You are on page 1of 144

Tailor-made poly(meth)acrylate-based networks

- Preparation, Structure and Properties -

Vincent Lima

CIP-DATA LIBRARY TECHNISCHE UNIVERSITEIT EINDHOVEN


Lima, Vincent
Tailor-made poly(meth)acrylate-based networks : preparations, structure and properties / by
Vincent Lima. Eindhoven: Technische Universiteit Eindhoven, 2005.
Proefschrift. ISBN 90-386-2996-6
NUR 913
Subject headings: radical polymerization / crosslinking / chain transfer agents; RAFT / mass
spectrometry; MALDI-TOF / liquid chromatography / telechelic acrylic polymers / coating
materials / Youngs modulus; micro-indentation
Trefwoorden: radicaalpolymerisatie / vernetting / ketenoverdracht ; RAFT / Massaspectroscopie;
MALDI-TOF / vloeistofchromatografie / lineaire acrylaatpolymeren / deklagen / Youngs
modulus; microhardheid

2005, Vincent Lima


Printed by PrintService Ipskamp, The Netherlands.
Cover designed by Vincent Lima, Martien Frijns
This research forms part of the research programme of the Dutch Polymer Institute (DPI),
Technology Area Coating Technology, DPI project #205.
An electronic copy of this thesis is available from the site of the Eindhoven University Library in
PDF format (www.tue.nl/bib).

Tailor-made poly(meth)acrylate-based networks


- Preparation, Structure and Properties -

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de


Technische Universiteit Eindhoven, op gezag van de
Rector Magnificus, prof.dr.ir. C.J. van Duijn, voor een
commissie aangewezen door het College voor
Promoties in het openbaar te verdedigen
op mandag 23 mei 2005 om 16.00 uur

door

Vincent Lima

geboren te Noisy-le-Sec, Frankrijk

Dit proefschrift is goedgekeurd door de promotoren:

prof.dr. R. van der Linde


en
prof.dr.ir. P. J. Schoenmakers

Copromotor:
dr. J.C.M. Brokken-Zijp

A mes parents

CHAPTER 1

Introduction

I) POLYMERS ............................................................................................................................... 3
II ) COATINGS .............................................................................................................................. 4
III) OUTLINE OF THIS THESIS ..................................................................................................... 5

CHAPTER 2

Literature review

I) CONTROLLED RADICAL POLYMERIZATION .......................................................................... 9


I.1 INTRODUCTION ....................................................................................................................... 9
I.2 NITROXIDE-MEDIATED POLYMERIZATION ........................................................................ 10
I.3 ATOM TRANSFER RADICAL POLYMERIZATION (ATRP) ................................................... 11
I.4 THE REVERSIBLE ADDITION-FRAGMENTATION CHAIN TRANSFER (RAFT)
POLYMERIZATION ...................................................................................................................... 12
II) METHODS OF SYNTHESIS FOR ,TELECHELIC POLYMERS ........................................... 16
II.1 POLYCONDENSATION .......................................................................................................... 16
II.2 IONIC POLYMERIZATION .................................................................................................... 17
II.3 RADICAL POLYMERIZATION............................................................................................... 17

II.3.1 Introduction............................................................................................... 17
II.3.2 Iniferters .................................................................................................... 18
II.3.3 Nitroxide-Mediated Polymerization.......................................................... 18
II.3.4 Atom Transfer Radical Polymerization..................................................... 20
II.3.5 Reversible Addition-Fragmentation chain Transfer polymerization ........ 21
II.3 CHARACTERIZATION OF TELECHELIC POLYMERS ........................................................... 21
II.4 CONCLUSION ....................................................................................................................... 22

CHAPTER 3

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

I) INTRODUCTION ...................................................................................................................... 29
II) MATERIALS .......................................................................................................................... 29
III) CHARACTERIZATION TECHNIQUES ................................................................................... 31
IV) RESULTS AND DISCUSSION ................................................................................................. 33
V) CONCLUSION ........................................................................................................................ 51

CHAPTER 4

Synthesis of carboxy-telechelic polyacrylates by RAFT

I) INTRODUCTION ...................................................................................................................... 55
II) MATERIALS .......................................................................................................................... 55
III) CHARACTERIZATION TECHNIQUES ................................................................................... 56
IV) RESULTS AND DISCUSSION ................................................................................................. 58
IV) CONCLUSION ....................................................................................................................... 68

CHAPTER 5

MALDI-TOF-MS analysis of RAFT polymers

I) INTRODUCTION ...................................................................................................................... 71
II) MALDI-TOF-MS: EXPERIMENTAL PROCEDURE ............................................................. 71
III) ANALYSIS OF THIOCARBONYL THIO-CONTAINING POLYMERS BY MALDI-TOF-MS... 72
IV) INFLUENCE OF THE LASER INTENSITY ON THE FRAGMENTATION .................................. 79

V) ANALYSIS OF TRITHIOCARBONATE-CONTAINING POLY(BUTYL ACRYLATE) BY MALDITOF-MS .................................................................................................................................... 82


V) CONCLUSIONS ...................................................................................................................... 85

CHAPTER 6

Networks applications: synthesis and properties

I) INTRODUCTION ...................................................................................................................... 87
II) EXPERIMENTAL.................................................................................................................... 88
II.1) THE ATTENUATED TOTAL REFLECTION FOURIER-TRANSFORM INFRARED
SPECTROSCOPY .......................................................................................................................... 88
II.2) TRANSMISSION FOURIER-TRANSFORM INFRARED SPECTROSCOPY .............................. 89
II.3) TEMPERATURE MODULATED DIFFERENTIAL SCANNING CALORIMETRY (TMDSC)... 89
II.4) MICRO-INDENTATION ........................................................................................................ 89
II.5) SOLID-STATE NUCLEAR MAGNETIC RESONANCE ........................................................... 93
III) STARTING COMPOUNDS ..................................................................................................... 94
IV) RESULTS AND DISCUSSION ................................................................................................. 96
IV.1) INFRA-RED RESULTS ......................................................................................................... 96

IV.1.1) Curing mechanism .................................................................................. 96


IV.1.2) Kinetics differences: random vs. telechelic .......................................... 102
IV.2) DETERMINATION OF E-MODULUS ................................................................................. 105

IV.2.1) Telechelic poly(butyl acrylate) micro-indentation measurements........ 105


IV.2.2 Influence of mono-functional chains on the mechanical properties of
poy(butyl acrylate)networks.............................................................................. 109
IV.2.3) Random poly (butyl acrylate-co-acrylic acid) polymers micro-indentation
measurements.................................................................................................... 110
VI.3) MTDSC MEASUREMENTS ............................................................................................. 112
VI.4) SOLID-STATE NUCLEAR MAGNETIC RESONANCE RESULTS ....................................... 115

VI.4.1) T1- relaxations experiments .................................................................. 115


VI.4.2) T2- relaxation experiments.................................................................... 119

VII) CONCLUSION ................................................................................................................... 122

CHAPTER 7

Conclusion and recommendations

I) CONCLUSION........................................................................................................................ 127
II) RECOMMENDATIONS ......................................................................................................... 129

Introduction

CHAPTER 1
INTRODUCTION

I) Polymers
Polymer is a word invented in the twentieth century from Greek words (poly) and
(mer) that can be literally translated by many parts. This term is employed to describe
macromolecular entities based on the repetition of one or several moieties. The definition of a
polymer was introduced by Staudinger1, who suggested that polymers were large molecules
containing long sequences of chemical units linked by covalent bonds. Interestingly, this
statement was issued 10 years after the first manufacturing of synthetic polymer (polymer made
out of phenol and formaldehyde under the trade name Bakelite2). Man knew how to synthesize
polymers before understanding what he was doing!
Polymeric materials are an important part of our modern life. They are generally regrouped in
the common language under the improper term of plastics (derived from thermoplastic), and
have generally to fight a negative image of oil-derivative products usually associated with
pollution images in the societys collective mind. The truth is rather different, as recycling is
possible for many types of polymers. The incineration of them is also a way to generate energy.
Misunderstanding takes place often in the public opinion: polymers are generally not biodegradable, which does not mean that they are polluting. Efficient waste disposal and selfresponsibility from the citizens can greatly reduce the impact of the polymers on our
environment.
The often forgotten part of polymeric material is the natural polymers. Compounds like wool,
silk, cellulose belong as well to the polymer family. The genetic code of every living species
(including humans) is stored in DNA (Desoxyribose Nucleic Acid) molecules. All the fine
organic chemistry occurring in living organisms is regulated through equilibrium reactions
involving proteins, which are nothing less than aminoacids copolymers.

Chapter 1

Few levels below the exceptional complexity of life mechanisms, mankind tried to develop its
own high-tech systems. Polymers are an integrant part of those inventions. Display panels,
memory chips, and semiconductor technologies based exclusively on polymers are nowadays
taking increasing importance in product technologies in detriment of metals. Pharmaceutical
applications have been found as well in the drug delivery systems, which show that
biocompatible polymers can be used for the controlled delivery of medication in vivo. The many
application areas of polymers extend from bulk materials, adhesives, foams and packaging
materials to textile (synthetic polymers such as polyesters or polyacrylics to replace natural ones
like silk, fur or wool), industrial fibers, composites
It is only in the last few decades of the past millennium that polymers have been recognized as
materials that can truly form unique and intelligent materials, applicable to areas where metals
and ceramics would not be usable. The reason behind it was the shift from a research for new
exotic monomers or combination of monomers, to a research for gaining control over the
microscopic structure of the polymer. The chain-length distribution, the monomer sequence
distribution (for copolymers), the tacticity, the functionality distribution and the degree of
branching for cross-linked materials, were for instance recognized as major factors influencing
the macroscopic properties of the polymeric material. It is this understanding of these
(micro)structure properties relationships that will certainly be the base for polymer
applications that will be discovered in the future.

II ) Coatings
In order to protect the materials from the degradation induced by the exterior environment
(atmospheric moisture, UV-light, everyday life use), or for decorative purposes (paints,
lacquers) the great majority of products have to be topped with a thin layer at their surfaces.
This layer is defined as a coating. A coating must be a completely continuous film in order to
fulfill its function. Any imperfection will lower its efficiency (worst chemical protection or
esthetic outlook).
The polymer-based coatings generally contain several components:
- The binder: this is the material that forms the continuous film that adheres to the
substrate. For polymeric coatings, the polymer can be prepared and incorporated in
the formulation before the application, or the polymerization can occur after the
application. The binder governs, to a large extent, the properties of the film.

Introduction

- The pigments: they are insoluble powders dispersed in the coating mixture. Their main
function is to give color to the coating and to provide its eventual hiding power. They
can also stabilize the coating, as they are partially absorbing or reflecting the visible
light, suppressing the damaging effect that the latter can have on the binder.
- The solvents: incorporated in many formulations, their role is to adjust the viscosity of
the coating mixture for the required application mode. The solvents evaporate during
and after application. Due to environmental regulations, the importance of organic
solvents and in a more general term of Volatile Organic Components (VOC) in coatings
formulations is constantly reduced. The trend in industrial research is to develop new
systems based on the use of water as a solvent or even to suppress the use of any solvent
(powder coatings).
- The additives: they are used to avoid defects (e.g. foam bubbles, bad leveling,
sedimentation) or to provide special properties (e.g. UV stability, improved surface slip)
to the coating.
Concerning the future of coatings, the tendency is to use them more and more as smart
material and not only as passive barriers. Biologically-active coating, designed to release
biocides (such as organotin compounds to prevent marine fouling) or electricity-conductive
coatings for electronic display applications are good examples of what smart coatings can be.

III) Outline of this thesis


Conventional

poly(meth)acrylate-based

coating

technology

relies

on

random

copolymerizations of acrylic, methacrylic and functional comonomers. Thermosetting acrylics


have typically a number average molecular mass of 10,000-20,000 g/mol and a polydispersity
index between 2 and 33. As an example, the hydroxyl-functional comonomers employed are
usually 2-hydroxyethyl methacrylate (HEMA) and 2-hydroxypropyl methacrylate (HPMA). The
polymers are then reacted with a multifunctional crosslinker, e.g. a triisocyanate in the case of
hydroxyl-functional resins, in order to obtain the three-dimensional network. The resulting film
obtained from such systems can be qualified as non-ideal, as the network contains dangling
ends, rings and even free polymer chains. These defects in the network structure have in general
a negative effect on material properties such as the Young modulus, the hardness and the scratch
resistance. Highly-functional chains are responsible for early gelation of the system and
brittleness.

Chapter 1

In the 1980s, the use of chain-transfer agents with functional groups was developed. For
example adding 2-mercaptoethanol4 or 2-mercaptopropionic acid5 in the polymerization mixture
will result, after transfer reactions involving those compounds, in initiating radicals bearing
hydroxyl or carboxyl groups. This will lead to a high fraction of molecules with a hydroxyl or a
carboxyl function on one end, reducing dramatically the fraction of non- and mono- functional
chains. The polydispersity index was also reduced, dropping to about 1.74. After cross-linking,
the films obtained exhibited improved mechanical properties but the strong odor of thiolcontaining compounds caused handling problems.
In order to investigate the influence of the functionality per polymer chain distribution and the
molecular weight distribution (MWD) on the formations and properties of poly(meth)acrylate
networks, we employed the Reversible Addition-Fragmentation chain Transfer (RAFT)6
polymerization technique to synthesize linear well-defined telechelic poly(meth)acrylates. The
obtained polymers will have to be telechelic and with a polydispersity index below 1.2. After the
synthesis of those molecules, the network formation and network properties will be studied,
establishing differences in networks made out of poly(meth)acrylates synthesized via random
radical copolymerization and the ones made out of telechelic poly(meth)acrylates.
A general introduction on polymers and coatings has been given in Chapter 1. The main
objectives of the investigation described in this thesis have also been detailed.
Chapter 2 will give a general overview about the controlled radical polymerization technique
and the ways to produce telechelic polymers, based on a literature survey.
Chapter 3 will treat the synthesis and characterization of linear hydroxyl-functional
polymethacrylates.
Chapter 4 will deal with the synthesis and characterization of linear carboxyl-functional
polyacrylates.
The discussion in Chapter 5 will develop around the Matrix Assisted Laser DesorptionIonisation Time Of Flight Mass Spectrometry (MALDI-TOF-MS) analysis of the polymers
prepared by RAFT polymerization.
In Chapter 6, the kinetics and mechanisms of the cross-linking reactions will be studied via
Fourier Transformed InfraRed (FT-IR) spectroscopy. Comparison between polyacrylates
synthesized via random radical copolymerization and telechelic polyacrylates with low
polydispersities will be made. The properties of the cross-linked networks will be investigated
using solid-state Nuclear Magnetic Resonance (NMR) spectroscopy, Temperature-modulated
Differential Scanning Calorimetry (TMDSC) and micro-indentation experiments.

Introduction

Chapter 7 will summarize the main findings of this thesis and consider the potential future
work that can be done on this subject.

Chapter 1

Reference List
1.

Staudinger, H.; Ber. 1920, 53B, 1073.


2.

Baekeland, L. H.; J. Ind. Eng. Chem 1909, 1, 149.

3.

Wicks Jr., Z. W., Jones, F. N., Pappas S.P.; Organic Coatings: Science and Technology
Volume I: Film Formation, Components and Appearance. New York: John Wiley &
Sons, Inc.; 1992

4.

Gray, R. A.; J. Coating. Technol. 1985, 57, 83.

5.

Buter, R.; J. Coating. Technol. 1987, 59, 37.

6.

Mayadunne, R. T. A., Rizzardo, E., Chiefari, J., Chong, Y. K., Moad, G., Thang, S. H.;
Macromolecules 1999, 32, 6977.

Literature review

CHAPTER 2
LITERATURE REVIEW
I) Controlled radical Polymerization
I.1 Introduction
In the past half-century, radical polymerization has enjoyed considerable interest from
industry. It still is, particularly due to its high tolerance for impurities. However, one major
drawback is the poor control over the characteristics of the synthesized polymers, in particular
over the polymer molecular weight (distribution) and functionality. The main reactions
occurring in the free-radical polymerization are depicted in Figure 2.1.
Initiation
I

kd

2R
ki

R + M

RM

Propagation
RM

nM

kp

RMn+1

Termination
2 RM

kt

dead chains

Transfer
RM

+ R'-X

ktr

RMX + R'

Figure 2.1. Schematic representation of the free-radical polymerization


In Figure 2.1, the term dead chains refers to chains that have undergone termination reaction
(either by combination or by disproportionation), and that can not add further monomer in

Chapter 2

propagation steps. Those dead chains are created by the reaction of two propagating radicals,
thus the termination rate can be expressed by the following equation:

Rt = 2k t [ M . ]2
The propagating step is governed by the following kinetic equation:

R p = k p [ M . ][ M ]
In the past twenty years, numerous research groups have devoted their efforts towards the
control of this polymerization process. The early techniques employed were based on reversible
termination of the propagating polymeric radicals, in Nitroxide-Mediated Polymerization
(NMP)1 and Atom Transfer Radical Polymerization (ATRP)2,3. Compared to classical radical
polymerizations, controlled radical polymerizations exhibit a low fraction of terminated polymer
chains. This is explained by the introduction of a reversible deactivation reaction. A small
number of radicals and a large number of dormant chains are present in the system. An
increase of the polymer chain lifetime in controlled radical polymerization is observed, the
growing polymer chains being reversibly deactivated by either a nitroxide (NMP) or a halide
(ATRP). In both cases, most of the polymer chains are reversibly deactivated, so a large number
of dormant chains and a small number of growing chains are obtained. As the termination
reaction between radicals is a second order in radical concentration and propagation is a first
order reaction, the effect of the radical concentration decrease will be much more pronounced on
the irreversible termination events. At full monomer conversion, the majority of the chains will
bear a nitroxide or a halide as an end-group that may allow a possible reactivation. Fresh
addition of monomer will result in a second growth of the chains.

I.2 Nitroxide-Mediated Polymerization


The first report of nitroxide-mediated polymerization was published in 1984 by Solomon et
al.1 NMP is related to the reversible trapping of propagating carbon-centered radicals by stable
nitroxide species (Figure 2.2). This produces an alkoxyamine with a carbon-oxygen bond that
can break at relatively high temperatures (typically around 120C). The chemical structure of the
nitroxide group employed will influence the equilibrium between propagating radicals and
dormant species. Nitroxide radicals are remarkable in many ways, as they react only with
carbon-centered radicals and do not dimerize.
Practically, two experimental procedures may be employed to conduct NMP. The first one is
the use of a classical initiator in combination with a nitroxide, e.g. 2,2,6,6-tetramethylpiperidine-

10

Literature review

N-oxyl (TEMPO)4. The presence of TEMPO in the system will ensure the control over the
polymerization under high temperature conditions (typically at 130C for styrene
polymerizations). The second experimental procedure is the addition of a selected alkoxyamine
to a monomer mixture. At high temperature, the carbon-oxygen bond will break, forming a
carbon-centered radical suitable for initiating the polymerization, and a nitroxide which will
control the radical polymerization by reversible termination. The situation is then identical to the
first one described.
O

O CH2 CH O N

O CH2 CH

O N

TEMPO
O

O
+

O CH2 CH

n CH2 CH

kp
O CH2 CH

n+1
O

kd
+

O CH2 CH

O N

ka

O
O CH2 CH O N

Figure 2.2. Example of styrene nitroxide-mediated polymerization


The control over the polymerization will depend on many parameters, including the activation
(ka) - deactivation (kd) rate constants, the concentration of alkoxyamines, the nature of the
monomer and the temperature. Although remarkable achievements have been reported in the
polymerizations of styrene monomer, the NMP technique showed some limitations with the
control of acrylic or methacrylic monomers polymerizations with the TEMPO nitroxide. The
low activation constant for the acrylates5 and the H-abstraction of the - hydrogen from the methyl group for the methacrylates6 are the two major factors that prevent the production of
materials with low polydispersities by NMP of those monomers. However, newly designed
nitroxides have recently been used to control the polymerizations of 1-3 dienes and acrylates7-9.

I.3 Atom Transfer Radical Polymerization (ATRP)


Atom Transfer Radical Polymerization was introduced in 1995 by Matyjazewski3 and
Sawamoto2. The principle is based on the Karash reaction: the Atom Transfer Radical Addition,
which is widely used by organic chemists for carbon-carbon bond formation.

11

Chapter 2

The control in an ATRP system is induced by the presences of an organic halide initiator and a
transition metal complex in a monomer solution. The reversible exchange of the halogen atom
between the growing polymer chain and the transition metal complex (in its higher oxidation
state) ensure the control over the polymerization (Figure 2.3).

R X

kact

Mt n

kdeact

Mtn+1X

kp
+M
Figure 2.3. General scheme of reversible termination in ATRP with R-X an alkyl halide, Mtn a
metal complex, kact the activation kinetic constant, kdeact the deactivation kinetic constant, kp the
propagating kinetic constant
The most commonly used transition metal couple (activator/deactivator) is the Cu(I)/Cu(II)2,10
system, although Ru(II)/Ru(III)3,11,12, Mo(V)/Mo(VI)13, Fe(II)/Fe(III)14-16, Ni(II)/Ni(III)17-19 and
Rh(I)/Rh(II)20 combinations are also reported in literature. The catalyst has to be complexed in
order to be active and be able to control the polymerization. This is achieved using some
nitrogen-containing ligands, e.g. (substituted) 2,2-bipyridines2,16, Schiff bases21,22, multidentate
tertiary amines23. The role of the ligand is to solubilize the metal in order to create a
homogeneous system and to establish a proper equilibrium between active and dormant species.
ATRP appeared to be more versatile than NMP, since it is applicable to the polymerization of
styrenic10,16,24 and (meth)acrylic monomers3,16,25-27. Major limitations are still encountered in
polymerizations of 1,3-diene, vinyl acetate28 and monomers that can complex and poison the
metal, e.g. acid group containing monomers.
The application of ATRP in industry is still hindered by the post-polymerization purification
(in order to remove the metal complex), often necessary for esthetic, environmental and stability
reasons.

I.4 The Reversible Addition-Fragmentation chain Transfer (RAFT) polymerization


The origin of Reversible Addition-Fragmentation chain Transfer polymerization can be found
in 1998 with the studies of Rizzardo et al.29,30 on polymerizations of acrylates, methacrylates

12

Literature review

and styrene, in the presence of thiocarbonyl thio compounds. The obtained polymers exhibit low
polydispersities and predictable molecular weights.
The mechanism was partially elucidated via ESR (Electron Spin Resonance) spectroscopy31
and the extra reactions that are added to a classical radical polymerization are depicted in Figure
2.4.
I

ki, kp

n (M)

Pn

kadd

Pn

Pm

+m

Z
ki, kp

(M)

k-add

Z
Pn

Pn

kadd
Pn

Pm

Pm

k-add

S
Z

Pn

Pm

S
+

Pn

Z
M

Figure 2.4. Extra reactions occurring in a RAFT polymerization process


After a classical initiation step, a growing oligomeric chain reacts with a thiocarbonyl thio
compound (RAFT agent), via addition to the sulfur-carbon double bond. A carbon-centered
intermediate radical is then formed. This intermediate radical has the ability to regenerate by
fragmentation the initial oligomeric radical/RAFT agent couple or to form a new propagating
radical R. and an oligomeric thiocarbonyl thio compound (dormant polymer species). The R.
radical should reinitiate the polymerization, forming a new propagating chain. The newlyformed thiocarbonyl thio-terminated chain acts as a chain transfer agent, like the initial RAFT
agent. This will ensure by repetitive addition-fragmentation steps on the polymeric thiocarbonyl
thio compounds, the exchange of the radical among all the chains present in the polymerization
system.
The leaving group of the thicarbonyl thio compounds (R, figure 2.4) has to be chosen in such
a way that it is a more stable radical than the oligomeric polymer chain in order to shift the
RAFT pre-equilibrium (Figure 2.4) towards the release of R radicals. The latter must rapidly
reinitiate the polymerization. In general, R is a substituted alkyl group, whose stability can be
increased with stabilizing functional groups32.
The activating group of the thicarbonyl thio compound (Z, figure 2.4) influences the stability
of the intermediate radical and thus the rate of addition of the growing radical to the
thiocarbonyl thio compounds.

13

Chapter 2

It has to be emphasized that transfer reactions do not influence the concentration of


propagating radicals. This concentration is dictated by the steady state approximation, assuming
that the production of radicals by initiation balances their disappearance by termination
reactions. In typical RAFT recipes, this concentration is in the order of magnitude of 10-8 mol.L1

. The concentration of dormant species is equal to the initial concentration of RAFT agent

(conservation of the thiocarbonyl thio moiety), in the order of magnitude of 10-3 mol.L-1. Thus, it
can be estimated that there is only one chain active for 100.000 dormant chains in a typical
RAFT polymerization system.
The rapid exchange of a small number of radicals over a large number of polymeric chains
allow all the chain to grow at the same rate. This is the reason why the RAFT process produces
polymers with masses that increase linearly with conversion.
One could argue that the constant production of radicals, creating new chains, via the initiator
decomposition is not compatible with the definition of a living process. It is true that a living
process implies a constant number of chains throughout the polymerization. However, if the
total amount of decomposed initiator is kept at a low level compared to the RAFT agent
concentration, the number of chains can be assumed to remain constant throughout the
polymerization since:

[ chains ]t = [ RAFT ]0 + 2 f ([ I ]0 - [ I ]t )
In typical RAFT recipes, the second term in the previous equation can be neglected and the
number of chain can be considered constant throughout the reaction.
The scientific community does not unanimously agree on the reasons of the reaction rate
retardation which is sometimes witnessed in the RAFT polymerization. The stability and fate of
the intermediate radical is the central point of the argument. A slower rate of polymerization
compared to the rate of the equivalent conventional free radical polymerization is due to a lower
concentration of propagating radicals in the presence of the RAFT agent.
A first group of scientists described the intermediate radical as a radical sink

33,34

. A low

constant of fragmentation k is the explanation given to justify the lower concentration of


propagating radicals. After addition to the carbon-sulfur double bond of the RAFT agent, the
radical is stored in the intermediate radical and fragments very slowly, lowering the propagating
radical concentration.
The other theory explaining the retardation in the RAFT polymerization was introduced by
Monteiro et al.35 An additional reaction was added to the original RAFT mechanism scheme
from Rizzardo et al.29, i.e. the intermediate radical termination (Figure 2.5). According to this

14

Literature review

theory, the carbon-centered radical can terminate in a similar way as a propagating chain. Due to
the presence of intermediate radicals and propagating radicals in the reactive system,
termination events between two intermediate radicals or between an intermediate radical and a
propagating radical might occur. Three-arm and four-arm shaped polymers might then be
formed. These events reduce the concentration of propagating radicals in the polymerization
system, decreasing the overall polymerization rate, causing retardation.

Pi
Pi

Pm

Pn

Pm

Pn

Z
Pm
Pi

S
Z

Pj

Pm

Pn

Pi

Pn
Pj

Figure 2.5. Possible reactions for he intermediate radical termination, forming 3-arm and 4arm polymer chains
The aim of this introduction on RAFT is not to solve this problem of retardation (which is not
of great importance in this thesis, as the RAFT polymerizations we will use for our synthesis of
polymers for network formation studies are not prone to retardation), but to give an introduction
about the hottest debate in the RAFT literature. Arguments were exchanged in papers (involving
computer simulations33, ab initio calculation on molecular orbitals36, SEC separation on nonpolymerizing model systems37,38), without a definitive answer that would convince the whole
community of scientists working on RAFT. One of the major problems is to find a way to access
all the kinetic parameters of the reactions involved in RAFT.
Compared to the numerous advantages of the RAFT process (possibility to control the
polymerization of numerous monomers, application to water-borne emulsion39-43, easy synthesis
of complex architectures), the problem of retardation occurring with specific RAFT agents
has to be considered as a minor problem. This is of course of interest for a complete
understanding of the mechanism, but retardation can be easily avoided by selecting the
appropriate activating group. It has been found that replacing the phenyl activating group by a

15

Chapter 2

benzyl in thiocarbonyl thioesters44 suppresses the retardation in styrene polymerization. This


replacement reduces the stability of the intermediate radical, increasing its fragmentation rate
constant and thereby reducing its concentration.

II) Methods of synthesis for ,telechelic polymers


End-functional polymers are extremely important for their possible application as building
blocks for instance as block copolymers, surfactants, macromonomers45-47. Functionality at
every chain-ends of a macromolecular chain defines a telechelic polymer. In the polymer
science community, the term telechelic commonly refers to a linear chain having the same
functionality at both chain-ends. Those materials are of great interest for their ability to form triblock polymers or tailor-made polymer networks by reaction with cross-linkers. It is for the
latter application that the first part of this thesis will focus on synthesis and characterization of
linear ,telechelic polymers, with narrow molecular weight distributions.

II.1 Polycondensation
One of the most often used methods to produce telechelic polymers is the polycondensation
technique. For example, reacting a bi-functional monomer A2 with another bi-functional
monomer B2 in slight excess will result in telechelic linear polymer chains bearing B-functional
end-groups (Figure 2.6). A slight excess of A2 will produce in a similar way A-terminated
telechelic linear polymers (Figure 2.6).

A A

+ (1+)

B B

B A B

B B

+ (1+)

B A B

A A

A B A

A B A

Figure 2.6. Preparation of telechelic polymers via polycondensation reactions

16

Literature review

II.2 Ionic polymerization


Another well-known method to prepare telechelic polymers is living ionic polymerization.
Using a functional initiator and a terminal end-capper, telechelic polymers can be produced
(Figure 2.7). Reports of telechelic polystyrene48-50, polybutadiene51 or polyisoprene48,50 can be
found in literature.
Initiation

Propagation

AMM...MM

End-capping

AMM...MMB

Figure 2.7. Preparation of telechelic polymers via living ionic polymerizations


Ionic polymerization leads to well-defined telechelic polymers with very low polydispersities.
However, the stringent conditions required for ionic polymerizations coupled with the limited
choice of monomers available restrains the use of ionic polymerizations for the production of
telechelic material in the industry.

II.3 Radical polymerization


II.3.1 Introduction
The most attractive polymerization technique for the industry is radical polymerization for its
relative high tolerance towards impurities, and the wide range of monomers that can be
polymerized. In order to control the structure of the produced chains, chemicals interfering in
the chain architecture-determining events (initiation, termination and chain-transfer) have to be
employed. The simplest way one can imagine to archive the synthesis of telechelic linear
polymer is to use a functional initiator and to ensure that the only chain stopping event is
termination via combination (Figure 2.8).
The ideal case of exclusive combination as chain-stopping event is extremely rare in the
polymer world. Even poly(styrene) and poly(acrylate)s, polymers that are know to terminate via
combination, exhibit a significant percentage of chains that terminate via disproportionation52,53.

17

Chapter 2

Initiation

Propagation

Termination

AMM...MM

AMM...MMMM...MMA

Figure 2.8. Preparation of telechelic polymers via radical coupling following a free radical
polymerization
The resulting hydrogen-terminated and double bond-terminated chains will lower the average
functionality per chain of the final polymer, and may lead to branching, due to their reactivity as
macromonomer.
II.3.2 Iniferters
An extra technique has to be mentioned in the list of controlled radical processes. In the early
80s, Otsu et al.54,55 obtain a certain degree of control over the polymerization of styrene and
methyl methacrylate by adding disulfide or dithiocarbamate compounds to the polymerization
media. Photochemically, those sulfur-containing agents decompose, forming sulfur-centered
radicals. The latter can act as initiators, chain transfer agent and termination agent. The term
iniferter given to those additives recalls their triple role in the polymerization. The term living
could not be employed to describe such systems. The blame is mainly due to the low transfer
constant of the dormant polymer species to the propagating ones, and to the occurrence of many
side reactions. The iniferter method was employed with the idea to obtain telechelic polymers.
However, a significant amount of non-functional chains were found in the reaction system56.
II.3.3 Nitroxide-Mediated Polymerization
A logical step forward towards the production of telechelic polymers via a radical process is
the use of controlled radical polymerization processes, which were described in section II.1.
Nitroxide-mediated polymerization was reported as a successful method to obtain telechelic
chains.1 One of the possible methods involves in a first step the polymerization of styrene in the
presence of a hydroxyl-functional alkoxyamine. In one step, -functional chains are obtained,
with an alkoxyamine group in the position. The successful reduction of the latter group in an
alcohol, using an acetic acid / zinc mixture1,57, allows in principle the preparation of welldefined ,-hydroxyl telechelic poly(styrene) with a functionality close to 2 (Figure 2.9).

18

Literature review

HO

O N

O N
NMP

OH

HO

CH3COOH / Zn

Styrene

HO

80 C

Figure 2.9. Preparation of telechelic polymers via NMP followed by a reduction of the
alkoxyamine
Another method combines the use of TEMPO with a hydroxyl-functional initiator (Figure
2.10). At the end of the controlled polymerization, the average functionality is slightly above 1,
probably due to termination reactions via combination occurring during the polymerization57.
This result was expected, since all the chains were initiated by a hydroxyl-functional peroxide
fragment. In a post-polymerization treatment similar to the one described previously, the
reduction of the alkoxyamine into an alcohol allows the synthesis of ,-hydroxyl telechelic
poly(1,3-butadiene).
H O O H

HO

1,3-butadiene
TEMPO

HO

CH2

CH CH CH2

O N

CH3COOH / Zn
80oC

HO

CH2

CH CH CH2

OH

Figure 2.10. Preparation of telechelic polymers via NMP using an hydroxyl-functional


initiator
Another method consists to heat up the -functional chains to 165C in order to thermally
cleave the carbon-oxygen bond57. Compared to a NMP process, the absence of monomer
suppresses the propagation possibility, and the chains are forced to terminate or to recombine
with the alkoxyamine, which would lead to another potential cleavage of the carbon-oxygen
bond. The yield of obtained telechelic material depends on the proportion of termination by
combination for the polymer chains, and the occurrence of side reactions. The results published
by Pradel et al.57 indicate that the average functionality per polymer chain below 2 and that side
reactions are the probable reason for these results.

19

Chapter 2

The methods described are in principle not restricted to the specific monomer systems
described. They are in general applicable to the monomers, whose polymerization can be
controlled via the NMP process.
II.3.4 Atom Transfer Radical Polymerization
In order to obtain ,-telechelic chains with ATRP, the usual first step is to employ functional
halides as initiators. End-group functionality such as hydroxyl27,58, ester2,27, phenyl59, amine59,
aldehyde59, anhydride58 or carboxylic acid58,60 can easily be introduced though the proper choice
of

initiator.

At

the

end

of

classical

ATRP

process, the functional group and

the halide atom will be found in the and positions. A modification of the halide group to
give the same functionality will result in linear telechelic polymer chains bearing the same
function at both ends. One of the main techniques used to achieve this is the nucleophilic
substitution. Using functional azides24,61, amines25,62 or phosphines63, chain-end functionality is
easily changed from halide to the desired one. In Figure 2.11 the end-chain functionalization
reaction described by Coessens et al.63 is presented.
Another possibility reported by Coessens et al.25 is an end-capping reaction with a functional
monomer that can not homopolymerize (e.g. allyl alcohol25, ethene6, C6064, maleic anhydride65)
under ATRP conditions, but can copolymerize with the monomer used in the ATRP. Only one
monomer unit adds to the already-existing chain, leading to an efficient end-capping.
CH3
HO CH2

CH2

Br
O

ATRP
MMA

CH3
HO CH2

CH2

Br

O
O

O
n

H2N C4H9

OH

CH3
HO CH2

CH2

NH

O
O

C4H9

OH

O
n

HBr

Figure 2.11. Preparation of telechelic polymers via ATRP followed by a nucleophilic


substitution

20

Literature review

II.3.5 Reversible Addition-Fragmentation chain Transfer polymerization


The most promising report of synthesis of telechelic polymers via RAFT has been published
in literature by Lai et al.66 Using symmetrical trithiocarbonates, telechelic polymers can be
easily obtained in one step (Figure 2.12). One limitation of this technique is the poor results with
methacrylic monomers, due to the relatively low stability of the leaving group compared to the
one of a methacrylic radical.
HOOC

COOH

HOOC

COOH

N N
CN

CN

Acrylate

HOOC

(Polyacrylate)

(Polyacrylate)

COOH

Figure 2.12. Production of telechelic poly(acrylate) in one step using the RAFT
polymerization with trithiocarbonates as RAFT agents
It can be noted that the use of an initiator bearing the same functionality as the leaving group,
ensures that all the chains will be initiated by a COOH-containing group (neglecting the chain
transfer events to e.g. monomer and solvent). This method has been employed before by
Mayadunne et al.67, but the functionalities introduced at the end-chains were less attractive
(CH2Ph, CH2(CH3)Ph).
A recent article of Lui et al.68 describes the synthesis of di-hydroxyl-terminated telechelic
polystyrene and poly (methyl acrylate) using a symmetrical di-hydroxyl-functional
trithiocarbonate. The characterization of the polymers in terms of functionality was limited to IR
and NMR.

II.3 Characterization of telechelic polymers


In the literature, numerous claims about telechelic polymer synthesis have been published,
without thorough characterization works on the resulting polymer. Many examples of
affirmations based on the theoretical chemistry involved in the synthesis, or non-adapted
analytical techniques, can be found.

21

Chapter 2

Often, titration of the functional groups is the method of choice to prove that the produced
polymers are telechelic. However, the titration technique gives an average number for the
number of functions per polymer chain. One of the most commonly encountered mistakes is the
confusion between the average functionality per polymer chain and the functionality
distribution. If a batch of polymer consists of a batch of 50% of tri-functional polymers, and
50% of mono-functional polymers, the average functionality would be estimated by titration to
be similar to the one of a sample containing 100% of bi-functional polymer. The average
functionality is the same, but the distributions are different. This problem of giving an average
functionality rather than a functionality distribution is encountered also in InfraRed (IR),
UltraViolet (UV) or Nuclear Magnetic Resonance (NMR) spectroscopies.
Other doubts about the relevance of results can rise when mass-spectrometric techniques are
used

to

prove

the

Desorption/Ionization

production
Time-Of-Flight

of

telechelic
Mass

polymers.

Spectrometry

Matrix-Assisted

Laser

(MALDITOFMS)

and

Electrospray Ionization Mass Spectrometry (ESIMS) provide good qualitative information, but
they can not be used for quantitative results, due to variations in the ionization, separation, and
detection efficiencies.
Liquid chromatography is nowadays one of the most popular methods of the analysis for the
Functionality-Type Distribution (FTD). In a more specific class of liquid chromatography,
Liquid Chromatography under Critical Conditions (LCCC), the retention times of polymer
chains are only based on the number of functionalities present in the polymer chains, and are
independent of their molar masses69-71. More details on LCCC will be given in Chapter III and
Chapter IV, when the methods of characterization of synthesized polymers will be discussed.

II.4 Conclusion
The need for well-defined telechelic polymers with a low polydispersity dictated our choice to
apply a CRP technique for the synthesis of our poly(meth)acrylate. The stringent conditions
required for ionic polymerizations seemed to limit its use. The nitroxide-mediated
polymerization technique was discarded, due to major drawbacks in the polymerization of
methacrylates, which would have prevented the synthesis of tailor-made polymers. The choice
between RAFT and ATRP was made primarily based on the versatility of the RAFT process,
and its ability to produce telechelic carboxyl-functional polyacrylates in one step with an
appropriate trithiocarbonate. Moreover, the absence of metal complexes in RAFT

22

Literature review

polymerization is also beneficial, as those compounds can have a major impact on the crosslinking process that will take place after the synthesis.
In parallel with carboxyl-functional poly(acrylate), hydroxyl-functional poly methacrylate will
be also synthesized via a two-step post-polymerization procedure. Using a hydroxyl-functional
RAFT agent, hydroxyl-functional polymethacrylate can be synthesized. Changing after
polymerization the thiocarbonyl thio into a hydroxyl function should allow the production of
telechelic polymethacrylates. The application of the synthesis process described by Liu et al.68
could have been suitable for our study, but it was published in the last year of this project,
preventing its utilization due to time constraints. For both hydroxyl-polymethacrylate or
carboxyl-polyacrylate, the LCCC technique will be the preferred method to analyze the
polymers synthesized, and check the functionality-type distribution.

23

Chapter 2

Reference List
1. Solomon, D. H.; Rizzardo, E.; Cacioli, P. 84-304756 EP 135280 July 11, 1984
2. Wang, J. S., Matyjaszewski, K.; Macromolecules 1995, 28, 7901.
3. Kato, M., Kamigaito, M., Sawamoto, M., Higashimura, T.; Macromolecules 1995, 28,
1721.
4. Georges, M. K., Veregin, R. P. N., Kazmaier, P. M., Hamer, G. K.; Macromolecules
1993, 26, 2987.
5. Bon, S. A. F.; PhD Thesis 1998
EindhovenUniversityofTechnology,Eindhoven,TheNetherlands
6. Bon, S. A. F., Steward, A. G., Haddleton, D. M.; J. Polym. Sci. Pol. Chem. 2000, 38,
2678.
7. Benoit, D., Chaplinski, V., Braslau, R., Hawker, C. J.; J. Am. Chem. Soc. 1999, 121,
3904.
8. Benoit, D., Harth, E., Fox, P., Waymouth, R. M., Hawker, C. J.; Macromolecules 2000,
33, 363.
9. Grimaldi, S., Finet, J. P., Le Moigne, F., Zeghdaoui, A., Tordo, P., Benoit, D.,
Fontanille, M., Gnanou, Y.; Macromolecules 2000, 33, 1141.
10. Matyjaszewski, K., Patten, T. E., Xia, J.; J. Am. Chem. Soc. 1997, 119, 674.
11. Simal, F., Demonceau, A., Noels, A. F.; Tetrahedron Lett. 1999, 40, 5689.
12. Simal, F., Demonceau, A., Noels, A. F.; Angew. Chem. Int. Edit. 1999, 38, 538.
13. Brandts, J. A. M., van de Geijn, P., van Faassen, E. E., Boersma, J., Van Koten, G.; J.
Organomet. Chem. 1999, 584, 246.
14. Teodorescu, M., Matyjaszewski, K.; Macromolecules 1999, 32, 4826.
15. Ando, T., Kamigaito, M., Sawamoto, M.; Macromolecules 1997, 30, 4507.
16. Matyjaszewski, K., Wei, M., Xia, J., McDermott, N. E.; Macromolecules 1997, 30,
8161.
17. Granel, C., Dubois, P., Jerome, R., Teyssie, P.; Macromolecules 1996, 29, 8576.
18. Uegaki, H., Kotani, Y., Kamigaito, M., Sawamoto, M.; Macromolecules 1997, 30, 2249.
19. Uegaki, H., Kotani, Y., Kamigaito, M., Sawamoto, M.; Macromolecules 1998, 31, 6756.
20. Percec, V., Barboiu, B., Neumann, A., Ronda, J. C., Zhao, M.; Macromolecules 1996,
29, 3665.
21. Haddleton, D. M., Jasieczek, C. B., Hannon, M. J., Shooter, A. J.; Macromolecules
1997, 30, 2190.

24

Literature review

22. Haddleton, D. M., Duncalf, D. J., Kukulj, D., Crossman, M. C., Jackson, S. G., Bon, S.
A. F., Clark, A. J., Shooter, A. J.; Eur. J. Inorg. Chem 1998, 1799.
23. Xia, J., Gaynor, S. G., Matyjaszewski, K.; Macromolecules 1998, 31, 5958.
24. Matyjaszewski, K., Nakagawa, Y., Gaynor, S. G.; Macromol. Rapid Comm. 1997, 18,
1057.
25. Coessens, V., Matyjaszewski, K.; Macromol. Rapid Comm. 1999, 20, 127.
26. Coessens, V., Pyun, J., Miller, P. J., Gaynor, S. G., Matyjaszewski, K.; Macromol.
Rapid Comm. 2000, 21, 103.
27. Haddleton, D. M., Waterson, C., Derrick, P. J., Jasieczek, C. B., Shooter, A. J.; Chem.
Commun. 1997, 683.
28. Xia, J., Paik, H. J., Matyjaszewski, K.; Macromolecules 1999, 32, 8310.
29. Chiefari, J., Chong, Y. K., Ercoles, 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.
30. Rizzardo, E.; Thang, S. H.; Moad, G. 98-AU569 WO 9905099 July 20, 1998
31. Hawthorne, D. G., Moad, G., Rizzardo, E., Thang S.H.; Macromolecules 1999, 32,
5457.
32. Chong, Y. K., Krstina, J., Le, T. P. T., Moad, G., Postma, A., Rizzardo, E., Thang, S.
H.; Macromolecules 2003, 36, 2256.
33. Vana, P., Davis, T. P., Barner-Kowollik, C.; Macromol. Theor. Simul. 2002, 11, 823.
34. Barner-Kowollik, C., Vana, P., Quinn, J. F., Davis, T. P.; J. Polym. Sci. Pol. Chem.
2002, 40, 1058.
35. Monteiro, M. J., de Brouwer, H.; Macromolecules 2001, 34, 349.
36. Coote, M. L., Radom, L.; J. Am. Chem. Soc. 2003, 125, 1490.
37. Kwak, Y., Goto, A., Tsujii, Y., Murata, Y., Komatsu, K., Fukuda, T.; Macromolecules
2002, 35, 3026.
38. Kwak, Y., Goto, A., Fukuda, T.; Macromolecules 2004, 37, 1219.
39. Butte, A., Storti, G., Morbidelli, M.; Macromolecules 2000, 33, 3485.
40. Monteiro, M. J., Sjoberg, M., Van der Vlist, J., Gottgens, C. M.; J. Polym. Sci. Pol.
Chem. 2000, 38, 4206.
41. de Brouwer, H., Tsavalas, J. G., Schork, F. J., Monteiro, M. J.; Macromolecules 2000,
33, 9239.
42. Gaillard, N., Guyot, A., Claverie, J.; J. Polym. Sci. Pol. Chem. 2003, 41, 684.
43. Smulders, W., Gilbert, R. G., Monteiro, M. J.; Macromolecules 2003, 36, 4309.

25

Chapter 2

44. Quinn, J. F., Davis, T. P., Rizzardo, E.; Chem. Commun. 2001, 1044.
45. Odian G.; Principles of polymerization. New York: Wiley; 1991
46. Liu, Y., Pan, C.; European Polymer Journal 1998, 34, 621.
47. Lou, X., Detrembleur, C., Jerome, R.; Macromolecules 2002, 35, 1190.
48. Peters, M. A., Belu, A. M., Linton, R. W., Dupray, L., Meyer, T. J., DeSimone, J. M.; J.
Am. Chem. Soc. 1995, 117, 3380.
49. Quirk, R. P., Lizarraga, G. M.; Macromolecules 1998, 31, 3424.
50. Tohyama, M., Hirao, A., Nakahama, S., Takenaka, K.; Macromol. Chem. Physic. 1996,
197, 3135.
51. Reed, S. F., Jr.; J. Polym. Sci. Pol. Chem. Ed. 1972, 10, 2493.
52. Zammit, M. D., Davis, T. P., Haddleton, D. M., Suddaby, K. G.; Macromolecules 1997,
30, 1915.
53. Vana, P., Davis, T. P., Barner-Kowollik, C.; Aust. J. Chem. 2002, 55, 315.
54. Otsu, T., Yoshida, M.; Makromol. Chem. Rapid. 1982, 3, 127.
55. Otsu, T., Yoshida, M., Tazaki, T.; Makromol. Chem. Rapid. 1982, 3, 133.
56. Otsu, T.; J. Polym. Sci. Pol. Chem. 2000, 38, 2121.
57. Pradel, J. L., Boutevin, B., Ameduri, B.; J. Polym. Sci. Pol. Chem. 2000, 38, 3293.
58. Malz, H., Komber, H., Voigt, D., Hopfe, I., Pionteck, J.; Macromol. Chem. Physic.
1999, 200, 642.
59. Haddleton, D. M., Waterson, C.; Macromolecules 1999, 32, 8732.
60. Zhang, X., Matyjaszewski, K.; Macromolecules 1999, 32, 7349.
61. Li, L., Wang, C., Long, Z., Fu, S.; J. Polym. Sci. Pol. Chem. 2000, 38, 4519.
62. Coessens, V., Matyjaszewski, K.; J. Macromol. Sci Pure 1999, A36, 811.
63. Coessens, V., Matyjaszewski, K.; J. Macromol. Sci Pure 1999, A36, 653.
64. Zhou, P., Chen, G. Q., Hong, H., Du, F. S., Li, Z. C., Li, F. M.; Macromolecules 2000,
33, 1948.
65. Koulouri, E. G., Kallitsis, J. K., Hadziioannou, G.; Macromolecules 1999, 32, 6242.
66. Lai, J. T., Filla, D., Shea, R.; Macromolecules 2002, 35, 6754.
67. Mayadunne, R. T. A., Rizzardo, E., Chiefari, J., Krstina, J., Moad, G., Postma, A.,
Thang, S. H.; Macromolecules 2000, 33, 243.
68. Liu, J., Hong, C. Y., Pan, C. Y.; Polymer 2004, 45, 4413.
69. Mengerink, Y., Peters, R., van der Wal, S., Claessens, H. A., Cramers, C. A.; Journal of
Chromatography, A 2002, 949, 337.
70. Gorbunov, A., Trathnigg, B.; Journal of Chromatography, A 2002, 955, 9.

26

Literature review

71. Philipsen, H. J. A., Klumperman, B., van Herk, A. M., German, A. L.; Journal of
Chromatography, A 1996, 727, 13.

27

28

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

CHAPTER 3
SYNTHESIS OF HYDROXYL-TELECHELIC
POLYMETHACRYLATES BY RAFT
I) Introduction
The first step into the thorough investigation of the network formation mechanisms and final
network properties of poly(meth)acrylates-based coatings is the synthesis of the building blocks
of those networks. The production of well-defined telechelic polymers will be mandatory. The
previous chapter summarized the techniques available for synthesis and concludes that the
RAFT polymerization process should be used in order to achieve our purposes.
In this chapter, the synthesis and characterization of RAFT poly(meth)acrylates with a
hydroxyl end-group and a two-step post-polymerization procedure, modifying the thiocarbonyl
thio into an hydroxyl end-group, will be discussed.

II) Materials
Monomers (methyl methacrylate, butyl methacrylate) of the highest purity available were
purchased from Aldrich. They were distilled over CaH2 prior to utilization in order to remove
the inhibitor. Toluene (Biosolve) and acetone (Aldrich) were distilled over CaH2 before use.
Tetrahydrofuran (THF, Biosolve) was refluxed over LiAlH4 (Aldrich) prior to use.
Azobis(isobutyronitrile) (AIBN, Merck) was recrystallized from methanol. 4,4-Azobis(4cyanopentanol) (ACP) was prepared according to the method described by Clouet et al.1 with a
purity superior to 99% (no other peaks than the expected ones were observed by 1H-NMR in
CDCl3). 1H-NMR (CDCl3): (ppm) 3.7 (m, 4H) 2.1-2.4 (m, 4H) 2.0 (s, 2H) 1.5-1.8 (m, 4H) 1.7
(s, 6H). 2-Cyanoprop-2-yl dithiobenzoate (RAFT-AIBN) was synthesized according to the
procedure described by Le et al.11 with a purity exceeding 99% (no other peaks than the
expected ones were observed by 1H-NMR in CDCl3) 1H-NMR (CDCl3): (ppm) 7.4-7.9 (m, 5H)

29

Chapter 3

1.9 (s, 6H). The synthesis of (4-cyano-1-hydroxypent-4-yl) dithiobenzoate (RAFT-ACP)


followed the route proposed by Rizzardo et al.18 A purity of 97% (1H-NMR in CDCl3) was
obtained. 1H-NMR (CDCl3): (ppm) 7.4-7.9 (m, 5H) 3.7 (m, 2H) 2.1-2.4 (m, 4H) 1.9 (s, 3H)
1.6 (s, 1H). The major byproduct of the synthesis is the combination product of two ACPderived radicals which give various peaks in the 1.0 - 3.0 ppm region. Other chemicals were
purchased from Aldrich and used without further purification.
Polymerization procedure. All polymerizations were carried out in three-neck round bottom
flasks, heated in an oil bath. The reaction solution was degassed with argon and three freezepump-thaw cycles were applied. A typical polymerization was performed as follows.
Polymerization of methyl methacrylate. (Step 1 Scheme 1) Methyl methacrylate (30.1 g, 3.01
10-1 mol), ACP (0.076 g, 3.02 10-4 mol), RAFT-ACP (0.40 g, 1.51 10-3 mol) and toluene
(89.4 g, 103.4 mL) were mixed in a three-neck round bottom flask. Oxygen was removed and
the flask was immersed in an oil bath that was preheated at 70 C. After 8 hrs of polymerization,
PMMA (PMMA1) was isolated by precipitation in heptane (conversion (GC) : 78%).
1

H-NMR (CDCl3): (ppm) 7.4-7.9 (phenyl group of the thiocarbonyl thio ester), 3.7 (CH3 of the

methyl ester group), 1.7-2.1 (CH2 of the polymer backbone), 1.6 (CH3 of the leaving group of
the RAFT agent), 1.2-1.6 (CH2 of the leaving group of the RAFT agent), 0.9-1.1 (CH3 of the
methacrylic group).
UV-VIS: Absorbance bands 220 nm (ester group, very strong), 300nm (thiocarbonyl thio group,
strong)
Aminolysis of the thiocarbonyl thio end groups in the RAFT polymers. Thiol-functional
polymers were obtained under basic conditions by cleavage of the dithioester. Oxygen was
excluded from the reactive system by three freeze-pump-thaw cycles. In some experiments, a
few drops of an aqueous solution of Na2S2O4 were added, in order to prevent the conversion of
the corresponding thiols to disulfides.
Aminolysis of poly(methyl methacrylate). PMMA1 (Mn = 17,000g.mol-1, 5.0 g, 2.94 10-4 mol)
was dissolved in 50 mL of THF followed by the addition of 1-hexylamine (0.032 g, 3.24
10-4 mol) at room temperature. The reaction was allowed to proceed during an hour. A fast and
noticeable color change took place (discoloration from pink to yellow), and the final colorless
polymer (PMMA2) was isolated by precipitation in heptane.

30

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

H-NMR (CDCl3): (ppm) 3.7 (CH3 of the methyl ester group), 1.7-2.1 (CH2 of the polymer

backbone), 1.6 (CH3 of the leaving group of the RAFT agent), 1.2-1.6 (CH2 of the leaving group
of the RAFT agent), 0.9-1.1 (CH3 of the methacrylic group).
UV-VIS: Absorbance bands at 220 nm (ester group, very strong), 300nm (thiocarbonyl thio
group, very weak)
As will be discussed below in this chapter, in some experiments, a few drops of an aqueous
solution of Na2S2O4 were added to the solution. The procedure with addition of Na2S2O4
performed on PMMA1 lead to PMMA3
Michael addition on the thiol-functional PMMA. Thiol-functional polymers were treated
under basic conditions in the presence of hydroxylethyl acrylate and a few drops of an aqueous
solution of Na2S2O4 were added, in order to prevent the conversion of the corresponding thiols
to disulfides. Oxygen was excluded from the reactive system by three freeze-pump-thaw cycles.
Michael addition on thiol-functional PMMA PMMA3 (Mn = 17,000 g.mol-1, 5.0 g, 2.94 10-4
mol) was dissolved in dimethyl sulfoxide (50 mL). Benzyltrimethylammonium hydroxide (40
wt. % solution in methanol) (1.22 10-3 g, 7.31 10-6 mol) and hydroxylethyl acrylate (0.034 g,
3.23 10-3mol) were added. A few drops of an aqueous solution of Na2S2O4 were mixed with
the solution. The reaction was allowed to proceed at room temperature overnight under argon.
The polymer (PMMA4) was then isolated by precipitation in heptane. The removal of the
catalyst was achieved by washing with water.

III) Characterization techniques


1

H-NMR and 13C-NMR spectra were recorded on a Varian-400 spectrometer using TMS as an

internal standard.
Monomer conversion was determined from the concentration of residual monomer, measured
using a Hewlett-Packard (HP 5890) GC, equipped with an AT-Wax capillary column (30 mm
0.53 mm 10 m); toluene was used as internal reference.
SEC (Size-Exclusion Chromatography) was carried out using a Waters model 510 pump and a
model 410 refractive-index detector (at 313 K). The columns used were a PL-gel guard column
(5 m particles) 50 7.5 mm, followed by two PL-gel mixed-C (5 m particles) 300 7.5 mm
columns (313 K). THF was used as the eluent at a flow rate of 1 mL/min. Lowpolydispersity
polystyrene standards (Polymer Labs) with molecular weights ranging from 580 to 7.1 106
gmol-1 were used for calibration of the columns. Molecular weights were recalculated using the

31

Chapter 3

Mark-Houwink parameters of PS, PMMA, PBMA and PBA in THF (PS: K = 1.140 10-4 dL/g,
a = 0.716; PMMA: K = 0.944 10-4 dL/g, a = 0.719; PBMA: K = 1.480 10-4 dL/g, a = 0.664;
PBA: K = 1.220 10-4 dL/g, a = 0.700). The samples (1 mg/mL THF) were filtered through a
0.2 m syringe filter prior to injection. Data acquisition and processing were performed using
Waters Millennium32 (v3.00) software.
MALDI-TOF-MS (Matrix-Assisted Laser-Desorption Ionization Time-Of-Flight Mass
Spectrometry) measurements were performed on a Voyager-DE STR instrument (Applied
Biosystems, Framingham, MA, USA) equipped with a 337-nm nitrogen laser. Positive-ion
spectra were acquired in the reflector mode. 2-[(2E)-3-(4-tert-butylphenyl)-2-methylprop-2enylidene]malononitrile was used as the matrix. All samples were hand-spotted on the target.
Sodium trifluoroacetate (CF3CO2Na) was added.
A Waters (Milford, MA, USA) 2690 Alliance liquid chromatography system was used to
perform the isocratic LC experiments. This HPLC instrument contained a built-in auto-injector
with a sample loop allowing injection of variable sample volumes, and was equipped with a
Waters 996 PDA (photodiode-array detector) and a Sedex 55 evaporative light-scattering
detector (ELSD, temperature 62 C, N2 pressure 2.2 bar). The mobile phase was prepared in-situ
using the solvent-mixing capability of the instrument. The data collection and the data analysis
were handled by Waters Millennium 3.2 software. The columns used (150 mm x 4.6 mm i.d.)
were packed in-house with Hypersil Silica (3 m particles; 100 pore size; Shandon, Runcorn,
UK). PMMA separations were performed at 25C using the following mobile phase: 42 v/v%
acetonitrile in dichloromethane, at a flow rate of 0.5 mL/min.
Thermogravimetric analysis (TGA) traces of polymers were recorded on a Pyris 6 TGA
instrument from Perkin Elmer with a temperature ramp of 10C/min from 40 to 500C under
nitrogen atmosphere.

32

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

IV) Results and discussion


The results of the RAFT polymerizations of methacrylates (first step of Scheme 1) are
reported in Table 3.1. All polymerizations exhibited the characteristics of controlled systems,
e.g. a linear increase of Mn with conversion and low polydispersity throughout the reaction.

Monomer

Initiator

CTA

Conversion

Mn,exp (Mn, th)

Mw/Mn

MMA

AIBN

RAFT-AIBN

82%

2500 (2000)

1.27

MMA

ACP

RAFT-ACP

78%

2400 (2000)

1.21

MMA

ACP

RAFT-ACP

73%

17000 (15000)

1.16

BMA

ACP

RAFT-ACP

78%

3200 (4000)

1.22

All polymerizations were carried out at 70C, for 8 hrs, in toluene under argon, with [Monomer] = 2.2 M,
[Initiator] = 0.035 M and [Initiator] : [CTA] = 1:5 ; Mn,

th

= (([Monomer]0/[CTA]0 ) conversion

Mmonomer ) + MCTA

Table 3.1. Conversions and molecular weights of the polymethacrylates synthesized by RAFT
polymerization
In order to achieve the synthesis of hydroxyl end-functionalized telechelic polymers, the
thiocarbonyl thio group had to be modified. A possible procedure to achieve this is proposed in
Scheme 3.2. According to the existing literature, thiocarbonyl thio groups undergo fast reactions
with amines17, leading to their reduction to thiols.
After polymerization, all the polymers synthesized by RAFT were treated with a slight excess
of 1-hexylamine. The efficiency of this reaction was determined by 1H-NMR. As shown in
Figure 3.3, no more aromatic protons of the thiocarbonyl thio group were detected in the 1HNMR spectrum, indicating a complete cleavage of the thiocarbonyl thio function.

33

Chapter 3

OH
CN

RAFT-ACP

+
HO

N N

OH

CN

ACP

CN
methacrylate,
Toluene, 70C

OH

(Polymethacrylate)
CN

hexylamine,
THF, RT

HS

(Polymethacrylate)

OH
CN

Michael addition
with HEA
O
HO

(Polymethacrylate)

OH
CN

Scheme 3.2. Proposed mechanism for the synthesis of hydroxyl-telechelic polymethacrylates


This complete cleavage was confirmed by other techniques.

13

C-NMR showed the

disappearance of the peak around 220 ppm characteristic of C6H5CS2. UV-absorbance spectra of
the RAFT polymer before and after aminolysis were recorded as well. The absorbance band
around 300 nm observed for the polymer prior to aminolysis, (due to the thiocarbonyl thio
group) was barely seen after reaction.

34

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

(1)

8.0

7.9

7.8

7.7

7.6

7.5

7.4

Chemical shifting (ppm)

8.0

7.9

7.8

7.7

7.6

7.5

7.4

Chemical shifting (ppm)

(2)
8

Chemical shifting (ppm)


Figure 3.3. 1H-NMR spectra in CDCl3 of a PMMA (general formula: CNC(CH3)2(C5H8O2)n CS2
C6H5) prepared by RAFT polymerization, before (1) and after (2) aminolysis
It was also noted that this post-polymerization treatment removed the red color, which is
characteristic of the presence of a thiocarbonyl thio group in the polymer chain.
However, some questions arose following the SEC characterization of the polymers. In the
chromatograms of the polymers obtained after cleavage of the thiocarbonyl thio group,
secondary populations were detected in the higher molecular weight region (Figure 3.4).
HS

OH

(PMMA)

(PMMA)

2.8

OH
CN

CN

3.0

3.2

3.4

3.6

3.8

4.0

4.2

4.4

4.6

Log M

Figure 3.4. SEC chromatograms of PMMA synthesized by RAFT before and after aminolysis

35

Chapter 3

This phenomenon has also been recently reported by Wang et al.20 These shoulders
correspond to polymers with molecular weights approximately twice higher than the expected
values. The presence of these species is most likely due to the oxidative coupling of thiols,
which results in the formation of a disulfide bridge between two polymer chains.
In order to confirm this hypothesis, a critical-liquid-chromatography method was developed6.
In a general sense, critical-liquid-chromatography methods allowed separation of polymer
chains according to their functionality. By definition, in critical chromatography, retention is
independant of molecular weight2-5,12-16. The application of this method to our study is governed
by the need of a polymer separation based on the hydroxyl functionality. With a suitable specific
interaction between the column and the hydroxyl groups, an elution time dependant on the
number of OH functional groups should be obtained. The method should be rather robust even
though it operates under critical conditions, and should not be dramatically dependant on
parameters such as solvent composition, flow rate or temperature, for example. Due to the
presence of thiocarbonyl thio or thiol group, the method should also be relatively insensitive to
the presence of other functional groups in the polymer chains.
The solvent mixture should typically consist of a poor eluent for the polymer, and another
solvent capable of desorbing the polymer chain from the column. For its specific interaction
with the polar group (e.g. hydroxyl), bare silica is chosen as the column material.
25

PMMA 1680
PMMA 3800
PMMA 6950
PMMA 13900
PMMA 28300

Retention time (min)

20

15

(A)
10

0
34

39

44

49

54

59

64

69

v/v % Acetonitrile

36

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

25
PMMA-OH 500
PMMA-OH 3300
PMMA-OH 6680

Retention Time (min)

20

PMMA-OH 13900
PMMA-OH 20000
PMMA-2OH-1450

15

PMMA-2OH-5718

(B)

10

0
34

39

44

49
54
v/v % Acetonitrile

59

64

69

Figure 3.5. (A) Dependence of retention time on the mobile-phase composition for PMMA
standards with different molecular weights. (B) Dependence of retention time on the mobilephase composition for PMMA-OH and HO-PMMA-OH samples with different molecular
weights. ELSD Note: Data points with retention times larger than 10 min in figure A (15 min in
figure B). were recorded and used to construct the curved lines. However, they are not shown in
the figures.
Dichloromethane is known to be a poor eluent for PMMA, it will then be mixed with
acetonitrile, a much more polar solvent, for the desorption of PMMA chains from the column.
The optimal ratio of the two for operating in critical condition has to be found via a systematic
variation as shown below.
Some non-functional, mono-functional and bi-functional PMMA samples of different molar
masses were used for the determination of critical conditions. It can be seen in Figure 3.5 (A)
that the retention time is independent of the molar mass of the analyzed sample for nonfunctional PMMA when the acetonitrile content is between 43% and 55%. On Figure 3.5 (B), it
is observed that around 43% of acetonitrile content, the retention time is also independent of the
molar mass for hydroxyl-mono-functional and di-functional PMMA. It is then concluded that
the critical mobile phase composition for the separation of hydroxyl-functional PMMA should
be a 43:57 ratio mix of acetonitrile:dichloromethane.

37

Chapter 3

4.6
4.4
Acetonitrile
(%)

4.2

34
36
38
40
43
48
55
60
70

Log Mp

4
3.8
3.6
3.4

(A)

3.2
3
2

10
12
14
Retention time (min)

16

18

20

4.3
4.1

2.9

Acetonitrile
(%)
34
36
38
40
43
48
55
60
70

2.7

(B)

3.9

Log Mp

3.7
3.5
3.3
3.1

2.5
2

12

17

22

Retention time (min)

Figure 3.6

(A) Calibration plots of log Mp vs. retention time for PMMA standards in
different mobile phase mixtures.

(B)Calibration plots of log Mp vs. retention time for PMMA samples with one OH end-group in
different mobile phase mixtures (The values as indicated in the figure refer to the volume
percentage of acetonitrile in dichloromethane).

38

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

Another display of the critical condition determination experiments is presented in Figure 3.6.
The same conclusion applies to this graph: non-functional and mono-functional PMMA are
separated based on their functionality, independent of their molar mass when a 43:57 ratio mix
of acetonitrile : dichloromethane is employed
Using this critical-liquid-chromatography method, we were able to confirm the presence of
polymer chains with two hydroxyl end groups in PMMA2. These can be formed by the coupling
of two thiol-terminated polymer chains during aminolysis, resulting in the formation of a
disulfide bond. As depicted in Figure 3.7 (Trace 2), the reaction of a hydroxyl-functional
PMMA synthesized by RAFT polymerization with 1-hexylamine resulted in a polymer with a
noticeable fraction of difunctional chains, confirming the hypothesis of the formation of a
disulfide bond. The small fraction of non-functional polymer might be the result of transfer by
hydrogen abstraction to solvent or polymer chain during the polymerization.
0-OH

1-OH

2-OH

(4)

(3)

(2)

(1)
0

50

100

150

200

250

300

Elution time (s)

Figure 3.7. Isocratic Liquid-chromatography separations of PMMA samples based on their OHfunctionality (1) PMMA1: PMMA (Mn=17.000 g/mol, PDI=1.16), synthesised by RAFT
polymerization (2) PMMA2: PMMA1 after reaction with 1-hexylamine in the absence of
sodium bisulfite; (3) PMMA3: PMMA1 after reaction with 1-hexylamine in the presence of
sodium bisulfite; (4) PMMA4: PMMA3 after Michael addition reaction with HEA in the
presence of DABCO and Na2S2O4

39

Chapter 3

Although all required precautions were taken (distillation of the solvent, removal of the
oxygen by freeze-pump-thaw cycles), all RAFT polymers made by us seemed to be prone to
such an oxidative coupling after the treatment with 1-hexylamine. The observation of shoulders
on the GPC chromatograms is the basis for this affirmation. Nevertheless, the coupling of the
thiols can be suppressed by the addition of an anti-oxidant (aqueous solution of sodium bisulfite
Na2S2O4) to the reactive medium. This is confirmed by the suppression of the shoulder on the
SEC chromatograms (not shown) and the detection of only monofunctional polymer in the LC
(Figure 3.7, Trace 3). It has to be noted that Wang et al.20 proposed a reductive system with zinc
and acetic acid to prevent the oxidative coupling of thiols to disulfides. However, the necessity
of maintaining the alcohol functions in our system prohibits the use of a compound with a
carboxylic acid function, and therefore we had to employ a different method to suppress the
coupling reaction.
Another side reaction of the nucleophilic cleavage of thiocarbonyl thio groups was revealed
by MALDI-TOF-MS experiments on the polymers after treatment with a base (Figure 3.8) and
by ESI-QTOF-MS (Electrospray-ionization-quadrupole-TOF)7.

OH

(PMMA)

H
HS

CN

OH

(PMMA)

OH

(PMMA)

CN

CN

2938.78

2970.67

2920

2940

2960

2980

3000

3038.88

3020

3040

3060

m/z

Figure 3.8. MALDI-TOF-MS spectra of PMMA RAFT-polymer after aminolysis; counter ion =
Na+
As expected, the family of peaks corresponding to the thiol-terminated polymers was found,
with a difference between the peaks of 100 mass units, corresponding to the MMA moiety.
However, an unexpected second family of peaks was also observed, higher than the previous
one by 32 mass units. The assignment of the peaks reveals a hydrogen-atom as an end group

40

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

instead of the expected thiol-group. Following the conclusions of Ladaviere et al.10, a hydrogenradical transfer occurring during the treatment with the base can partially explain the presence of
this family of peaks. Another possible explanation will be given below. Another family was
detected with a difference of 40 m/z relative to the hygrogen-terminated one. So far, we cannot
propose a satisfactory structure assignment for this family. The disulfide-containing chains
could not be observed by mass-spectrometry techniques. We assume that the weak sulfur-sulfur
bond was broken during the analysis.
The second end group modification step of Scheme 3.2 relies on the ability of thiol
compounds to bind to a single molecule of a vinylic compound with an activated double bond
(e.g. acrylates). This Michael addition reaction is known to be essentially quantitative for
organic compounds9,19. In our polymerizations using ACP and RAFT-ACP as initiator and
RAFT agent, respectively, almost all the chains bearing a thiocarbonyl thio function possess an
alcohol function at the other end of the chain (according to the mechanism described by
Rizzardo5, and confirmed by our analysis, Figure 3.7). After the aminolysis of these RAFT
polymers, the thiol end groups obtained can be further reacted with a hydroxyl-functional
acrylate (e.g. 2-hydroxyethyl acrylate - HEA). The side reaction of the aminolysis that results in
hydrogen-terminated chains will have a major impact on the ultimate yield of bifunctional
hydroxyl-terminated polymers: the hydrogen-terminated chains can not be converted to a
hydroxyl-functional group during the Michael addition. This results in a lower fraction of
telechelic polymers. Some dead material resulting from the RAFT polymerization itself can be
an obstacle as well to the production of these telechelic material in a high yield.
In the second step of our two-step post-polymerization procedure (Scheme 3.1), the thiolfunctional polymer reacts with a hydroxyl-functional acrylate in a Michael addition reaction.
The yield of this Michael reaction could not be determined accurately by 1H-NMR, due to the
low intensities of the signals arising from the protons of the end groups and/or their overlap with
signals corresponding to the protons of the polymer backbone. Quantitative results were
obtained with liquid-chromatography (LC) techniques. Figure 3.5 shows the LC chromatograms
of a RAFT PMMA after each of the two steps of the post-polymerization treatment. The
cleavage of the thiocarbonyl thio function in the presence of sodium hydrosulfite (Trace 3)
results in a monofunctional polymer (as described above). The Michael addition of HEA on the
thiol-functional polymer in the presence of Na2S2O4 is reflected in the chromatogram by the
appearance of a large peak in the 2-OH region. However, the yield achieved (66.7 % - Table
3.9) was relatively low compared to those obtained with model compounds (for example, the
addition of HEA on dodecanthiol was nearly quantitative).

41

Chapter 3

Sample

% 0-OH

% 1-OH

% 2-OH

PMMA1

0.8

99.2

0.0

PMMA2

0.8

65.2

34.0

PMMA3

0.7

99.3

0.0

PMMA4

0.2

33.1

66.7

PMMA1: PMMA (Mn=17.000 g/mol, PDI=1.16), synthesised by RAFT polymerization; PMMA2:


PMMA1 after reaction with 1-hexylamine in the absence of sodium bisulfite; PMMA3: PMMA1 after
reaction with 1-hexylamine in the presence of sodium bisulfite; PMMA4: PMMA3 after Michael addition
reaction with HEA in the presence of DABCO and Na2S2O4.

Table 3.9. Quantitative differentiation of PMMA-functional chains, according to their OHfunctionality, by icocratic liquid-chromatography experiments.
This low yield can be explained by several factors. The termination by disproportionation that
occurs during the RAFT polymerization leads to dead chains without the thiocarbonyl thio
function as an end group. The two-step post-polymerization treatment will not affect those
chains, which will consequently remain monofunctional. A solution to partly circumvent this
problem would be to stop the polymerizations at a lower conversion, limiting the fraction of
dead chains in the polymerization product.
Another explanation for the low yield is the formation of hydrogen-terminated chains as a side
reaction during the aminolysis of the thiocarbonyl thio groups. These cannot be functionalized
further by the Michael addition. The accessability of the SH group may also be a point of
concern. A coiling effect may prevent the reaction from occurring quantitatively, lowering the
percentage of , - telechelic polymer obtained. As a final remark, we could not accurately
estimate the yield of the Michael addition of the thiol-functional poly(methacrylate). After the
aminolysis, the hydrogen-terminated and the thiol-terminated chains could not be differentiated
by an analytical technique in a quantitative way. Some side-reactions might also occur during
this Michael addition reaction, lowering the ultimate yield of di-hydroxy telechelic chains.
As an interesting remark, we want to add that the two-step post polymerization procedure
resulted into telechelic polymers without a labile thiocarbonyl thio function. Apart from the
disappearance of the red color, the absence of this function enhances the thermal stability (the
thermogravimetric traces of a RAFT-PMMA and a telechelic PMMA are presented in Figure
3.10).

42

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

100

(2)
80

(1)

60
40
20
0
100

200
300
Temperature (C)

400

500

Figure 3.10. Thermogravimetric analysis (TGA) traces of RAFT PMMA (1)


(HO(CH2)3CNC(CH3)(C5H8O2)nCS2C6H5) and telechelic PMMA (2)
(HOCH2CH2OOCCH2CH2S(C5H8O2)nCCN(CH3)(CH2)3OH)

Due to the lack of availability of reference materials, a similar liquid-chromatography


technique could not be developed yet for PBMA. In this case, MALDI-TOF-MS was employed
to investigate the outcome of the same two-step post-polymerization procedure. The MALDITOF-MS spectrum obtained for a PBMA RAFT-polymer after aminolysis and Michael addition
with HEA is presented in Figure 3.11. Although it is impossible to draw any quantitative
conclusions from the MALDI-TOF-MS, we can clearly observe that a significant amount of
telechelic polymer chains with two hydroxyl end groups, corresponding to the expected
structure, were obtained.

43

Chapter 3

2000

3000

4000

m/z (Da)

2559.01
2416.93

(1)

O
HO

S (PMMA)

OH
CN

2400

2420

2440

2460

2480

2500

2520

2540

2560

m/z (Da)

2416.57

2558.66

(2)

2400

2420

2440

2460

2480

2500

2520

2540

2560

m/z (Da)

Figure 3.11. Experimental (1) and theoretical (2) MALDI-TOF-MS spectra of PBMA RAFTpolymer after reaction with 1-hexylamine and HEA; counter ion = Na+

44

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

Comprehensive 2D LCxSEC separation


In the course of synthesizing RAFT PMMA, some preliminary experiments were made using
AIBN as an initiator rather than ACP. The polymer chains obtained from a AIBN/RAFT-ACP
initiator/RAFT agent combination could have initiated during the polymerization by two types
of radicals (neglecting chain transfer or side-reactions): either the leaving group of the RAFT
agent (bearing a hydroxyl function) or the fragment derived from the AIBN (non-hydroxy
functional). The LC method was employed for the analysis of this polymer (VL37A - Mn= 2.600
g/mol, PDI = 1.29 Figure 3.12).
0-OH

0.7

1-OH

2-OH

0.6

ELSD response

0.5
0.4
0.3

VL37 A

0.2
0.1

VL37B

0
0

50

100

150

200
Time (s)

250

300

350

Figure 3.12. Separation of functional RAFT polymers according to the number of OH groups.
Evaporative light-scattering detector, mobile phase 43 % acetonitrile in dichloromethane, homepacked silica column (150 mm 4.6 mm i.d.; 3-m particles; 10nm pore size), flow rate 1
mL/min.
A step further was taken with the decision to analyze this sample in a second dimension. The
LC analysis provided useful information on the functionality distribution, the chains being
separated according to the number of hydroxyl groups they were bearing, however no
information on the molar mass averages could be collected with this method. The determination
of common characteristics of synthetic polymers such as the different molar masses (Mn, Mw,
Mz) or the polydispersity index required an extra chromatographical analysis (SEC). A method
consisting of two-dimensional liquid chromatography providing information on both the
functionality-type distribution (FTD) and the molecular weight distribution (MWD) was then
8

developed .

45

Chapter 3

The details on the elaboration of the method can be found in reference 8. In this chapter, the
focal point of the discussion will be given to the interpretation of the chromatograms obtained
by LCxSEC.

(a)

(b)

Figure 3.13. Two-dimensional LCSEC chromatogram of sample VL37A. (a) UV detection at


300 nm. (b) ELSD. The LC mobile phase was 50 % ACN in DCM and the flow rate was 4
L/min.
In Figure 3.13, the UV detection reveals the presence of 4 distinct species in polymer
VL037A. The following assignment can be made:
-

Peak #1: polymers having a relatively high molecular weight and no OH-functionality.
Polymer chains initiated by a non-functional AIBN fragment are probably detected with
this peak. The growth of the chain has been controlled by the thiocarbonyl thio
compounds in a RAFT process, after the initiation by the AIBN-derived radical. Some
chains initiated by an alkyl group after a chain transfer reaction might as well be present
in the population represented by this peak. It corresponds mainly to the peak in the 0-OH
region on Figure 3.12.

Peak #2: relatively high molar masses are observed with hydroxyl functionality. The
typical RAFT recipes result in majority into polymer chains initiated by the leaving group
of the RAFT agent. In our polymerization system, the RAFT agent bears a hydroxyl

46

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

group, hence it is logical to find a peak in this position. The population responsible for this
peak is mainly the one causing the peak in the 1-OH region on Figure 3.12.
-

Peak #3: low molar masses compounds with no hydroxyl-functionality are observed.
Several explanations might be given for this compound. During the reaction, the AIBN
decomposes but has only an initiation efficiency of 0.6-0.7 (depending on the exact
conditions), which means that 30 to 40 % of the produced radicals will not initiate new
growing chains, but will terminate form TetraMethyl SuccinodiNitrile (TMSN).

CN CN
TMSN
However, the absence of a chromophore prevents the TMSN molecule from being
decteted. Another possibility to explain the presence of peak #3 is the termination of very
short chains (a couple of monomer units added) at the beginning of the polymerization.
Following the initiator decomposition profile, the concentration of radicals is the highest
at the beginning of a radical polymerization, before the establishment of a steady state. It
is then the moment when the termination probability is the highest. The transition region
between peak #1 and peak #3 can be explained by termination of chains throughout the
polymerization. Once again, the termination procedure results in chains without a
thiocarbonyl thio function at the end. Hence, it forbids the use of this argument as a valid
explanation.
This peak might also be caused by the recombination of a thiocarbonyl thio radical and a
radical derived from the AIBN. This forms a RAFT-AIBN compound.
-

Peak #4: some low molar masses compounds with an hydroxyl function are detected. LCESI-MS experiments showed the presence of unreacted RAFT agent in the final product
obtained after polymerization. Short chains with one or two monomer units were also
detected via mass spectrometry.

The presence of peak #1 and peak #2 were expected and are in line which what should be
expected from a RAFT polymerization using a hydroxyl-functional RAFT agent and a nonhydroxyl functional initiator. The peaks #3 and #4 are more intriguing as it seems improbable
that a RAFT agent (with a transfer constant that can be estimated to be around 100) can remain
untouched after 8 hours of polymerization at 70C. No satisfactory explanation can be provided
so far.

47

Chapter 3

Quantitative results concerning molar masses and concentration for peak #1 and peak #2 were
determined by both ELSD and UV detection. Data for peaks #3 and #4 are not presented, due to
the low molar mass of the chains, that were below the lowest standard used for calibration of the
SEC (Mn = 620 g/mol). Their combined molar concentrations represent approximately 1% in
sample VL037A. The quantitative results are presented in Table 3.14 (ELSD calibration details
presented in Reference 8).

UV, 300nm
Peak name

Calibrated ELSD
Conc.*

Mn,

Mp,

mol %

kg/mol

kg/mol

1.28

16

2.1

2.5

1.24

3.2

1.28

84

2.4

2.9

1.27

91

3.0

1.30

N/A

2.3

2.9

1.28

N/A

Mn,

Mp,

kg/mol

kg/mol

Peak #1

2.3

2.3

Peak #2

2.7
2.6

PDI

PDI

Conc.*
mol %

Peaks #1 and
#2
(combined)
*

Combined molar concentration of peak 3 and 4 represents less than 1 mol % of the total chains

Table 3.14. Quantitative results for two-dimensional LCxSEC separation of sample VL037A
using UV detection at 300 nm and calibrated ELSD.
Small differences in the relative concentrations of peak #1 and peak #2 can be noted by
comparing the results obtained with ELSD detection and UV detection at 300nm. A possible
explanation can be the presence in the population represented by peak #2 of chains not
terminated by the thiocarbonyl thio chromophoric group. Termination by disproportionation
occurring during the polymerization will produce hydrogen-terminated and double-bondterminated chains. If those chains were initiated by the RAFT leaving group, they will be
hydroxyl-functional and will contribute to the peak #2. However the absence of a chromophore
will prevent them from being detected by the UV detector, resulting in an underestimation by
ELSD of the concentration of the chains producing peak #2.
The sample VL037A was treated with 1-hexylamine in order to reduce the thiocarbonyl thio
function into a thiol. Some coupling between thiols occurs, resulting in the formation of
disulfides. Coupling of hydroxyl-functional chains will be noticed in the chromatogram by the
appearance of a peak in the dihydroxyl-functional region. (VL37B - Mn= 2.800 g/mol, PDI =

48

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

1.35 Figure 3.12). The 2D-LC x SEC method was used for characterization of this sample
(Figure 3.15).

Figure 3.15. Two-dimensional LCSEC chromatogram of sample VL37B detected by ELSD.


The analytical conditions were chosen identical to those of Figure 3.13
The detection method for the SEC analysis was restricted to ELSD due to the fact that nearly
all the chains that were bearing a chromophoric group (thiocarbonyl thio group) in VL037A
have lost this fragment in the aminolysis process. The lack of chromophores forbids the use of
UV detection in order to draw quantitative conclusions. Using ELSD, the quantitative analysis
of sample VL37B is presented on table 3.16.

Peak Name

Mn, kg/mol

Mp, kg/mol

PDI

Conc. (mol %)

Peak #1

2.4

2.5

1.25

Peak #2

2.6

3.0

1.31

86

Peak #3

3.2

5.2

1.35

2.7

3.0

1.32

N/A

Peaks #1, #2 and


#3 (combined)

Table 3.16. Quantitative results for LCxSEC separation of sample VL037B without Na2S2O4
using calibrated ELSD.
The main conclusion from the 2D-LCxSEC experiment is the straightforward observation of
the disulfide-containing chains. The double hydroxyl functionality recorded coupled with the
approximate double mass of the Mp gives in a single chromatographic experiment a double

49

Chapter 3

proof backing up the disulfide formation hypothesis. This conclusion is similar to the one drawn
after the individual SEC and LCCC experiments, as was described earlier in this chapter. After
these analytical investigations, the mechanism proposed in Scheme 3.2 can be updated with our
findings. The Scheme 3.17 acknowledges the presence of the side reactions occurring during the
aminolysis.

OH

HO

CN

N N

CN

OH
CN

ACP

RAFT-ACP
methacrylate,
Toluene, 70C

S (Polymethacrylate)

OH
CN

hexylamine,
THF, RT

(Polymethacrylate)

HO
HS (Polymethacrylate)

OH
CN

S S

CN

CN

OH

(Polymethacrylate)

H (Polymethacrylate)

OH
CN

Michael addition
with HEA

O
HO

S (Polymethacrylate)

OH
CN

Scheme 3.17. Synthesis of hydroxyl-telechelic polymethacrylates

50

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

V) Conclusion
In this chapter, a synthetic path to obtain , functional linear polymers employing the
RAFT polymerization process was proposed. With a proper choice of the initiator - RAFT agent
system, the two-step post-polymerization procedure (namely, an aminolysis of a thiocarbonyl
thio group, followed by a Michael addition on the resulting thiol) was demonstrated to yield
telechelic polymers. Liquid-chromatography separations proved the successful synthesis of
telechelic PMMA to a significant extent. The absence of the thiocarbonyl thio moiety results in
a colorless material with an enhanced thermal stability. However, the fraction of dihydroxylfunctional PMMA chains in the final product was limited by side reactions, occurring during the
RAFT polymerization of MMA and during the aminolysis procedure. The main undesirable
reaction was found to be the formation of disulfide bridges from two thiol-functional polymer
chains. Addition of an aqueous solution of sodium bisulfite suppressed this oxidative coupling
reaction. The efficiency of this procedure was demonstrated with the help of liquidchromatographic separations. It was also found, using mass-spectrometric techniques, that the
reduction of thiocarbonyl thio groups into thiols is accompanied by the production of hydrogenterminated polymer chain. MALDI-TOF-MS spectra were used to demonstrate that RAFT
polymerization of BMA, followed by the two-step end group modification procedure, yielded
hydroxyl-functional telechelic poly(butylmethacrylate). The proposed method to synthesize
telechelic hydroxyl polymers showed disappointing results in terms of average functionality for
PMMA. The synthesized polymers will not be of satisfactory purities for the network formation
studies.
A comprehensive two-dimensional liquid chromatography technique (LCxSEC) was
developed in order to simultaneously measure the molecular-weight distribution (MWD) and
functionality-type distribution (FTD) for hydroxyl-functional PMMA. Quantitative results were
obtained for a RAFT-PMMA before and after aminolysis.

51

Chapter 3

Reference List
1. Clouet, G., Knipper, M., Brossas, J.; Polym. Bull. (Berlin) 1984, 11, 171.
2. Cools, P. J. C. H., Van Herk, A. M., German, A. L., Staal, W.; Journal of Liquid
Chromatography 1994, 17, 3133.
3. Falkenhagen, J., Much, H., Stauf, W., Mueller, A. H. E.; Macromolecules 2000,
33, 3687.
4. Gorbunov, A., Trathnigg, B.; J. Chromatogr. A 2002, 955, 9.
5. Gorshkov, A. V., Much, H., Becker, H., Pasch, H., Evreinov, V. V., Entelis, S. G.;
Journal of Chromatography 1990, 523, 91.
6. Jiang, X., Lima, V., Schoenmakers, P. J.; J. Chromatogr. A 2003, 1018, 19.
7. Jiang, X., Schoenmakers, P. J., van Dongen, J. L. J., Lou, X., Lima, V., BrokkenZijp, J.; Anal. Chem. 2003, 75, 5517.
8. Jiang, X., van der Horst, A., Lima, V., Schoenmakers, P. J.; J. Chromatogr. A
2005, accepted.
9. Kharasch, M. S., Fuchs, C. F.; J. Org. Chem. 1948, 13, 97.
10. Ladaviere, C., Doerr, N., Claverie, J. P.; Macromolecules 2001, 34, 5370.
11. Le, T. P.; Moad, G.; Rizzardo, E.; Thang, S. H. 97-US12540 WO 9801478
January 15, 1998
12. Macko, T., Hunkeler, D., Berek, D.; Macromolecules 2002, 35, 1797.
13. Mengerink, Y., Peters, R., de Koster, C. G., van der Wal, S., Claessens, H. A.,
Cramers, C. A.; J. Chromatogr. A 2001, 914, 131.
14. Mengerink, Y., Peters, R., van der Wal, S., Claessens, H. A., Cramers, C. A.; J.
Chromatogr. A 2002, 949, 337.
15. Pasch, H., Rode, K., Chaumien, N.; Polymer 1996, 37, 4079.
16. Philipsen, H. J. A., Klumperman, B., van Herk, A. M., German, A. L.; J.
Chromatogr. A 1996, 727, 13.
17. Rizzardo, E., Chiefari, J., Chong, B. Y. K., Ercole, F., Krstina, J., Jeffery, J., Le, T.
P. T., Mayadunne, R. T. A., Meijs, G. F., Moad, C. L., Moad, G., Thang, S. H.;
Macromol. Symp. 1999, 143, 291.

52

Synthesis of hydroxyl-telechelic polymethacrylates by RAFT

18. Rizzardo, E.; Thang, S. H.; Moad, G. 98-AU569 WO 9905099 July 20, 1998
19. Szabo, J. L., Stiller, E. T.; J. Am. Chem. Soc. 1948, 70, 3667.
20. Wang, Z., He, J., Tao, Y., Yang, L., Jiang, H., Yang, Y.; Macromolecules 2003,
36, 7446.

53

54

Synthesis of carboxy-telechelic polyacrylates by RAFT

CHAPTER 4
SYNTHESIS OF CARBOXY-TELECHELIC
POLYACRYLATES BY RAFT

I) Introduction
As described in the previous chapter, the quantitative preparation of hydroxyl-functional
telechelic PMMA was unsuccessful. Another strategy was investigated in order to prepare
telechelic polymers by RAFT polymerization with a higher purity. The previous study
concluded that the end-group modification procedure has to be avoided in order to obtain high
yields. With the use of bi-functional trithiocarbonates in conjugation with functional initiators, a
high percentage of telechelic polyacrylate is expected to be produced readily after the
polymerization, without any further modification needed. This expectation is based on a
publication from Lai et al.1 Production of telechelic polyacrylates and polystyrene was claimed
by them although the quantitative analytical characterization of those polymers was lacking. In
this chapter, we will reinvestigate the synthetic route proposed by Lai et al.1, and a particular
attention will be dedicated to the quantitative characterization for the end-groups of the
produced carboxyl-functional telechelic polymers, using liquid chromatography under critical
conditions.

II) Materials
Butyl acrylate (99 %) was purchased from Aldrich and was distilled over CaH2 prior to
utilization in order to remove the inhibitor. Toluene (Biosolve) and acetone (Aldrich) were
distilled over CaH2 before use. Azobis(isobutyronitrile) (AIBN, Merck) and 4,4-azobis(4cyanovaleric acid) (ACVA, Aldrich) were recrystallized from methanol. S,S-Bis(,dimethyl--acetic

acid)trithiocarbonate

and

S-1-dodecyl-S'-(,-dimethyl--acetic

55

Chapter 4

acid)trithiocarbonate were prepared according to the route described by Lai et al., yielding
purities above 99% (1H-NMR in CDCl3).
Polymerization procedure. All RAFT polymerizations were carried out in three-neck round
bottom flasks, heated in an oil bath. The reaction solution was degassed with argon and three
freeze-pump-thaw cycles were applied. The final product was isolated by rotary evaporation.
Polymerization of butyl acrylate (Figure 4.1) Butyl acrylate (42.5 g, 3.32 10-1 mol), ACVA
(0.72 g, 2.56 10-3 mol), S,S-bis(,-dimethyl--acetic acid)trithiocarbonate (6 g, 2.12
10-2 mol) toluene (41.1 g, 47.5 mL) and acetone (37.6 g, 47.5 mL) were mixed in a three-neck
round bottom flask. Oxygen was removed and the flask was immersed in an oil bath that was
preheated at 80 C. After 4 hrs of polymerization, PBA was isolated by rotary evaporation.

III) Characterization techniques


Monomer conversion was determined from the concentration of residual monomer, measured
using a Hewlett-Packard (HP 5890) GC, equipped with an AT-Wax capillary column (30 m
0.53 mm 10 m); toluene was used as internal reference.
SEC (Size-Exclusion Chromatography) was carried out using a Waters model 510 pump and a
model 410 refractive-index detector (at 313 K). The columns used were a PL-gel guard column
(5 m particles) 50 7.5 mm, followed by two PL-gel mixed-C (5 m particles) 300 7.5 mm
columns (313 K). THF was used as the eluent at a flow rate of 1 mL/min. Lowpolydispersity
polystyrene standards (Polymer Labs) with molecular weights ranging from 580 to 7.1 106
gmol-1 were used for calibration of the columns. Molecular weights were recalculated using the
Mark-Houwink parameters of PS, PMMA, PBMA and PBA in THF (PS: K = 1.140 10-4 dL/g,
a = 0.716; PMMA: K = 0.944 10-4 dL/g, a = 0.719; PBMA: K = 1.480 10-4 dL/g, a = 0.664;
PBA: K = 1.220 10-4 dL/g, a = 0.700). The samples (1 mg/mL THF) were filtered through a
0.2 m syringe filter prior to injection. Data acquisition and processing were performed using
Waters Millennium32 (v3.00) software.
MALDI-TOF-MS (Matrix-Assisted Laser-Desorption Ionization Time-Of-Flight Mass
Spectrometry) measurements were performed on a Voyager-DE STR instrument (Applied
Biosystems, Framingham, MA, USA) equipped with a 337 nm nitrogen laser. Positive-ion
spectra were acquired in the reflector mode. 2-[(2E)-3-(4-tert-butylphenyl)-2-methylprop-2enylidene]malononitrile was used as the matrix. All samples were hand-spotted on the target.
Sodium trifluoroacetate (CF3CO2Na) was added.

56

Synthesis of carboxy-telechelic polyacrylates by RAFT

A Waters (Milford, MA, USA) 2690 Alliance liquid chromatography system was used to
perform the isocratic LC experiments. This HPLC instrument contained a built-in auto-injector
with a sample loop allowing injection of variable sample volumes, and was equipped with a
Waters 996 PDA (photodiode-array detector) and a Sedex 55 evaporative light-scattering
detector (ELSD, temperature 62 C, N2 pressure 2.2 bar). The mobile phase was prepared in-situ
using the solvent-mixing capability of the instrument. The data collection and the data analysis
were handled by Waters Millennium 3.2 software. The columns used (150 mm x 4.6 mm i.d.)
were packed in-house with Hypersil Silica (3 m particles; 100 pore size; Shandon, Runcorn,
UK). PBA separations were performed at 25C using the following mobile phase: 6 v/v%
acetonitrile, 1 v/v % acetic acid (added as 10 v/v % acetic acid in dichloromethane) in
dichloromethane, at a flow rate of 0.5 mL/min.
The comprehensive two-dimensional LC experiments (LCSEC) were carried out using a
Shimadzu -LC 10ADvp pump at a flow rate of 20 and 40 l/min (Shimadzu, s Hertogenbosch,
The Netherlands). A Rheodyne two-position six-port injection valve (Berkeley, CA, USA)
equipped with a 20 l injection loop was used for the LC separation. The stationary phase used
in the LC was Hypersil silica (Shandon, Runcorn, UK), particle size 3-m, pore diameter 120 ,
column dimensions were 150 4.6 mm I.D., column temperature was kept constant at 55C in a
(Millipore) Waters temperature-control module. The 2D-SEC separation system consists of a
Kratos Spectroflow 400 pump (ABI, Ramsey, NJ, USA) equipped with two 50 4.6 mm I.D.
PL-columns (Polymer Laboratories, 5-m particles with 100 pore size and/or 6 m oligoPore
particles with 100 pore size, Church Stretton, Shropshire, UK). The SEC was coupled to a
Kratos Spectroflow 757 UV-absorbance detector (ABI), = 230 nm, and a Varex II A
evaporative light-scattering detector (ELSD) (Burtonsville, Maryland, USA), using N2 (3 bar) as
nebulizer gas at 65 C, the flow rate used in the SEC was 0.9 ml/min. For comprehensive
LCSEC, the LC and SEC systems were coupled by an air-actuated VICI two-position ten-port
injection valve (Valco, Schenkon, Switzerland). This valve was operated using a high-speed
switching accessory (switching-time of 20 ms using 5 bar N2) and dual injection loops of equal
volume (40 or 80 l depending on the LC flow rate, t0 SEC 2 min.). Data was collected using
a Keithley KNM-DCV 12 Smartlink interface (Cleveland, OH, USA). Two-dimensional plots
and distribution data were calculated with an in-house program written in a Matlab (Natick,
MA, USA) software environment. This program allowed us to extract LC and SEC
chromatograms at any position in the LCSEC contour plot. The Matlab software also
controlled the valve switching and recorded exact switching times for accurate slicing.

57

Chapter 4

IV) Results and discussion


The procedure described by Lai et al.1 was employed to synthesize telechelic carboxylfunctional polybutylacrylate. This route has the advantage that it leads in principal to telechelic
material in one step (Figure 4.1).
HOOC

COOH

HOOC

COOH

N N
CN

CN
ACVA

Acrylate
Toluene/acetone
0
80 C

HOOC

(Polyacrylate)

(Polyacrylate)

COOH

Figure 4.1. Synthesis of carboxyl-telechelic polyacrylates


We decided to polymerize butyl acrylate rather than methyl methacrylate. This choice was
dictated by the poor control over MMA polymerization using S,S-bis(,-dimethyl--acetic
acid)trithiocarbonate and S-1-dodecyl-S'-(,-dimethyl--acetic acid) trithiocarbonate as
RAFT agents. In MMA polymerizations using these RAFT agents, the stability of the leaving
group is not sufficient to ensure a fast fragmentation of the intermediate radical in the preequilibrium of the RAFT process. This will result in polydispersities that are too high for our
purpose (approximately 1.7). The BA polymerization conditions and main results are presented
in Table 4.2.
Lai et al. claimed to have synthesized linear telechelic polymers1 with COOH end-groups.
However, the analytical characterization was limited to SEC analysis, which does not prove the
production of telechelic polymers. In order to determine the actual fraction of telechelic chains
in our polymers, we developed a new liquid-chromatography method, suitable for separation of
poly(butyl acrylates) according to their carboxyl end group-functionality, under critical
conditions (i.e. retention independent of molecular weight)2-6. The procedure to establish the
critical conditions was similar to the one exposed in Chapter III. Systematic variation of the
mobile phase composition was used in order to find the dichloromethane : acetonitrile ratio
suitable for performing liquid chromatography in critical conditions.

58

Synthesis of carboxy-telechelic polyacrylates by RAFT

Sample

Initiator

CTA

[RAFT]/[Ini]

Conversion

Mn,exp (Mn, th)

Mw/Mn

VL103

AIBN

95 %

2100 (2200)

1.13

VL068

AIBN

20

99 %

2200 (2200)

1.12

VL123

ACVA

97 %

2800 (3200)

1.11

VL127

ACVA

99 %

2200 (2200)

1.11

VL131

ACVA

20

99 %

2300 (2200)

1.14

All polymerizations were carried out at 80C, in a toluene/acetone (1:1 vol/vol) mixture under an argon
atmosphere
A=

HOOC

C12H25

B=

COOH

COOH

Table 4.2. Conversions and molecular weights characteristics of the poly(butyl acrylate)
samples synthesized by RAFT polymerization
The critical conditions were more troublesome to find than in the case of hydroxyl-functional
PMMA. Using pure acetonitrile (known as a good desorbing solvent for PBA on silica columns)
as mobile phase resulted in full retention of mono and di-functional chains on the bare silica
column. We circumvent this problem by adding some acid (10 v/v % formic acid or acetic acid
in dichloromethane) to the mobile phase.

PBA 600
PBA 7390

Retention time (min)

PBA 32333

(a)

4
3
2
4

10

11

Acetonitrile (%)

59

Chapter 4

PBA 1-COOH
2610
PBA 1-COOH
13260

Retention time (min)

PBA 1-COOH
19090

(b)

4
3
2
4

10

11

Acetonitrile (%)

10

PBA 2-COOH 2400


PBA 2-COOH 3200

Retention time (min)

PBA 2-COOH 5540

PBA 2-COOH 11450

(c)

7
6
5
4
3

10

11

Acetonitrile (% )

5
PBA 0-COOH
PBA 1-COOH

4.5

PBA 2-COOH

(d)

LogMp

4
3.5
3
2.5
2
3

Retention time (min)

Figure 4.3. Dependence of retention time on the mobile-phase composition (at 0.5 % HAc) for
PnBA samples with different molecular weights.
(a) Non-functional PnBAs. (b) Mono-functional carboxyl PnBAs. (c) Di-functional carboxyl
PnBAs. (d) Molecular-weight effect on retention time for non-, mono- and di-carboxyl
functional PnBAs under near-critical conditions (6 % ACN and 0.5 % HAc in DCM).

60

Synthesis of carboxy-telechelic polyacrylates by RAFT

ELSD detector, home-packed Hypersil silica column (150 mm x 4.6 mm i.d.; 3-m particles;
10nm pore size), flow rate 0.5 mL/min, column temperature 55C.
It can be noticed in Figure 4.3 that the critical point, where the retention time is independent
of the molar mass of the poly(butyl acrylate) chains (determined by the crossing points of the
different curves for non-, mono- and di-COOH-functional chains), is slightly different for the
three types of polymer. For the non-functional chains, the critical conditions are observed for 6.6
v/v % acetonitrile in DCM, while it is 5.8 v/v % for mono-functional and 5.5 v/v % for difunctional. This fact is consistant with the observations of Gorbunov et al.7 and Skvortsov et al.8
They stated that strong interactions between a functional polymer and a stationary phase will be
dependant of the molar mass, the interaction decreasing with an increasing molar mass, resulting
in a lower retention. We encountered this situation in our separations, with the strong interaction
between the carboxyl group and the silica column. To counterbalance the effect of the molecular
weight on the retention of carboxyl-functional poly (butyl acrylate), the mobile phase mixture
should contain less desorbing solvent, acetonitrile in our case. That explains that the amount of
acetonitrile required decreases with increasing number of end-groups. It can be seen in Figure
4.3 that the retention time for nonfunctional poly (butyl acrylate) increased with increasing
molecular weight, which is the opposite of the behaviour observed for mono- and di-functionalCOOH-functional poly (butyl acrylate). As a final note, we observed a much clearer separation
for low molar mass functional polymers than for high molar mass polymers. In this chapter, we
will focus on polymers with Mn < 5,000 g/mol. The established conditions (6 v/v % acetonitrile,
0.5 v/v % acetic acid in dichloromethane) are then very suitable for separations of poly (butyl
acrylate) chains based on carboxyl functionality, independent of the molar masses of the chains.
The chromatograms acquired are depicted in Figure 4.4, the quantitative results are presented
in Table 4.5. The ELSD calibration curves for mono - and di-carboxyl-functional PBAs were
established using known injected amount of polymers VL131 and VL127 (Table 4.2),
respectively. The calibration curves fit very well with a power law.
As seen in Figure 4.4, the products of polymerizations VL103 and VL068 (Table 4.2) using
AIBN as an initiator showed major peaks in the 2-COOH region. However, a significant peak in
the 1-COOH region can also be observed. The peak in the 2-COOH region can be attributed to
structure 1 (Figure 4.6, expected RAFT polymer structure) or 4 (Figure 4.6, dead chain resulting
from the combination of two chains initiated by the leaving group of the RAFT agent). The peak
in the 1-COOH region can be explained by the structures 2 (Figure 4.6, RAFT polymer
structure, one chain initiated by AIBN) and 5 (Figure 4.6, dead chain resulting from the

61

Chapter 4

combination of two chains, one initiated by the leaving group of the RAFT agent, and the other
by AIBN). The relative intensity of the 1-COOH peak compared with the 2-COOH one is higher
in the case of polymer VL103 as for VL068. This is consistent with the amount of initiator in
VL103, which is more than twice as large than in VL068. No signals in the 0-COOH region is
observed in either chromatogram. This indicates that in both experiments, the amounts of
materials 3 (Figure 4.6, RAFT polymer, with two chains initiated by AIBN) and 6 (Figure 4.6,
dead chain resulting from the combination of two chains initiated by AIBN) in the samples are
negligible. This was expected, considering the relatively low concentrations of initiator used.

0-COOH

1-COOH

2-COOH

ELSD response

VL131

VL127
VL123
VL068
VL103

10

Time (min)
Figure 4.4. Liquid-chromatography separations of PBA samples bases on their COOHfunctionality

62

Synthesis of carboxy-telechelic polyacrylates by RAFT

Sample

% Mono-functional

% Di-functional

Initiator used

VL103

97

AIBN

VL068

99

AIBN

VL123

99

ACVA

VL127

99

ACVA

VL131

99

ACVA

Table 4.5. Quantitative analysis of carboxyl-functional PnBAs by LC-ELSD. Isocratic


conditions and column as in Figure 4.3 (samples as indicated in Table 4.2).
HOOC

(PBA)

(PBA)

COOH

HOOC

(PBA)

(PBA)

CN

(PBA)
CN

(PBA)

CN

3
HOOC

(PBA) (PBA)

COOH

HOOC

(PBA) (PBA)
CN

(PBA) (PBA)
CN

CN

HOOC

(PBA)

CN

(PBA)

COOH

HOOC

(PBA)

CN

(PBA) (PBA)

COOH

(PBA)

7
HOOC

CN

HOOC

COOH

CN

CN

COOH

(PBA) (PBA)
CN

10

Figure 4.6. Possible structures resulting from the butyl acrylate-polymerizations using S,SBis(,-dimethyl--acetic acid)trithiocarbonate as RAFT agent and AIBN or ACVA as
initiator.

63

Chapter 4

MALDI-TOF-MS was used as an analytical technique to study the RAFT polymer VL068
despite considerable debate in the literature (notably concerning chain fragmentation during
analysis). The spectra that were acquired for trithiocarbonate-containing polymers showed no
evidence of chain fragmentation during the analysis (Figure 4.7). In this respect, analysis by
MALDI-TOF-MS seems easier for the RAFT polymers with trithiocarbonate groups than for
those with thiocarbonyl thio functions. In the spectrum, groups of three peaks separated by 22
mass units can be seen, each triplet being separated from the next one by a mass of 128.2
mass units (corresponding to the butyl acrylate monomer mass). The separation of each peak in
the triplet by 22 mass unit can easily be explained, considering the experimental procedure
employed for the sample preparation in the MALDI-TOF-MS analysis. The samples are mixed
with a sodium salt, in order to facilitate the ionisation of the macromolecule. The sodium can
easily exchange with the hydrogen atom of the carboxylic acid group, leading to an increase in
mass for the polymer chain of 22 g/mol. The majority of the polymer chains in VL068 contains
two carboxyl groups. The exchange of zero, one or two hydrogen(s) of the carboxylic acid
groups explains the presence of a family of three peaks separated by 22 mass units. MALDITOF-MS, despite its non-quantitative character, indicates the overwhelming presence of
polymer chains with structure 1 (Figure 4.6), confirming the results obtained by liquid
chromatography.

1500

2000

2500

3000

3500

m/z (Da)

Figure 4.7. MALDI-TOF-MS spectra of PBA RAFT-polymer VL068. Counter ion= Na+

64

Synthesis of carboxy-telechelic polyacrylates by RAFT

The liquid-chromatography separations performed on samples VL123 and VL127 (Table 4.2)
revealed the dominant presence of dicarboxyl polymer chains (Table 4.5). Assuming that
termination of propagating poly(butyl acrylate) chains occurs only by combination, the expected
structures in experiments VL123 and VL127 are 1 (Figure 4.6, expected RAFT polymer
structure), 4 (Figure 4.6, dead chain resulting from the combination of two chains initiated by
the leaving group of the RAFT agent), 7 (Figure 4.6, RAFT polymer structure, one chain
initiated by ACVA), 8 (Figure 4.6, RAFT polymer, with two chains initiated by ACVA), 9
(Figure 4.6, dead chain resulting from the combination of two chains, one initiated by the
leaving group of the RAFT agent, and the other by ACVA) and 10 (Figure 4.6, dead chain
resulting from the combination of two chains initiated by ACVA). All those structures are
dicarboxyl-functional and cannot be differentiated with the liquid-chromatography method that
was employed. The presence of small amounts of monofunctional chains is probably the result
of hydrogen-transfer reactions. Since such reactions are inherent to any kind of radicalpolymerization process, the presence of this peak does not require too much attention. It has to
be noted that these transfer reactions also provide a reasonable explanation for the presence of
monofunctional peaks in the chromatograms of VL103 and VL068.
S-1-Dodecyl-S'-(,-dimethyl--acetic acid)trithiocarbonate was also used as a RAFT
agent in the polymerization of butyl acrylate. The use of this particular RAFT agent was dictated
by the desire to obtain monofunctional carboxyl chains, with a controlled molecular weight and
a low polydispersity. The possible structures of the polymer-chains at the end of this
polymerization (VL131 Figure 4.2) are presented in Figure 4.8, and the analysis of the
polymer obtained by liquid-chromatography in Figure 4.4, Trace 5.
The liquid chromatogram showed that the RAFT polymerization resulted exclusively in
monofunctional carboxyl chains. These monofunctional polymers can have the following
structures: 1 (Figure 4.8, expected RAFT polymer structure) or 2 (Figure 4.8, RAFT polymer,
with a chain initiated by ACVA). If the termination reaction is again assumed to occur only by
combination, the chromatogram shows no terminated product in the case of VL131. This was
unexpected, as termination reactions are not suppressed or limited in RAFT polymerization,
contrary to ATRP or NMP. The structures that result from termination by combination are: 3
(Figure 4.8, dead chain resulting from the combination of two chains initiated by the leaving
group of the RAFT agent), 4 (Figure 4.8, dead chain resulting from the combination of two
chains, one initiated by the leaving group of the RAFT agent, and the other by AIBN) and 5
(Figure 4.8, dead chain resulting from the combination of two chains initiated by ACVA)

65

Chapter 4

C12H25

(PBA)

COOH

C12H25

(PBA) (PBA)

COOH

(PBA)

(1)

HOOC

CN

(2)

COOH

HOOC

COOH

(PBA) (PBA)
CN

(3)

(4)
HOOC

COOH

(PBA) (PBA)
CN

CN

(5)
Figure 4.8. Possible structures resulting from the butyl acrylate-polymerization using S-1Dodecyl-S'-(,-dimethyl--acetic acid)trithiocarbonate as RAFT agent and ACVA as
initiator.
The absence of a small signal in the dicarboxyl region for sample VL131 may be partially
explained by the chromatographic broadening at high elution times. Apparently the fraction of
dead chains is below the detection limit of the present method and equipment. Some modelling
work has been conducted in order to determine the theoretical amount of dead chains that should
be present in our polymerization system. This modelling work was extremely tough to conduct,
due to the impossibility to get access to all the kinetic constants involved in the RAFT
mechanism, through literature or simple experiments. With all the precautions needed for the
results of this modelling work, it seems that at the end of the polymerization the proportion of
dead chains (difunctional - terminated by combination) accounts for less than 1 % of the total
amount of chains. This would explain the non-detection of difunctional chains in sample VL131.
Nevertheless, even with the question mark surrounding the absence of difunctional chains in
VL131, the quantitative production of monofunctional chains is very suitable for our purpose.
Comprehensive 2D LCxSEC separation
Similarly to what was performed in Chapter 3 on hydroxyl-functional poly (methyl
methacrylate), the LC separation technique was coupled to a SEC separation in order to

66

Synthesis of carboxy-telechelic polyacrylates by RAFT

determine in a single experiment both the functionality-type distribution and the molecular
weight distribution. Due to time constraints, no quantitative results were obtained. The
following will then only provide qualitative informations.

1
2
2

(a)

(b)

(c)
Figure 4.9. Two-dimensional LCSEC chromatogram of sample VL103. (a) UV detection at
230 nm. (b) UV detection at 230 nm. (zoom on the 2-3-4 peaks) (c) ELSD. The LC mobile
phase was 50% ACN in DCM and the flow rate was 40 L/min.
The two-dimensional separation spectra exhibits 4 main peaks as can been seen in Figure 4.9.
Peaks 2, 3 and 4 are respectively peaks corresponding to polymer chains with 0, 1 and 2 COOH
functionality. The explanations for the presence of those 3 peaks are the same as the ones
provided for the one-dimensional LC-separation of VL103:
-

Peak #2: polymers having a relatively high molecular weight and no COOH functionality.
Two polymer chains initiated by a non-functional AIBN fragment are probably detected
with this peak. The growth of the chain has been controlled by the thiocarbonyl thio
compounds in a RAFT process, after the initiation by the AIBN-derived radical. Some
chains initiated by an alkyl radical after a chain transfer reaction might as well be present

67

Chapter 4

in the population represented by this peak. It corresponds mainly to the peak in the 0COOH region on Figure 4.4.
-

Peak #3: relatively high molar masses are observed with 1 COOH functionality. The
polymers chains might have been initiated by a non-functional AIBN-derived radical or
during the RAFT polymerization some transfer reaction have occurred, resulting in
trithiocarbonate-containing chains with one COOH-functional chain on one side of the
CS3 moiety and hydrogen- or AIBN- terminated chain on the other side.

Peak #4: relatively high molar masses are observed with 2 COOH functionalities. The
typical RAFT recipes result in majority into polymer chains initiated by the leaving group
of the RAFT agent. Having employed di-functional trithiocarbonate, this peak belongs to
the expected population after a RAFT experiment. Both sides of the chains bear a COOHfunctionality.

However, similarly to what was found in the previous chapter, a peak at low molar masses is
detected (Peak #1). Although more investigations would be required, it is logical to assume that
it may be some remaining RAFT agent. The surprising finding of the last chapter seems to occur
with the trithiocarbonate as RAFT agent as well. Further work will be required to explain fully
the presence of this low molecular weight compound in the final polymer.

IV) Conclusion
In this chapter, a synthetic path to obtain , carboxyl functional linear poly(butyl acrylate)
employing the RAFT polymerization process was proposed. An existing procedure proposed in
literature was employed, and a new advanced chromathographic characterization was employed
to confirm the purity of the expected structures. The use of trithiocarbonate-containing RAFT
agents in combination of functional initiator was demonstrated to yield telechelic polymers and
monofunctional polymers with a very high purity concerning the end-group functionality (above
99%). MALDI-TOF-MS analyses were conducted on telechelic polymers, that confirmed the
structure of di-functional polymers. Contrary to the PMMA RAFT polymers discussed in the
previous chapter, no evidence of fragmentation during the analysis was observed (see Chapter 5
for more details).

68

Synthesis of carboxy-telechelic polyacrylates by RAFT

The highly pure mono- and di-functional poly(butyl acrylate) will be the materials of choice
for the network formation and network properties studies that will be described in Chapter 6.
The high purity and the stability of those compounds are the main reasons behind this choice.

69

Chapter 4

Reference List
1. Lai, J. T., Filla, D., Shea, R.; Macromolecules 2002, 35, 6754.
2. Cools, P. J. C. H., Van Herk, A. M., German, A. L., Staal, W.; Journal of Liquid
Chromatography 1994, 17, 3133.
3. Gorshkov, A. V., Much, H., Becker, H., Pasch, H., Evreinov, V. V., Entelis, S. G.; Journal
of Chromatography 1990, 523, 91.
4. Pasch, H., Rode, K., Chaumien, N.; Polymer 1996, 37, 4079.
5. Falkenhagen, J., Much, H., Stauf, W., Mueller, A. H. E.; Macromolecules 2000, 33, 3687.
6. Macko, T., Hunkeler, D., Berek, D.; Macromolecules 2002, 35, 1797.
7. Gorbunov, A., Trathnigg, B.; Journal of Chromatography, A 2002, 955, 9.
8. Skvortsov, A. M., Fleer, G. J.; Macromolecules 2002, 35, 8609.

70

MALDI-TOF-MS analysis of RAFT polymers

CHAPTER 5
MALDI-TOF-MS ANALYSIS OF RAFT POLYMERS

I) Introduction
Matrix Assisted Laser Desorption-Ionisation Time Of Flight Mass Spectrometry (MALDITOF-MS) is a powerful technique for the qualitative end-group characterization of polymers1-3.
During the course of this project, the determination of end-groups was of primary importance.
Hence, MALDI-TOF-MS was used on the different synthesized polymers in order to get
information on the nature of the end-groups. In this technique, the polymer is dispersed in a
solid matrix, desorbed by a UV-laser, and ionized by the presence of alkali salts. The main
advantages of the MALDI-TOF-MS technique versus the other mass spectrometric techniques
are the almost exclusive production of single-charged polymer molecules after ionization and
the high resolution of the obtained spectra (especially for low molar masses polymers, typically
with Mn < 10.000 g/mol). The main disadvantage is the sensitivity for fragmentation of certain
polymer chains during the analysis. Several reports in literature present RAFT polymers as
compounds prone to fragmentation under laser irradiation4-9.
In this chapter, MALDI-TOF-MS analysis of poly(butyl methacrylate) and poly(butyl
acrylate) RAFT polymers will be described. Conclusions will be drawn on the eventual
fragmentation during the analysis. The influence of the employed laser intensity has been
investigated. Also, the effect of the monomer nature and the type of RAFT agent used during the
synthesis will be highlighted.

II) MALDI-TOF-MS: Experimental procedure


Measurements were performed on a Voyager-DE STR (Applied Biosystems, Framingham,
MA, USA) instrument equipped with a 337 nm nitrogen laser. Positive-ion spectra were

71

Chapter 5

acquired

in

reflector

mode.

DCTB

(trans-2-[3-(4-tert-butylphenyl)-2-methyl-2-

propenylidene]malononitrile) was chosen as the matrix. Sodium trifluoroacetate (CF3CO2Na)


was added as the cationic ionization agent. The matrix was dissolved in THF at a concentration
of 40 mg/mL. Sodium trifluoroacetate (CF3CO2Na) was added to THF at a concentration of 1
mg/mL. The dissolved polymer concentration in THF was approximately 1 mg/mL. In a typical
MALDI experiment, the matrix, salt, and polymer solutions were premixed in the following
ratio: 5 L of sample: 5 L of matrix: 0.5 L of salt. Approximately 0.5 L of the obtained
mixture was hand spotted on the target plate. For each spectrum, 1000 laser shots were
accumulated.

III) Analysis of thiocarbonyl thio-containing polymers by MALDITOF-MS


PBMA (Mn = 2.500 g/mol, PDI = 1.26) was analyzed using the experimental procedure
described above. The influence of the laser intensity was investigated, acquiring spectra with
intensities of 1950, 2100 and 2300 Lux.
The RAFT agent used for the synthesis was (4-cyano-1-hydroxypent-4-yl) dithiobenzoate
(RAFT-ACP) (purity of 97% determined by 1H-NMR in CDCl3) and the initiator was 4,4azobis(4-cyanopentanol) (ACP) with a purity superior to 99% (1H-NMR in CDCl3). The
polymerization was carried out in toluene at 70C and the final product was isolated by rotary
evaporation of the solvent and the unreacted monomer.

OH
CN

HO

N N
CN

(4-cyano-1-hydroxypent-4-yl) dithiobenzoate
RAFT-ACP

OH
CN

4,4-Azobis(4-cyanopentanol)
ACP

72

MALDI-TOF-MS analysis of RAFT polymers

Voyager Spec #1[BP = 2423.2, 64]

2423.1970

100

64

2707.5936
2565.4097

90
80

1710.2227
1995.6326

70

3417.5312
3133.1492
3560.7662

% Intensity

1569.0460
60
50
40
30
20
10

1558.1203
2268.1144
2980.0560
3702.9235
2125.9027
2560.2762
3270.9129
823.1484
4271.7502
1414.9229
2695.7069
751.9689
1983.7309
3406.6591
4265.6484
4965.5708
2843.3222
656.0670
3696.4157
1841.5657
4412.8402
801.1282
1271.7540
3138.2636
2295.9647
3855.1865
4592.1421
5233.1306
664.7383
1177.4519
2871.7995
2142.9179
3423.7266
4028.5272
4666.1270
5486.6006
880.4635
1471.1112
2817.0926
2301.0723
3340.9071
4034.1308
4822.4203
5507.2971
674.8573
1170.4264
1685.9418
2830.3766
3680.0091
4204.4732
4876.9659
2212.9217
5454.8694
1350.9061
2782.7518
1141.4652

0
609.0

1707.4

2805.8

3904.2

0
6101.0

5002.6

Mass (m/z)

B1

Voyager Spec #1[BP = 2423.2, 64]

100
90

A1

70

A1

B1 2281.0024
C1

80

2423.1970 1

2280.0118

64

2422.2086

2409.2851

% Intensity

2278.9991
60

2424.2364

2268.1144

50

2410.3006
2411.2970

40
2267.1308

30

2275.0270
2310.9808
2236.7972
2274.5880
2205.9354
2241.0672
2276.5727
2309.7541
2204.0262
2257.2180
2288.1599
2315.9166
2215.4963
2258.2832
2305.4263

20
10
0
2193.0

2252.2

2417.1913
2426.2525
2386.5513
2394.1232
2381.4948
2384.5376

2352.6472
2348.3605
2341.0117

2311.4

2370.6

2439.2277
2430.6182
2461.5158
2421.6378
2457.6620
2420.3192
2448.9501
0
2489.0

2429.8

Mass (m/z)

Figure 5.1. MALDI spectra of PBMA RAFT polymer with Laser Intensity = 1950 Lux
Voyager Spec #1[BP = 655.6, 762]

2408.6238
2551.7106

100
90

2124.4516
2136.3077
1851.1275

80

% Intensity

1566.9518
1424.8460

50

20
10

2841.7799

3410.7922

4121.3522

3830.5214

1282.7464

40
30

2693.8207

3552.9742
2835.9175
2563.5916
3694.7244
3131.9593
2558.6602
3405.1979
3547.3303
2273.8630

70
60

564.0

1989.7402

3842.4764

4406.0614
4548.0187

4690.7893
4126.6538
2425.6115
1846.8712
3004.8400
4832.2468
1705.0853
2284.5100
3846.7316
2852.9151
4303.3715
4837.0588
2039.3919
1298.7091
2857.4012
3581.6868
4095.7210
4699.2350
2109.3880
1088.5244
1639.1373
2671.7338
3226.4493
3820.4002
4333.9929
1117.1203
1893.2532
2448.0141
1248.8904

1036.7138

0
1035.0

2028.6

3022.2

4015.8

5009.4

5260.9570
5326.9286
0
6003.0

Mass (m/z)

73

Chapter 5

Voyager Spec #1[BP = 655.6, 762]

2408.6238

100
90

2267.5349

80

B2

70

B2 2420.5002

2278.3967

A2

2265.5355
2268.5560

A2

60

% Intensity

564.0

2266.5291

2273.8630

50

2419.4801
2407.6231

C2

C2

2416.2352
2422.5052
2417.6869

2272.8476
40

2280.4142

30
20

2275.9834
2271.8271
2324.5755
2296.4215
2251.4823
2334.3102
2288.2618
2248.0058
2335.8783
2308.0850

2209.4955
2211.5340

10
0
2193.0

2252.8

2423.5060
2414.0287
2411.1850
2428.5725
2440.1045
2404.1754

2363.4639
2367.5589

2312.6

2372.4

2461.5393
0
2492.0

2432.2

Mass (m/z)

Figure 5.2. MALDI spectra of PBMA RAFT polymer with Laser Intensity = 2100 Lux
Voyager Spec #1[BP = 2124.4, 2468]

2124.3861

100
90

1698.1298

80

2835.8270

1555.0391
1413.9500

70

3119.9821
2841.6675

60

% Intensity

2468.0
2550.6469
2693.7392

986.6792

2699.6007

3268.6154
3694.8862
3546.2775
3979.5186

50
2273.5968
40
30
20

895.4131
900.6172
906.8252

10

1989.5840

3832.3479

2136.2708
2425.5515
1897.2506
2436.4596
1936.1144
2387.4460

1709.9998
1282.7132
1440.8240
1475.4604

0
853.0

1767.2

2846.7628
2993.9339
3000.3587
2899.6739

3415.0685
3564.3138
3518.3700

2681.4

Voyager Spec #1[BP = 2124.4, 2468]

4262.9652

4116.6408
3971.4804
3991.3680
3957.4466

3595.6

4546.9804
4542.8283

4422.7869

4976.0427
4860.4160
0
5424.0

4509.8

Mass (m/z)

2266.4674

100
90

A3

A3

2267.4773

2409.5687

80
70
% Intensity

2265.4754

2410.5603

B3

2468.0

2408.5628

B3

60
2415.6256

50
2273.5968
2274.5804

40
30
20
10
0
2154.0

2416.5931

C3

2270.5064
2271.7413
2276.7117
2286.4502
2237.4612
2255.4847

2182.4138
2179.5015
2222.6

C3

2412.5381

2324.5407
2318.5689

2291.2

2380.5213

2359.8

2413.6792
2419.4738
2422.5130
2431.6696

2464.6522

2428.4

0
2497.0

Mass (m/z)

Figure 5.3. MALDI spectra of PBMA RAFT polymer with Laser Intensity = 2300 Lux
Figure 5.1, Figure 5.2 and Figure 5.3 represent the spectra acquired at various laser intensities
and enlargements in the 2200 2400 Da region. It is obvious, from the enlargements of the
spectra that the laser intensity strongly influences the observed peak patterns. In the three

74

MALDI-TOF-MS analysis of RAFT polymers

spectra, three patterns can be observed that are repeated every 142.2 mass units (this value
corresponds to the mass of a single butyl methacrylate unit). Although the spectrum of Figure
5.1 suffers from a general poor signal-to-noise ratio, due to the low laser intensity used for the
measurement, a striking difference in resolution can still be noticed. While the peaks with an A
or C label have a relatively good resolution, the B peaks in the three spectra exhibit very poor
resolution, with non-resolved isotopic distributions.
Assigning the A and C peaks leads to the following structures:

CH3

CH2
C4H9

A=

CH

CH2

O
O

OH

, Na+

OH

, Na+

CN

C4H9

CH3
S

C=

CH2
O

C
O

C4H9

CN
n

The assignment for the C cluster corresponds to the expected structure after a RAFT
polymerization. The thiocarbonyl thio moiety is attached to the polymer chain. The structure
related to the A cluster can be related to several phenomena that may have happened to the
polymer chain prior to or during the analysis:
-

Termination by disproportionation is the preferred termination mode for methacrylic


monomers. Simulation performed on these systems indicates that roughly 10% of the
polymer chains present at the end of the polymerization are the result of termination
reactions. It is then logical that we observe the peaks of the A family.

Fragmentation during the MALDI-TOF-MS analysis is known to happen for certain


RAFT agents/polymers8,10. The cleavage of the weak carbon-sulfur bond of the
thiocarbonyl thio compound may result in chains terminated by a double bond following
the reaction scheme displayed in Figure 5.4.

75

Chapter 5

CH3
HO

CH3

CH2
CN

CH2

C4H9

C4H9

CH2
CN

C4H9

CH2

C4H9
H

CH2

CH2
CN

CH2

CH3
HO

CH2

CH3
HO

C4H9
n

C4H9

Figure 5.4. Fragmentation of the thiocarbonyl thio-containing polymer during MALDI-TOFMS experiments
Inspection of the isotopic distribution reveals a noticeable deviation from the theoretical
isotopic distribution (Figure 5.5).

% Intensity

Voyager Spec #1[BP = 2124.4, 2468]

100
90
80
70
60
50
40
30
20
10
0
2264.77807

2266.4674

2468.0

2267.4773

2265.4754

2268.4743

(a)

2269.4725
2270.5064

2266.08587

2267.39366

2268.70145

0
2271.31704

2270.00925

Mass (m/z)

% Intensity

ISO:Na(CH2CCH3COOC4H9)15CCH3CNC3H6O

100
90
80
70
60
50
40
30
20
10
0
2264.77807

2266.5517
2265.5483

100
2267.5550

(b)

2268.5581
2269.5612
2266.08587

2267.39366

2268.70145

2270.00925

2270.5642
0
2271.31704

Mass (m/z)

Figure 5.5. MALDI-TOF-MS spectrum of PBMA RAFT polymer with laser intensity of 2300
Lux (a) and the simulated spectrum corresponding to a polymer A with n=15 (b)

76

MALDI-TOF-MS analysis of RAFT polymers

Despite a small shift of the baseline, it seems that the peaks at positions 2267.5 0.1 Da and
2268.5 0.1 Da are more intense in the acquired spectrum than in the theoretically calculated
spectrum. The reason is probably the presence underneath those peaks of another family of
peaks (higher by 2 Da) corresponding to the hydrogen-terminated chains with the following
formula:

CH3
D=

HO

CH2
CN

H
O

C4H9
n

These hydrogen-terminated chains can have several origins:


-

Termination by disproportionation will result, as discussed above, not only in double


bond-terminated chains, but also in hydrogen-terminated chains.

Transfer reaction during the radical polymerization may occur. Toluene is known to
have a relatively low transfer constant for methacrylates (e.g. Ctr = 5.10-5 for methyl
methacrylate at 70C11, the transfer constant for butyl methacrylate is expected to be in
the same order of magnitude). In the eventuality of transfer to toluene, some polymer
chains should be found with a benzyl end-group. In the MALDI-TOF-MS spectra, those
chains should be displayed around 2247.0 Da. Evidence for the presences of those peaks
could not be found. One may wonder whether the benzyl radicals are suitable for butyl
methacrylate initiation. According to Fischer and Radom12, the addition of benzyl
radicals to methyl methacrylate is faster than the one of 2-cyano-2-propyl (the AIBNderived radical). Assuming a minor difference in reactivity between the methyl
methacrylate and the butyl methacrylate, we can conclude that the absence of benzylterminated polymer chains in the spectrum suggests that the hypothesis of transfer to
solvent can be discarded. Branched polymer chains, resulting from transfer to polymer,
can not be detected directly by MALDI-TOF-MS. However, these transfer reactions
should not be occurring during the polymerizations13 (Ctr = 2.48 10-4), considering the
relatively low molecular weight of the final polymer. Transfer to monomer can also be
rejected as a cause of the formation of H-terminated chains, as the transfer constant to
methyl methacrylate is Ctr = 3.10-5 11.

Fragmentation during the MALDI-TOF-MS analysis produces chains with a terminal


radical. Two of those radicals may terminate by disproportionation, in a similar fashion
to what occurs in radical polymerization. However, the proximity of a thiocarbonyl thio

77

Chapter 5

radical after the carbon-sulfur bond cleavage will ensure a predominant occurrence of
the double-bond forming mechanism shown in Figure 5.4.
The termination by disproportionation of growing chains during the polymerization may be an
acceptable explanation for the combined presences of double-bond-terminated chains and
hydrogen-terminated chains. However, making the reasonable assumption that the ionization
efficiencies for both types of chains are similar, and assuming a 1:1 ratio between H-terminated
and double bond-terminated chains, the isotopic distribution in the 2264-2271 Da region should
be as shown in Figure 5.6 (b), which is clearly not the case.

% Intensity

Voyager Spec #1[BP = 2124.4, 2468]

100
90
80
70
60
50
40
30
20
10
0
2265.00381

2266.4674

2468.0

2267.4773

2265.4754

2268.4743
2269.4725
2270.5064

2266.25206

2267.50031

2268.74856

(a)
0
2271.24506

2269.99681

Mass (m/z)

% Intensity

ADD (ISO:Na(CH2CCH3COOC4H9)15CCH3CNC3H6O , ISO:NaH2(CH2CCH3COOC4H9)15CCH3CNC3H6O)

100
90
80
70
60
50
40
30
20
10
0
2265.00381

2267.5593

290.0

2268.5647

2266.5517

2269.5689

(b)

2265.5483
2270.5726

2266.25206

2267.50031

2268.74856

2269.99681

0
2271.24506

Mass (m/z)

Figure 5.6 MALDI-TOF-MS spectrum of PBMA RAFT polymer with laser intensity of 2300
Lux (a) and the simulated spectra corresponding to a 1:1 ratio of polymer A with n=15 and
polymer D with n=15 (b)
This evidence shows that the disproportionation-termination mechanism, either during the
polymerization or during the analysis, is not the main explanation for the detection of polymer
chains with a terminal double bond. The fragmentation of the weak carbon-sulfur bond of the
thiocarbonyl thio-containing chains and the hydrogen abstraction by the thiocarbonyl thio
radical formed (Figure 5.4) must be the main cause. The spectrum displayed does not match
with the hypothetical spectrum expected from the structure predicted by the RAFT
polymerization theory. The fragmentation during the analysis is responsible for the appearance
of the extra peaks corresponding to polymer chains of general structure A.
The main reason that can be proposed for the high tendency of thiocarbonyl thio-containing
polymers to undergo fragmentation during the MALDI-TOF-MS analysis is the absorbance
band around 300 nm that the PBMA polymer exhibits. The laser used in the MALDI-TOF-MS

78

MALDI-TOF-MS analysis of RAFT polymers

to irradiate the matrix in which the polymer is dispersed, has a wavelength of 337 nm, close to
this absorbance band. It is easy to imagine that the thiocarbonyl thio end-group can absorb this
wavelength and reach a higher energy level. The compound can then lower its energy by
fragmentation, yielding a relatively stable tertiary carbon-centered radical and a thiocarbonyl
thio radical, which is heavily stabilized through resonance with the carbon-sulfur double bond
and the aromatic ring.
Another explanation for the fragmentation is the local heating that the polymer sample
undergoes at the moment of the irradiation. It is known that the temperature can rise up to more
than 500C for a short period of time (in the order of a nanosecond) at the location of the laser
impact14,15. This temperature can be sufficient to induce a thermal decomposition in the case of
RAFT polymers, by breaking up the weak carbon-sulfur bond.

IV) Influence of the laser intensity on the fragmentation


Despite the impossibility to draw absolute quantitative conclusions from mass spectrometric
analysis, it is legitimate to compare the analysis of the same sample recorded with various laser
intensities. In Figures 5.1, 5.2 and 5.3, it is obvious that the relative intensity of the cluster A
increases with an increase of the laser intensity and, that the intensity of the cluster C follows an
opposite trend, decreasing with increasing intensity. The hypothesis of the laser-induced
fragmentation is thus supported by the variation of the peak patterns with intensity.
The bad resolution of the B cluster in the three spectra presented is rather unusual and seems to
be the result of a fragmentation during the flight of the polymeric ions. The measured masses
cannot be ascribed to any of the possible structures expected from the synthesis. The better
resolution of the cluster C seems to point toward a fragmentation during the ionization. An
explanation can be based on the hypothesis of the excitation of the carbon-sulfur bond to a
higher energy level. It seems that some chains relax by fragmenting during the ionization (peaks
C), while some other ones fragment during the flight (peaks B). The remaining chains are
unaffected by any fragmentation steps and are detected with the thiocarbonyl thio moiety still
attached to one of their chain ends (peaks A).
Poly(butyl acrylate) (Mn = 1.800 g/mol, MWD = 1.29) was analyzed following the
experimental protocol described above. The mode of synthesis was a RAFT polymerization with
2-cyanoprop-2-yl dithiobenzoate as RAFT agent and AIBN as initiator. The polymerization was
performed in toluene at 80C and the final product was isolated by evaporation of the solvent

79

Chapter 5

and the unreacted monomer. The spectrum acquired with a laser intensity of 1400 Lux is
presented in Figure 5.7
Voyager Spec #1[BP = 2295.3, 272]

2295.3378

100

272.0

2167.2151

90

2039.0867

2423.4625

80
70

2551.6004

1782.8271

% Intensity

60
2679.7278

50
1653.7040

40

2808.8565
2936.9762

1525.5779

30

1397.4435
1269.3055
1141.1709

20
10

1129.3114

1535

2682.7221

2170.2539

1913.9565
1886.0568

1629.7985

1386.5763

0
1005

2426.4742
1656.6941

2439.4292

2183.2293

2065

2687.7446

2595

3065.1197
2939.9059

3193.1741
3321.3125
3199.1952
0
3655

3125

Mass (m/z)

Figure 5.7. MALDI spectrum of PBA RAFT polymer with Laser Intensity = 1400 Lux
The main population can clearly be assigned to the structure expected after a RAFT
polymerization:
S

E=

CH2 CH
O

C4H9

, Na+

CN
n

% Intensity

Voyager Spec #1[BP = 2295.3, 272]

100
90
80
70
60
50
40
30
20
1888.0525
10
0
1887.0

1910.9545

215.0

1909.9603
1911.9652
1912.9689
1913.9565

1898.0548
1899.2

1919.9893

1911.4

1923.6

(a)

1926.9846
1935.8

0
1948.0

Mass (m/z)

% Intensity

ISO:CCH3CH3CN(C7H12O2)13NaCS2C6H5

100
90
80
70
60
50
40
30
20
10
0
1887.0

1910.1143

100

1909.1110
1911.1164
1912.1180

(b)

1913.1196
1899.2

1911.4

1923.6

1935.8

0
1948.0

Mass (m/z)

Figure 5.8. MALDI-TOF-MS spectrum of PBA RAFT polymer with laser intensity of 1400 Lux
(a) and the simulated spectrum corresponding to polymers E with n =13 (b)

80

MALDI-TOF-MS analysis of RAFT polymers

A small difference in the peak positions between the theoretical and experimental spectra (less
than 1 mass unit) can be noticed in Figure 5.8. It was assumed that the calibration of the
spectrometer was the main reason for this difference and no further attention was paid to it.
Similarly to the previous series of experiments with thiocarbonyl thio-containing poly(butyl
methacrylate), the laser intensity was increased. Figure 5.9 represents a spectrum of the same
poly(butyl acrylate) recorded with a laser intensity of 2400 Lux.
Voyager Spec #1[BP = 1001.0, 12663]

1129.1329

100

1.0E+4

90
80

1257.2584

70
% Intensity

60

1385.3829

50
40

1141.0568

30

1132.1405
1125.0842
1058.9389

20

1513.4978
1397.3080
1653.5720
1388.3836
1413.3500

10
0
1034

1910.8437

1770.7558
1898.8970
1644.6053
1894.8233
1626.4803

1529

2167.1240
2423.4084
2148.6512
2143.1502

2404.6710

2024

2519

2680.6366
2936.9950
3014

0
3509

Mass (m/z)

Figure 5.9. MALDI spectrum of PBA RAFT polymer with Laser Intensity = 2400 Lux

% Intensity

Voyager Spec #1[BP = 1001.0, 12663]

100
90
80
70
60
50
40
30
20
10
0
1735

1782.7124

2964.1

1783.7091
1770.7558
1771.7638
1766.7004
1755.7363

1757

1784.7126
1785.6851

1779

1798.6449

1801

(a)

1813.6874

1823

0
1845

Mass (m/z)

% Intensity

ISO:CCH3CNCH3(C7H12O2)12CS2C6H5Na

100
90
80
70
60
50
40
30
20
10
0
1735

1782.0306

100

1783.0325

(b)

1784.0340
1785.0355
1757

1779

1801

1823

0
1845

Mass (m/z)

Figure 5.10. MALDI-TOF-MS spectrum of PBA RAFT polymer with laser intensity of 2400
Lux (a) and the simulated spectrum corresponding to polymers E with n =12 (b)
Figure 5.10 shows evidence that some thiocarbonyl thio end-functionalized poly(butyl
acrylate) chains are still detected, despite the high laser intensity used. The other peaks recorded
could not be assigned to any structures expected from a RAFT polymerization. Their absence in
the spectrum recorded at low laser intensity indicates that they are probably the result of

81

Chapter 5

complex fragmentation/rearrangement reactions. The main conclusion of this experiment is that


the nature of the monomer greatly influences the overall stability of the chain. The thiocarbonyl
thio-containing poly(butyl methacrylate) chains could no longer be detected at high laser
intensity, while their equivalents in a poly(butyl acrylate) sample could be observed. The radical
stability of the carbon-centered poly(butyl acrylyl) radical (secondary radical) is lower than that
of the poly(butyl methacrylyl) radical (tertiary radical). It is then logical that less fragmentation
occurs after the laser irradiation in the acrylate case compared to the methacrylate case.

V) Analysis of trithiocarbonate-containing poly(butyl acrylate) by


MALDI-TOF-MS
As mentioned in Chapter 4, the polymerization of methacrylic monomers is troublesome with
S,S-Bis(,-dimethyl--acetic acid)trithiocarbonate, which was the preferred RAFT agent
employed in this project, due to the stabilities of the propagating radical and the leaving group
radical. The study on trithiocarbonate-containing polymers was therefore limited to poly (butyl
acrylate).
Poly(butyl acrylate) (Mn = 2.100 g/mol, MWD = 1.12) was analysed by MALDI-TOF-MS.
The

RAFT

agent

used

for

the

synthesis

was

S,S-Bis(,-dimethyl--acetic

acid)trithiocarbonate (purity superior to 99% determined by 1H-NMR in CDCl3), and the


initiator used was azobis(isobutyronitrile) (recrystallized from methanol). The polymerization
was performed in an acetone/toluene 1:1 mixture at 80C and the final product was isolated by
rotary evaporation of the solvent and the unreacted monomer. The spectrum acquired with a
laser intensity of 2200 Lux is presented in Figure 5.11
Voyager Spec #1[BP = 2400.6, 18377]

2400.5872

100

1.8E+4

2528.7406

90
2144.2828

80

2656.8945

% Intensity

70

50

2913.2013
2378.6419
2484.8373
2228.5150
3042.3545
2531.7365
2100.3575
2275.4292
2869.3050
3170.5044
1972.2015
2916.2036
1866.0154
3298.6447
2509.7984
2766.1167
3148.5638
2253.4794
3426.7948
1890.9837
2662.9136
3404.8327
2919.1576
2407.6455
2138.2850
1887.9856

40
30
20
10

2785.0491

2016.1332

60

1318.5452

0
1307.0

1630.7084
1608.7380

1841.4

2375.8

2910.2

3444.6

3683.0863
0
3979.0

Mass (m/z)

82

MALDI-TOF-MS analysis of RAFT polymers

Voyager Spec #1[BP = 2400.6, 18377]

2400.5872

100

1.8E+4
2528.7406

90
2401.5854

80

2529.7421

2399.5867

70
% Intensity

2527.7399
60

2402.5812

2530.7364

50
2378.6419

40

2506.8011
2507.8017

2379.6406
2377.6367

30

2403.5744

2505.7988

2380.6374
20
10

2404.5738
2405.5909

2359.6740
2360.6698

2324.6014

0
2311

2531.7365

2486.8326

2364

2452.8054

2417

2532.7339

2487.8352
2488.8245

2470

2533.7157
2522.7562
0
2576

2523

Mass (m/z)

Figure 5.11. MALDI spectra of PBA RAFT polymer with Laser Intensity = 2200 Lux
The assignments of the peaks have already been discussed in Chapter 4 and the conclusion
was that the three families of peaks corresponded to one original compound, with the following
structure:

F=

HOOC

CH2
O

CH
C
O

S
C4H9

S
S

CH2
O

CH
C
O

COOH
C4H9
m

% Intensity

Voyager Spec #1[BP = 2400.6, 18377]

100
90
80
70
60
50
40
30
20
10
0
2477.0

2528.7406

1.6E+4

2527.7399
2484.8373
2483.8358

2531.7365

2505.7988

2532.7339

2509.7984
2511.7926

2487.8352
2490.8

(a)

2506.8011

2504.6

2522.7562
2518.4

2532.2

0
2546.0

Mass (m/z)

% Intensity

ADD (ISO:Na2OOCNaCCH3CH3(C7H12O2)17CS3COOHCCH3CH3 , ISO:HOOCNaCCH3CH3(C7H12O2)17CS3COOHCCH3CH3) , ISO:NaOOCNaCCH3CH

100
90
80
70
60
50
40
30
20
10
0
2477.0

2483.4214
2484.4235
2482.4181

2505.4033
2506.4054

2528.3931
2529.3952

2504.4001

2486.4267
2487.4281
2489.4309
2490.8

199.9

2527.3899

2508.4086
2509.4100
2511.4128
2504.6

(b)

2531.3984

2518.4

2532.3998
2534.4025
2532.2

0
2546.0

Mass (m/z)

Figure 5.12. MALDI-TOF-MS spectrum of PBA RAFT polymer with laser intensity of 2200
Lux (a) and the simulated spectrum corresponding to polymers F with the substitutions of 0, 1
and 2 hydrogen atoms from the carboxylic groups by sodium atoms with n + m =17 (b)

83

Chapter 5

A difference of 1.4 mass unit is present between the theoretical spectra and the experimental
one in Figure 5.12. This difference was assumed to be caused by a calibration problem and no
further attention was paid to it. MALDI-TOF-MS analysis of the trithiocarbonate-containing
poly(butyl acrylate) showed no evidence of fragmentation. The laser intensity was raised to
2500 Lux, without any fragmentation patterns observed.
The explanation for this low tendency towards fragmentation is again based on the radical
stabilities of the compounds that might be created after fragmentation. Breakage of the carbonsulfur bond between the trithiocarbonate and the polymer would produce the following radical
species:

S
HOOC

(PBA)

CH CH2
O

(PBA)

COOH

O
C4H9

A comparison of the stabilities of the radicals produced upon fragmentation of


trithiocarbonate-containing poly(butyl acrylate) with the ones of thiocarbonyl thio-containing
poly(methyl methacrylate) leads to two major observations:
-

The sulfur-centered radical is stabilized much more in the case of thiocarbonyl thiocontaining polymer, due to the delocalization of the unpaired electron on the aromatic
ring. Resonance delocalization occurs in the trithiocarbonate radical as well, but it is
limited to the carbon-sulfur double bond.

The carbon-centered radical is more stable in the case of the polymethacrylate (tertiary
carbon) than in the case of polyacrylate (secondary carbon)

These two observations explain that the trithiocarbonate-containing poly(butyl acrylate) is less
prone to fragmentation than the thiocarbonyl thio-containing poly(butyl methacrylate) and
poly(butyl acrylate). In terms of energetic levels, the polymer molecules will reach a higher
energy level after irradiation by the laser, but contrary to the polymethacrylates, they will not
relax via a fragmentation reaction, but will return to their initial state.

84

MALDI-TOF-MS analysis of RAFT polymers

V) Conclusions
Thiocarbonyl

thio-containing

poly(butyl

methacrylate),

poly(butyl

acrylate)

and

trithiocarbonate-containing poly(butyl acrylate) were analyzed by MALDI-TOF-MS. The


thiocarbonyl thio-containing poly(butyl methacrylate) showed evidence of chain fragmentation
during the analysis, with almost complete disappearance of the peaks corresponding to the
chains bearing a thiocarbonyl thio end-group at high laser intensities. A mechanism relying on
abstraction of a hydrogen from the carbon-centered poly(methacrylyl) radical by the
thiocarbonyl thio radical was proposed. This mechanism was supported by isotope-distribution
considerations. Thiocarbonyl thio-containing poly(butyl acrylate) chains showed less
fragmentation. Even at high laser intensities, some intact thiocarbonyl thio-containing chains
were observed.
The trithiocarbonate-containing poly(butyl acrylate) proved to be much more stable than the
two other polymers. No evidence was found for fragmentation even at high laser intensities. The
lower stability of the poly(acrylyl) radical versus the poly(methacrylyl) radical and the lower
stability of trithiocarbonate radical in comparison with the thiocarbonyl thio radical are the main
causes for the greater stability of the trithiocarbonate-containing poly(butyl acrylate).

85

Chapter 5

Reference List
1. Scrivens, J. H., Jackson, A. T.; Int. J. Mass. Spectrom. 2000, 200, 261.
2. Hanton, S. D.; Chem. Rev. 2001, 101, 527.
3. McEwen, C. N., Peacock, P. M.; Anal. Chem. 2002, 74, 2743.
4. Vosloo, J. J., Wet-Roos, D., Tonge, M. P., Sanderson, R. D.; Macromolecules 2002, 35,
4894.
5. Destarac, M., Charmot, D., Franck, X., Zard, S. Z.; Macromol. Rapid Com. 2000, 21,
1035.
6. D'Agosto, F., Hughes, R., Charreyre, M. T., Pichot, C., Gilbert, R. G.; Macromolecules
2003, 36, 621.
7. Ganachaud, F., Monteiro, M. J., Gilbert, R. G., Dourges, M. A., Thang S.H., Rizzardo, E.;
Macromolecules 2000, 33, 6738.
8. Schilli, C., Lanzendoerfer, M. G., Mueller, A. H. E.; Macromolecules 2002, 35, 6819.
9. Vana, P., Albertin, L., Barner, L., Davis, T. P., Barner-Kowollik, C.; J. Polym. Sci. Pol.
Chem. 2002, 40, 4032.
10. Favier, A., Ladaviere, C., Charreyre, M. T., Pichot, C.; Macromolecules 2004, 37, 2026.
11. Gopalan, M. R., Santhappa, M.; J. Polym. Sci. 1957, 25, 333.
12. Fischer, H., Radom, L.; Angew. Chem. Int. Edit. 2001, 40, 1340.
13. Morton, M., Piirma, I.; J. Am. Chem. Soc. 1958, 80, 5596.
14. Koubenakis, A., Frankevich, V., Zhang, J., Zenobi, R.; J. Phys. Chem. A 2004, 108, 2405.
15. Sadeghi, M., Wu, X., Vertes, A.; J. Phys. Chem. B 2001, 105, 2578.

86

Networks applications: synthesis and properties

CHAPTER 6
NETWORKS APPLICATIONS: SYNTHESIS AND
PROPERTIES

I) Introduction
To study the influence of molecular weight distribution and functionality-type distribution of
starting polymers on the network formation, network structure and material properties in cured
thermoset resins, a comparison between end-functionalized telechelic polymers with a narrow
molecular weight distribution and random co-polymers with about the same average
functionality by number and molecular weight by number are needed. The carboxylic acidfunctional poly(butyl acrylate) polymers, whose synthesis and characterizations were described
in chapter 4 are used as telechelic starting compounds. In addition to those RAFT polymers,
several random copolymers of butyl acrylate and acrylic acid were synthesized. The numberaverage molecular weight and average functionality per polymer chain of those random
copolymers were nearly identical to the telechelic ones. A comparison between the two kinds of
polymers will allow us to have a clear view on the influences of the distributions of both
parameters on the network structure and on the mechanical properties of the cured coatings
made from them.
This chapter will first focus on the mechanism of cure and the rate of cure of these polymers
using triglycidyl isocyanurate (TGIC) as a cross-linker and 1,4-diazabicyclo[2.2.2]octane
(DABCO) as a catalyst. Attenuated Total Reflection Fourier-Transform InfraRed (ATR-FT-IR)
spectroscopy will be used to study these aspects for both types of polymers (random copolymers
and telechelic ones). The elastic moduli of the final cured coatings will then be determined by
micro-indentation. A particular attention will be given to the influence of the dangling chains on
these moduli. Information about the final network structures will be provided by analysis of
these cure materials with solid-state nuclear magnetic resonance spectroscopy (NMR) and
temperature modulated differential scanning calorimetry (TMDSC).

87

Chapter 6

II) Experimental
II.1) The Attenuated Total Reflection Fourier-Transform InfraRed spectroscopy
Attenuated Total Reflection Fourier-Transform InfraRed spectroscopy technique relies on the
fact that reflection occurs when a beam radiation passes from a denser medium to a less dense
medium. Beyond a certain critical angle of incidence the reflection is almost complete. It has
been proven both theoretically and experimentally, that in fact the beam penetrates into the less
dense medium. The depth of penetration can vary from a fraction of a wavelength to several
wavelengths, depending of the wavelength used, the incidence angle, and the difference of the
density of the two media. The penetrating radiation eventually interacts with the less dense
medium, attenuating the beam at wavelengths of absorbance bands. The beam radiation after
interaction with the media is recorded, and by comparison with the original radiation, an
absorption spectrum is obtained.

Polymer
Heating
plate

Diamond
crystal
To detector

From radiation
source

Figure 6.1. Schematic representation of the trajectory an infrared beam in an ATR crystal
coated with a polymer
In the setup used here, a single reflection of an IR beam was used to determine the change in
the absorption during cure (Figure 6.1).
One difference between ATR-IR and classical absorption infrared spectroscopy is that the
spectra acquired are not thickness-dependant, as the radiation penetrates only a few micrometers
into the sample. The peaks observed are similar to the ones observed in ordinary absorption
infrared. The main advantage of ATR-IR is that absorption spectra are readily available for a
variety of samples with a minimum of preparation. However, homogenous materials are needed.

88

Networks applications: synthesis and properties

The curing reactions were followed by attenuated total reflection (ATR)-FTIR on a Bio-Rad
Excalibur FTS3000MX infrared spectrophotometer using the golden gate setup (20 scans per
spectrum with a resolution of 4 cm1) equipped with an ATR diamond unit.

II.2) Transmission Fourier-Transform InfraRed spectroscopy


A thin film of the reaction mixture was applied on a zinc selenide (ZnSe) window.
During the curing reaction IR spectra were recorded using a Biorad UMA 500 infrared
microscope that was coupled to a Biorad FTS 6000 FTIR spectrometer. The microscope was
equipped with a broad band MCT detector. This window was heated to the reaction temperature
by means of a Linkam THMS 600 hot stage. The heating rate was 30 C/min.
The kinetic mode of the Biorad WinIR pro software was used to record the spectra during the
reaction. The reaction time was set to 30 min., the time resolution was 20 sec. Spectra were
recorded with a resolution of 4 cm-1

II.3) Temperature Modulated Differential Scanning Calorimetry (TMDSC)


TMDSC experiments were carried out employing a modified Perkin-Elmer DSC-7 apparatus
calibrated for cell constant and temperature using indium standards. The instrument was
modified using a commercial precision function generator to allow a sinusoidal temperature
modulation. The measurements were carried out under nitrogen atmosphere to prevent
degradation of the polymer samples upon heating. The masses of the pans on the reference side
and the sample side were balanced to give a zero signal for the baseline. The experiments were
carried out with a global temperature ramp of 2 K/min with an oscillatory frequency of 12.5
mHz and a temperature amplitude of 0.06 K.

II.4) Micro-indentation
Indentation experiments were carried out at room temperature and ambient atmosphere using a
home built apparatus1. In general, in one run, 25 indentations were made using a maximum
loading of 1, 2 or 4 mN with a penetration rate of 10 nm/s. The calibration procedure of Oliver
and Pharr2 was used to correct for the load frame compliance of the apparatus and the imperfect
shape of the indenter tip. The compliance of the system was determined to be 0.3 nm/mN and

89

Chapter 6

from the projected area of the indenter A, the contact depth hc was calculated using the relation
[6.1] with a=24.5 and b=5.71 m.
A=ahc2+bhc

[6.1]

The tip radius was 0.9 m. The area function of this indenter was calibrated using B270 glass
(Schott, Jena, Germany), with an elastic modulus that was determined independently to be
751 GPa using the pulse-echo method. The calibration was performed using B270 glass for a
depth range of 0.12.9 m, where the maximum indentation depth was restricted by the load
limitation. A typical load-displacement curve for a viscoelastic materialobtained after a microindentation experiment is presented in Figure 6.2.
The effective moduli (Eeff) of our coatings at different loadings were calculated using relation
[6.2].

E eff =

1 S
2
A

[6.2]

where is a constant, related to the geometry of the indenter ( = 1.034 for a Berkovich
indenter, used here) and S the unloading stiffness.
P
Pmax

P
hmax

a b c d

hc

Loading

Unloading

hres

hc hmax h

Figure 6.2. Schematic representation of load P versus indenter displacement h. (a) Initial
surface; (b) surface profile after load removal; (c) indenter; (d) surface profile under load.

90

Networks applications: synthesis and properties

In the case of a micro-indentation experiment performed on a coating, the obtained E value


can represent not only the intrinsic properties of the top layer, but a combination of the
mechanical properties of both the coating and the substrate, especially if the penetration depth is
high compared to the coating thickness, and if a large difference exists between the elastic
modulus of the coating and the one of the substrate. It has been shown that even for rubbery
materials, the influence of the substrate on the Eeff can be neglected if the penetration depth of
the indenter is less than 1/10 of the total thickness of the coating3. The elastic modulus and the
hardness values can be typically represented as in Figure 6.3.

E, H

EEs,effHs,s H s

Eapp
Happ

Eeffc,EHf, cHf

Happ
s
s
EEs,effH,sH

Eapp
hc/hf

Figure 6.3. Schematic representation of the apparent elastic modulus Eapp and hardness Happ
versus relative indenter displacement hc/hf (Eeffc, H c, hf = effective elastic modulus and hardness
of the substrate; Eeffs, H s)
As can be seen in Figure 6.4, The Eeff measured within the accuracy of measurement is at least
not influenced by the substrate for our telechelic and random copolymer-based networks when
the total penetration depth (hc) is less than 10% of the total thickness of the coating (hf).
The elastic modulus of the cross-linked coating itself was determined using relation [6.3],
assuming that the telechelic-based networks and the random copolymer-based networks behave
like a perfectly elastic material ( = 0.5).

1
1 - 2 1 - i
=
+
E eff
E
Ei

[6.3]

where E and are Youngs modulus and Poissons ratio for the sample analyzed, Ei and i
Youngs modulus and Poissons ratio for the indenter.

91

Chapter 6

Load = 1 mN
Load = 2 mN
Load = 4 mN

0
0.00

0.05

0.10

0.15

0.20

0.25

0.30

hc / hf

0.35

0.40

0.45

0.50

Eeff (MPa)

Eeff (MPa)

(a)

(b)

Load = 1 mN
Load = 2 mN
Load = 4 mN

0
0.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

0.40

0.45

0.50

ht / hf

Figure 6.4. Evolution of effective elastic modulus versus the penetration depth of the indenter :
thickness of the film for coating formulations ratio with [epoxy]:[COOH] ratio of 1.4 :1
catalyzed by DABCO (7 w/w % of TGIC) for telechelic poly(butyl acrylate) networks (a) and
random poly(butyl acrylate-co-acrylic acid) networks (b) (hf being the thickness of the coating)
In practice, in this chapter, the Eeff values were used to compare results. These values were
obtained from measurements with a total penetration depth of less than 10% of the total coating
thickness. The values presented were values averaged over three calculated unloading
stiffnesses S using loads of 1, 2 and 4 mN. The error of measurement in these average values
was 10%. The data obtained are presented in Figure 6.14 and 6.17. The data presented are for
different [epoxy] : [COOH] ratios in the coatings formulations. The [epoxy] : [DABCO] ratio
was always kept the same (1 : 0.07 in weight). Phenomena such as pile-up and sink-in, which
may lead to some overestimation or underestimation of the contact area and consequently to
errors for the elastic moduli and the hardness, were not taken into account.
Figure 6.5 shows that both types of cross-linked coatings are purely elastic and that the value
of 0.5 for the Poissons ratio could be used in equation [6.3]. The variation observed in the
loading and unloading curve can be explained by hysteresis. The larger amount of hysteresis
observed in the cross-linked random copolymers will be explained later in the Results and
Discussion part (Section IV).

92

Networks applications: synthesis and properties

2.5
2.0

2.0

Load, mN

Load, mN

1.5

1.0

0.5

(a)

0.0

1.5

1.0

(b)

0.5

0.0
0

10

15

20

25

30

Displacement, m

10

15

20

Displacement, m

Figure 6.5. Load-displacements curves for coating formulations ratio with [epoxy]:[COOH]
ratio of 1.4 :1 catalyzed by DABCO (7 w/w % of TGIC) for random poly(butyl acrylate-coacrylic acid) networks (a) and telechelic poly(butyl acrylate) networks (b)

II.5) Solid-state Nuclear Magnetic Resonance


Proton-decoupled solid-state
1

13

C-NMR spectra were recorded on a Bruker DMX500

spectrometer, operating at a H and

13

C frequency of 500.13 and 125.13 MHz respectively. A

4mm magic angle-spinning (MAS) probe head was used in static mode as well as with a sample
rotation rate of 6 kHz. The radio-frequency power was adjusted to obtain 5s 90 pulses both for
the 1H and 13C nuclei. The 38.56 ppm resonance of adamantine was used for external calibration
of the 13C chemical shift.
Proton spin-lattice relaxation in the laboratory and in the rotating frame T1(1H) and
T1(1), respectively, were measured for each component of the polymers separately both via
cross-polarization (CP) to the 13C nuclei and via direct excitation of the 1H nuclei. Relaxation
delays in CP-derived experiments, D1, were 1 and 3 s, and the number of experiments per
relaxation data set, NE, was 20.
Proton transverse relaxation, T2, was determined at several temperatures ranging from -100C
to + 180C, using a BVT-3000 variable temperature unit. The decay of the transverse
magnetization was measured with the Hahn-echo pulse sequence (HEPS), 90--180-(acquisition), where 2.5 s. An echo signal is formed after the second pulse in the HEPS
with a maximum at time t = 2 after the first pulse. By varying the pulse spacing in the HEPS,
the amplitude of the transverse magnetization, A(2), is measured as a function of time t = 2. T2
was determined by computer, fitting each 1H-NMR decay with a bi-exponential function.

93

Chapter 6

III) Starting compounds


Telechelic poly(butyl acrylate): the sample VL103 was used for the experiments involving
telechelic polymers. Its average functionality per chain was determined to be 1.985 by liquidchromatography separation, and its number-average molecular weight is equal to 2.100 g/mol
with a polydispersity of 1.13. More details for the synthesis and characterization can be found in
Chapter 4.
Random poly (acrylic-acid-co-butyl acrylate): In order to determine the effect of the molecular
weight distribution on the network formation and the network properties, some
poly(butylacrylate)/poly(acrylic acid) copolymers were synthesized. The recipes of the
polymerizations were designed in order to obtain number-average molecular weights and
number-average COOH-functionality as similar as possible to the ones of the poly(butyl
acrylate) synthesized by RAFT.
Butyl acrylate (20.0 g, 0.156 mol), acrylic acid (1.4 g, 2.10 10-3 mol) and AIBN (6.4 g, 3.89
10-2 mol) in toluene (50 mL) were added dropwise to a 250 mL round-bottom flask containing
50 mL of toluene. The flask was heated to 80C. The reaction was allowed to proceed for 8 hrs.
Isolation of the copolymer was achieved by rotary evaporation.
Overall conversion: 99+% (GC); Mn = 2200 g/mol, MWD = 1.75 (GPC)
Based on the almost complete conversion, the average molar mass and the initial butyl acrylate /
acrylic acid ratio, the average carboxylic acid functionality per chain is assumed to be close to 2.
Triglycidyl isocyanurate (TGIC): the TGIC cross-linker was received from Aldrich and was
claimed to be 98% pure. However, Liquid-Chromatography Mass Spectrometry (LC-MS) and
1

H-NMR spectroscopy reveal the presence of impurities. It was estimated by 1H-NMR that up to

15 % of the expected epoxy functionalities were not present in the product. The interpretation of
the spectrum in order to determine the chemical structures of the impurities was not successful.
Purification by recrystallisation was known to be very difficult to perform, it was then decided
to use the TGIC as such. It the rest of this chapter the ratios [carboxylic acid] : [epoxy] are
expressed assuming a 100% purity of the TGIC. We were fully aware of the presence of
impurities, but the same batch of TGIC was applied to all the systems, which allows comparison
among the various experiments. The results presented in this chapter will then be semiquantitative rather than quantitative.

94

Networks applications: synthesis and properties

N
N

O
N

O
O

TGIC
1,4-diazabicyclo[2.2.2]octane (DABCO): DABCO (98+%, Merck) was used as received.

DABCO
CH3COOH (Acetic acid, 99+%, Aldrich) and acetonitrile (HPLC grade, 99+%, Biosolve) were
used as received.
Sample preparation
In general, mixtures of polymer TGIC DABCO were prepared using the following ratios:
-

[epoxy] : [COOH] = 1.6 : 1 (molar ratio)

[epoxy] : [DABCO] = 1 : 0.07 (weight ratio)

Variation of the [epoxy] : [COOH] ratio were employed. The same ratios were employed for
the model experiment using acetic acid.
ATR-FT-IR experiments: the polymer, the cross-linker and the catalyst were initially dissolved in
acetonitrile in order to obtain a homogenous solution. The acetonitrile is then removed under
reduced pressure at room temperature. The solventless mixture is then directly applied on the
ATR-FT-IR crystal, preheated at the desired temperature under a nitrogen atmosphere. The time
zero point corresponds to the time of application of the sample on the golden gate. The layer
thickness was a few microns.
Transmission FT-IR experiments: the sample preparation was similar to the one of the ATR-FTIR samples. A thin film of reaction mixture was then applied on a zinc selenide (ZnSe) window.
The time zero point corresponds to the time when the heating up of the ZnSe window starts.

95

Chapter 6

TMDSC: Typically, for every measurement, about 2 mg of cured polymer coating was weighted
out using a precision balance. Standard aluminum pans of similar predetermined masses were
used for all the measurements. The polymer, the cross-linker and the catalyst were initially
dissolved in acetonitrile (50 w/w %) and applied on an aluminum substrate. The samples were
then cured overnight at 120C, and the coating was scratched out of the plate with a scalpel.
Micro-indentation experiments: the polymer, the cross-linker and the catalyst were initially
dissolved in acetonitrile (50 w/w %) and applied on an aluminum substrate. The plate is placed
under a ventilated fume hood for 30 min, in order to evaporate some of the solvent and avoid
crater formation during the cure. The aluminum plates are then placed overnight in an oven at
120C for cross-linking to occur. The thickness of the coatings after cure was about 100 m.
Solid-state NMR experiments: the polymer, the cross-linker and the catalyst were initially
dissolved in acetonitrile (50 w/w %) in a 10 mL glass vial. The vial was then placed overnight in
an oven at 120C to cross-linking purposes. The cured material was scraped out of the vial and
about 100 mg of it was placed in a solid-state NMR rotor for analysis.

IV) Results and Discussion


IV.1) Infra-red results
IV.1.1) Curing mechanism
The acid curing of epoxy is widely used in the coatings industry. Although the reaction can
proceed at high temperature, without the addition of an extra ingredient, often catalysts are used
in order to lower the curing time and/or temperature. Through the years various catalytic
systems were employed: ammonium salts4, metal salts5-7, salts of carboxylic acids8 and tertiary
amines9-17. A lot of the studies used anhydrides as the acid component and therefore, the
mechanism proposed in some of these articles must be different from our situation where we use
carboxylic acid groups. Based on our experiments done so far and on the literature available, we
propose the following mechanism for the reaction of epoxy groups with carboxylic acids in the
presence of DABCO, for our TGIC, telechelic polymer and DABCO mixtures:

96

Networks applications: synthesis and properties

O
N

R
O H

OH

H O

H O

k1

R
O H

NH ,

k-1

OOC

Complex A

NH ,

OOC

k2

O
R'

k-2

,
R'

OOC

R , N

NH

Complex B
O

,
R'

0.35

OOC

R , N

k3
NH

O CH2 CH R'
O

OH

20 min

0.30
Absorbance

0.25
0.20
0.15
0.10
0.05
0.00
3500

3000

2500
W

2000
b -1

1500

1000

Wavelength (cm )

Figure 6.6. ATR-FT-IR absorption spectra during the curing of telechelic poly(butyl acrylate)
by TGIC ([epoxy] : [COOH] = 1.6:1), catalyzed by DABCO (7 w/w % of TGIC), at a
temperature of 100C.
As it can be seen in Figure 6.6, a change in absorption during cure of the mixture of telechelic
polymer, TGIC and DABCO can clearly be observed by transmission FT-IR. A decrease in
absorption of a shoulder at 1700 cm-1 and a peak at 900 cm-1 are observed. Both are known to be

97

Chapter 6

characteristic absorptions for the C=O of the COOH dimer and the C-O of the epoxy groups
respectively, which are present in the reacting mixture, showing that at the temperature used
here, the cross-linking of the components occurs. No absorption for the formation of the new
ester group was found. The C=O stretching absorption of this new ester group is expected at
about 1730 cm-1 which coincides with the large C=O stretching vibration of the acrylate group
present in the telechelic polymer.
Interestingly, following the thermal curing with ATR-FT-IR, we were able to notice a band
related to an intermediate at about 1560 cm-1 which appeared and disappeared during cure
(Figure 6.7). The following reasoning always relies on data obtained following the curing of
telechelic poly(butyl acrylate), but the same was observed for random poly(acrylic acid-co-butyl
acrylate).
Fedoseev et al.18, who studied the cure of acrylic acid and ethylene oxide observed the
formation of a band during cure. They employed pyridine as a catalyst, used NaCl and KBr
pellets as their medium in the IR experiments. They assigned the band at about 1550 cm-1 to the
product of the reaction between the carboxylic acid and the KBr medium. However, in our
studies, the ATR diamond crystal used is inert and can not react with the carboxylic acid,
forming a carboxylate anion. Hence, this band has to be attributed to an intermediate that is
produced during the reaction. The position of the intermediate band at 1560 cm-1 strongly
suggests that the intermediate formed contains a carboxylate anion. The change in absorption of
this band over cure time at various cure temperatures is presented in Figure 6.8.

40 min
20 min
10 min
5 min
0 min

1700

1650

1600

1550

1500

1450

-1

wavenumber (cm )

Figure 6.7. ATR-FT-IR absorption spectra during the curing of telechelic poly(butyl acrylate)
by TGIC ([epoxy] : [COOH] = 1.6:1), catalyzed by DABCO (7 w/w % of TGIC), at a
temperature of 100C.

98

Networks applications: synthesis and properties

In order to get more information about the structure of this intermediate, test experiments were
performed. No band at 1560 cm-1 was observed during the heating at 100C of mixtures of
telechelic polymer and DABCO, TGIC and DABCO, or telechelic polymer and DABCO. The
ratios of the different components in these mixtures were kept the same as in a typical cure
reaction. No reaction at all was observed for the first two mixtures over a heating period of 1
hour. In the last mixture, a decrease in adsorption of the C=O band of the acid dimer end-group
was observed at 1700 cm-1. However, the rate of disappearance was much slower compared to a
curing experiment in the presence of DABCO. Apparently, the band at 1560 cm-1 is only
observed when the 3 components (acid dimer cross-linker catalyst) are present. This
suggests that a complex between the 3 species is formed during the cure reaction and that the
band at 1560 cm-1 may be attributed to the C=O band of the COO- group in complex B of the

100
o

80 C
80

-1

Absorption change in the 1560 cm band A.U.

proposed mechanism.

60

100 C

40

20

120 C

0
0

20

40

60

80

100

120

Time (min)

Figure 6.8. Change in absorption over time of the 1560 cm-1 band at different curing
temperatures of a telechelic poly(butyl acrylate), TGIC ([epoxy] : [COOH] = 1.6:1), DABCO (7
w/w % of TGIC) mixture
It is interesting to notice in Figure 6.8 that after the building up of this intermediate, this
compound is still present at the end of the cure. The level of this absorption seems to vary with
the reaction temperature. However, when it is taken into account that generally the intensity of
the C=O band of a carboxylate anion group is 2-3 times stronger compared to the one of a
carboxylic acid, the final amount left may be very small (a few percent of the original acid

99

Chapter 6

group) and the reaction of this final amount may be slowed down considerably by the decrease

100

-1

Absorption change in the C=O band at 1700 cm A.U.

of chain mobility by network formation at the end of the cure reaction.

80

60

80 C

40
o

100 C
20
o

120 C
0
0

20

40

60

80

100

120

Time (min)

Figure 6.9. Decrease of the 1700 cm-1 band at different temperatures for the curing of telechelic
poly(butyl acrylate) by TGIC ([epoxy] : [COOH] = 1.6:1), catalyzed by DABCO (7 w/w % of
TGIC)
A continuous decrease is observed for the intensity of the 1700 cm-1 characteristic of the
carboxylic acid dimers (Figure 6.9), and also for 900 cm-1 band, characteristic of the C-O in the
epoxy group by ATR-FT-IR (Figure 6.10). However, it was more difficult to follow its
disappearance, as the region contains other peaks that interfere with the one belonging to the

100

80

-1

Absorption change in the 900 cm band A.U.

epoxy group.

80 C

60

40

20

100 C

120 C
0

10

20

30

40

50

60

70

80

90

100 110 120 130

Time (min)

Figure 6.10. Decrease of the 900 cm-1 band at different temperatures for the curing of telechelic
poly(butyl acrylate) by TGIC ([epoxy] : [COOH] = 1.6:1), catalyzed by DABCO (7 w/w % of
TGIC).

100

Networks applications: synthesis and properties

Comparing the first part of the reaction in Figure 6.8 and 6.10, the spectra suggest that the
intermediate responsible for the band at 1560 cm-1 also influences the 900 cm-1 band. The
changes in the slopes at different curing temperatures in Figure 6.10 coincide with the maxima
observed in Figure 6.8. The proposed intermediate molecular structure contains also an epoxy
group which is expected to absorb at about 900 cm-1. The slow initial decrease in absorption at
900 cm-1 over time strongly suggests that initially the complex B is formed, while the epoxy is
slowly consumed. We suggest that when most of the catalyst has disappeared, the reaction
accelerates, provoking the changes in Figure 6.8 and 6.10. More investigation is needed to
confirm this speculation.
A model experiment was conducted, following the reaction at 100C of acetic acid, TGIC
([epoxy] : [COOH] = 1.6:1) and DABCO (DABCO (7 w/w % of TGIC). The IR-spectra were
easier to analyze in the C=O band region, as the massive band of the butyl acrylate of our
polymers was not present. We could clearly see that a great majority of the carboxylic acid
groups are presents as dimers at time zero of the reaction (C=O=1700 cm-1), while the presence
of monomer (C=O=1750 cm-1) was very small. We can see in Figure 6.11 that the intermediate
at 1560 cm-1 is also formed in this model reaction and that its concentration varies similarly to

Absorbance A.U.

the intermediate found in the curing of the telechelic polymer.

-1

1578 cm

-1

3006 cm

10

15

20

25

Time (min)

Figure 6.11. Variation of various bands at 90C for the reaction of acetic acid with TGIC
([epoxy] : [COOH] = 1.6:1), catalyzed by DABCO (7 w/w % of TGIC).

101

Chapter 6

This IR experiment was performed in the transmission mode, which allowed us to obtain a
stronger signal in interesting regions of the spectra. The variations of the following bands were
followed:
-

1578 cm-1: Symmetric stretching of the C=O band of the carboxylate anion

3006 cm-1: Symmetric stretching of the C-H band of the epoxy group

The disappearance of the epoxy group seems to proceed in two steps similarly to what was
found previously. Acceleration in the epoxy consumption happens after a maximum in the
intermediate concentration is reached. The same reasoning that was made on the curing of the
telechelic polymers by TGIC seems to be applicable for this model reaction as well
In the curing of random copolymers, the intermediate at 1560 cm-1 was also detected. We can
then propose the mechanism presented for our TGIC, telechelic polymer and DABCO mixtures
as a general reaction mechanism for epoxy groups with carboxylic acids in the presence of
tertiary amines.
IV.1.2) Kinetics differences: random vs. telechelic
As it can been seen in Figure 6.12, the disappearance of the acid dimer peak is much faster in
the case of telechelic polymer than for the random copolymer case. In both cross-linking
reactions, the concentration of the acid groups, epoxy groups and catalyst are the same, within
the error of measurement. Several explanations for the difference observed can be given using
steric considerations. For the telechelic poly (butyl acrylate), the distance between carboxylic
acid groups is roughly 2.000 g/mol. Once one carboxylic acid has reacted with an epoxy, the
other carboxylic acid of the chain is still available, in a reasonably mobile form, being far away
from the reacted one. For the random copolymer chains bearing 2 or more acid groups, the
distance between the two carboxylic acid groups can be much smaller. It is on average, largely
inferior to the one encountered in the telechelic polymer case. After reaction of a carboxylic acid
group, the mobility of the closest unreacted acid group is on average reduced, in a much more
pronounced fashion than for the telechelic one. When a copolymer chain contains more than two
acids groups, this effect is even more pronounced. This lack of mobility/availability of the
carboxylic acid groups may be one of the reasons behind the difference of the cure reaction rate
shown in Figure 6.12.
Another explanation for the difference in rate might be the concentration of acid dimers
between two carboxylic acid end-groups. There might be differences in the dimer formation
probability, possibly leading to a lower concentration of acid dimers for the random poly

102

Networks applications: synthesis and properties

(acrylic acid-co-butyl acrylate). This ultimately may lower the formation of complex B, and may
explain the rate difference observed in the FT-IR studies. Unfortunately, the ratio between
monomers of acid group and dimers of acid could not be verified, because the absorption at
about 1750 cm-1, where the band for the C=O of the acid monomer lies, was overlapped by the

1.0

-1

Absorption change in the C=O band at 1700 cm A.U.

very strong C=O band of the acrylate group present in the polymers.

0.8

0.6

(a)
0.4

-1

80 C

0.2
o

120 C
0.0
0

Absorption change in the C=O band at 1700 cm A.U.

100 C

50

100

150

200

Time (min)
1.0

0.8

0.6

(b)

0.4
o

80 C
0.2

100 C
o

120 C

0.0
0

100

200

300

400

500

600

700

800

Time (min)

Figure 6.12. Decrease of the 1700 cm-1 band at different temperatures for the curing of carboxyfunctional polymer by TGIC ([epoxy] : [COOH] = 1.2:1), catalyzed by DABCO (7 w/w % of
TGIC); (a) telechelic poly (butyl acrylate); (b) random poly(acrylic acid-co-butyl acrylate)
In literature, a mechanism for the catalysis by tertiary amines of the carboxylic acid epoxy
reaction was proposed by Kucharski and Rubczak12 (Figure 6.13):

103

Chapter 6

HOOC

k1

NH

OOC

k-1

NH

,
R'

OOC

OOC

k2

R'

k3

k-2

O CH2 CH
O

O CH2 CH
O

R'

NH

k-4

,
R'

OOC

NH

R'

k4
R

O CH2 CH
O

R'

OH

Figure 6.13. A proposed mechanism for the catalyze by DABCO of the carboxylic acid epoxy
reaction
The authors stated using a steady-state approximation on the carboxylic anion quaternary
ammonium cation complex, and assuming that k 2 <<k1 , k -1 that:

[ DABCO, H + , R - COO - ] = K 1 [ R - COOH][DABCO]


where K 1 =

[6.4]

k1
k -1

Unfortunately, we can affirm that the proposed mechanism or the assumptions made were
incorrect. The disappearance of the acid groups can be expressed by:

d[COOH]
= k -1 [ DABCO - H + , R - COO - ] - k 1 [ R - COOH][DABCO]
dt

[6.5]

By substituting equation [6.5] in equation [6.4], we obtain:

d[COOH]
=0
dt

which can obviously not be true. It seems that the steady-state approximation used by the
authors is not realistic. The band at 1560 cm-1 surely represents a complex that is formed during
the cross-linking reaction. Figure 6.8 clearly shows that the steady-state approximation can not

104

Networks applications: synthesis and properties

be applied to this compound. We believe that this is the point where Kucharski and Rubczaks
kinetic investigation is incorrect. However, their proposed mechanism for the epoxy-carboxylic
acid reaction is more detailed than the one we have proposed at the beginning of this chapter. It
was impossible for us to investigate thoroughly the mechanism with the help of techniques such
as FT-IR or NMR, due to time constraints.

IV.2) Determination of E-modulus


IV.2.1) Telechelic poly(butyl acrylate) micro-indentation measurements
The Eeff for coatings based on telechelic poly(butyl acrylate) cross-linked with TGIC with be
referenced as Eefftel and are presented in Figure 6.14. The effect of the [polymer] : [cross-linker]
ratio on this mechanical property appeared to be large.
30

25

Eeff (MPa)

20

1
1.2
1.4
1.6
2
2.4
2.8
3.2

15

10

1mN

2mN

4mN

Loading force

Figure 6.14. Eeff tel values for VL103 telechelic polymer cross-linked with TGIC with various
[epoxy] : [COOH] ratios.
It can be noticed in Figure 6.14 that the Eefftel increases continuously when an increasing
amount of epoxy functions is present in the formulation before cross-linking.

105

Chapter 6

As it was shown before (equation [6.3])

1
1 - 2 1 - i
=
+
E eff
E
Ei

[6.3]

and considering that Ei >> E and that the cross-linked polymer behaves like a perfectly elastic
material ( = 0.5) we obtain:

E tel = 0.75Eeff tel

[6.6]

For the sake of clarity, all the reasoning will be done on Eefftel (unless otherwise mentioned).
The conclusions will of course also apply to Etel as well.
In a different form, a part of the data of Figure 6.14 is presented again in Figure 6.15.
25

Eeff (MPa)

20

15

10

0
0.8

1.2

1.4

1.6

1.8

2.2

2.4

2.6

2.8

3.2

3.4

[Epoxy]:[COOH] ratio

Figure 6.15. Evolution of the effective elastic modulus with various amounts of
[epoxy]:[COOH] ratios for telechelic-based poly(butyl acrylate) networks. Loading force = 1
mN
The curve displayed in Figure 6.15 was derived from Figure 6.14, using the values obtained
with a loading force of 1 mN. It is usually typical for a coating to have values for E increasing
up to a certain [polymer]:[cross-linker] ratio (typically 1:1), and then to observe a decrease of
the E modulus when the cross-linker is in excess. This is easily understandable, as for ratios
different from 1:1, the structure of the network will be imperfect, and defects such as dangling
ends will have a tremendous effect in decreasing the E modulus. The molecular weight between
cross-links might also be different. However in our case, the effective modulus increases

106

Networks applications: synthesis and properties

continuously. For [epoxy] : [COOH] ratios above 2, the Eeff values reaches numbers that are
difficult to assign to soft poly(butyl acrylate) networks. The well-known polyetherification of
TGIC, catalyzed by bases is probably the reason behind this huge increase in the modulus at
high [epoxy] : [COOH] ratios
In order to confirm this, the curing conditions were applied to a mixture of only TGIC and
DABCO, with no polymer. After curing a yellow brittle film was obtained. The TGIC reacts
with itself in a poly-condensation reaction, forming a polyether. With three functionalities per
organic molecule, the reaction will ultimately lead to the formation of a three-dimensional
insoluble film. The film obtained contains a lot of hard isocyanurate rings, which confer a high
E modulus to the film.
Interpreting this experiment, it seems likely to conclude that a competitive reaction (autopolymerization of TGIC) occurs next to the carboxylic acid epoxy reaction. When an excess of
TGIC is used, clusters of hard polyethers are present in the matrix of poly(butyl acrylate). When
the [epoxy] : [COOH] ratio was raised to 3.2, a phase separation was visually noticeable. It is
then logical to assume that for the [epoxy] : [COOH] ratios superior to 2 (when the E values rise
to improbable values), the Eefftel measured is not a parameter only influenced by the structure of
the poly(butyl acrylate) network (formed by epoxy-carboxylic acid reactions), but rather a
number relative to the modulus of a poly(butyl acrylate) network, with polymerized TGIC
included as network domains and/or clusters inside it.
It is also known that trans-esterification might occur between the alcohol group resulting from
the epoxy-carboxylic acid reaction and the butyl acrylate-side chain. Although this phenomenon
may occur, it is difficult to explain the high values (for example 25MPa for the 3.2:1 ratio) only
with this reaction, as the chemical structure of the network should be roughly the same. It is only
hard domains of polymerized TGIC that can cause the elastic modulus to rise to such values.
The principal question that remains concerns the influence of this autopolymerization sidereaction on the values obtained for the elastic modulus for [epoxy] : [COOH] ratios below 2. We
will answer this question in the last part of this chapter using solid-state NMR experiments.

Rubber elasticity theory


In the case of perfectly alternative elastomeric networks, resulting in a Poissons ratio of
nearly 0.5, the modulus, E, can be expressed as the following equation using the Gaussian
rubber elasticity theory19-21.

107

Chapter 6

E = 3RT e

[6.7]

where R is the gas constant and T the absolute temperature and e the density of elastically
active chains.
For our telechelic polymer we can assume that the density of cross-link points can be related
to the molecular weight of the polymer, and we obtain:

E=

3RT
Mc

[6.8]

where is the density of the poly(butyl acrylate) gel, Mc, the molar mass between junctions,
R, the gas constant and T, the absolute temperature. This formula is valid well-above Tg. The Tg
of our polymers was experimentally determined by DSC and are in the range of -10C
(experimental data will be provided in the next section of this chapter). Performing the microindentation experiments at room temperature, we may assume that this rubber elasticity theory is
applicable to our coatings. It is generally assumed than a minimum of 20 monomer units in the
polymer chain is needed for the rubber elasticity theory to be applicable. We are with our
systems a little bit below this limit, but we still consider that comparing our experimental results
to the theoretical ones is relevant.
When we assume that the molar mass between cross-links is equal to the molar mass of the
telechelic polymer, that is 2000 g/mol and a value of 1 for the density, the theoretical value for
the E modulus of 3.7MPa can be calculated using equation [6.8]. From this value, and assuming
a Poisson's ratio of 0.5, we calculated a theoretical value for Eefftel of 4.9 MPa. This theoretical
value is very close to the value observed for [epoxy] : [COOH] ratios of 1.4, 1.6 and 2. Hence,
the most perfect network in our coatings is probably at a much higher [epoxy] : [COOH] ratio
than 1:1. Further evidence for this will be presented in the solid-state NMR part of this chapter.
One could argue that due to the complexity of the network topology, it is not possible to
compare our experimental values with the ones obtained using the Gaussian rubber elasticity
theory. It is however interesting to notice that the obtained experimental values are very similar
with those theoretically expected for a perfectly alternating network. A much stronger difference
would have been expected beforehand. The Gaussian model used here does not take into

108

Networks applications: synthesis and properties

account network defects such as loose ends, loops and entanglements. The first two factors will
lower the Eefftel value whereas the last factor will enlarge it. The difference between the
experimental and the theoretical values for the Eefftel value can then be explained by the network
defaults.
The deviations at ratios above 2 might be attributed to the presence of the polyether network
produced by the autopolymerization of TGIC. The deviation with the rubber elasticity theory at
lower [epoxy] : [COOH] ratios are possibly caused by the larger Mc values between cross-links
and/or the presence of more dangling ends in the networks. The influence of these factors on the
elastic modulus of the network will be discussed in more details in the next section of this
chapter.
IV.2.2 Influence of mono-functional chains on the mechanical properties of poy(butyl
acrylate)networks
The synthesis of well-defined linear mono-functional poly(butyl acrylate) allowed us to
introduce a controlled amount of defects (dangling ends) in the network structure. We select the
[epoxy] : [COOH] ratio of 1.6 for our next experiments. This choice was governed by the global
reliable results that were obtained in our previous investigations, and the belief that the influence
of the polyether formation is not yet predominant for this ratio. The coating formulations
prepared contained 5 w/w %, 10 w/w %, 20 w/w % of the total weight of polymers of monofunctional polymer VL131 (Mn= 2.300 g/mol, MWD = 1.14). Due to the small differences
between the molar masses of the mono- and di-functional poly(butyl acrylate), the weight
percentage can roughly be approximated to number percentage.
As shown in Figure 6.16, the Eefftel decreased continuously with an increasing amount of
dangling ends. This effect is more severe when the first few percents of mono-functional chains
are introduced.
The plasticizing effect of the dangling ends is a known phenomenon in the literature. This
gives a reasonable explanation why the Eefftel values measured for [epoxy] : [COOH] ratios of 1
and 1.2 are lower than the one predicted by the rubber elasticity theory. Impurities in the crosslinker lower the actual amount of epoxy groups in the formulation (about 15% less), and lower
the [epoxy] : [COOH] below stoechiometry at experimental ratios below 1.2.

109

Chapter 6

Eeff (MPa)

4.5

3.5

3
0

10

15

20

25

wt. % of dangling ends added

Figure 6.16. Evolution of the effective elastic modulus with various amounts of dangling ends
for telechelic-based poly(butyl acrylate) cross-linked with TGIC ([epoxy] : [COOH] = 1.6).
This may result in non-reacted COOH functions, leaving dangling ends in the network
structure. The dangling ends can also be obtained in another manner. At high conversion of the
cross-linking reaction, the remaining COOH functional groups might be hard to access for the
epoxy groups still present, for entropic reasons. The mobility of the chains is reduced by the
cross-linking points, allowing some dangling ends to persist, although macroscopically some
epoxy and carboxylic acid groups are still present in the network.
By addition of mono-functional polymer to the coating formulation, not only the amount of
dangling ends but also Mc increases. This increase will lower the Eefftel value and is another
explanation for the lowering of the moduli reported in Figure 6.16.
IV.2.3) Random poly (butyl acrylate-co-acrylic acid) polymers micro-indentation measurements
The Eeff for coatings based on random poly (butyl acrylate-co-acrylic acid) cross-linked with
TGIC will be referred to as Eeffran and are presented in Figure 6.17. The effect of the [polymer] :
[cross-linker] ratio on the effective elastic modulus was investigated.

110

Networks applications: synthesis and properties

Eeff (MPa)

1
1.2
1.4
1.6
2

2mN

1mN

4mN

Loading force

Figure 6.17. Eeffran values for random copolymer cross-linked with TGIC with various [epoxy]
: [COOH] ratios.
Similar to the telechelic poly(butyl acrylate), all the reasoning will be done on Eeffran (unless
otherwise mentioned). The evolution of Eeffran with different [polymer]:[cross-linker] ratios for
random poly(butyl acrylate-co-acrylic acid) is presented on Figure 6.18.
As seen in Figure 6.18, we observe for the effective modulus of the random copolymers-based
networks a trend similar to the one of the telechelic polymers-based networks. Unfortunately no
coating at high [epoxy] : [COOH] ratios were prepared, it was impossible to compare both
systems for high ratios. For the random copolymers, the effective modulus continuously
increases with increasing amounts of cross-linker, similar to what was observed for the
telechelic polymers. The same explanation can be given, i. e. the formation of a more complete
network with a lower Mc and the presence of less-dangling ends at [epoxy] : [COOH] ratios of
1.6:1 and 2:1.
We observed in our experiments that the values of the effective modulus are lower for the
networks based on the random copolymers compared to the telechelic polymer-based ones. This
difference can be estimated at 20%. The lower values found are in contradiction with the
22

theoretical predictions of Dusek et al . Their calculations show that random copolymers-based


networks have a higher elastic modulus compared to the telechelic polymers-based networks,
the average functionality per chain being the same.
The main factors explaining these low values may be the incompletion of the cross-linking
reactions, and the larger Mc. Another factor explaining this lowering might be the non

111

Chapter 6

randomness of the copolymer chains. Those defects in the random copolymers-based networks
are possibly responsible for the hysteresis phenomenon that was observed on the loaddisplacement curves during the micro-indentation experiments for the random copolymer-based
coatings(Figure 6.5). The adhesion between the indenter tip and the coating can be another
reason.
30

Random copolymers
Telechelic polymers

25

Eeff (MPa)

20

15

10

0
0

0.5

1.5

2.5

3.5

[epoxy] : [COOH] ratio

Figure 6.18. Evolution of the effective elastic modulus with various amounts of
[epoxy]:[COOH] ratios for for random poly(butyl acrylate-co-acrylic acid) networks and
telechelic poly(butyl acrylate) networks. Loading force = 1 mN

VI.3) MTDSC measurements


-5

-10

Tg (C)

-15

-20

-25

-30
Telechelic polymers (Tg uncured = - 61.2C)
Random polymers (Tg uncured = - 46.9C)

-35
0.8

0.9

1.1

1.2

1.3

1.4

1.5

1.6

1.7

[epoxy] : [COOH]

Figure 6.19. Evolution of the glass transition temperature with various amounts of
[epoxy]:[COOH] ratios for for random poly(butyl acrylate-co-acrylic acid) networks and
telechelic poly(butyl acrylate) networks.

112

Networks applications: synthesis and properties

For both systems (telechelic poly (butyl acrylate) and random copolymers) an increase of the
glass transition temperature with an increasing amount of the [epoxy] : [COOH] ratio is
observed (Figure 6.19). Those measurements are in line with the conclusions drawn from the
micro-indentation experiments.
We observe a decrease of the glass transition temperature with an increasing amount of
dangling ends as expected (Figure 6.20), confirming the results of the micro-indentation
measurements. The decrease in the glass transition temperature is more severe when the first
few percents of mono-functional chains are introduced as it was observed for the Eefftel.
-8
-8.5
-9

Tg (C)

-9.5
-10
-10.5
-11
-11.5
-12
0

10

15

20

25

wt. % of dangling ends added

Figure 6.20. Evolution of the glass transition temperature with various amounts of dangling
ends for telechelic-based poly(butyl acrylate) networks ([epoxy] : [COOH] = 1.6).
When the different values obtained for the telechelic cross-linked coatings (with and without
extra-added dangling chains) were plotted against the corresponding glass transition temperature
values, a linear relation was found within the accuracy of the measurement (Figure 6.21). No
such linear relation was found for the random copolymers- based coatings (Figure 6.22).

113

Chapter 6

5
4.5
4

Eeff (MPa)

3.5
3
2.5
2
1.5
1
0.5
0
-18

-16

-14

-12

-10

-8

-6

Tg (C)

Figure 6.21. Evolution of the effective elastic modulus with the glass transition temperature for
telechelic-based poly(butyl acrylate) networks ([epoxy] : [COOH] = 1.6).
4
3.5

Eeff (MPa)

3
2.5
2
1.5
1
0.5
0
-32

-30

-28

-26

-24

-22

-20

Tg (C)

Figure 6.22. Evolution of the effective elastic modulus with the glass transition temperature for
random-based poly(butyl acrylate) networks ([epoxy] : [COOH] = 1.6).
It is known that for cross-linked coatings, the glass transition temperature and the elastic
modulus can both be described as a function of the molecular weight between cross-links23. This
can explain the linear relation observed in Figure 6.21. These data also suggest that the
variations found in the elastic modulus and in the glass transition temperature of the telechelic
polymer-based coatings are mainly determined by variations in Mc. In general, no linear relation
between the E modulus and the glass transition temperature is found. The numerous

114

Networks applications: synthesis and properties

imperfections in the random copolymer-based networks (especially the presence of free-polymer


chains with a very low elastic modulus) may be the cause for the deviation from linearity shown
in Figure 6.22.

VI.4) Solid-state Nuclear Magnetic Resonance results


VI.4.1) T1- relaxations experiments
Networks based on telechelic poly (butyl acrylate) cross-linked with TGIC (with [epoxy] :
[COOH] ratios of 1:1 and 1.6:1) were analyzed with solid-state NMR. Figure 6.23 shows the 13C
spectra of the starting compounds and of the cross-linked material.
The assignments of the different peaks of the

13

C NMR spectrum of the VL103 telechelic

poly (butyl acrylate) are presented in Table 6.24. Figure 6.23.c strongly suggests that in this
formulation most of the epoxy (at 50 ppm) and acid groups (at 182 ppm) have disappeared after
cross-linking.
We have investigated proton spin-lattice relaxation in the rotating frame (T1) and in the
laboratory (T1) via cross-polarization of the 13C nuclei to study possible domain formation.
1

H-NMR T1 and T1 measurments yield mobility information of the considered nuclei at

respectively the nanosecond and millisecond time scale. This information can provide
indications about the homogeneity of the networks, and the possible presence and miscibility of
polymerized TGIC in our coatings.
As a result of the proton-proton dipolar coupling, protons in a polymer continuously exchange
their polarization (nuclear magnetization). This process, usually referred to as spin diffusion,
tends to average out local differences in NMR properties, such as relaxation. Spin diffusion is
fast in rigid polymers with closely interspaced protons, and slow in mobile polymers with low
proton density.
At a standard spin diffusion coefficient D of approximately 1 nm2.ms-1 (although this value
can be the subject of discussion considering the high mobility of the polymer chains), protonproton spin diffusion averages out any T1 or T1 relaxation difference for average domain size in
a polymer blend smaller than about 1nm,. All protons then decay with the same effective T1 or
T1. In contrast, if the domain size is larger than about 50 nm, spin diffusion is too slow to
average out such differences, and each phase will decay with its intrinsic, probably different T1
or T1 values. Furthermore, since T1 tends to be 10 to 100 times longer than T1, the effective

115

Chapter 6

diffusion path length is longer in T1 experiments. Therefore, in the intermediate range, for
domain sizes ranging from 1 to 50 nm, different effective T1 values and a single effective T1 are
expected. Spin diffusion is still able to average out T1 differences, although it is not able to
homogenize T1.
For the domain analysis proton using T1 and T1 relaxation via cross-polarization to the 13C
nuclei, the signals in the aliphatic regions (0-70 ppm) and in the carbonyl region (140-180 ppm)
were studied.

(a)

200 180 160 140 120 100

80

60

40

ppm

80

60

40

ppm

80

60

80

60

(b)

200 180 160 140 120 100

(c)

180

160

140

120

100

200 180 160 140 120 100

40

40

20

ppm

ppm

Figure 6.23. Solid-state 13C NMR spectra obtained via direct excitation for (a) telechelic
poly(butyl acrylate) (b) TGIC (c) cross-linked network with [epoxy] : [COOH] ratio of 1:1.6
(* = solvent peaks)

116

Networks applications: synthesis and properties

HOOC

CH2

CH

CH2

4
2

CH

CH CH2

CH CH2

11
COOH

10

3
1

Carbon number

Description

Chemical shift (ppm)

CH3 butyl group (CH3CH2CH2CH2)

14

CH2 butyl group (CH3CH2CH2CH2)

20

CH2 butyl group (CH3CH2CH2CH2)

31

CH2 butyl group (CH3CH2CH2CH2)

64.5

COO acrylate group

174

COO acrylate group, close to CS3

170

CS3 trithiocarbonate

Theoretically: 220

CH backbone

40

CH2 backbone

37

Quaternary carbon of the end-groups

41

10

CH3 of the end-groups

25.5

11

End-group COOH

182

Table 6.24. 13C NMR peak attribution for VL103

117

Chapter 6

(a)

*
*

(b)

*
*

*
150

100

50

-50 ppm

Figure 6.25. 13C-NMR spectra obtained at different T1-filter times via cross-polarization to
the 13C nuclei, for cross-linked network with [epoxy] : [COOH] ratio of 1:1 (* = sideband) (a)
1

H T1-filter = 10 s (b) 1H T1-filter = 6 ms

The analysis of two 13C spectra with different proton T1 filters, as presented in Figure 6.25,
indicates that the poly (butyl acrylate) and the TGIC decay similarly. Hence, both compounds
show identical T1 behavior. The homogeneity is then proven on the nanometer scale. The T1relaxation experiments were then unnecessary, as they are giving information on the micrometer
scale.
The main conclusion of this T1-relaxation experiment is the homogeneity of the formed
network after the epoxy-carboxylic acid reaction. Unfortunately the lack of time did not allow us
to investigate T1-relaxation for formulations with higher epoxy content. It would have been
interesting to perform relaxation experiments on formulations that gave unrealistic high values
for the elastic modulus (such as 3.2 : 1 for example). Nevertheless, we are able to conclude that
the polyetherification (autopolymerization of the TGIC) reaction does not occur in an extended

118

Networks applications: synthesis and properties

fashion, as no phase of TGIC could be recorded. We can then affirm that, when there is no
excess of epoxy groups versus carboxylic acid groups, the polyetherification is not competitive
to the epoxy carboxylic acid reaction. The rate of this polyetherification is probable orders of
magnitude lower than the one of the epoxy COOH reaction. The data obtained so far suggest
that the polyetherification only significantly occurs when almost all the carboxylic acid groups
have reacted.
VI.4.2) T2- relaxation experiments
T2-relaxation was investigated for telechelic-based poly(butyl acrylate) networks, with
[epoxy] : [COOH] ratios of 1:1 and 1.6:1 in order to collect information on molar masses
between cross-linking points. As reported by Litvinov24,25, the T2-relaxation time for elastomer
networks is sensitive to rotational motion of the polymer chains at temperatures about 100C150C above the glass transition temperature. More specifically, T2 depends on the amplitude of
collective segments motions. The mobility of a chain is dependant of the temperature. The
higher the temperature, the more mobile the chain is. However, beyond a certain point, the
mobility of the chain will not vary with the temperature anymore, its motions being restricted by
physical or chemical entanglements. In our case, it is naturally the cross-linking points that will
prevent the chain from moving freely. The T2-relaxation time becomes then independent of the
temperature, and on the T2 temperature, a plateau value is obtained. At low temperatures, the
chain motion is characterized by local segment motion. When no cooperative motion occurs, the
value of the T2-relaxation time reaches a constant value. The T2 temperature graph exhibits
then another plateau at low temperatures determined by the static dipolar proton-proton couple.
At higher temperature, the chain motion increases, and the dipole interactions become more and
more averaged.
In our proton T2-relaxation experiments, the observed decays of the so-called Hahn-echo as
a function of the echo time, 2 did not follow a mono-exponential behavior. However, it was
possible to fit these decays with a bi-exponential function, described by the following
equation26:

2
2
A(2 ) = K exp
+ (1 ) exp

2a
2b

[6.9]

where K represents a constant related to the amplitude of the chain motions, is the fraction
of fast-relaxing segments in the networks, T2a and T2b are the T2-relaxation values respectively

119

Chapter 6

related to fast and slow-relaxing components in the networks. Due to the structure of the poly
(butyl acrylate) chain, it is logical to assign the T2a proton relaxation to the protons of the poly
(butyl acrylate) backbone, and the T2b proton relaxation to the protons of the butyl side-chains.
The parameter is then independent of the temperature.
The number of statistical segments, Z, may be determined using the following equation24,25:

Z=

T2p
aT2rl

[6.10]

where T2p is the value of T2 at the high temperature plateau and T2rl is the value of T2,
independent of the network structure, in the glassy state. A geometrical coefficient, a, is
introduced and is dependant on the angle between the segment axis and the internuclear vector
between protons in the main chain. For polymers containing aliphatic protons in the main chain,
the coefficient a is close to 6.2 0.7. The presence of the trithiocarbonates CS3 in the middle of
our polymer chains probable influences the value of this coefficient.
[epoxy] : [COOH] 1:1
10000

[epoxy] : [COOH] 1.6:1

T2A (s)

1000

100

10

1
150

200

250

300

350

400

450

Temperature (K)

Figure 6.26. Evolution of T2a-relaxation time with temperature for networks based on telechelic
poly (butyl acrylate) with different amount of cross-linker
Using the number of backbone bonds in the statistical segment, C, the molar mass of network
chains between chemical network junctions, Mc is calculated:

Mc =

ZCM u
n

[6.11]

120

Networks applications: synthesis and properties

Where Mu is the molar mass per elementary chain unit for the polymer chain and n is the
number of backbone bonds in an elementary unit.
For our polymers, the values for the a and C coefficients are not known. Instead of
determining the absolute masses between cross-linking points, analyzing the T2a-relaxations
curves, we can estimate a ratio for the masses between cross-links in the two systems.

T2p

(1:1)
M c (1:1)
T2rl
=
M c (1:1.6) T2p
(1.6 :1)
T2rl

[6.12]

As it can be seen in Figure 6.26, the T2a-relaxation time curve reaches at high temperatures a
plateau value for both systems. This evidences the presences of networks points in the systems,
limiting the motions of the polymer chains, and acknowledges the formation of a threedimensional network in both samples. The gap between the two plateau values at high and low
temperature is wider in the case of the coating prepared with a 1:1 [epoxy] : [COOH] ratio,
indication of a looser network.
Using the values obtained in Figure 6.26 an the relation [6.12], we obtain a value of 2.8 10%
for the ratio between the distances between cross-links in the sample with the [epoxy] : [COOH]
ratio of 1:1 and in the one with the [epoxy] : [COOH] ratio of 1:1.6.
Using equation [6.8] it is easy to write that:
tel
Eeff
(1: 1.6)
M c (1: 1)
= tel
M c (1:1.6)
Eeff (1:1)

[6.13]

The values for the elastic moduli of these coatings presented of Figure 6.15 leads to a ratio of
2.4 20 %. This value is, within the error of measurement, equal to the value observed with
solid-state NMR. T2-relaxation measurements on both coatings. Hence, these solid-state NMR
experiments confirm that the differences of elastic modulus values obtained for the cross-linked
telechelic polymers using these two ratios in the formulations presented on Figure 6.14 and 6.15
can be explained by a looser network, not taking into account the dangling ends. The topological
distance between two neighboring cross-linking points is approximately three times larger in the
sample with a 1:1 [epoxy] : [COOH] ratio.

121

Chapter 6

VII) Conclusion
In this chapter, the telechelic poly (butyl acrylate), whose synthesis was described in Chapter
4 were employed to form three-dimensional networks, via an epoxy-carboxylic acid reaction
between the acid functionalities present at the ends of the linear polymer chains and the epoxy
groups of the TGIC cross-linker, catalyzed by the tertiary amine functions of DABCO. In order
to determine the influence of the functionality distribution and the molecular weight distribution
of the polymer, some random poly (acrylic acid-co-butyl acrylate) were synthesized with the
number-average molecular weight and number-average functionality very similar to the ones of
the telechelic polymer.
The mechanism of this epoxy-carboxylic reaction was studied by ATR-FT-IR using telechelic
polymers for cross-linking experiments. A band at 1560 cm-1 was detected, and its growth and
disappearance over time were recorded. We attributed this band to the stretching vibration of a
carboxylate anion group present in an intermediate complex involving the COO- group derived
from the carboxylic acid functions present in the polymer, the epoxy group and DABCO. No
presence of this band was recorded when only two of the three components were mixed and
heated up to the curing temperature. A new mechanism to explain the cross-linking reaction was
proposed. More investigations were done to check the validity of the mechanism. The structure
of the intermediate was confirmed by studying the rate of disappearance of the acid dimers, the
epoxy band at 900 cm-1 and the rate of formation and disappearance of the intermediate at 1560
cm-1 at different cure temperature. Further information was obtained by replacing the polymer
by acetic acid, while keeping the same [epoxy] : [COOH] ratio. The same variations in the band
adsorptions of the epoxy and the intermediate were found in this experiment.
The random copolymers were found to cross-link similarly to the telechelic ones but at a
slower rate. Two main reasons were proposed to explain this phenomenon. First of all, the lack
of mobility of random copolymers chains compared to the telechelic ones, once the crosslinking process has begun. Another possibility is that the smaller amount of COOH dimers
between two close COOH groups in the random copolymer might as well reduce the reaction
rate of the cross-linking process. The cross-linked coating made from telechelic and random
polymers always had a glass transition temperature well below room temperature and appeared
to be fully elastic.
The elastic moduli of coatings based on telechelic poly (butyl acrylate) and random poly
(acrylic acid-co-butyl acrylate) were estimated using micro-indentation measurements. The

122

Networks applications: synthesis and properties

evolutions of the E modulus versus the [epoxy] : [COOH] ratio used in the cross-linking
formulation were presented for both systems. It was found that with a similar [epoxy] : [COOH]
ratio in the starting formulations, the moduli of the coatings based on the telechelic polymer
were higher than the coatings prepared from random copolymers. We attributed this difference
to two different phenomena. The incompletion of the cross-linking reaction in the random
copolymer is a first possible explanation. The second reason is the difference in cross-link
density between both systems. The network based on random copolymer is probably less
perfect and is looser than the one based on telechelic polymers. TMDSC confirms these
results with evolutions of glass transition temperatures in line with the evolutions of the E
moduli. A linear relation between glass transition temperature and elastic modulus was found
for the telechelic polymers. This relation is not linear for random copolymers. We attributed this
difference to the presence of non-functional chains in the network.
Influence of the dangling chains on the E modulus of the coatings based on telechelic
poly(butyl acrylate) was demonstrated by micro-indentation, employing various amounts of
well-defined mono-functional chain in the curing of telechelic polymer. It has been shown that
the E modulus decreases with an increasing amount of dangling chains. This effect is more
severe when the first few percents of mono-functional chains are introduced.
For high [epoxy] : [COOH] ratios (above 2), a dramatic increase of the E modulus was
observed in the telechelic polymer-based coatings, those values being unrealistic for
polyacrylate-based networks. The reason behind it is the formation of hard clusters, whose
presence is justified by the polyetherification of the epoxy moieties of the TGIC. This sidereaction is competitive to the epoxy-carboxylic acid one but its kinetics is probably much
slower. It occurs to a significant extent, only when the epoxy groups are in excess, as no phase
separation is observed for low [epoxy] : [COOH] ratios by solid-state NMR.
Finally, solid-state NMR was employed to investigate the homogeneity of the networks made
from telechelic polymers for the lower [epoxy] : [COOH] ratios. T1-relaxation experiments
revealed the homogeneity of the formed network after the epoxy-carboxylic acid reaction on the
nanometer scale, ruling out the presence of TGIC-polymerized clusters in the final network. T2relaxation experiments showed that the coating with a 1:1 [epoxy] : [COOH] ratio has an
average molecular weight between cross-linking points roughly three times higher than the one
of the 1.6:1 [epoxy] : [COOH]. This ratio for molecular weight between cross-links was
compared to the one calculated from the rubber elasticity theory, using the E modulus values
determined by micro-indentation. A satisfactory agreement between the two methods of
calculation was found.

123

Chapter 6

Reference List
1. Soloukhin, V. A., Posthumus, W., Brokken-Zijp, J. C. M., Loos, J., de With, G.;
Polymer 2002, 43, 6169.
2. Oliver, W. C., Pharr, G. M.; J. Mater. Res. 1992, 7, 1564.
3. Li, Z., Brokken-Zijp, J. C. M., de With, G.; Polymer 2004, 45, 5403.
4. Percy, E. J. GB 1120301
5. Bearden, C. R. US 3215731
6. FR 1485764
7. Takayama, Y. JP 44002685
8. Ito, H. JP 43026606
9. Caldwell, J. R. US 2484487
10. Simpson, A. J. GB 998394
11. Hamamoto, Y. JP 45017662
12. Kucharski, M., Lubczak, R.; J. Chem. Technol. Biot. 1998, 72, 117.
13. Shechter, L., Wynstra, J.; J. Ind. Eng. Chem. 1956, 48, 86.
14. Kakiuchi, H., Tanaka, Y.; J. Org. Chem. 1966, 31, 1559.
15. Madec, P. J., Marechal, E.; Makromol. Chem. 1983, 184, 323.
16. Matejka, L., Pokorny, S., Dusek, K.; Polym. Bull. 1982, 7, 123.
17. Matejka, L., Lovy, J., Pokorny, S., Bouchal, K., Dusek, K.; J. Polym. Sci. Pol.
Chem. Ed. 1983, 21, 2873.
18. Fedoseev, M., Gurina, M., Sdobnov, V., Kondyurin, A.; J. Raman Spectrosc.
1996, 27, 413.
19. Dossin, L. M., Graessley, W. W.; Macromolecules 1979, 12, 123.
20. Pearson, D. S., Graessley, W. W.; Macromolecules 1980, 13, 1001.
21. Langley, N. R., Polmanteer, K. E.; Journal of Polymer Science, Polymer Physics
Edition 1974, 12, 1023.

124

Networks applications: synthesis and properties

22. Dusek, K., Duskova-Smrckova, M., Lewin L.A., Huybrechts J., Barsotti R.J.; 27th
FATIPEC Congress, Aix-en-Provence 2004, Congress Proceedings, Vol 1 2004,
209.
23. Porter, D.; Group interaction modeling of polymer properties. New York: 1995
24. Litvinov, V. M., Dias, A. A.; Macromolecules 2001, 34, 4051.
25. Litvinov, V. M.; Macromolecules 2001, 34, 8468.
26. Fry, C. G., Lind, A. C.; Macromolecules 1988, 21, 1292.

125

Chapter 6

126

CHAPTER 7
CONCLUSION AND RECOMMENDATIONS
I) Conclusion
The rapid development in the past few decades of living controlled radical polymerization
allowed the possibility to synthesize well-defined polymers in a quite straightforward manner.
The molecular weights can easily be controlled, low polydispersities obtained and
functionalization of the polymer chain is not much of a problem anymore. One question arises:
is the synthesis of such well-defined structures only an academic scientific achievement or can it
leads to breakthrough findings in the field of material properties?
In order to determine the influence of the functionality type distribution and the molecular
weight distribution for polymers used as building block for the formation and properties of
poly(meth)acrylate networks, we have synthesized well-defined telechelic polyacrylates using
the RAFT polymerization technique, and random copolymer randomly introducing
functionalities in polyacrylate chains.
In this thesis, a chapter of introduction was written, in which an update on controlled radical
polymerization (with an emphasis on Nitroxide Mediated Polymerization, Atom Transfer
Radical Polymerization and Reversible Addition Fragmentation chain Transfer technique) and a
literature review on available methods to obtain telechelic polymers were given.
The first step towards the synthesis of well-defined networks concerned the synthesis of linear
end-functionalized polymers with a controlled molecular weight and a low polydispersity. The
RAFT polymerization technique was employed to obtain hydroxy end-functionalized poly
(methyl methacrylate). A two-step post-polymerization procedure was proposed in order to
modify the thiocarbonyl thio end-group, transforming it successively into a thiol end-group (via
aminolysis) and a hydroxyl end-function (via a Michael addition). Liquid-chromatography
separations proved the successful synthesis of telechelic PMMA to a significant extent.
However, attempts to quantitatively synthesize telechelic PMMA proved to be unsuccessful. We

127

demonstrated that formation of disulfide functions via oxidative coupling during the aminolysis
and hydrogen-transfer to the solvent were the main reasons lowering the final yield of ,poly(methyl methacrylate). MALDI-TOF-MS and two-dimensional LC-SEC separations
confirm these findings.
The selection of butyl acrylate as a monomer and trithiocarbonates as RAFT agents resulted in
the straightforward synthesis of well-defined carboxylic acid-functional poly (butyl acrylate).
Liquid-chromatography separations proved the nearly quantitative production of mono- and difunctional poly(butyl acrylate) chains, those compounds being suitable for network formation
studies.
The end-group analysis by MALDI led to interesting observations on fragmentation
phenomena during the analysis of thiocarbonyl thio-containing poly(meth)acrylate. The
importance of the monomer was proven, as well as the effect of the stabilizing group of the
RAFT agent in the fragmentation probability. The effect of the laser intensity on the
fragmentation was investigated as well.
In the final part of this thesis, the mechanism for the reaction of the carboxylic acid groups of
the polymers with a tri-functional epoxy-crosslinker is described. The cure reaction was studied
with FT-IR and a new mechanism is proposed to explain the experimental data. The network
structure and several properties of the cured networks were then determined by microindentation, DSC and solid-state NMR. Comparison between networks based on the telechelic
polymers that have a narrow functionality distribution and a narrow molecular weight
distribution, and random copolymers with a broad functionality distribution and a broad
molecular weight distribution was made. It was found that the presence of dangling ends has an
effect on the properties. Even only a few percent of dangling chains in a network structure
lowers the elastic moduli of the cured coatings. However, the main factor influencing the elastic
modulus and the glass transition parameters seems to be the molecular weight between crosslinks. A linear relationship between those two parameters was found for telechelic polymers.
The elastic modulus values for coatings based on telechelic polymers and prepared with high
[epoxy]: [COOH] values (above 2:1) suggest the presence of local domains of high elastic
moduli, formed by the autopolymerization of TGIC.

128

II) Recommendations
The production of telechelic poly (butyl acrylate) was relatively easy once trithiocarbonates
were used as RAFT agents. The paper describing the synthesis of those trithiocarbonates and
their successful application as controlling agents for the polymerization of butyl acrylate was
published one and a half year after the start of this project. Chapter 3 relates the struggles that
were encountered trying to use thiocarbonyl thio compounds as RAFT agents in order to obtain
telechelic polymers. Contrary to what was claimed at this time in the literature, the aminolysis
reaction is far from being quantitative. We evidenced at least two side-reactions occurring at the
same time as the reduction of thiocarbonyl thio functions into thiols: formations of disulfide
bridges, and hydrogen-abstraction from the solvent. If end-group functionalities are of primary
importance for polymer chains, we advice to avoid any manipulation on the polymer after the
RAFT polymerization. The use of di-functional RAFT agent, with functionality on both the
leaving group and the activating group is encouraged in order to obtain telechelic polymers after
one step. The trithiocarbonates are a specific class of such compounds, as the activating and
leaving group are degenerated.
Numerous claims in the literature concerning functionalities of telechelic polymers suffer
from a general lack of quantitative analytical work. The structures are often confirmed by mass
spectrometry analysis, without the required quantitative aspect of the characterization. We have
demonstrated in Chapter 3 and 4 that the combination of synthetic and analytical works are
required to seriously confirm the synthesis of telechelic chains. In order to complete the
analytical work, more work will be performed on the two-dimensional LC-SEC separation of
the telechelic poly (butyl acrylate) in order to quantify the amounts of the different populations
detected (Chapter 4).
The time spent on the synthesis and characterization of telechelic chains was lacking at the
end of this project, when the comparison between telechelic polymers and random copolymers
as building blocks for networks was made. Several points could not to be investigated due to
time constraints:
- Influence of the molecular weight on the E modulus: following the synthetic path
presented in Chapter 4, telechelic polymers with various molecular weights could easily
be synthesized. The determination of the elastic modulus by micro-indentation could bring
information of the influence of the cross-link density on the mechanical properties.

129

Comparison of experimental values with the ones obtained by various theoretical models
could then be made.
- Nature of the cross-linking agent: after synthesizing carboxylic acid functional polymers,
the choice was made to use the epoxy chemistry in order to obtain tri-dimensional
networks. The TGIC cross-linker was up to our knowledge the tri-functional epoxy
compound with the most documentation in the literature. Unfortunately the analysis
performed on it revealed the presence of impurities, making the investigation on the crosslinking itself and on the properties of the final coatings impossible to be quantitative. The
amount of epoxy group can be determined, but the functionality distribution was
unknown. Two choices in order to improve this study can be proposed: either the TGIC
can be purified by preparative HPLC for example, or the epoxy COOH chemistry has to
be replaced by another one. That can imply a change of cross-linking agent, or the
modification of the COOH-functionality of the polymer end-groups. As we wrote it in the
previous paragraph, this modification should preferentially be performed on the RAFT
agent itself before the polymerization, rather than after it.
- Mechanism of the carboxylic acid epoxy reaction, catalyzed with tertiary amines: the
mechanism we have proposed in chapter 6 is probably over-simplified, although it should
be sufficient to describe the cross-linking reaction with an acceptable agreement with
experimental data. Reactions at fixed [epoxy] : [COOH] ratios and variable DABCO
amount should be performed in order to have a look at the catalyst concentration
influence. In the same way, curing formulations with a fixed amount of DABCO and
different [epoxy] : [COOH] ratios should be prepared. These experiments should validate
the mechanism we have proposed and possibly could gain access to the kinetic constants.
- Solid-state NMR: networks with formulations containing high amounts of TGIC should
be looked at with T1-and T1-relaxations techniques, in order to prove the presence of hard
clusters in the networks, formed by the autopolymerization of the TGIC. We proposed this
theory in order to explain the amazingly high values obtained for the elastic modulus of
coatings based on telechelic poly (butyl acrylate) when high TGIC contents were used in
the preparation.

130

Summary
The research described in this thesis is a part of project #205 of the DPI (Dutch Polymer
Institute). One of the challenges was to synthesize telechelic poly(meth)acrylayes with welldefined masses and low polydispersities, in order to evaluate the potential benefit that can be
obtain in the coating formulations by using such polymers. Complete characterization of the
synthesized polymers was done, via several analytical techniques. The second part of the project
concerned the comparison between the newly-designed polymers and the more classical random
copolymers, both types of polymers being used as building blocks for three-dimensional
networks.
The RAFT (Reversible Addition-Fragmentation chain Transfer) polymerization technique has
been chosen for the production of polymer chains with predicable molar masses and low
polydispersities. A hydroxyl-functional thiocarbonyl thio RAFT agent was initially used. In
order to obtain an hydoxy functionality at both ends of a linear polymethacylate, a new two-step
post-polymerization procedure, involving an aminolysis of the thiocarbonyl thio group of the
RAFT polymer, followed by a Michael addition on the resulting thiol was proposed. Contrary to
what was claimed in the literature at the time the research was conducted, the aminolysis
reaction was found to be prone to side-reactions that lowered its yield. We evidenced transfer to
solvent and oxidative coupling of thiols as side reactions by using MALDI-TOF-MS and liquid
chromatography under critical conditions as analytical techniques. A polymer batch containing
66% of telechelic OH-functional chains was produced, following the proposed method, but
attempts to increase this percentage were unsuccessful.
Using symmetrical COOH-functional trithiocarbonates as RAFT agents, it was possible to
synthesize telechelic poly(butyl acrylate) in one step. The molar masses of these polymers were
predictable and low polydispersity indexes were obtained. MALDI-TOF-MS and liquid
chromatography under critical conditions confirmed the structure of the polymer, and
quantitative results proved that some batches of poly(butyl acrylate) synthesized via this route
contain more than 99% of bi-functional chains. Considering the high purity of those polymers in
terms of functionality, those polymers were chosen as building blocks for the network formation
and networks properties studies.
The MALDI-TOF-MS analysis of the RAFT polymers pointed out the tendency for RAFT
polymers to fragment under laser irradiation. It was found that three main parameters influence
the fragmentation probability of a RAFT polymer chain during a MALDI analysis:
-

the laser intensity

the type of RAFT agent used during the synthesis (thiocarbonyl thio-containing
compounds are more prone to fragment than the trithiocarbonate-containing ones)

the chemical nature of the backbone (polymethacrylates fragment more easily than
polyacrylates)

Finally, a comparison between telechelic polymers and random copolymers as building blocks
for networks was made. The chemistry used for the cross-linking reaction was the carboxylic
acid epoxy reaction, catalyzed by tertiary amine. A mechanism was proposed to describe the
reaction. FT-IR measurements prove the existence of a intermediate involving the 3 components
(COOH of the polymer cross-linker catalyst) during the reaction. A characteristic band at
1560 cm-1 was attributed to the C=O band of the COO- group of an intermediate in the proposed
mechanism. The telechelic polymers reacted faster than their random equivalent.
The elastic moduli of the networks were determined using micro-indentation technique.
Contrary to what could be expected from a theoretical model, the telechelic-based coatings
exhibited higher elastic modulus values than the random co-polymers-based ones. A linear
relation between the elastic modulus and the glass transition temperature of the networks was
established for networks based on telechelic polymers.
Solid-state NMR T1relaxation experiments showed that at low [epoxy} : [COOH] ratio in
the formulation of telechelic chains-based polymers, the network obtained after cross-linking
was homogeneous at nanometer scale. T2relaxation experiments confirmed the conclusions
obtained from micro-indentation, showing that the molar mass between cross-kinks is the major
parameter influencing the elastic modulus, in the case of telechelic chains-based networks.

Samenvatting
Het onderzoek, uitgevoerd in het kader van dit proefschrift, was een onderdeel van project 205
van het Dutch Polymer Institute. Een van de uitdagingen van dit onderzoek was de synthese van
lineaire (meth)acrylaat polymeren met een goedgekarakteriseerde eindstandige reactieve
groepen

(functionaliteit

2),

een

voorspelbaar

moleculairgewicht

en

een

nauwe

moleculairgewichtsdistributie. Een tweede uitdaging van het onderzoek was om deze polymeren
te gebruiken als uitgangsmateriaal voor het maken van doorgeharde coatings en een aantal
eigenschappen van deze coatings te bepalen. Een derde uitdaging van dit onderzoek was om
deze coatings te vergelijken met coatings die gemaakt waren van meer klassieke polymeren met
een zeer vergelijkbaar gemiddelde moleculair gewicht en reactieve groep functionaliteit maar
met een veel bredere moleculair gewicht distributie en een veel bredere en hangende
functionaliteit van de reactieve groepen over de ketens. Het doel van dit deel was om na te gaan
in hoeverre het gebruik van de hierboven beschreven eindstandig functionele polymeren
aantrekkelijker eigenschappen aan de doorgeharde coatings geven door de aanwezigheid van
een ander driedimensioneel netwerkstructuur.
Voor het synthetiseren van polymeer ketens met een voorspelbare moleculair gewicht en een
smalle moleculair gewichtsdistributie werd gebruik gemaakt van de RAFT (Reversible
Addition-Fragmentation chain Transfer) polymerisatiemethode. Initieel werd in dit onderzoek
een hydroxygefunctionaliseerde dithioester RAFT agens gebruikt. Om een hydroxyfunctionaliteit te verkrijgen aan beide uiteinden van een lineair polymethacrylaat, werd er een
nieuwe twee-staps synthese toegepast na de polymerisatie reactie. De eerste stap was een
aminolyse van de dithioester groep van het RAFT polymeer. In de tweede stap werd de
verkregen eindstandige thiol groep omgezet in een HO- groep (Michael addition). In
tegenstelling tot wat werd beweerd in de literatuur op het moment dat dit onderzoek werd
uitgevoerd traden er nevenreacties op tijdens de aminolyse, die de uiteindelijke de opbrengst
verlaagde. Het verkregen polymeer product bestond slechts voor 66% uit polymeren met aan
beide uiteinden een hydroxy groep. Door gebruik te maken van MALDI-TOF-MS en vloeistof
chromatografie onder kritische condities werd aangetoond dat oxidatieve koppeling van de thiol
groepen en de reactie met het oplosmiddel verantwoordelijk waren voor deze nevenreacties.
Pogingen om een zuiverder product te krijgen mislukten.

Door gebruik te maken van een symmetrisch trithiocarbonaat RAFT agens met aan beide
zijden een eindstandige COOH-functionaliteit werden er

in een stap poly(butyl)acrylaat

polymeren gemaakt met aan beide zijden een eindstandige COOH groep. De verkregen
moleculairgewichten van deze polymeren kwamen overeen met de berekende waarden.
Bovendien was de variatie in moleculairgewicht over de verschillende ketens zeer laag.
MALDI-TOF-MS en vloeistof chromatografie onder kritische condities bevestigden de structuur
van deze polymeren en kwantitatieve resultaten toonden aan dat sommige poly(butyl)acrylaten,
die langs deze weg verkregen waren, voor meer dan 99% bestonden uit polymeren met aan
beiden zijden een -COOH groep. Door deze hoge zuiverheid in functionaliteit bleken deze
polymeren geschikte uitgangsstoffen te zijn voor het maken van goed gedefinieerde
driedimensionele moleculaire netwerken en voor het bestuderen van de eigenschappen van de,
na doorharding, uit deze polymeren verkregen coatings.
Uit onze MALDI-TOF-MS analyse bleek dat de RAFT polymeren geneigd zijn te
fragmenteren onder invloed van het laserlicht dat gebruikt wordt tijdens de analyse. . De
volgende drie parameters benvloeden de waarschijnlijkheid van fragmentatie tijdens de analyse:
-

de licht intensiteit

het type RAFT agens dat gebruikt werd gedurende synthese van deze polymeren
(polymeren verkregen met behulp van dithiocarbonyl agentia zijn meer gevoelig voor
fragmentatie dan trithiocarbonyl verbindingen).

De chemische samenstelling van de polymeer keten (polymethacrylaten fragmenteren


makkelijker dan polyacrylaten)

Tenslotte werden de goed gekarakteriseerde en gedefinieerde lineaire poly(butyl)acrylaten met


COOH eindgroepen die hierboven beschreven zijn doorgehard met behulp van een trifunctionele
epoxydoorharder

onder

invloed

van

een

tertiaire

aminekatalysator.

Zowel

de

doorhardingsreactie als enkele materiaaleigenschappen van de verkregen coatings werden


bestudeerd. De verkregen resultaten werden vergeleken met de resultaten die verkregen werden
door in plaats van deze polymeren lineaire polybutyl copolymeren te gebruiken met eenzelfde
gemiddelde

functionaliteit

en

moleculair

gewicht

maar

met

een

veel

bredere

moleculairgewichtsdistributie en een veel bredere en hangende functionaliteit van de COOH


groepen over de polymeerketens. Deze laatste polymeren werden gesynthetiseerd door gebruik
te maken van standaard radicaal polymerisatietechnieken.

Voor de doorhardingsreactie van beide typen van polymeren werd eenzelfde nieuw
mechanisme voorgesteld. Metingen met behulp van FT-IR toonden het bestaan van een
intermediair complex aan tussen de eindgroep van het polymeer de doorharder en de katalysator.
Een karakteristieke absorptie bij 1560 cm-1 werd waargenomen voor dit intermediair. Deze band
werd toegewezen aan de COO- groep van het polymeer. De snelheid van de doorhardingsreactie
was langzamer voor de polymeren met de hangende COOH functionaliteit.
De elasticiteitsmoduli van de coatings gemaakt uit beide typen hierboven besproken
polymeren werden bepaald met behulp van microindentatie metingen. In tegenstelling tot wat
verwacht kon worden op basis van de theorie bleken de elasticiteitsmoduli lager uit te vallen
voor de coatings gemaakt uit polymeren met hangende functionaliteit. Alleen voor de coatings
gemaakt uit de polymeren met eindstandige COOH functionaliteit werd er een lineaire relatie
gevonden tussen de E-moduli en de Tg, zelfs als het percentage loseindige groepen in het
netwerk groot was.
Vaste stof NMR T1

relaxatie experimenten lieten zien dat de netwerkstructuur na

doorharding van de eindstandige COOH polymeren homogeen is op nanometer schaal wanneer


de [COOH]/[epoxy] gelijk is aan een. De vaste stof NMR T2 relaxatiemetingen bevestigden de
conclusies van de microindentatie metingen namelijk dat het moleculairgewicht tussen de
crosslinks de belangrijkste parameter is voor de bepaling van de grootte van de E-modulus van
de coatings gemaakt uit eindstandige COOH polymeren

Acknowledgments
Completing a Ph.D. thesis on its own is a utopia. During the four years spent in Eindhoven, I
was lucky to have great people on my side always friendly and ready to help me.
I deeply thank Prof. Rob van der Linde, my first promoter, for the opportunity he gave me to
start a Ph.D. research in his group. Rob, despite your retirement in 2002, you have always
followed closely my struggles and progresses during these years. Your energy and positive
attitude was always of great help when things seemed to go badly. I am extremely proud to have
you as a promoter.
My second promoter, Prof. Peter Schoenmakers is acknowledged for the great help he
provided. Peter, I have very much appreciated my frequent meetings with you in Eindhoven or
in Amsterdam. There were always a lot a fun and a lot of science. Your extreme attention to
every word in scientific discussions in particular made me more precise/accurate about my
speaking and my writing. Thanks a lot for everything.
I would like to thank my co-promoter Dr. Jos Brokken for the daily supervision she provided
me. Jos, your endless efforts to improve the quality of my various manuscripts resulted in
better papers and a better thesis. We confronted on numerous occasions our views on different
aspects of my research. The long discussions often gave me new ideas and extra motivation. I
have always had the feeling that if I could convince you, I could convince anyone.
I would like to thank all the members of my examination committee for their presence and
their valuable comments during the revision of this thesis.
The Dutch Polymer Institute is thanked for the financial support that was provided for this
project. I would like especially to thank the industrial contact people whose remarks were of
great help throughout this project.
During those 4 years, I met a lot of people who help me with analytical techniques that
seemed a little bit difficult to me at first sight. Xulin Jiang was the master of Liquid
Chromatography for me. All my synthetic work would have been useless without a proper
characterization. Xulin, thanks to your great work, we were able to show that the goals assigned
to both our projects were reached. It has been a pleasure to collaborate with you. After Xulins
project was over, Aschwin van der Horst was of great help for the characterization of COOHfunctional chains. Thank you Aschwin !
Wieb Kingma and Marion van Straten are thanked for their help with SEC and MALDI-TOFMS. Bastiaan Staal is also thanked for the amazing amount of time he spent helping me. Bas,
without your help, Chapter 5 would never have been as it stands now.

Chapter 6 essentially exists due to the help of five people. Without proper help, I would never
have been able to write it. Otto van Asselen, I am greatly grateful for all the time you have tried
to help me determining a proper mechanism for my cross-linking reaction, using FT-IR. Zhili
Li, your expertise in micro-indentation experiment combined with an extreme rapidity
producing results was a true relief when time was winding down. Prof. Guenther Hoehne is
thanked for his precious help with TMDSC measurements. Dr. Pieter Magusin and Brahim
Mezari, thanks a lot for the solid-state NMR experiments. That was a short but fruitful and
enjoyable collaboration.
Many thanks to Dr. Marille Wouters for all the rheology experiments performed at TNO.
Unfortunately, none of them could appear in this booklet, this is how science is sometimes.
I would also like to thank Patrick Terregino, whose help during three months resulted in the
finishing of the RAFT polymer synthesis. The VL103 sample you have synthesized Pat is the
central part of the network formation study. Thanks a lot!
The STO 1.28 office was a place of fun and scientific arguments through the years. Thanks to
Auke, Huiqi and Willem-Jan for having contributed to the nice atmosphere during working
hours. The SPC group has always considered me as one of them despite the groups shuffling
that happened at the beginning of my stay at the TU/e. I would like to thank particularly Wouter,
Rajan, Jelena, Jens, Raf, Ellen, Maarten and Robin for their friendship and good moments. Dr.
Bert Klumperman is also acknowledged for the guidance he provided me in the RAFT world.
The old SVM gang will always have a special place for me. Marshall, Willem, Okan, Chouab
Gabrille, Jos, Victor and Robert, it was a lot of fun during my stay in Eindhoven, and I hope we
will keep in touch in the future.
Despite living outside France, I have never felt lonely. Merci tous mes amis franais qui ont
toujours rpondu prsent quand le besoin sen faisait sentir. Julien, Laurent(s), Francis,
Michael(s), Emmanuel(s), Ronan, Stphane(s), Grald, Pierre et Alexandre, merci pour votre
amiti.
Merci galement ma famille et ma belle-famille pour tout le soutien ncessaire
laccomplissement dune thse ltranger. Merci tout particulirement mes parents qui mont
toujours donn la possibilit de suivre la voie que javais choisie. Sans lenvironnement dans
lequel jai t lev, mon cursus naurait pas pu tre ce quil a t.
Et au final, je remercie grandement Delphine pour tout ce quelle ma apport au quotidien
pendant ces 4 ans. Vivre ensemble et travailler ensemble nest pas toujours facile, mais ton
soutien tait indispensable pour la russite de cette thse. Le meilleur reste venir

Curriculum Vitae
Vincent Georges Robert Lima was born on the 9th of June 1977 in Noisy-le-sec, France. He
obtained in 1994 his high school diploma (Baccalaurat) in Lyce Gustave Monod in Enghienles-bains (France). He then graduated in 1999 from the Institut de Science et Technologie
Chimie des Matriaux (Universit Pierre et Marie Curie Paris, France) after completing an
graduation project in the North-American research center of ATOFINA in King of Prussia (PA,
USA), working on Synthesis of core-shell poly(acrylate)s impact modifiers in emulsion, using
the ATRP technique. After fulfilling his military obligations, he started in 2001 a Ph.D. project
under the guidance of Prof. Rob van der Linde at the Technological University of Eindhoven
(The Netherlands). This thesis is the results of 4 years of research. Since March 2005, he has
been working in the research department of Oc Technologie in Venlo (The Netherlands).

You might also like