Transformations: Block Copolymer

170

Figure 5.3. 1H NMR spectrum of difunctional 1-(3-azidopropyl)pyrrole-PIB, with peak integrations, prepared by reaction of difunctional 1-(3-bromopropyl)pyrrole-PIB with sodium azide (Table 1, Entry 8). The product is a mixture of isomers with PIB substituted at the 3- (major) and 2- (minor) positions of the pyrrole ring.

7.0 6.0 5.0 4.0 3.0 2.0 1.5 1.0 0.5

d'

h

d

f

l

j

g

ee'

i

5'

5 24

4'3' c'

a'

c a

b

b'

k

CH2 CH2 CH2N N3b'c' a'

d'

e'

3'

4' 5'

ne

f

ac

2

54

d

CH2 N3N CH2 CH2b

g

hk

kk

l

i

j

δ ppm

1.53

3.00

1.52

5.83

2.41

2.39

6.6 6.4 6.2 6.0 4.5 4.0 3.5 3.0 2.5 2.0

Page 190: Polyisobutylene Chain End Transformations: Block Copolymer

171

Figure 5.4. 1H NMR spectrum of difunctional 1-(2-azidoethyl)pyrrole-PIB prepared by post-polymerization reaction of the corresponding difunctional 1-(2-bromoethyl)pyrrole-PIB with sodium azide. The product is a mixture of isomers with PIB substituted at the 3- (major) and 2- (minor) positions of the pyrrole ring (Table 1, Entry 10).

7.0 6.0 5.0 4.0 1.6 1.2 0.8

k

j

d'

dh

e

e'

i

a'

a

c'

c3

4

4'

25

5'

CH2N CH2 N3c' a'

d'

e'

3'

4' 5'

ne

f

ac

2

54

d

CH2 N3N CH2g

hk

kk

l

i

jl

g

δ ppm

f

6.6 6.3 6.0 4.2 4.0 3.8 3.6 3.4

Page 191: Polyisobutylene Chain End Transformations: Block Copolymer

172

Figure 5.5. Reaction conversion vs. time for reaction of difunctional 1-(2-bromoethyl)pyrrole-PIB with sodium azide: 50/50 (v/v) heptanes/DMF, 90°C, [CE] = 0.10 mol/L, [NaN3] = 0.30 mol/L; C3 isomer (triangles), C2 isomer (squares), overall (circles).

0 2 4 6 8 10 120

20

40

60

80

100

NCH2 CH2 N3

NCH2 CH2 N3

N CH2 CH2 N3 C3 isomer

C2 isomer

overall

conv

ersi

on (

%)

time (h)

Page 192: Polyisobutylene Chain End Transformations: Block Copolymer

173

Figure 5.6. Reaction conversion vs. time for reaction of monofunctional 1-(2-bromoethyl)pyrrole-PIB with sodium azide: 50/50 (v/v) heptanes/DMF, 90°C, [CE] = 0.10 mol/L, [NaN3] = 0.30 mol/L; C3 isomer (circles), C2 isomer (squares), overall (triangles).

0 2 4 6 8 10 120

10

20

30

40

50

60

70

80

90

100

N CH2 CH2 CH2

NCH2 CH2 CH2

NCH2 CH2 CH2

N3

N3

N3

con

vers

ion

(%)

time (h)

C3 isomer

C2 isomer

overall

Page 193: Polyisobutylene Chain End Transformations: Block Copolymer

174

Figure 5.7. 1H NMR spectrum of difunctional 1-(2-azidoethyl)pyrrole-PIB, with peak integrations, prepared by end-quenching of difunctional tert-chloride-terminated PIB with N-(2-Azidoethyl)pyrrole in 60/40 (v/v) hexane/MeCl at -70°C (Table 5.2, Entry 1).

7.0 6.0 5.0 4.0 1.6 1.2 0.8

2.39

j

d'

d

h

e

e'

i

a'a

c'c

3

4

4'

25

k

5'

β α'

α α'

CH2N CH2 N3c' a'

d'

e'

3'

4' 5'

ne

f

ac

2

54

d

CH2 N3N CH2g

hk

kk

l

i

j lg

δ ppm

f

β

α

1.60

1.602.41

5.80

3.00

0.06

0.05

5.2 5.0 4.8 4.6

Page 194: Polyisobutylene Chain End Transformations: Block Copolymer

175

Figure 5.8. 13C NMR spectrum of difunctional 1-(2-azidoethyl)pyrrole-PIB prepared by end-quenching of difunctional tert-chloride-terminated PIB with N-(2-Azidoethyl)pyrrole in 60/40 (v/v) hexane/MeCl at -70°C (Table 5.2, Entry 1).

140 120 60 50 40 30

hgi

CH2N CH2 N3c' a'

e'

f'

3'

4' 5'

d' 2'

nf

h

ac

2

54

e

CH2 N3N CH2i

k

l

n

dg

m

j 3n

e'

f

dd'

f '

e

3'4

4'2

5'

5

3 a'

lk

j

m2'

a

c'c

δ ppm

122 120 118 108 106 60 58 56

33 32 31 30 29

Page 195: Polyisobutylene Chain End Transformations: Block Copolymer

176

Figure 5.9. GPC traces of difunctional tert-chloride-terminated PIB before (dotted line) and after (solid line) end-quenching with 1-(2-azidoethyl)pyrrole (Table 5.4, Entry 1).

12 13 14 15 16 17

time (min)

Page 196: Polyisobutylene Chain End Transformations: Block Copolymer

177

Figure 5.10. 1H NMR spectrum of 1-[2-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)ethyl] pyrrole from reaction of 1-(2-azidoethyl)pyrrole with propargyl alcohol catalyzed with Cu(I)Br/PMDETA (Table 5.4, Entry 2).

8 7 6 5 4 3

l

h

f

c a

3

N CH2 CH2 NNN

CH2 OHac

23

l

m n

δ ppm

2

Page 197: Polyisobutylene Chain End Transformations: Block Copolymer

178

Figure 5.11. Reaction kinetics of azide-terminated PIB (2- or 3-carbon tether) with propargyl alcohol in THF at 25°C with initial [alkyne]/[azide] = 2.0: A) Cu(I)Br/PMDETA catalyst with 3-carbon tether; B) bromotris(triphenylphosphine)copper(I) with 2-carbon tether (solid circles) or 3-carbon tether (open circles).

0 10 15 20 250

20

40

60

80

100

conv

ersi

on (

%)

time (min)

0.00 0.02 0.04 0.060.0

0.5

1.0

1.5

2.0

Ln(

(M-p

)/(M

(1-p

)))

[PIB-N3]

0(M-1)t

0 4 8 12 16 20 240

20

40

60

80

100

conv

ersi

on (%

)

time (h)

0.0 0.5 1.0 1.50.00

0.25

0.50

0.75

1.00

1.25

Ln(

(M-p

)/(M

(1-p

)))

[PIB-N3]

0(M-1)t

B

A

Page 198: Polyisobutylene Chain End Transformations: Block Copolymer

179

Figure 5.12. 1H NMR spectrum of 1-[3-(4-hydroxymethyl-1H-1,2,3-triazol-1-yl)propyl] pyrrole-PIB prepared by Click chemistry (Table 5.5, Entry 1).

8 7 6 5 4 3 2 1

e

3

4',425

5'

5'

4'3'

54

2

bb'

a

a'

cc'

c b a

c' b' a'

d'

d'

d

d

e

δ ppm

e

7.6 7.5 6.6 6.3 6.0 5.7

5.0 4.5 4.0 2.6 2.4 2.2

N

CH2 CH2 CH2 N

NN

CH2 OH

N CH2 CH2 CH2 N

NN

CH2 OH

Page 199: Polyisobutylene Chain End Transformations: Block Copolymer

180

Figure 5.13.

1H NMR spectrum of 1-[3-(4-vinylcarbonyloxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole-PIB prepared by Click chemistry (Table 5.5, Entry 2).

7 6 5 4 3 2 1

gf

f g

gf

N

CH2 CH2 CH2 N

NN

CH2 C

O

CH CH2

g4',425

5'

5'

4'3'

54

2

bb'

a

a'

c

c'

c b a

c' b' a'

d'

d'

d

d

e

e

δ ppm

e

4.5 4.0 2.50 2.25

N CH2 CH2 CH2 N

NN

CH2 C

O

CH CH2

6.6 6.4 6.2 6.0 5.8

Page 200: Polyisobutylene Chain End Transformations: Block Copolymer

181

Figure 5.14.

1H NMR spectrum of 1-[3-(4-glycidyloxymethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole-PIB prepared by Click chemistry (Table 5.5, Entry 3).

7 6 5 4 3 2 1

f

N CH2 CH2 CH2 N

NN

CH2 O CH CH2

O

CH2

hh

h

h

gf

N

CH2 CH2 CH2 N

NN

CH2 O CH CH2

O

CH2

gf

f g

3

4',425

5'

5'

4'3'

54

2

bb'

a

a'

c

c'

c b a

c' b' a'

d'

d'

d

d

e

e

δ ppm

e

6.5 6.0 5.0 4.5

4.5 4.0 3.5 3.0 2.5

Page 201: Polyisobutylene Chain End Transformations: Block Copolymer

182

Figure 5.15.

1H NMR spectrum of 1-[3-(4-dimethylaminomethyl-1H-1,2,3-triazol-1-yl)propyl]pyrrole-PIB prepared by Click chemistry (Table 5.5, Entry 4).

7 6 5 4 3 2 1

N

CH2 CH2 CH2 N

NN

CH2 NCH3

CH3

aa

a

34',425

5'

5'

4'3'

54

2

bb'

a

a'c

c'

c b a

c' b' a'

d'

d'

d

d

e

e

δ ppm

e

N CH2 CH2 CH2 N

NN

CH2 NCH3

CH3

4.5 4.0 3.52.6 2.4 2.2

Page 202: Polyisobutylene Chain End Transformations: Block Copolymer

183

CHAPTER VI

FACILE POLYISOBUTYLENE FUNCTIONALIZATION VIA THIOL-ENE CLICK

CHEMISTRY

Introduction

Polyisobutylene (PIB) has commercial utility as a stabilizing fuel and motor oil

additive, packaging elastomer, adhesive and sealant, and more recently, as a biomaterial.

This completely saturated hydrocarbon elastomer has excellent thermal and oxidative

stability, gas-barrier properties, and biocompatibility. Due to these and other unique

characteristics, PIB remains the subject of continued research; e.g. as a self–separating

homogeneous catalyst support,1 a matrix for quantum dot-polymer composities for inks

with enhanced photoluminensce,2 and a biostable/biocompatible thermoplastic elastomer

(TPU) for the drug-eluting coating on coronary stents.3

Efficient methods for the synthesis of PIB carrying functional end groups have

long been sought to facilitate the creation of new PIB-based materials. In the past, post-

polymerization modification procedures were employed, which were often laborious,

time-consuming, and multistep.4 To circumvent these difficult procedures, in-situ

quenching has recently been developed to conveniently functionalize PIB through direct

reaction of quasiliving PIB with a nucleophilic quenching or capping agent. Various

classes of compounds with demonstrated effectiveness include olefins such as

alkenylsilanes5 and butadiene,6 hindered nucleophiles,7 and activated aromatic

compounds such as 2-alkylfurans, thiophene, N-substituted pyrroles, and alkoxy

benzenes.8 Although in-situ quenching is attractive due to its simplicity, the quenching

agents are often expensive or unavailable commercially. Also, in-situ quenching is

Page 203: Polyisobutylene Chain End Transformations: Block Copolymer

184

generally limited to soft nucleophilic (π) quenching agents, since hard nucleophiles (σ)

inevitably react with the Lewis acid.9

Since adoption of the thiol-ene reaction into the class of Click chemistries,10

numerous researchers have utilized this powerful synthetic tool for polymer

modification,11 polymer-protein conjugate synthesis,12 network formation,13 and creation

of complex polymeric architectures.10a,14 Thiol-ene chemistry offers many advantages

including mild reaction conditions, tolerance to oxygen and water, simple purification,

high reaction rates, modularity, and quantitative conversion in the absence of metal

catalyst. Moreover, the vast array of commercially available thiol and alkene

functionalities provides an incredibly versatile selection of reagents.

Monofunctional PIB possessing a high proportion (up to 85-90%) of exo-olefin

(methyl vinylidene) end groups is available commercially (e.g., Glissopal, BASF;

Ultravis, BP Chemicals) and both monofunctional and difunctional PIBs with ~100%

exo-olefin termini are readily produced via in-situ quenching of quasiliving PIB.7

Application of thiol-ene chemistry to exo-olefin PIB would thus provide a practical

means to a variety of functional PIBs. However, only a few literature reports have

appeared on this topic, and reaction conversions were typically low and/or reaction times

were long. For example, Blackborow et al.15 reported the radical addition of thiols

carrying a number of functional groups including hydroxyl, carboxylic acid, and

alkoxysilane to Ultravis (52 or 84% exo-olefin end groups), using either photo- or

thermal initiation. High conversions were qualitatively claimed but reaction times were

long (12-17 h for photoinitiation, 20-50 h for thermal initiation). Gorski et al. reported

the radical addition of alkyl and hydroxyl functional thiols to Glissopal; although reaction

Page 204: Polyisobutylene Chain End Transformations: Block Copolymer

185

kinetics with Glissopal were not reported, a model compound for exo-olefin PIB, 2,4,4-

trimethyl-1-pentene, required long reaction times even when reacted neat in the presence

of a ten-fold excess of thiol (p ≈ 85% @ 6 h).16 Later, based partly on their earlier

patent,15 Blackborow et al. reported the radical addition to Ultravis of thiols with various

functional groups including hydroxyl, methoxyethoxyethyl, carboxylic acid, and

organosilane.17 Conversions, determined using FTIR, ranged from 35 to 80%, with

reactions times of 5 h or more; in the few examples where high conversions were reached

(p ≥ 90%), reaction times between 44 and 88 h were necessary.17

The PIB modifications described in these previous literature reports do not meet

the qualifications of a Click reaction. Herein, we show that the thiol-ene reaction, under

appropriate conditions, can be used to simply, rapidly, and quantitatively functionalize

exo-olefin PIB with an array of functional thiols (Figure 6.1). These reactions achieve

near-quantitative conversion of the exo-olefin in less than 10 min, are applicable to both

mono and difunctional PIBs, and require minimal equipment, and the products can be

purified by either a simple precipitation or wash.

Experimental

Materials

All materials were purchased from Aldrich at the highest available purity and

used as received unless otherwise stated. Free radical photoinitiator, dimethoxy-2-

phenylacetophenone (DMPA), was purchased from CIBA and used as received.

Instrumentation

FTIR spectra were recorded using a Bruker Equinox 55 spectrometer. Samples

were sandwiched between two sodium chloride plates. Each spectrum was collected as

Page 205: Polyisobutylene Chain End Transformations: Block Copolymer

186

an average of 64 scans. The data were analyzed using the Bruker OPUS/IR Version 4.0

software.

Real-time FTIR was used to monitor exo-olefin conversion kinetics (1640 cm-1).

Exo-olefin-terminated PIB (0.501 g, 0.40 mmol), DMPA (0.012 g, 1 wt%), and 0.40 mL

CH2Cl2 were charged to a 20 mL glass scintillation vial. After dissolution, 2-(tert-

butoxycarbonylamino)ethanethiol (100.9 µL, 0.60 mmol) was added, and the contents

were shaken for 20 min. The resulting reaction mixture was sandwiched between sodium

chloride plates resulting in a sample thickness of approximately 250 microns. The light

intensity of the high pressure mercury lamp delivered to the sample via a light pipe was

∼28.3 mW/cm2.

NMR spectra were acquired using a Varian Mercuryplus 500 MHz NMR

spectrometer. Samples were dissolved in chloroform-d (3−7%, w/v) and analyzed within

5 mm NMR tubes.

The number-average molecular weight (Mn) and polydispersity index (PDI) of the

polymeric materials were measured using gel permeation chromatography (GPC). The

GPC system, operating at 35 °C, consisted of a Waters Alliance 2695 separations

module, an online multiangle laser light scattering (MALLS) detector (MiniDAWN,

Wyatt Technology Inc.), an interferometric refractometer (Optilab rEX, Wyatt

Technology Inc.), an online differential viscometer (ViscoStar, Wyatt Technology, Inc.),

and two mixed E (3 µm bead size) PL gel (Polymer Laboratories Inc.) GPC columns

connected in series. Freshly distilled THF served as the mobile phase at a flow rate of

1.0 mL/min. Sample concentrations were 10−12 mg/mL, with an injection volume of

100 µL. The detector signals were recorded using ASTRA software (Wyatt Technology

Page 206: Polyisobutylene Chain End Transformations: Block Copolymer

187

Inc.) and molecular weights were determined using dn/dc values calculated from an

equation relating dn/dc of PIB in THF as a function of PIB molecular weight.18

Synthesis of exo-olefin-terminated PIB

Mono- and difunctional exo-olefin-terminated PIBs were prepared and

characterized according to literature using 2-chloro-2,4,4-trimethylpentane (TMPCl) or

1,3-di(1-chloro-1-methylethyl)-5-tert-butylbenzene, respectively, as initiators and

1,2,2,6,6-pentamethylpiperidine (PMP) as a hindered-amine quencher.19 A representative

procedure to prepare monofunctional exo-olefin-terminated PIB was as follows:

Quasiliving polymerization of isobutylene(IB) with TMPCl as an initiator was carried out

within a N2 atmosphere glovebox, equipped with an integral cryostated hexane/heptanes

bath. Into a round bottom flask with a mechanical stirrer, infrared probe, and

thermocouple were added 572 mL CH3Cl, 860 mL hexane, 2.50 mL (2.19 g, 0.0147 mol)

TMPCl, and 0.86 mL (0.79 g, 0.0074 mol) of 2,6-lutidine. The mixture was allowed to

equilibrate to -60 °C, and then 29.0 mL (19.9 g, 0.354 mol) of IB was added to the

reactor and allowed to reach thermal equilibrium. To begin the polymerization, 4.8 mL

(8.3 g, 0.044 mol) of TiCl4 was charged to the reactor. Full monomer conversion (≥ 98

%) was achieved in 90 min, after which time 8.0 mL (6.9 g, 0.044 mol) PMP and an

additional 4.8 mL TiCl4 (8.3 g, 0.044 mol) were added to the polymerization. PMP was

allowed to react with the living chain ends for 90 min. Finally, the reaction was

terminated by addition of excess prechilled methanol. The contents of the reaction flask

were allowed to warm to room temperature, and the polymer in hexane was immediately

washed with methanol and then precipitated into methanol from hexane. The precipitate

was collected by dissolution in hexane; the solvent was washed with water, dried over

Page 207: Polyisobutylene Chain End Transformations: Block Copolymer

188

MgSO4, and concentrated on a rotary evaporator. Residual solvent was removed under

vacuum at 40 °C.

Synthesis of Mono-functional Primary Chloride-terminated PIB

A representative procedure was as follows: exo-olefin-terminated PIB (0.251 g,

0.20 mmol), DMPA (0.006 g, 1 wt%), and 0.22 mL dichloromethane (CH2Cl2) were

charged to a 20 mL glass scintillation vial. After dissolution, 3-chloro-1-propanethiol

(29.2 µL, 0.30 mmol) was added, and the contents were shaken for 20 min. The sample

was then irradiated using a medium pressure Hg lamp (light intensity ∼6.68 mW/cm2) for

approximately 3.5 min. CH2Cl2 was removed under reduced pressure, and the crude

reaction mixture was twice precipitated into methanol from hexane. The final precipitate

was collected and put under reduced pressure until a constant weight was achieved.

One-pot Synthesis of Mono-functional Primary Amine-terminated PIB

A representative procedure was as follows: exo-olefin-terminated PIB (0.251 g,

0.20 mmol), DMPA (0.006 g, 1 wt. %), and 0.26 mL CH2Cl2 were charged to a 20 mL

glass scintillation vial. After dissolution, 2-(tert-butoxycarbonylamino)ethanethiol (50.7

µL, 0.30 mmol) was added, and the contents were shaken for 20 min. The sample was

then irradiated using a medium pressure Hg lamp (light intensity ∼6.68 mW/cm2) for

approximately 3.5 min. For deprotection, 4.28 mL of a triflouroacetic acid and CH2Cl2

mixture (50:50 v/v) was injected into the scintillation vial and agitated for 30 min.

CH2Cl2 was then removed under reduced pressure, and the crude reaction mixture was

dissolved in hexane. The resulting solution was washed twice with brine and then twice

with deionized water, neutralized with sodium bicarbonate, filtered, and put under

reduced pressure until a constant weight was achieved.

Page 208: Polyisobutylene Chain End Transformations: Block Copolymer

189

Synthesis of Mono-functional Carboxylic Acid-terminated PIB

A representative procedure was as follows: exo-olefin-terminated PIB (0.126 g,

0.10 mmol), DMPA (0.005 g, 1 wt%), and 0.23 mL chloroform (CHCl3) were charged to

a 20 mL glass scintillation vial. After dissolution, thiol glycolic acid (20.8 µL, 0.30

mmol) was added, and the contents were shaken for 20 min. The sample was then

irradiated using a medium pressure Hg lamp (light intensity ∼6.68 mW/cm2) for

approximately 8 min. CHCl3 was removed under reduced pressure, and the crude

reaction mixture was dissolved in hexane. The resulting solution was washed three times

with water and put under reduced pressure until a constant weight was achieved.

Synthesis of Mono-functional Primary Alcohol-terminated PIB

A representative procedure was as follows: exo-olefin-terminated PIB (0.157 g,

1.25 mmol), DMPA (0.061 g, 1 wt%), and 0.26 mL CHCl3 were charged to a 20 mL

glass scintillation vial. After dissolution, 1-mercapto-6-hexanol (68.4 µL, 5.00 mmol)

was added, and the contents were shaken for 20 min. The sample was then irradiated

using a medium pressure Hg lamp (light intensity ∼6.68 mW/cm2) for approximately 6

min. CHCl3 was removed under reduced pressure, and the crude reaction mixture was

dissolved in hexane. The resulting solution was washed three times with methanol and

put under reduced pressure until a constant weight was achieved.

Results and Discussion

Telechelic mono- and difunctional exo-olefin PIBs were synthesized via

quasiliving cationic polymerization followed by quenching with the hindered amine,

1,2,2,6,6-pentamethylpiperidine.7 Low molecular weight polymers were targeted to

facilitate end group characterization by NMR spectroscopy. Molecular weight,

Page 209: Polyisobutylene Chain End Transformations: Block Copolymer

190

polydispersity index (PDI), and functionality of the exo-olefin PIBs are summarized in

Table 6.1. Both polymers exhibited very low PDIs with unimodal, symmetrical elution

peaks in GPC (Figures 6.2 and 6.3) and high chain end functionality (≥ 97 %), calculated

as previously reported in literature.7 Excellent agreement between GPC and 1H NMR

molecular weights was observed for both PIB precursors.

A variety of photo-initiated thiol-ene reactions were conducted, and the results are

summarized in Table 6.2. Particular emphasis was placed on primary halogen and amine

functionalities due to their novelty and utility. The clear, transparent nature and inert

hydrocarbon backbone of PIB facilitated efficient photochemical reactions. Typical

synthetic procedures involved charging a scintiallation vial with an appropriate mass of

polymer, thiol, and 1 wt% photoinitiator, and then dissolving the mixture in a suitable

solvent. Solvent selection (CH2Cl2 or CHCl3) and reagent concentrations were crucial

parameters, and were chosen to ensure a homogenous reaction mixture. Reaction

conversions were improved by reducing reaction temperature, in agreement with prior

literature for thiol-ene functionalization of PIB.17 In the absence of external cooling,

internal reaction temperatures were found to rise as high as 40 °C; whereas by placing the

reactor in an ice bath, reaction temperatures were maintained below 15 °C.

Thiol-ene reactions involving 3-chloropropane thiol in 50% excess (Exp. 1-3)

achieved high conversions in less than 5 min, with essentially quantitative conversion

achieved with cooling. Figure 6.4 depicts 1H NMR spectra of the monofunctional exo-

olefin PIB precursor and resulting 3-chloropropane thioether functionalized PIB (Exp. 2).

Resonances associated with the exo-olefin terminus of the precursor, consisting of

methylene (2.00 ppm), methyl (1.78 ppm), and olefinic protons (4.64 and 4.85 ppm),

Page 210: Polyisobutylene Chain End Transformations: Block Copolymer

191

were all absent in the final product. New resonances in the product associated with the 3-

chloropropane thioether moiety were located at 3.66 (triplet), 2.66 (triplet), 2.04 (quintet),

and 1.74 ppm. In addition, diastereotopic protons, d1 and d2, were observed at 2.52 and

2.35 ppm, each appearing as a doublet of doublets due to the adjacent chiral center.

Thiol-ene reactions with cysteamine and cysteamine hydrochloride, Exp. 4 and 5,

resulted in no thioether formation. These results were theorized to be a consequence of

thiolate formation, due to amine basicity, rendering the thiol group incapable of thiyl

radical formation. Furthermore, incompatibility of these reagents in a common reaction

medium, creating a heterogeneous reaction mixture, impeded their ability to react.

Typical solvents for cysteamine, including water, methanol, and DMF, were incompatible

with the hydrophobic PIB; THF and THF/H2O solvent systems were explored with

thermal and photo initiators, respectively, with limited success.

Conversion of the olefin increased dramatically when the amine group of

cysteamine was protected with a tert-butoxycarbonyl (Boc) group (Exp. 6). The Boc

group produced a completely homogenous reaction mixture and considerably reduced the

basicity of the amine, diminishing thiolate formation. Although the Boc-protected

cysteamine yielded the desired product, complete conversion was still not achieved in

Exp. 6. However, near-quantitative conversion was achieved for both the mono (Exp. 7)

and difunctional (Exp. 8) exo-olefin polymers by cooling the reaction with an ice-bath.

Thiol-ene reaction kinetics by FTIR for cysteamine-Boc are shown in Figure 6.5.

GPC analysis of the PIB-Boc and PIB-Cl products of Table 2 revealed

symmetrical traces with no high molecular weight coupled species. Slight decreases in

Page 211: Polyisobutylene Chain End Transformations: Block Copolymer

192

elution volume were observed, indicative of increased molecular weight, with respect to

the exo-olefin PIB. Figure 6.6 shows GPC traces of PIB-exo, PIB-Boc, and PIB-Cl.

Sunlight-activated radical generation was also attempted (Exp. 9); however,

extended reaction times were required to achieve complete olefin conversion and

substantial byproducts were detected. Prolonged exposure to sunlight (3 h) may have

caused unwanted photo-oxidization to occur. The susceptibility of thiols and various

sulfur derivatives to photo-oxidation is well known in literature.20

In an effort to obtain amine functionality directly from Boc-protected cysteamine,

a procedure was developed in which the thiol-ene reaction and subsequent deblocking

were carried out in one pot. Deblocking was accomplished by simply charging a 50 vol%

trifloroacetic acid (TFA) solution in CH2Cl2 to the reactor after the thiol-ene reaction was

complete (Exp. 10). No by-products were detected while still maintaining near

quantitative conversion and simple purification. 1H NMR spectra of Boc-protected and

deprotected amine functionalized PIB (PIB-Boc and PIB-NH2) are shown in Figure 6.7.

Peaks due to residual exo-olefin were not observed. PIB-Boc resonances appeared at

4.92 (g′), 3.31 (f), 2.62 (e), 1.72 (c), and at 2.51 (d1) and 2.37 (d2) ppm (diastereotopic

protons, doublet of doublets). As expected, after Boc deprotection disappearance of the

secondary amine (g′) and tert butyl (h′) proton peaks occurred. Furthermore, the ultimate

methylene protons adjacent to nitrogen (f) shifted slightly upfield; whereas the

penultimate methylene protons (e) shifted downfield. Further confirmation was provided

by comparison of FTIR spectra of PIB-exo, PIB-Boc, and PIB-NH2 as shown in Figure

6.8.

Page 212: Polyisobutylene Chain End Transformations: Block Copolymer

193

To further demonstrate the versatility and modularity of the thiol-ene approach,

two additional functionalities were selected, carboxylic acid (Exp. 11) and hydroxyl

(Exp. 12). Both reactions achieved high conversions (≥ 98) in minutes (< 10 min).

Structural evidence by 1H NMR, FTIR, and GPC are provided in the supporting

information for the carboxylic acid (Figures 9-11) and hydroxyl (Figures 12-14)

functional groups.

Conclusion

To conclude, the above described is a simple, rapid synthesis of primary halogen,

amine, carboxylic acid, and hydroxyl-functional PIBs (mono- and difunctional) via thiol-

ene click chemistry. The methods produce near-quantitative functionalization within 10

min without difficult purification or reaction conditions. Reduced reaction temperature

(ice bath) facilitates increased conversion. Functionalization using cysteamine requires

Boc protection; however, the Boc protecting group may be easily removed in one pot.

Page 213: Polyisobutylene Chain End Transformations: Block Copolymer

194

References

1. Bergbreiter, D. E.;Hamilton, P. N.;Koshti, N. M. J. Am. Chem. Soc. 2007, 129, 10666.

2. Wood, V. Panzer, M. J.; Chen, J.;Bradley, S.;Halpert, J. E.; Bawendi, M. G.; Bulovic,

V. Adv. Mater. 2009, 21, 2151.

3. Pinchuk, L.; Wilson, G. J.; Barry, J. J.; Schoephoerster, R. T.; Parel, J-M; Kennedy, J.

P. Biomaterials 2008, 29, 448.

4. Kennedy, J. P.; Ivan, B.In Designed Polymers by Carbocationic Macromolecular

Engineering: Theory and Practice, Hanser, New York, 1992.

5. (a) Wilczek, L.; Kennedy, J. P. J. Poly. Sci., Part A: Polym. Chem. 25, 3255; (b)

Nielsen, L. V.; Nielsen, R. R.; Gao, B. G.; Kops, J.; Iván, B. Polymer 38, 2529; (c) Roth,

M.; Mayr, H. Macromolecules 1996, 29, 6104.

6. (a) US Pat., 5 212 248, 1991. (b) De, P.; Faust, R. Macromolecules 2006, 39, 6861.

7. (a) Simison, K. L.; Stokes, C. D.; Harrison, J. J.; Storey, R. F. Macromolecules 2006,

39, 2481; (b) Magenau, A. J. D.; Martinez-Castro, N.; Storey, R. F. Macromolecules

2009, 42, 2353.

8. (a) Hadjikyriacou, S.; Faust, R. Macromolecules 1999, 32, 6393; (b) Martinez-Castro,

N.; Lanzendörfer, M. G.; Müller, A. H. E.; Cho, J. C.; Acar, M. H.; Faust, R.

Macromolecules 2003, 36, 6985; (c) Martinez-Castro, N.; Morgan, D. L.; Storey, R. F.

Macromolecules 2009, 42, 4963; (d) Morgan, D. L.; Storey, R. F. Macromolecules 42,

6844.

9. (a) Faust, R.; Kennedy, J. P. J. Macromol. Sci., Chem. 1990, A27, 649; (b) Iván, B.;

Kennedy, J. P. J. Poly. Sci., Part A: Polym. Chem. 1990, 28, 89-104.

Page 214: Polyisobutylene Chain End Transformations: Block Copolymer

195

10. (a) Killops, K. L.; Campos, L. M.; Hawker, C. J. J. Am. Chem. Soc. 2008, 130, 5062;

(b) Becer, C. R.; Hoogenboom, R.; Schubert, U. S. Angew. Chem. Int. Ed. 2009, 48,

4900; (c) Lodge, T. P. Macromolecules 2009, 42, 3827.

11. (a) Hordyjewicz-Baran, Z.; You, L.; Smarsly, B.; Sigel, R.; Schlaad, H.

Macromolecules 2007, 40, 3901; (b) Yu, B.; Chan, J. W.; Hoyle, C. E.; Lowe, A. B. J.

Poly. Sci., Part A: Polym. Chem. 2009, 47, 3544 (c) Li, M.; De, P.; Gondi, S. R.;

Sumerlin, B. S. J. Poly. Sci., Part A: Polym. Chem. 2008, 46, 5093.

12. (a) Aimetti, A. A.; Machen, A. J.; Anseth, K. S. Biomaterials 2009, 30, 6048; (b)

Jones, M. W.; Mantovani, G.; Ryan, S. M.; Wang, X.; Brayden, D. J.; Haddleton, D. M.

Chem. Commun. 2009, 5272.

13. (a) Hoyle, C. E.; Lee, T. Y.; Roper, T. J. Poly. Sci., Part A: Polym. Chem. 2004, 42,

5301; (b) Shin, J.; Nazarenko, S.; Hoyle, C. E. Macromolecules 2009, 42, 6549. (c)

Khire, V. S.; Yi, Y.; Clark, N. A.; Bowman, C. N. Adv. Mater. 2008, 20, 3308.

14. Chan, J. W.; Yu, B.; Hoyle, C. E.; Lowe, A. B. Chem. Commun. 2008, 4959.

15. Blackborow, J. R.; Boileua, S.; Mazeaud, B. Eur. Patent EP 342 792 A1, 1989.

16. Gorski, U.; Maenz, K.; Stadermann, D. Angew Makromol Chem. 1997, 253, 51.

17. Boileau, S.; Mazezeaud-Henri, B.; Blackborrow, R. European Polymer Journal 2003,

39, 1395.

18. Martinez-Castro, N.; Morgan, D. L.; Storey, R. F. Macromolecules 2009, 42, 4963.

19. (a) Simison, K. L.; Stokes, C. D.; Harrison, J. J.; Storey, R. F. Macromolecules 2006,

39, 2481 (b) Magenau, A. J. D.; Martinez-Castro, N.; Storey, R. F. Macromolecules

2009, 42, 2353.

Page 215: Polyisobutylene Chain End Transformations: Block Copolymer

196

20. Capozzi, G.; Modena, G. In The Chemistry of the Thiol Group, ed. S. Patai, John

Wiley & Sons, Bristol, vol.2, ch. 17, pp. 832-833.

Page 216: Polyisobutylene Chain End Transformations: Block Copolymer

197

Table 6.1. Mono- and difunctional exo-olefin PIB precursors Sample Mn, GPC

(g/mol) Mn, NMR

(g/mol) Mw/Mn

f

a

(%) PIB-exo 1,300 1,268 1.05 ≥ 97

exo-PIB-exo 3,600 3,433 1.03 98-99 a – Determined by 1H NMR.

Page 217: Polyisobutylene Chain End Transformations: Block Copolymer

198

Table 6.2. Thiol-ene reactions with exo-olefin PIB Exp Thiol [ene] [SH]:[ene] t

(m) pene

e (%)

1 HS(CH2)3Cl 0.80 1.50 3.5 97 2a HS(CH2)3Cl 0.80 1.50 3.5 > 99 3a,b HS(CH2)3Cl 0.80 1.50 3.5 > 99 4 HS(CH2)2NH2 0.80 1.50 3.5 0 5 HS(CH2)2NH2 HCl 0.80 1.50 3.5 0 6 HS(CH2)2NH-Boc 0.80 1.50 3.5 93 7a HS(CH2)2NH-Boc 0.80 1.50 3.5 > 99 8a,b HS(CH2)2NH-Boc 0.80 1.50 3.5 > 99 9a,c HS(CH2)2NH-Boc 0.80 1.50 180 > 99 10a,d HS(CH2)2NH-Boc 0.80 1.50 3.5 > 99 11a HSCH2COOH 0.40 3.00 8 > 99

12a HS(CH2)5OH 0.38 4.00 6 98 a Ice bath. b Difunctional PIB. c Sunlight. d One-pot deprotection e NMR.

Page 218: Polyisobutylene Chain End Transformations: Block Copolymer

199

Figure 6.1. Photo-initiated radical addition of functional thiols (halo, NH-Boc, carboxylic acid, hydroxyl) to exo-olefin-terminated PIB and subsequent Boc deprotection.

S Rn

nhv, MeCl2

S

HN

O

On

SNH2

n

H+

HS-(CH2)1-5-R

R = Cl, NH-Boc, COOH, OH

-or-

1-5

Page 219: Polyisobutylene Chain End Transformations: Block Copolymer

200

Figure 6.2. GPC traces of mono-functional PIB before (solid line, tert-chloride-terminated PIB) and after quenching with PMP (dashed line, exo-olefin-terminated PIB).

11 12 13 14 15 16 17 18 19

time (min)

Page 220: Polyisobutylene Chain End Transformations: Block Copolymer

201

Figure 6.3. GPC traces of difunctional PIB before (solid line, tert-chloride-terminated PIB) and after quenching with PMP (dashed line, exo-olefin-terminated PIB).

10 11 12 13 14 15 16 17

time (min)

Page 221: Polyisobutylene Chain End Transformations: Block Copolymer

202

Figure 6.4. 1H NMR spectra of exo-olefin PIB (top) and PIB-Cl (bottom).

Sn Cl

g

f

e

c

fd1

d

a

b

d

d1,2 e

a

b

g

c

n

d2

5 4 3 2 1ppm

Page 222: Polyisobutylene Chain End Transformations: Block Copolymer

203

Figure 6.5. PIB exo-olefin conversion vs. time for reaction with 2-(tert-butoxycarbonylamino)ethanethiol, monitored by observing diminution of the C=C stretch at 1645 cm-1 using real-time FTIR.

0 50 100 150 200 250

0.0

0.2

0.4

0.6

0.8

1.0

1.2

exo-o

lefi

n co

nver

sion

time (s)

1660 1650 1640 1630 1620

0.2

0.3

0.4

0.5

0.6 0 6 12 24 48 96 498

a.i.

wavenumber (cm-1)

time (s)

Page 223: Polyisobutylene Chain End Transformations: Block Copolymer

204

Figure 6.6. GPC traces of monofunctional PIB before (dashed line, exo-olefin-terminated PIB) and after thiol-ene reaction (solid line, Boc-protected and halogen-terminated PIB).

12 14 16 18

time (min)

PIB-Boc

PIB-Cl

PIB-exo

Page 224: Polyisobutylene Chain End Transformations: Block Copolymer

205

Figure. 6.7. 1H NMR spectra of PIB-Boc (top) and PIB-NH2 (bottom).

7 6 5 4 3 2 1ppm

S

HN

O

On

c

f

c

c

f

ee

f

e

c ef

g′

g′

h′

h′h′

g

g

1.5 1.4 1.3 1.2 1.1 1.0 0.9

h′

SNH2

n

d1

d1,2

d2

d2

d1

d1,2

Page 225: Polyisobutylene Chain End Transformations: Block Copolymer

206

Figure 6.8. FTIR spectra of exo-olefin-terminated PIB (top), Boc-protected PIB (middle), and amine-terminated PIB (bottom).

wavenumber (cm-1)

4000 3500 3000 2500 2000 1500 1000 500

R-(NH)-(stretching)

R-(NH2)(scissoring)

1700 1650 1600

PIB-exo

PIB-Boc

PIB-NH2

=C-H(stretching)

C=C(stretching)

=C-H(bending)

C=O(stretching)

C-O(stretching)

3125 3100 3075 3050

R-(NH2)(stretching)

Page 226: Polyisobutylene Chain End Transformations: Block Copolymer

207

Figure 6.9. 1H NMR spectra of exo-olefin-terminated PIB (top) and carboxylic acid-terminated PIB (bottom).

6 5 4 3 2 1 0

ppm

d

a

b

d

a

b

c d1,2 e

e

d2d1 c

n

SnO

OH f

f

Page 227: Polyisobutylene Chain End Transformations: Block Copolymer

208

Figure 6.10. FTIR spectra of exo-olefin-terminated PIB (top) and carboxylic acid-terminated PIB (bottom).

3500 3000 2500 2000 1500 1000

wavenumber (cm-1)

C = O

PIB-COOH

PIB-exo

Page 228: Polyisobutylene Chain End Transformations: Block Copolymer

209

Figure 6.11. GPC traces of monofunctional PIB before (dashed line, exo-olefin-terminated PIB) and after thiol-ene reaction (solid line, acid-terminated PIB).

12 14 16 18

time (min)

PIB-COOH

PIB-exo

Page 229: Polyisobutylene Chain End Transformations: Block Copolymer

210

Figure 6.12. 1H NMR spectra of exo-olefin-terminated PIB (top) and hydroxyl-terminated PIB (bottom).

5 4 3 2 1

ppm

a

b

d

a

b

n

d

Sn OH3

cd1,2 e

cd2

e, d1

g

g

f

f

Page 230: Polyisobutylene Chain End Transformations: Block Copolymer

211

Figure 6.13. FTIR spectra of exo-olefin-terminated PIB (top) and hydroxyl-terminated PIB (bottom).

3500 3000 2500 2000 1500 1000

wavenumber (cm-1)

PIB-OH

PIB-exo

OH

Page 231: Polyisobutylene Chain End Transformations: Block Copolymer

212

Figure 6.14. GPC traces of monofunctional PIB before (dashed line, exo-olefin-terminated PIB) and after thiol-ene reaction (solid line, hydroxyl-terminated PIB).

12 14 16 18

time (min)

PIB-OH

PIB-exo

Page 232: Polyisobutylene Chain End Transformations: Block Copolymer

213

CHAPTER VII

TRANSFORMATIONS OF α,ω - THIOLPOLYISOBUTYLENE: EFFICIENT

ROUTES TO A CHAIN TRANSFER AGENT, TELECHELICS, AND PIB BASED

THIOURETHANES.

Introduction

Thiol based chemistries are vast and encompass a broad range of reactions.1

Utility of the thiol functional group in materials/polymer synthesis has been extensive, as

it provides a practical means for many efficient transformations and complex

architectures.2 Chain transfer agents (CTAs) for reversible addition fragmentation

transfer (RAFT) are commonly synthesized through sequential treatment of anionic

species (e.g. thiolates, alkoxides) with carbon disulfide and an alkylating agent.3,4 Shipp

et al. illustrated successful conversion of ω-bromine poly(tert-butyl acrylate) (tBA) into a

RAFT macroinitiator through substitution of the terminal bromide with the a potassium

xanthate.5 Subsequent RAFT polymerization of vinyl acetate (Vac) was conducted,

demonstrating a convenient method for synthesis poly(tBA-b-Vac) block copolymers, in

which the monomers possess large reactivity differences.

Thiol-ene “click” reactions, through either a radical or nucleophile mediated

mechanism, provide efficient hydrothiolation routes across virtually any double bond.2,6,7

Numerous examples are available in the literature for polymer end group8,9 and

backbone10 modification, and many of which are covered in two excellent reviews.6,7

The highly efficient thiol-Michael addition was shown by Finnik and coworkers to be a

facile transformation strategy.8 This was accomplished in a consecutive fashion, by

reduction of terminal CTAs of RAFT polymers and subsequent addition to an acrylate

Page 233: Polyisobutylene Chain End Transformations: Block Copolymer

214

(i.e. activated ene). Campos et al. prepared ene functional initiators and monomers, for

controlled RAFT, ATRP, and ring opening polymerization, and modified the resulting

polymer through the photoinitiated thiol-ene reactions.10 More recently, photoinitiated

radical 1,2-diaddition of thiols to alkynes (i.e. thiol-yne) has been recognized as a means

for synthesis of highly functional materials.11,12,13,14,15 Chan et al. demonstrated the

modularity and efficiency of the thiol-yne reaction to produce 16 and 8 arm

polyfunctional materials11 and Hensarling et al. functionalized polymer surface brushes

with an array of functionalized thiols.13 Other research groups have shown utility of this

chemistry for third generation dendrimers (i.e. 192 functional groups),13 highly

crosslinked networks,12 and hyperbranched polymers.15

Thiols react quickly and efficiently with isocyanates under basic catalysis, and

this reaction was been used for polymer functionalization through thiocarbamate linkages

16 and synthesis of polythiourethane.17,18 Lowe and coworkers reacted thiol-terminated

poly(N,N-diethylacrylamide), obtained by in-situ reduction of the terminal CTA of

RAFT-derived polymers, with functional isocyanates in order to study the effect of

various attached functional groups on phase behavior.16 Segmented polythiourethane

elastomers have been synthesized via the thiol-isocyanate reaction, and were reported to

exhibit microphase separation between hard and soft segments.17 Two glass transition

temperatures were observed, corresponding to the hard and soft-segment phases

respectively. The hard segment Tg was observed at 132 °C, and the soft segment Tg

ranged from -57 to -23 °C, depending on the thiourethane composition.

Preparation of thiol functional polymers has been accomplished by a variety of

synthetic pathways. A popular approach has been the attachment to a polymer or

Page 234: Polyisobutylene Chain End Transformations: Block Copolymer

215

initiator, for example via an ester linkage, of a molecule such as mercaptoacetic acid or 2-

mercaptoethanol, in which the thiol group has been protected with a 2,4-dinitrophenyl

moiety. Protection is typically carried out by prior reaction of the thiol with Sangers

reagent, i.e. 2,4-dinitrofluorobenzene, in the presence of a base catalyst. In this way,

protected thiol initiators for ATRP19 and ring opening polymerization of ε-caprolactone,20

were developed, and the presence of the protected thiol was reported to have no

observable deleterious effects to polymerization. Using a similar approach, poly(ε-

caprolactone) and poly(ethyleneoxide) were fitted with thiol end groups via reaction of

the terminal hydroxyl group on the polymer with 2,4-dinitrophenylmercaptoacetic acid,

following by deblocking.21 Tsarevsky and Matyjaszewski developed a thio-functional

ATRP initiator based on 2-mercaptoethyl 2-bromopropionate, in which the thiol group

was blocked by simple disulfide formation (dimerization). After polymerization,

polymers containing a thiol head group were obtained by reduction.22 Terminal thiol

functionality is inherent to RAFT-prepared polymers. CTAs in the presence of primary

amines readily reduce into the corresponding thiol.8,9,23 Disulfide formation has been

observed in these systems, but can be overcome by incorporation of phosphorus

oxidizing agents.23 Successful substitution of ATRP prepared ω-bromo polystyrene (p ≈

95 %),24 and a pendent chlorine functional polystyrene derivative with thiol functionality

through thiourea has also been demonstrated.25 However, thiols have yet to be attached

to PIB chain ends and preparation of this telechelic polymer has essentially been

overlooked. Owing to the high nucleophilicity of the sulfur atom, thiols can be produced

by the nucleophilic substitution of a halogen, using NaSH26, dimethylthioformamide,27

and thiourea28. Herein, we report the use of thiourea to produce near quantitative

Page 235: Polyisobutylene Chain End Transformations: Block Copolymer

216

difunctional thiol-terminated PIB from a bromine-terminated precursor. Subsequent

utility of this telechelic polymer is demonstrated with a variety of chemistries for

synthesis of a novel RAFT macroinitator, alkyne terminated and tetrahydroxyl terminated

PIB, and PIB based thiourethane polymers.

Experimental

Materials

Hexane (anhydrous, 95%), 2,6-lutidine (redistilled, 99.5%), TiCl4 (99.9%,

packaged under N2 in sure-seal bottles), 3-phenoxypropyl bromide (96%), chloroform-d

(99.8% atom % D), dimethylformamide (DMF) (99%), thiourea (≥ 99%), propargyl

acrylate (98%), dimethylphenylphosphine (DMPP) (97%), 6-mercapto-1-hexanol (97%),

carbon disulfide (≥ 99.9 %), 2-bromopropionic acid (≥ 99 %), triethylamine (≥ 99%)

(TEA), 1,6-hexanedithiol (Kosher grade, ≥ 97%), tetrahydrofuran-d8 (99.5 atom % D),

and 4,4′-methylenebis(phenyl isocyanate) (MDI) were purchased from Sigma-Aldrich

and used as received. Methanol (ACS-grade, 99.9%), hexanes (ACS-grade, 99.9%), and

heptane (HPLC-grade, 99.5%) were purchased from Fisher Scientific and used as

received. THF (HPLC-grade, 99.9%) was purchased from Fisher Scientific and distilled

over CaH2 prior to use. Free radical photoinitiator, 2.2-dimethoxy-2-phenylacetophenone

(DMPA), was purchased from CIBA and used as received. Isobutylene (IB) (BOC,

Gases, 99.5%) and CH3Cl (Alexander Chemical Corp., 99.95%) were dried by passing

the gaseous reagent through columns packed with CaSO4 and CaSO4/4 Å molecular

sieves, respectively, and condensed within a N2-atmosphere glove box immediately prior

to use. Cationic polymerization initiator 1,3-bis-(2-chloro-2-propyl)-5-tert-butylbenzene

Page 236: Polyisobutylene Chain End Transformations: Block Copolymer

217

(bDCC) and α,ω-bis[4-(3-bromopropoxy)phenyl]polyisobutylene (α,ω PIB-Br) were

synthesized as previously described9,14.

Synthesis of α,ω-Bis[4-(3Thiopropoxy)phenyl]polyisobutylene (α,ω PIB-SH)

To a dry one-neck 100 mL round-bottom flask equipped with a magnetic stir-bar

were added 0.7929 g (10.42 mmol) thiourea, 21 mL DMF, and 10 mL heptane. The flask

was placed under a slow N2 purge and immersed in an oil bath heated to 90°C. In a

scintillation vial, α,ω PIB-Br (Mn = 2,900 g/mol, 1.250 g, 0.43 mmol) was dissolved in

11 mL heptane and added to the reaction flask. The solution was turbid, yet remained

monophasic throughout the reaction. After 6 h, the oil bath temperature was raised to

110 °C, and the reaction became a clear homogenous solution. After equilibrating at the

higher temperature for 15 min, a solution of 0.324 g (8.1 mmol) NaOH and 2 mL

deionized water was injected into vessel, at which time vigorous gas evolution occurred

and a white precipitate formed. Approximately 4 h later, the vessel was removed from

the oil bath and allowed to cool with stirring for 20 min. Then, 0.8 mL neat H2SO4 was

charged to the reaction vessel. After 30 min, the biphasic reaction mixture was placed in

a separation funnel and the DMF/H2O layer was immediately drained. The remaining

organic layer was diluted with hexanes, washed three times with MeOH, and precipitated

into MeOH. The precipitate was collected in hexanes, and the solvent was stripped under

vacuum to yield the final, α,ω PIB-SH.

Synthesis of α,ω-bis{4-[3-(Carboxyethylidenesulfanylthiocarbonylsulfanyl)propoxy] -

phenyl}polyisobutylene (α,ω PIB-CTA)

To a scintillation vial, under an N2 atmosphere, were added 0.1972 g (Mn = 2900

g/mol, 0.068 mmol) α,ω PIB-SH and 0.19 mL carbon disulfide (3.2 mmol). After a

Page 237: Polyisobutylene Chain End Transformations: Block Copolymer

218

homogenous solution was obtained, 0.22 mL of TEA (1.6 mmol) was added resulting in a

clear yellow solution. The contents were allowed to stir for 7 h, at which time 71.0 µL

(0.79 mmol) of 2-bromopropionic acid followed by 0.42 mL chloroform were charged to

the vial. Upon addition, the solution briefly turned red and later returned to its original

yellow color. This solution was allowed to stir for 12 h, and afterward, the solvent was

removed under reduced pressure. The remaining contents were extracted with hexanes

(3X) and filtered, and the resulting organic solution was washed with 1 N HCl (3X) and

then deionized water (3X). Hexanes were removed under reduced pressure to obtain a

viscous yellow polymer.

Synthesis of α,ω-bis{4-[3-(Propargyloxycarbonylethylenesulfanyl)propoxy]phenyl}-

polyisobutylene (α,ω PIB-alkyne)

To a scintillation vial were added 0.2031 g (Mn = 2900 g/mol, 0.070 mmol) α,ω

PIB-SH and 1.0 mL THF (or chloroform). After a homogenous solution was obtained,

15.9 µL of propargyl acrylate (0.144 mmol) and then 1.0 µL DMPP (7.0 µmol) were

charged to the vial. The contents were allowed to stir for 30 min, at which time the

solvent was removed under reduced pressure. The remaining contents were taken up into

hexanes and precipitated twice into methanol. The polymer was then redissolved in

hexanes, and the solvent stripped under vacuum to yield the final, α,ω PIB-alkyne.

Sequential Nucleophilic Thiol-ene/Radical Thiol-yne Reactions to Produce Tetrahydroxy-

functional PIB (α,ω PIB-(OH)2)

To a scintillation vial were added 0.190 g (Mn = 2900 g/mol, 0.065 mmol) α,ω

PIB-SH and 1.0 mL d-chloroform. After a homogenous solution was obtained, 15.4 µL

of propargyl acrylate (0.140 mmol) and then 0.9 µL DMPP (6 µmol) were charged to the

Page 238: Polyisobutylene Chain End Transformations: Block Copolymer

219

vial. The contents were allowed to stir for 30 min and then transferred into another

scintiallation vial containing 0.029 g DMPA (∼ 1.6 wt%) and 91.0 µL 6-mercaptohexanol

(0.665 mmol). Afterward, the vial was irradiated using a medium pressure Hg lamp

(light intensity ∼6.00 mW/cm2) for 6 min. Chloroform was removed under reduced

pressure, and the crude reaction mixture was twice precipitated into chilled methanol

from hexanes, using centrifugation to aid settling. The final precipitate was then

collected in hexanes, and the solvent was stripped under vacuum to yield the final, α,ω

PIB-(OH)2.

Synthesis of α,ω-bis{4-[3-(Phenyliminocarbonylsulfanyl)propoxy]phenyl}polyisobutylene

(α,ω PIB-phenyl thiourethane)

Two drops of neat phenyl isocyanate were added to a solution, consisting of 0.091

g (Mn = 2900 g/mol, 0.031 mmol) α,ω PIB-SH, 1 µL TEA (7 µmol), and 0.3 mL d-THF,

within a 5 mm NMR tube, and the reaction was allowed to mix for 10 min. After 1H

NMR analysis, to ensure complete conversion of the thiol group, the solvent was

removed under reduced pressure. The remaining contents were taken up in hexanes, and

the resulting solution was filtered, washed with 1 N HCl (2X), and then precipitated into

methanol. The final precipitate was then collected in hexanes, and the solvent was

stripped under vacuum to yield the final, α,ω PIB-phenyl thiourethane.

Synthesis of Polyisobutylene-based Polythiourethane Without Chain-extending Dithiol

Under a dry nitrogen atmosphere, a solution of 0.297 g (Mn = 2,900 g/mol, 0.102

mmol) α,ω PIB-SH, 5 µL TEA (0.036 mmol), and 0.7 mL dry THF was prepared in a

scintillation vial equipped with a magnetic stir bar. In a separate scintillation vial, 0.026

g MDI (0.10 mmol) was dissolved in 0.3 mL dry THF, and after dissolution, the contents

Page 239: Polyisobutylene Chain End Transformations: Block Copolymer

220

were transferred into the first scinitillation vial. The sealed vial was then submerged into

an oil bath at 50 °C with stirring. After two hours, the vial was removed from the oil

bath. The contents were precipitated twice into methanol from a THF, and the final

precipitate was collected and dried under vacuum.

Synthesis of Polyisobutylene-based Polythiourethane with Chain-extending Dithiol (PIB-

20-thiourethane)

Chain-extended PIB-based polythiourethanes were prepared by a two-step, one-

prepolymer method. A solution of 0.083 g MDI (0.332 mmol) in 0.8 mL dry THF was

prepared in a 25 mL round bottom flask equipped with a magnetic stir-bar. A second

solution consisting of 0.5005 g (Mn = 2,900 g/mol, 0.17 mmol) α,ω PIB-SH and 10 µL

TEA (0.072 mmol) in 2.5 mL THF was then added dropwise to the flask over a period of

30 min at room temperature. A third solution consisting of 0.025 mL (0.163 mmol) 1,6-

hexanedithiol in 2.4 mL THF was then added dropwise over a period of 15 min. This

rate of addition was sufficiently slow to maintain the reaction temperature below 23 °C.

After addition of the dithiol solution, the reaction was allowed to stir for 1 h at room

temperature. Afterward, solvent was removed under reduced pressure. The contents

were taken up in THF, precipitated into hexanes, and centrifuged. A clear solid polymer

resulted.

Results and Discussion

Synthesis of α,ω PIB-SH

The initial objective of this research was to development a practical and

inexpensive method for synthesis of α,ω PIB-SH. Thiol functionality is commonly

obtained through alkyl halide substrates, thus difunctional primary bromide PIB was first

Page 240: Polyisobutylene Chain End Transformations: Block Copolymer

221

prepared. 3-Bromopropoxyphenyl-terminated PIB was synthesized as previously

reported;29 characterization of the resulting difunctional polymer revealed quantitative

conversion of tert-chloride end groups into primary bromide functionality. 1H NMR

spectrum (Figure 7.1) of the purified α,ω PIB-Br was devoid of resonances associated

with the gem-dimethyl and methylene protons adjacent to the tert-chloride moieties,

located at 1.68 and 1.96 ppm, respectively. New resonances characteristic of the

trimethylene tether of the 3-bromopropoxyphenyl end group were visible at 4.09 (c), 3.61

(a), and 2.32 (b) ppm. Integration of the latter resonances relative to those of the

aromatic initiator revealed essentially quantitative functionality (f = 1.00-0.99). In

addition, elimination products (i.e. exo and endo olefin) were absent. 13C NMR analysis

(Figure 7.2) bolstered the 1H NMR results, indicating quantitative capping of the polymer

chain ends. Molecular weight values for the 3-bromopropoxyphenyl-terminated PIB

determined by GPC and NMR were in excellent agreement: 2,900 g/mol (GPC-MALLS,

known dn/dc), 2,800 g/mol (GPC-MALLS, 100% mass recovery dn/dc), and 2,990 g/mol

(H NMR integration). For all subsequent experimentation a molecular weight value of

2,900 g/mol was used. GPC traces (Figure 7.3) of the pre-quench and post-quench

polymer display no detectable coupling and narrow, symmetrical molecular weight

distributions (PDI = 1.08).

Conversion of the terminal halogen into thiol functionality was first attempted

using NaSH in a heptane/DMF cosolvent system at 90 °C. Substitution with NaSH

would allow a direct one-step reaction and low material costs. Results of these

preliminary experiments indicated complete displacement of terminal bromine; however,

a significant fraction of the product was coupled through formation of disubstitued

Page 241: Polyisobutylene Chain End Transformations: Block Copolymer

222

sulfide (f mole ≥ 0.14). Reducing the reaction temperature to room temperature

significantly reduced sulfide formation (f mole ≤ 0.08) but required unreasonably long

reaction times (e.g. > 63 h) for full halide displacement.

The use of thiocarbonyl-based nucleophiles, particularly thiourea, has been

reported to effectively reduce or eliminate sulfide formation.1 Reaction of thiourea and

subsequent deblocking are shown in Scheme 7.1. First the alkylisothiouronium salt was

produced using a 1:1 (v:v) DMF:heptane cosolvent mixture at 90°C. Hydrolysis of the

salt by aqueous base produced thiolate chain ends, which were then acidified to form the

desired thiol functional group. Upon completion of the thiourea reaction, all three

methylene tether resonances of the α,ω PIB-Br shifted upfield (Figure 7.4, inset). A

strong shift from 3.61 to 2.74 ppm was observed for the methylene protons adjacent to

the halogen, due to the replacement of the inductively withdrawing bromine with the less

electronegative sulfur (i.e. increased shielding). In addition, the splitting pattern changed

from a triplet to a quartet. Inspection of the methylene protons two and three units away

from the substitution showed less pronounced upfield shifts of 2.32 to 2.08 ppm (b) and

4.09 to 4.06 ppm (c), respectively. 13C NMR analysis (Figure 7.5) showed a dramatic

shift of the methylene carbon adjacent to bromine from 30.2 ppm to 21.5 ppm (1);

whereas lesser shifts were observed for methylene carbons further away, from 32.7 to

33.7 ppm (2) and 65.4 to 65.7 ppm (3). Assignments for the precursor α,ω PIB-Br and

product α,ω PIB-SH were made using two dimensional gHSQC NMR experiments

(Figures 7.6 and 7.7).

Initial attempts using the thiourea approach were not successful. It was found that

reaction temperature had a significant influence on sulfide formation during the base

Page 242: Polyisobutylene Chain End Transformations: Block Copolymer

223

hydrolysis step. Preliminary experimentation involved cooling the reaction vessel, by

removal from the oil bath, prior to base addition. GPC analysis (Figure 7.8) showed a

significant fraction of high molecular weight species and multimodal molecular weight

distribution (solid line) when hydrolysis was carried out at low temperature. Maintaining

the reaction temperature at 90°C, or preferably raising it to 110°C, dramatically

suppressed sulfide formation, as evident by the resulting monomodal elution peak (dotted

line). Sulfide formation could also be detected by 1H NMR (Figure 7.8 inset), by the

appearance of a triplet due to the methylene protons adjacent to the sulfide linkage, which

appear slightly upfield from the quartet characteristic of the methylene protons adjacent

to the desired thiol group. 1H NMR integration revealed that a bath temperature of 110°C

(internal reaction temperature ≈ 99°C) reduced sulfide formation to approximately 1 mole

fraction (fmole ≈ 0.01).

Kinetics of Isothiouronium Salt Formation

Isothiouronium salt formation was probed to optimize reactions condition. In this

study, a reaction vessel was charged using the same reactant concentrations listed in the

experimental with aliquots drawn at regular intervals. Solvent was removed from the

aliquots in a vacuum oven at 40°C over several days. Disappearance of the α,ω PIB-Br

triplet at 3.61 ppm and the appearance of a new peak attributed to the isothiouronium salt

at 3.44 ppm was monitored by 1H NMR. Conversion of the bromine-terminated PIB to

isothiouronium salt PIB vs. time is shown in Figure 7.9. The data indicate that complete

conversion of the bromine chain ends occurred within ∼ 200 min. The reaction contained

a large excess of thiourea (i.e. 12.7 X bromine chain ends), and thus pseudo first-order

kinetics was expected. As shown in Figure 7.9 inset, a semi-logarithmic first-order plot

Page 243: Polyisobutylene Chain End Transformations: Block Copolymer

224

was indeed linear, with an observed rate constant of k = 0.025 s-1, confirming pseudo

first-order behavior.

Site Transformations of α,ω PIB-SH

The thiol group is a versatile functional group for various chain end

transformations. In this section, synthesis of a RAFT macro-CTA, alkyne terminal PIB,

tetrafunctional telechelic PIB, and PIB-based segmented thermoplastic polythiourethanes

(TPTUs) will be demonstrated. Telechelic PIBs and PIB-based block copolymers,

particularly segmented TPUs and TPTUs have many potential applications in material

and polymer science.

Earlier chapters in this document have focused on novel strategies to combine

quasiliving cationic polymerization of isobutylene and reversible addition fragmentation

(RAFT) polymerization. In an effort to directly functionalize PIB with a chain transfer

agent without using coupling chemistry, e.g., catalyzed esterification30 or click

chemistry,31 thiol-terminated PIB was of interest. Thiol substrates are commonly used in

synthesis of primary and secondary trithiocarbonates.3,4,32 Typically, trithiocarbonate

salts are formed in situ from a thiol and carbon disulfide in the presence of triethylamine,

and later reacted with an alkyating agent. Many examples of this are available in the

literature for synthesis of asymmetrical trithiocarbonates with different alkyating agents,

a few of which are bromopropionyloxy derivatives,33 1-(bromoethyl)benzene,34 and N-

(bromomethyl)phthalimide.35

α,ω PIB-SH was thus employed in a pseudo-bulk reaction with carbondisulfide in

the presence of triethylamine with 2-bromopropionic acid, using a modification of a

procedure given by Evans et al.33. Carbon disulfide served both as electrophile for the

Page 244: Polyisobutylene Chain End Transformations: Block Copolymer

225

thiolate chain ends and as reaction solvent. PIB was observed to have complete solubility

in TEA and carbondisulfide. Therefore, bulk conditions were selected to promote rapid

formation of the trithiocarbonate salt prior to addition of the 2-bromopropionic acid;

when more diluted conditions were used in chloroform, low chain end conversions

resulted. Upon dissolution of the α,ω PIB-SH in TEA and carbondisulfide the resulting

solution turned a clear yellow color, and during the alkylation step, the reaction briefly

turned a red color although eventually returning to its original yellow color. After

purification, 1H NMR showed a strong shift in the methylene protons adjacent to the thiol

group, 2.74 ppm (quartet) to 3.58 (triplet), and two additional resonances associated with

the S-propionic acid chain end (Figure 7.10). These two additional resonances, a quartet

at 4.88 ppm and a doublet at 1.64 ppm, corresponded with proton assignments for similar

molecular.33

Further utility of α,ω PIB-SH was shown by converting the thiol telechilic PIB

into α,ω PIB-alkyne with propargyl acrylate. Azide and alkyne functional groups are

desirable because they can be used in robust and efficient click chemistry reactions;

however, a practical means of PIB functionalization with the latter has yet to be

demonstrated. Synthesis of α,ω PIB-alkyne was accomplished by using the recently

published phosphine catalyzed thiol-ene Michael addition reaction.11 Mild reaction

conditions were used, i.e. without exclusion of oxygen and moisture, 3 % excess

propargyl acrylate, and catalytic concentration of DMPP, to produce 100 % alkyne

terminated PIB. Reaction rates were fast, with complete conversion in less than 5

minutes. 1H NMR was used to verify the structure of the reaction product, as shown in

Figure 7.11. Similar to the α,ω PIB-CTA spectrum, the signal of the methylene protons

Page 245: Polyisobutylene Chain End Transformations: Block Copolymer

226

adjacent to thiol (a) changed from a quartet to a triplet; however its chemical shift

changed very little. Four additional resonances located at 4.71 (singlet, -OCH2CCH,

2H), 2.83 (triplet, -SCH2CH2COO-, 2H), 2.68 (triplet, -SCH2CH2COO-, 2H), and 2.48

(singlet, -OCH2CCH, 1H) ppm were present. 13C NMR spectroscopy supplied further

structural evidence of the thiol-ene Michael addition, as shown in Figure 7.12. The

original α,ω PIB-SH resonances at 65.70 (3), 33.67 (2), and 21.50 (1) were absent, and

new resonances for the corresponding nuclei in the product were observed at 66.19 (3),

29.47 (2), and 28.90 (1) ppm. Peak assignments were made using a combination of

small-molecule analogy, NMR predictive software, and qHSQC NMR spectroscopy

(Figure 7.13).

An inherent advantage of the nucleophilic thiol-ene reaction (Michael addition) is

the ability to perform a one-pot, sequential thiol-ene/thiol-yne reaction to fabricate highly

functional materials.11 As demonstrated above, the Michael reaction is tolerant to the

presence of an alkyne, and subsequent irradiation of the alkyne and two equivalents of a

thiol, in the presence a photoinitiator, has been reported to yield exclusively36 and

quantitatively the 1,2-diaddition product.11 To demonstrate this procedure in the present

case, a formulation was prepared which closely mimicked that which was used for the

earlier-described alkyne-functionalization, excepting the solvent. d-Chloroform was used

here to facilitate simple 1H NMR characterization and provide a homogenous reaction

mixture. Once the thiol-Michael reaction was completed, 6-mercaptohexanol and DMPA

(i.e. photoinitiator) were charged to the reactor. The contents were then exposed to

irradiation for 6 min and purified. Figure 7.14 is a 1H NMR spectrum of the purified

polymer with insets providing an expansion and comparison of the α,ω PIB-alkyne

Page 246: Polyisobutylene Chain End Transformations: Block Copolymer

227

precursor (Figure 7.14 (a)) to the tetrahydroxyl functional PIB product (Figure 7.14 (b)).

Protons directly associated with the alkyne moiety (q and o) were absent in the final

product, whereas, a series of new protons associated with the 1,2-thiolether addition

product appeared. The assignments given in Figure 7.14 were based on those reported in

literature for an analogous polyol made from 6-mercaptohexanol, via the sequential thiol-

ene/thiol-yne reaction.11 Diastereotopic protons (q) were visible at 4.27 and 4.24 ppm,

due to the newly formed chiral center (*).

Arguably the most interesting and potentially useful application for a PIB-based

polythiol is as a soft segment in polythiolurethanes. PIB-based polyurethanes/polyureas

have received significant attention as an emerging biomaterial;37,38 with PIB as a key

component of these materials offering biocompatibility, and hydrolytic and oxidative

stability. The reaction of thiols with isocyanates, as explored by Hoyle and coworkers,

has been found to be efficient and rapid,16,17 while the resulting materials possessed many

unique mechanical and physical properties.17,18 Therefore, segmented PIB-based

thiourethanes may serve as a potential new class of TPU biomaterials.

Model reactions were first explored between α,ω PIB-SH and phenyl isocyanate.

Triethylamine, at approximately 0.2 wt%, was used as a catalyst in the presence of excess

phenyl isocyanate. Quantitative conversion of the thiol functional group was observed in

less than 10 min as determined by 1H NMR. A spectrum of the purified product (bottom)

and the α,ω PIB-SH (top) precursor is shown in Figure 7.15. Signals for each of the

tether protons (a, b, c) were visible, as well as characteristic peaks of the phenyl thioester

carbamate species (f, g, h, i).

Page 247: Polyisobutylene Chain End Transformations: Block Copolymer

228

After establishing the efficiency of the thiol-isocyanate reaction, PIB-based

thiourethane polymers were synthesized, with and without a small-molecule dithiol chain

extender. In the former case, stoichiometric amounts of α,ω PIB-SH and MDI, and

triethylamine (0.3 wt%), were reacted under an dry nitrogen atmosphere at 50°C for 2 h.

After the step growth polymerization was complete a substantially more viscous reaction

medium resulted. GPC was used to characterize the polythiourethane relative to the α,ω

PIB-SH precursor (Figure 7.16). The resulting chromatogram (b) indicated that fairly

monodisperse polymer was produced (PDI = 1.39) and a large increase in molecular

weight (Mn = 63,200 g/mol) relative to the precursor (a) was apparent from the decreased

elution time. Next, the reaction was conducted in the presence of a small-molecule

dithiol chain extender, in order to create a polythiourethane with phase-separated hard-

segment/soft-segment morphology. A typical two–step, one-prepolymer procedure was

employed. First, α,ω PIB-SH was reacted with excess MDI, followed by addition of 1,6-

hexanedithiol chain extender. The stoichiometry between isocyanate and total thiol was

1:1, and the recipe was designed to yield 20 wt% hard segment. The resulting segmented

polythiolurethane was insoluble in typical solvents suitable for PIB and was therefore

purified by precipitation from THF into hexanes. The purified polymer had a significant

increase in molecular weight, to 77,600 g/mole, and a fairly low PDI of 1.45 (Figure 7.16

(a)).

Page 248: Polyisobutylene Chain End Transformations: Block Copolymer

229

Conclusions

Synthesis of thiol-terminated PIB was accomplished through a three step, one-pot

process. Quantitative conversion of primary bromide to thiol was shown using 1H NMR

and 13C NMR spectroscopy. The utility of this telechelic polymer as an efficient

precursor to a RAFT CTA, various multifunctional telechelics, and PIB-based segmented

polythiolurethane was demonstrated. The RAFT CTA was synthesized by a base-

catalyzed alkylation reaction. Alkyne- and tetrahydroxy-functional PIB were synthesized

through the highly efficient thiol-ene and sequential thiol-ene/thiol-yne click reactions.

Polythiourethanes were synthesized through the base catalyzed thiol-urethane reaction.

Page 249: Polyisobutylene Chain End Transformations: Block Copolymer

230

References

1. Patai, S. The Chemistry of the Thiol Group, Parts 1-2, John Wiley and Sons, Inc., New

York, 1974.

2. Lowe, A. Polym. Chem., 2010, DOI: 10.1039/b9py00216b

3. Moad, G.; Rizzardo, E.; Thang, S. H. Aust. J. Chem. 2005, 58, 379-410

4. Moad, G.; Rizzardo, E.; Thang, S. H. Aust. J. Chem. 2006, 59, 669-692.

5. Petruczok, C. D.; Barlow, R. F.; Shipp, D. A. J. Polym. Sci.: Part A: Poly. Sci. 2008,

46, 7200-7206.

6. Sumerlin, B. S.; Vogt, A. P. Macromolecules 2009, ASAP Article, DOI:

10.1021/ma901447e

7. Iha, R. K.; Wooley, K. L.; Nystrom, A. M.; Burke, D. J.; Kade, M. J.; Hawker, C. J.

Chem. Rev. 2009, 109, 6520-5686

8. Qui, X-P.; Winnik, F. M. Macromol. Rapid. Commun. 2006, 27, 1648-1653.

9. Li, M. De, P.; Gondi, S. R.; Sumerlin, B. S. J. Polym. Sci., Part A: Polym. Chem.,

2008, 46, 5093-5100.

10. Campos, L. M.; Killops, K. L.; Sakai, R.; Paulusse, J. M. J.; Damiron, D.;

Drockenmuller, E.; Messomore, B. W.; Hawker, C. J. Macromolecules 2008, 41, 7063-

7070.

11. Chan, J. W.; Hoyle, C. E.; Lowe, A. B. J. Am. Chem. Soc., 2009, 131, 5751-5753.

12. Fairbanks, B. D.; Scott, T. F.; Kloxin, C. J.; Anseth, K. S.; Bowman, C. N.

Macromolecules 2009, 42, 211-217.

13. Chen, G.; Kumar, J.; Gregory, A.; Stenzel, M. H. Chem. Commun. 2009, 6291-6293.

Page 250: Polyisobutylene Chain End Transformations: Block Copolymer

231

14. Hensarling, R. M.; Doughty, V. A.; Chan, J. W.; Patton, D. K. J. Am. Chem. Soc.

2009, 131, 14673-14675.

15. Konkolewicz, D.; Gray-Weale, A.; Perrier, S. J. Am. Chem. Soc. 2009, 131, 18075-

18077.

16. Li, H.; Yu, B.; Matsushima, H.; Hoyle, C. E.; Lowe, A. B. Macromolecules 2009, 42,

6537-6542.

17. Shin, J.; Matsushima, H.; Chan, W. C.; Hoyle, C. E. Macromolecules 2009, 42, 3294-

3301.

18. Li, Q.; Zhou, H.; Wicks, D. A.; Hoyle, C. E.; Magers, D. H.; McAlexander, H. R.

Macromolecules 2009, 42, 1824-1833.

19. Carrot, G.; Hilborn, J.; Hedrick, J. L.; Trollsas, M. Macromolecules 1999, 32, 5171-

5173.

20. Carrot, G.; Hilborn, J. G.; Trollsas, M.; Hedrick, J. L. Macromolecules 1999, 32,

5264-5269.

21. Trollsas, M.;Hawker, C. J.; Hedrick, J. L. Macromolecules 1998, 31, 5960-5963.

22. Tsarevsky, N. V.; Matyjaszewski, K. Macromolecules 2002, 35, 9009.

23. Chan, J. W.; Yu, B.; Hoyle, C. E.; Lowe, A. D. Chem. Commun. 2008, 4959-4961

24. Garamszegi, L.; Donzel, C.; Carrot, G.; Nguyen, T. Q.; and Hilborn, J. Reactive &

Functional Polymers 2003, 55, 179-183.

25. Dong, Y.; Lu, J.; Xu, Q. Journal of Macromolecular Science, Part A, Pure and

Applied Chemistry 2008, 45, 37-43.

26. Ellis, L. M., Jr.; Reid, E. E. J. Am. Chem. Soc. 1932, 54, 1674-1687.

Page 251: Polyisobutylene Chain End Transformations: Block Copolymer

232

27. Yamashita, K.; Saba, H.; Tsuda, K. J. Macromol. Sci., Chem. 1989, A26, 1291-1304.

28. Cossar, B. C.; Fournier, J. O.; Fields, D. L.; Reynolds, D. D. J. Org. Chem. 1962, 27,

93-95.

29. Morgan, D. L.; Storey, R. F. Macromolecules 2009, 42, 6844-6847.

30. Magenau, A. J. D.; Martinez-Castro, N.; Storey, R. F. Macromolecules 2009, 42,

2353-2359.

31. Magenau, A. J. D.; Martinez-Castro, N.; Savin, D. A.; Storey, R. F. Macromolecules

2009, 42, 8044-8051.

32. Barner-Kowollik, C. In Handbook of RAFT Polymerization; Wiley: Weinheim, 2008;

Chapter 6.

33. Malic, N.; Evans, R. A. Aust. J. Chem. 2006, 59, 763-771.

34. Postma, A.; Davis, T. P.; Evans, R. A.; Li, G.; Moad, G.; O’Shea, M.

Macromolecules, 2006, 39, 5293-5306.

35. Mayadunne, R. A.; Moad, G.; Rizzardo, E. Tetrahedron Lett. 2002, 43, 6811-6814.

36. Saucer, J. C. J. Am. Chem. Soc. 1957, 79, 5314,

37. Jewrajka, S. K.; Yilgor, E.; Yilgor, I.; Kennedy, J. P. Journal of Polymer Science:

Part A: Polymer Chemistry 2009, 47, 38-48.

38. Umaprasana, O.; Kulkarni, P.; Faust, R. Polymer 2009, 50, 3448-3457.

Page 252: Polyisobutylene Chain End Transformations: Block Copolymer

233

Figure 7.1. 1H NMR Spectrum of α,ω-bis[4-(3-bromopropoxy)phenyl] PIB.

O Brn

8 7 6 5 4 3 2 1 0

gh

a

b

cde

fi

j

j

g

fi

h

e d c

b

a

δ ppm

a

Page 253: Polyisobutylene Chain End Transformations: Block Copolymer

234

Figure 7.2. 13C NMR Spectrum of α,ω-bis[4-(3-bromopropoxy)phenyl] PIB.

O Brn

160 140 120 100 80 60 40 20

13

12

1312

δ ppm

10

11

1110

3

2

74

2

1

5

4 1356

6

7

34 33 32 31 30 29A

Page 254: Polyisobutylene Chain End Transformations: Block Copolymer

235

Figure 7.3. GPC traces of prequenched (dotted line) and quenched (solid line) PIB

10 11 12 13 14 15 16 17 18

time (min)

Page 255: Polyisobutylene Chain End Transformations: Block Copolymer

236

Figure 7.4. 1H NMR spectra of α,ω-bis[4-(3-thiopropoxy)phenyl] PIB with comparison of the endgroup regions to the precursor bromine-terminated PIB.

O SHn

8 7 6 5 4 3 2 1 0

ba

'b

'a

c

'c

δ ppm

j

e

dc

a b

fi

h gc a

b

def

gj

i

4.0 3.5 3.0 2.5 2.0

HS-PIB-SH

Br-PIB-Br4.0 3.5 3.0 2.5 2.0

Page 256: Polyisobutylene Chain End Transformations: Block Copolymer

237

Figure 7.5. 13C NMR of α,ω-bis[4-(3-thiopropoxy)phenyl]polyisobutylene with comparison of the endgroup chemical shifts to those of the bromine-terminated PIB.

160 140 120 100 80 60 40 20

76 5

4

56

74

2

3

'3

'2

2

1

'1

3

δ ppm

1

66 65

HS-PIB-SH

Br-PIB-Br

66 65

34 32 30 22 21

34 32 30 22 21

O SHn

HS-PIB-SH

Br-PIB-Br

7

Page 257: Polyisobutylene Chain End Transformations: Block Copolymer

238

Figure 7.6. gHSQC of α,ω-bis[4-(3-bromopropoxy)phenyl] PIB.

O Brn

Page 258: Polyisobutylene Chain End Transformations: Block Copolymer

239

Figure 7.7. gHSQC of α,ω-bis[4-(3-thiopropoxy)phenyl] PIB.

O SHn

Page 259: Polyisobutylene Chain End Transformations: Block Copolymer

240

Figure 7.8. Temperature effects on sulfide formation during base hydrolysis step.

10 12 14 16 18

time (min)

3.0 2.9 2.8 2.7 2.6

3.0 2.9 2.8 2.7 2.6

δ ppm

Page 260: Polyisobutylene Chain End Transformations: Block Copolymer

241

Figure 7.9. Conversion of bromine terminated PIB to isothiouronium salt terminated PIB versus time.

0 50 100 150 200 250 3000.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

conv

ersi

on

time (min)

0 20 40 60 80 100 1200.0

1.0

2.0

3.0

Ln(

[Br 0]/

[Br]

)

Page 261: Polyisobutylene Chain End Transformations: Block Copolymer

242

Figure 7.10. 1H NMR of α,ω PIB-CTA.

7 6 5 4 3 2 1

O SH

O S S

S

COOH

de

abcde

e d

b

c

g

a

e dbc

g

f

a

δ ppm

f

7 6 5 4 3 2 1

bac

Page 262: Polyisobutylene Chain End Transformations: Block Copolymer

243

Figure 7.11. 1H NMR of thiol terminated PIB (top) and alkyne terminated PIB (bottom).

7 6 5 4 3 2 1

O SH

O

OO S

a

de

abcde

e db

c

a rs

e dbc

q

o

a s r q

δ ppm

o

7 6 5 4 3 2 1

bac

2.9 2.8 2.7 2.6

2.9 2.8 2.7 2.6

Page 263: Polyisobutylene Chain End Transformations: Block Copolymer

244

Figure 7.12. 13C NMR of alkyne terminated PIB with expansion of alkyne PIB (bottom) and expansion of thiol terminated PIB (top).

O

OO S

180 160 140 120 100 80 60 40 20

11

11

9

3

10

76 5

43

213

1210

8

δ ppm

1

35.0 32.5 30.0 27.5 22 21 20

HS-PIB-SH

1

2

9

8

56

74

212 13

1

78 77 76 67 66 65

8

78 77 76 67 66 65

3

35.0 32.5 30.0 27.5 22 21 20

3

Page 264: Polyisobutylene Chain End Transformations: Block Copolymer

245

Figure 7.13. gHSQC of alkyne terminated PIB.

2.32.83.33.84.34.8f2 (ppm)

25

30

35

40

45

50

55

60

65

f1 (ppm)

O S

O

O

Page 265: Polyisobutylene Chain End Transformations: Block Copolymer

246

Figure 7.14. 1H NMR of tetra hydroxyl functional PIB.

7 6 5 4 3 2 1

(b)

a

q

o

n'

n

n

b

b

a

m,l

c

c

δ ppm

mm

k

q

as,o r

o

n'p

clb

s r q p kl

rs(a)

5.0 4.5 4.0 3.0 2.5 2.0

5.0 4.5 4.0 3.0 2.5 2.0

S

O

O

S

SOH

OH

*

Page 266: Polyisobutylene Chain End Transformations: Block Copolymer

247

Figure 7.15. 1H NMR of phenyl thiourethane terminated PIB.

8 7 6 5 4 3 2 1 0

O SH

fh

i

i

h

O S NH

O

de

abcde

e d

bc

g

a

e dbc

gf

a

δ ppm

f

8 7 6 5 4 3 2 1 0

bac

7.4 7.2 7.0 6.8

7.4 7.2 7.0 6.8

Page 267: Polyisobutylene Chain End Transformations: Block Copolymer

248

Figure 7.16. GPC traces of (a) chain extended thiourethane with hexane dithiol (b) chain extended thiourethane without hexane dithiol (c) thiol terminated PIB.

8 10 12 14 16 18

(b) (c)

time (min)

(a)

Page 268: Polyisobutylene Chain End Transformations: Block Copolymer