Sie sind auf Seite 1von 49

Telechelic polymers derived

from natural resources as


building blocks for polymer
thermosets

Susana Torrón Timhagen

Licenciate thesis

KTH Royal Institute of Technology, Stockholm 2015


Department of Fiber and Polymer Technology
Copyright © 2015 Susana Torron
All rights reserved

Paper 1 © 2014 WILEY-VCH

TRITA-CHE Report 2015:13


ISSN 1654-1081
ISBN 978-91-7595-469-1
To my family
Abstract
The growing interest for replacing fossil-based materials with other
greener alternatives, leads to the search of feed stocks from sources
traditionally used for other purposes. Good examples of this are corn,
seeds or beans from which is possible to obtain diverse compounds
extensively used nowadays for the preparation of different polymeric
materials. One important source of raw materials is wood. Its main
component, cellulose, has been used for decades for the preparation of
materials with various applications. Another component of wood that
has gained a lot of interest due to its complex polyester structure is the
suberin present in the bark. Through an hydrolysis it is possible to cleave
those ester bonds and retrieve a large variety of epoxy fatty acids2.

In the present work a procedure for the isolation and purification of an


epoxy fatty acid proceeding from outer Birch bark3, the 9,10-epoxy-18-
hydroxydecanoic acid, is presented. The inbuilt epoxy-functionality of
this compound opens up a broad range of possibilities due to the
multiple chemistries it can be subjected to. Moreover, when functional
groups are introduced at the end of the polymer chain 4 we create a large
“tool box” with different options for network formation.
The polymerization of this monomer has been done using enzymatic
catalysis. With this not only is possible to preserve the epoxy group
intact during the whole process5, 6, but also to control the degree of
polymerization and introduce the end-functionalities in a “one-pot
synthesis”.
With an increased number of functional groups in the main chain, the
possibilities for network formation also increases. The different ways
these groups react with each other will then determine the properties of
the network. A control in the reactivity will then lead in the possibility of
material design which in terms of applications adds an additional value
to these biobased materials.
Sammanfattning
Det ökande intresset för att ersätta fossilbaserade material med andra
mer miljövänliga alternativ leder sökandet till nya råmaterial som
traditionellt sett används för andra ändamål. Några råmaterial som
redan används för framställning av olika polymera material är majs, frön
och bönor.

En viktig källa till råmaterial är träd. Dess huvudkomponent, cellulosa,


har använts i sekler för framställning av material för olika applikationer.
En annan komponent av träd som har fått ökad uppmärksamhet på
grund av dess komplexa polyesterstruktur är suberin som utvinns från
bark. Genom hydrolys är det möjligt att klyva esterbindningarna i
Suberin och separera ut olika fettsyror 2. Bland dem är 9,10-epoxi-18-
hydroxyoktadekansyra, som kan extraheras från björknäver3, av särskilt
intresse. Denna monomer med en inbyggd epoxifunktionalitet
representerar ett perfekt substrat för olika bindemedelstillämpningar.
Ett sätt att öka reaktiviteten hos 9,10-epoxi-18-hydroxyoktadekansyra är
att tillföra olika funktionella grupper i dess ändar, vilket resulterar i
"telekeliska polymerer". Dock är syntesen av telekeliska polymerer ofta
komplicerad med flera olika reaktionssteg där man ofta använder olika
skyddsgruppskemier. Ett sätt att kringgå detta problem är att använda
sig av en selektiv katalysator, som exempelvis ett enzym. Med hjälp av
dessa naturliga katalysatorer är det möjligt att införa olika
funktionaliteter och samtidigt bevara närvarande reaktiva grupper 5, dvs
epoxigruppen, samtidigt som polymerisationsgraden kan kontrolleras
genom reglering av de stökiometriska förhållandena.
Med ett ökat antal funktionella grupper i huvudkedjan, ökar också
möjligheterna till att bilda nätverk. De olika sätten dessa grupper
reagerar med varandra kommer då bestämma egenskaperna hos den
slutliga härdplasten. Alltså, genom att kontrollera reaktiviteten ges
möjligheten att skräddarsy den slutgiltiga produkten, vilket tillför ett
mervärde för dessa biobaserade material.
List of appended papers
This thesis is a summary of the following papers:

Paper I
“Polymer thermosets from multifunctional polyester resins based
on renewable monomers”, S. Torron, S. Semlitsch, M. Martinelle,
M. Johansson, Macromol. Chem. Phys. 2014, 215, 2198-2206.

Paper II
“Oxetane terminated telechelic epoxyfunctional-polyesters as
cationically polymerizable thermoset resins – tuning the reactivity
with structural design”, S. Torron and M. Johansson, Submitted.

Other publications that do not appear in this thesis:

“Telechelic polyesters and polycarbonates using enzyme


catalysis”, Susana Torron, Mats K G Johansson, Eva Malmström,
Linda Fogelström, Karl Hult, Mats Martinelle, Chapter 2 on
“Handbook on Telechelic Polyesters and Polycarbonates”, edited
by Pan Stanford Publishing, Submitted.
Abbreviations
ATR-FTIR Attenuated Total reflectance- Fourier Transform Infrared
CALB Lipase B from Candida Antarctica
CDCl3 Deuterated Chloroform
DA Dimethyl Adipate
DHB 2,5-dihydroxy benzoic acid
DMA Dynamic Mechanical Analysis
DP Degree of Polymerization
EGDMA Ethylene glycol Dimethacrylate
eROP Enzymatic Ring Opening Polymerization
HCl Hydrochloric acid
1H NMR Proton- Nuclear Magnetic Resonance
H2SO4 Sulphuric acid
LiBr Lithium Bromide
MALDI-ToF Matrix-Assisted Laser Desorption/Ionization- Time of Flight
MS Mass Spectrometry
Mw Molecular Weight
NaOH Sodium Hydroxide
PI Photoiniator
PMA Poly(methyl acrylate)
PMMA Poly(methyl methacrylate)
ROP Ring Opening Polymerization
RT-FTIR Real Time- Fourier Transform Infrared
SEC Size Exclusion Chromatography
Tg Glass Transition temperature
THF Tetrahydrofuran
TMPO Trimethylpropyl oxetane
UV Ultraviolet

Sample Code:
EFA Epoxy Fatty Acid (9,10-epoxy-18-hydroxydecanoic acid)
EFA-EGMA Epoxy Fatty Acid end-capped with Ethylene Glycol Methacrylate
EDP3 Epoxy-functional oligomer with Degree of Polymerization 3
ODP3 Oxetane-functional oligomer with Degree of Polymerization 3
EODP3 Epoxy/oxetane oligomer with Degree of Polymerization 3
EODP6 Epoxy/oxetane oligomer with Degree of Polymerization 6
Table of content
1. Purpose of the study………………………………..............…1
2. Introduction……………………………………………………….…2
2.1. Background…………………………………………….………..2
2.2. Monomers from renewable resources………..…..…2
2.2.1. Suberin……………………………………………………….…3
2.3. Polymers by enzymatic catalysis………………….....…5
2.3.1. Telechelic polymers by enzymatic catalysis………….…….7
2.4.Thermoset formation…………………………...………..…8
2.4.1. Photopolymerization………………………..…………..……9
Free-radical polymerization………………………..………9
Cationic ring opening polymerization/ Cationic UV-
curing…………………………………………................…11
3. Experimental………………………………………..………….…12
3.1. Materials…………………………………………………………12
3.2. Extraction of 9,10-epoxy-18-hydroxyoctadecanoic
acid from outer birch bark……………………….……..12
3.3. Synthesis of telechelic oligomers by enzyme
catalyzed esterification …………………………...........13
3.3.1. Synthesis of ,ω-methacrylate end-capped
epoxyfunctional oligomers (EFA-EGMA)…………….…..13
3.3.2. Synthesis of ,ω-oxetane end-capped epoxyfunctional
polyester DP3 (EODP3) and DP6 (EODP6)…………..….13
3.3.3. Synthesis of ,ω-hexane end-capped epoxyfunctional
polyester DP3 (EDP3)……………………………………....14
3.3.4. Synthesis of ,ω-oxetane end-capped aliphatic polyester
DP3 (ODP3)……………………………………………….….14
3.3.5. Synthesis of Dimethyl hexadecanedioate………….…..…14
3.4. Thermoset formation…………………………….......…..15
3.4.1. Radical polymerization of methacrylates…………………15
3.4.2. Cationic polymerization……………………………….…... 15
3.4.3. Dual curing……………………………………………………15
3.5. Analytical methods…………………………………...…....16
3.5.1. 1H NMR……………………………………..………………..16
3.5.2. MALDI-TOF………………………………………….…….…16
3.5.3. ATR-FTIR………………………………………………….….16
3.5.4. RT-FTIR…………………………………………………....…16
3.5.5. SEC…………………………………………………………..…17
3.5.6. DMA……………………………………………………………17
3.5.7. UV-Light sources…………………………………….…….…17
4. Results and discussion………………………………….……18
4.1. Extraction of 9,10-epoxy-18-hydroxyoctadecanoic
acid (EFA) from outer birch bark…………………....18
4.2.Synthesis of telechelic oligomers by enzyme
catalyzed esterification………………………………...…18
4.2.1. 1H NMR ……………………………………..…………….…21
4.2.2. MALDI-TOF MS…………………………………….…….…22
4.2.3. SEC………………………………………………………….…23
4.3. Thermoset formation………………………..………….…23
4.3.1. Polymerization performance. Fourier Transform Infrared
and Photo-Real Time Infrared Analysis…….…25
Epoxide/ methacrylate systems. Cationic polymerization
vs radical polymerization………………………………..…25
Epoxide/ oxetane systems. Homopolymerization vs
copolymerization in cationic polymerization ……….….26
4.3.2. Mechanical properties of the formed networks……….…29
5. Conclusions…………………………………………………………32
6. Future work……………………………………………..…………34
7. Acknowledgements……………………………………….……35
8. References……………………………………………………..……36
Purpose of study

1. Purpose of the study


The search of natural based raw materials as mile stones for material
formation together with the examination of new greener chemical routes
for their preparation, center nowadays the research of many material
scientists. In a county like Sweden where around 23 million hectares are
covered by forest, that trend is inevitably conducted to the search of raw
materials derived from wood.
The overall purpose of this work is to utilize a monomer derived from
outer birch bark, the 9,10-epoxy-18-hydroxydecanoic acid, to prepare
polymeric materials.

This thesis covers some especific goals as:

extraction and purification of a monomer proceeding from


outer birch bark,

definition of a synthetic route that will allow the formation of


various oligomers end-capped with different functional groups.
Focus on enzyme catalyzed esterifications,

preparation of polymer thermosets using different strategies,

analysis of the structural properties of the networks by mean of


their mechanical properties

1
Introduction

2. Introduction

2.1. Background
Traditionally, synthetic polymeric materials have been prepared using
raw materials obtained from fossil sources. However, their poor
degradability, high cost and inevitable extinction, have pushed the trend
towards the search of more environmentally friendly alternatives. In
these terms, the use of compounds obtained from e.g. annual crops,
wood or natural oils are seen as promising alternatives 7,8, 9.
The production of materials derived from natural resources has been in
continued development for many centuries. The first documented uses
of natural based materials are dated in 1600 BC, when the Mayan
civilization in Central America is assumed to be the first to use rubber
proceeding from the rubber tree to fabricate balls for their kids. Later in
1839, the discovery of vulcanization of rubber by Charles Goodyear
increased the uses of this raw material as the durability highly increases.
Contemporary to Goodyear, Braconnot, along with Christian Schönbein
and others, developed derivatives of the natural polymer cellulose,
producing new, semi-synthetic materials, such as celluloid and cellulose
acetate. The introduction of fossil resources during the 1900s took over
as a resource for synthetic polymers. The foreseen depletion of these
resources has however during the last decades reawakened the interest
for biobased resources. Nowadays, it is possible to see how the so called
“bioplastics” derived e.g. from corn starch, are implemented in the
market and produced in large scale10.

2.2. Monomers and polymers from renewable


resources

Depending on the natural source they proceed from, monomers and


polymers are generally divided into two groups: those proceeding from
vegetal biomass, as i.e. cellulose, terpenes, plant oils or sugars and those
proceeding from animal biomass, as i.e. chitin/ chitosan or casein.
Among the vegetal sources, wood is considered to be one of the most
important sources of raw materials due to its rich and variable

2
Introduction

composition. Wood comprises mainly three components: cellulose,


hemicellulose and lignin. These components have been exploited for
decades for the preparation of diverse biomaterials. One of the
components of wood that have not been exploited as much as the others
is the bark. However, it has been shown that the bark of some trees as
e.g. birch tree, is also an important source of raw materials11. Moreover,
it should be taken into consideration that outer birch bark has
traditionally been used for the only purpose of energy recovery 12 and
therefore finding an alternative use to it is of great interest.

2.2.1. Suberin

Suberin comprises 40 wt% of outer bark. Suberin acts as a hydrophobic


barrier to control the movement of water, gases, and solutes as well as
an antimicrobial barrier. Suberin consists of a crosslinked polymer
comprising aromatic and aliphatic components 12, 13 (respectively
primary cell wall and suberin lamellae on Figure 1). The main
components of the aliphatic domain are ω-hydroxy fatty acids, ,ω-
dicarboxylic acids and mid-chain di-hydroxy and epoxy homologous.
The primary cell wall also comprises of several triterpenoids and
aromatic structures.

3
Introduction

Figure 1. Model for the macromolecular structure of suberin13. The spheres


represent the points of union between the fatty acids and rest of the suberin.

Through a mild alkaline hydrolysis it is possible to depolymerize the


suberin to extract a large variety of long carbonated chain fatty acids. It
has been shown that outer bark of the birch tree is capable to undergo
this process yielding in a large variety of fatty acids2, 3, 14 (see Figure 2).

4
Introduction

Figure 2. Fatty acids present in majority in birch bark suberin

Among those fatty acids, the 9,10-epoxy-18-hydroxydecanoic acid is the


most abundant (100g per kg of bark). The presence of a carboxylic acid
together with an alcohol as end functional groups in the main chain,
make this monomer a suitable candidate for synthesis of polyesters. At
the same time, its intrinsic epoxy group opens a broad range of
possibilities for network formation or post-functionalization. However,
the reaction conditions usually needed for the synthesis of polyesters
might not always be suitable for the prevalence of the epoxide and
therefore new synthetic routes might be taken into consideration. In this
context, enzyme catalysed polyesterification is seen as a good alternative
to the “traditional” synthetic routes.

2.3. Polymers by enzymatic catalysis

At the same time as the interest for finding greener alternatives to the
traditional fossil based raw materials increases, increases the awareness
for trying to find greener solutions to the traditional catalytic methods.
The search for more environmentally friendly systems that would reduce
side products and avoid the use of harsh reaction conditions is, in this
context, of great interest15. In these terms, enzymes are anticipated as

5
Introduction

suitable candidates for this purpose16-21. Due to their high selectivity,


tedious protective/ deprotective steps are usually avoided. This
selectivity allows e.g. to form epoxy- functional polyesters from epoxy-
functional ω-hydroxy fatty acids keeping intact the epoxy group through
all the process1, 6, 22. This selectivity is consequence of the specificity of
the enzymes; every enzyme is responsible to catalyze a specific reaction.
One type of enzymes that have been widely used in polymer synthesis
are hydrolases. Hydrolases are responsible to catalyze the cleavage of
bonds by hydrolysis reactions, which then will allow the formation of
polyesters by condensation reactions23-26. One specific group of
hydrolases that has been of great importance in polymer synthesis are
lipases25, 27, 28. In order to facilitate the removal of the enzyme and
terminate the reaction, commercial enzymes are usually immobilized in
polymeric resins. Lipase B from Candida antarctica (CALB) is the
lipase that has been most extensively used for polyester synthesis and
commercial preparations are available, e.g. Novozyme 435.

In general terms, the use of enzymes as catalysts in polyester


synthesis offers many advantages:

- In a first term, as all catalysts, their use help to lower the activation
energy required to reach a certain transition state and thus increase
the reaction rate.
- Secondly, their specificity allows milder the reaction conditions
since less energy is required to break or form certain chemical
bonds. Milder reaction conditions also help to avoid the formation of
side products and backbone scission.
- Possibility to perform reactions in bulk, minimizing the use of
solvents.
- Possibility of tailoring the size of the polymer by introducing an end-
capping group. This point will be further discussed in point 2.3.1.
Telechelic polymers by enzymatic catalysis.

However, the use of lipases in polyesterification reactions is not always


trivial and some factors need to be controlled:

- In order to push the equilibrium towards product formation, the


water formed as biproduct needs to be removed. However, enzymes
require a small amount of water to retain their catalytic activity,
therefore is importance to balance the water content.

6
Introduction

- Lipase B from Candida antarctica (CALB) is thermally stable and


retain its activity for hours, however, it exhibit an upper temperature
limit29 around 90oC and therefore its use is limited to certain
monomers.
- Although preferable, the use of bulk conditions increases
dramatically the viscosity of the system. High viscosity limits the
diffusion of the reactants to the active site of the enzyme and the
evaporation of smaller molecular weight compounds. On the other
hand, the use of solvents favors intramolecular cyclations. These
factors restrict the obtainment of high molecular weight polyesters
through enzymatic catalysis. However it is possible to aim for high
molecular weight polyesters through the synthesis of telechelic
polymers.

2.3.1. Telechelic polymers by enzymatic catalysis

By definition, telechelic polymers are “prepolymers capable of entering


into further polymerization or other reactions through its reactive end-
groups”30. The possibility of further reactivity allows the obtainment of
high molecular weight linear polymers and polymer thermosets. The
introduction of an end-functional group allows also to tailor the length
of the polymer chain since the desired functional group will act as a
“stopper” in the propagation of the chain.
The synthesis of ,ω-functionalized telechelic polymers by enzyme
catalyzed reactions31, 32 is usually achieved by: Ring opening
polymerization (eROP)33-38 and polycondensation20, 39, 40. In eROP, the
-functionality is usually introduced by a functional initiator and the ω-
functionality by the terminator (see Scheme 1).

Scheme 1. Synthesis of ,ω-telechelic polyesters through enzymatic ring opening


polymerization.

7
Introduction

In polycondensation, the ,ω-functionalities are introduced by a


compound called “end-capper” that stops the condensation reaction by
reacting with the ends of the polymer chain (see Scheme 2).

Scheme 2. Synthesis of ,ω-telechelic polyesters through polycondensation

The key feature of enzyme catalyzed synthesis of telechelic polymers is


mild reaction conditions and selectivity that allow a large variety of
different end-groups to be incorporated without severe effects of side
reactions. These conditions render it possible to incorporate e.g.,
hydroxyl- or carboxylic acids as end-cappers in polyester backbones as
well as other functional groups e.g. acrylates/methacrylates 32, 34, thiols34,
35 5, epoxides38, or to conveniently produce epoxy-functional polyesters

directly from epoxy-functional ω-hydroxy fatty acids1, 22.

2.4. Thermoset formation


Through the formation of polymer thermosets it is possible to create
materials with higher mechanical and thermal properties than the
precursor polymer. Contrary to thermoplastics or gels, where the
networks are formed by physical entanglement of the polymer chains, in
thermosets the polymer chains are crosslinked, i.e. linked by covalent
bonds. This difference in their formation is also translated in a
difference in their properties, while thermosets cannot be dissolved or
melted, thermoplastics can41.

In order to form polymer thermosets it is necessary that the prepolymer


possess two or more functional groups capable to chemically react with
each other or with another compound called “crosslinker”. The more
reactive sites present in the polymer, the more possibilities for network

8
Introduction

formation, at the same time as high crosslinked networks will lead to


stronger materials. In this context, the election of the functional groups
present in the polymer and the way they react with each other will be
determining factors in terms of material design.

In general terms, polymerization reactions are triggered by heat or UV-


light. However, polyesters can undergo degradation at certain
temperatures42, therefore only photopolymerization reactions have been
considered in this work. Moreover photopolymerizations are considered
to be the easiest and fastest way to transform a liquid monomer or melt
into a crosslinked polymer network at room temperature 43, 44, which
increases the possibilities for industrial applications.

2.4.1. Photopolymerization

To the greatest extend, three components are needed to carry out a


photopolymerization45, 46 reaction: photoinitiator, UV-light and a
reactive monomer or oligomer. Photopolymerizations involves the
propagation of an active center by interaction with the monomer
through a chain wise mechanism. Among these, the photoinitiator plays
a crucial role since it will determine how the polymerization is
conducted. Upon excitation triggered by UV-light, the photoinitiator
dissociates to form either radicals or ions. However, it is the activated
monomer after reaction with the photoinitiator who truly initiates the
polymerization.

Attending to the scope of this thesis, only radical and cationic ring
opening polymerization will be considered.

Free-radical photopolymerization:

In general terms, radical polymerization follows a typical mechanism:

9
Introduction

In the first step (1), the photoinitiator (PI) forms a radical by


dissociation triggered by UV-light. This radical will then react with the
monomer to form another radical (2) that will then initiate the
propagation reaction that will yield in the formation of the polymer (3).
The termination of the polymerization is usually caused by combination
of two radicals or chain transfer.
Radical polymerizations are usually fast and yield in high molecular
weight polymers or highly crosslinked networks, however, except from
thiol-ene reactions47, they exhibit oxygen inhibition and inert
atmospheres are commonly required48, 49.
Acrylates and methacrylates are very reactive species towards radical
polymerization. Both species are vinyl esters which results in the
formation of very stable radicals due to conjugation with the ester
group50. This radical is more stable in methacrylates than in acrylates,
since it is formed in a tertiary carbon (see Scheme 3). This causes that
the propagation reaction in methacrylates is slower than in acrylates 51.

Scheme 3. Radical polymerization of methyl methacrylate and methyl acrylate.

The extra methyl group present in methacrylates also confers rigidity to


the polymer and while poly(methyl acrylate) (PMA) is soft and rubbery,
poly(methyl methcrylate) (PMMA) is a hard plastic.

10
Introduction

Cationic ring opening polymerization/ Cationic UV-curing:

Cationic UV-curing is one of the polymerization processes most


extended in industry52. The term “curing” is usually used in material
engineering to refer to the toughening or hardening of a polymeric
material by cross-linking of polymer chains53.
Some of the main advantages of cationic UV-curing are the lack of
oxygen inhibition and the excellent adhesion properties that some of the
compounds present when cured, e.g. epoxy resins.

As in the free radical photopolymerization, in cationic ring opening


photopolymerization, the reaction starts with the cleavage of the
initiator to form a protonic acid. Depending on how strong the formed
acid is, the ring opening will be caused by nucleophilic attack of the
oxerane to the acid or vice versa. The most common photoinitiators
used in cationic ring opening of cyclic ethers consist on diarylidonium 54
or triarylsulfonium55 salts. These compounds generate very strong
protonic acids by photolysis which will then triggered the formation of
an oxiranium ion that will then initiate the cationic polymerization. In
Scheme 4 is depicted the general reaction process for the cationic ring
opening of epoxides.

Scheme 4. Cationic ring opening polymerization of epoxides.

When epoxy-functional polymers are used as starting materials of the


cationic polymerization, the reaction yields in polymer networks.

11
Experimental

3. Experimental

3.1. Materials

All chemicals used are shown in Figure 3. 9,10-epoxy-18-


hydroxyoctadecanoic acid (EFA) was extracted from outer birch bark
(Betula Pendula) supplied by Holmen AB (c/o Holmen Energi, SE-
89810). The outer bark was used as given and finely ground using a ZM
200 (Retsch) grinder. Ethylenglycol methacrylate, dimethyl adipate,
hexanol, hexadecanedioic acid (99%), ethylene glycol and immobilized
CalB (Novozyme 435) were purchased from Sigma-Aldrich.
Trimethylpropyl oxetane (TMPO) was used as received from Perstorp
AB. Photoinitiators Uvacure 1600 and Irgacure 651 were supplied by
CYTEC and BASF respectively.

Figure 3. List of chemical compounds used in this thesis.

3.2. Extraction of 9,10-epoxy-18-


hydroxyoctadecanoic acid from outer
birch bark

100 mL of a solution 0.8M NaOH were poured into a round bottom flask
containing 10g of grinded, dried, outer birch bark. In order to
impregnate the bark with the solution, the flask was connected to a
water pump, under stirring for 10 min. After that, the mixture was

12
Experimental

refluxed for 1 hour, cooled to room temperature and the solid residue
was removed with the aid of a centrifuge. The resulting solution was
acidified to pH~6 with a solution 5% v/v H2SO4 and left in the fridge
overnight. The resulting precipitate was isolated by centrifugation and
recrystallized from toluene to yield a light-yellow powder (Yield: 60%).

3.3. Synthesis of telechelic oligomers by


enzyme catalyzed esterification
3.3.1. Synthesis of ,ω-methacrylate end-capped epoxyfunctional
oligomers (EFA-EGMA)

9,10-epoxy-18-hydroxyoctadecanoic acid (EFA) and ethylene glycol


dimethacrylate (EGDMA) were introduced in a round bottom flask using
a 3:1 (EFA monomer: dimethacrylate) ratio. A small amount of toluene
(500 wt% relative to the monomer) was used to dissolve the reactants.
Molecular sieves 4Å (25 wt% of the total amount of monomers) and
enzyme (20 wt% of the total amount of monomers) were then added to
the solution and the mixture was heated up to 60 oC for 24h. The catalyst
was then removed by filtration and the product used without further
purification.

3.3.2. Synthesis of ,ω-oxetane end-capped epoxyfunctional


polyester DP3 (EODP3) and DP6 (EODP6)

EFA monomer, dimethyl adipate (DA) and TMPO were put into a 25 mL
round bottom flask in a, respectively, 3:1:2 ratio for DP3 and 6:1:2 for
DP6, and heated up until they had reached the molten state at around
85oC. When the mixture had reached that temperature, CALB was added
(10% of the total amount of monomers) and the reaction was allowed to
proceed at open atmosphere for 2h and afterwards allowed to run under
vacuum overnight to remove low molecular weight compounds.
The reaction was stopped by dissolving the product in chloroform and
filtering off the immobilized enzyme. The product was characterized and
used without further purification (Yield EODP3: 94%, yield EODP6:
65%).

13
Experimental

3.3.3. Synthesis of ,ω-hexane end-capped epoxyfunctional


polyester DP3 (EDP3)

EFA, dimethyl adipate (DA) and hexanol were put into a round bottom
flask in a 3:1:2 ratio and heated up to 85 oC. After the reaction mixture
had reached that temperature the enzyme was added to the flask (10% of
the total amount of monomers) and the reaction was allowed to proceed
at open atmosphere for 2h and afterwards under vacuum overnight.
The reaction was stopped by filtering out the enzyme and the product
was characterized and used without further purification (Yield: 65%).

3.3.4. Synthesis of ,ω-oxetane end-capped aliphatic polyester


DP3 (ODP3)

Dimethyl hexadecanedioate, Ethylene glycol and TMPO were put into a


round bottom flask in a 3:2:2 ratio and heated up to 85 oC. After the
reaction mixture had reached that temperature the enzyme was added to
the flask (10% of the total amount of monomers) and the reaction was
allowed to proceed at open atmosphere for 2h and afterwards under
vacuum overnight. The reaction was stopped by filtering out the enzyme.
The product was purified by column chromatography using silica as
stationary phase and a gradient of eluents, starting with a mixture of
Ethyl acetate: Heptane, 30:70 and increasing to 100% Ethyl acetate.
(Yield: 11%).

3.3.5. Synthesis of Dimethyl hexadecanedioate

Hexadecanedioic acid (1eq.) was suspended in methanol (20mL).


Afterwards HCl (2.2eq.) was added dropwise at room temperature. After
4 hours the formed precipitate was filtered out and dried under vacuum
to yield a white powder (yield: 69%).

14
Experimental

3.4. Thermoset formation

3.4.1. Radical polymerization of methacrylates

Mixtures containing 25mg EFA-EGMA, Irgacure 651 (4 wt% of the


monomer) and 100mg chloroform were homogenized and casted on a
glass substrate. The solvent was allowed to evaporate before covering
with a quartz microscope slide. Free-standing films were obtained by
radical polymerization after 2min of light exposure with an intensity of
21mW/cm2. (Dose: 2.52 J/cm2)

3.4.2. Cationic polymerization

All oligomers subjected to cationic polymerization (EFA-EGMA, EODP3,


EODP6, EDP3, ODP3) were prepared by charging vials with 25mg of
each oligomer Uvacure 1600 (4 wt% of the monomer) and 100mg
chloroform. After homogenization of the mixture, the sample was casted
on a glass substrate and the solvent was allowed to evaporate before
proceeding with the curing. In order to obtain homogeneously cured
films, a quartz microscope slide was put on top of the mixture. Polymer
thermosets using cationic polymerization were obtained after irradiating
the sample with an intensity of ~21mW/cm2 during 2.5min. (Dose: 3.15
J/cm2)

3.4.3. Dual curing

Vials were charged with 25mg EFA-EGMA , Irgacure 651 (4 wt%),


Uvacure 1600 (4 wt%) and 100mg chloroform. After homogenization of
the mixture, the sample was casted on a glass substrate and the solvent
was allowed to evaporate before covering with a quartz microscope slide.
Cured resins were obtained after 2.5 min of light exposure with an
intensity of 21mW/cm2. (Dose: 3.15 J/cm2)

15
Experimental

3.5. Analytical methods


1
3.5.1. H NMR

The spectra were recorded at 400 MHz on a Bruker AM 400 using


deuterated chloroform (CDCl3) as solvent. The solvent signal was used
as reference.

3.5.2. MALDI-TOF

MS spectrum acquisitions were conducted on a Bruker UltraFlex


MALDI-TOF MS with SCOUT-MTP Ion Source (Bruker
Daltonics,Bremen) equipped with a N2-laser (337nm), a gridless ion
source and reflector design. All spectra were acquired using a reflector-
positive method with an acceleration voltage of 25 kV and a reflector
voltage of 26,3 kV. The detector mass range was set to 200-2500 m/z. A
THF solution of DHB was used as matrix. The obtained spectra were
analyzed with FlexAnalysis broker Daltonics, Bremen, version 2.

3.5.3. ATR-FTIR

Analysis was carried out using a Perkin-Elmer Spectrum 2000 FT-IR


instrument (Norwalk, CT) equipped with a single reflection (ATR:
attenuated total reflection) accessory unit (Golden Gate) from Graseby
Specac LTD (Kent, England) and a TGS detector using the Golden Gate
setup. Each spectrum collected was based on 16 scans averaged at 4.0
cm-1 resolution range of 600-4000 cm-1. Data were recorded and
processed using the software Spectrum from Perkin-Elmer.

3.5.4. RT-FTIR

The RTIR analysis were made using a Perkin-Elmer Spectrum 2000 FT-
IR instrument (Norwalk, CT) equipped with a single reflection (ATR:
attenuated total reflection) accessory unit (Golden Gate) from Graseby
Specac LTD (Kent, England). RT-FTIR continuously recorded the
chemical changes over the range 4000-600 cm-1. Spectroscopic data

16
Experimental

were collected at an optimized scanning rate of 1 scan per 1.67 seconds


with a spectral resolution of 4.0 cm-1 using Time Base® software from
Perkin-Elmer.

3.5.5. SEC

Using DMF ( 0.2 mL min-1) with 0.01M LiBr as the mobile phase was
performed at 50oC using a TOSOH EcoSEC HLC-8320GPC system
equipped with an EcoSEC RI detector and three columns (PSS PFG 5μm;
microguard, 100Å, and 300Å) (Mw resolving range: 100-300000 g/mol)
from PSS GmbH. A calibration method was created using narrow linear
polystyrenes standards. Corrections for the flow rate fluctuations were
made using toluene as an internal standard.

3.5.6. DMA

The study of the physical properties of the polymerized resins was


performed on a Q800 DMTA (TA instruments), equipped with a film
fixture for tensile testing. The measurements were done between -60°C
and 160°C, with heating rate of 5°C/min. The tests were performed in
controlled strain mode with a frequency of 1Hz, oscillating amplitude of
10μm and forcetrack of 125%.

3.5.7. UV-Light sources

The light source used for the crosslinking of the films was a Black Ray B-
100AP (100 W, λ=365 nm) Hg UV-lamp with an irradiance of
~21mW/cm2 as determined with an UVICURE Plus High Energy UV
integrating radiometer (EIT, USA), measuring UVA at 320 ≤ λ ≤390 nm.
The light source used for the photo-RTIR measurements was a
Hamamatsu L5662 equipped with a standard medium-pressure 200-W
L6722-01 Hg-Xe lamp and provided with optical fibers. The UV intensity
was measured using a Hamamatsu UV-light power meter (model
C6080-03) calibrated for the main emission line centered at 365nm. The
intensity of the lamp was 7mW/cm2.

17
Results and discussion

4. Results and discussion

4.1. Extraction of 9,10-epoxy-18-


hydroxyoctadecanoic acid (EFA) from
outer birch bark

The extraction of 9,10-epoxy-18-hydroxyoctadecanoic acid (EFA) from


outer birch bark can easily be accomplish through an alkaline hydrolysis
of the suberin. By lowering carefully the pH to 6 it is possible to isolate
this monomer by a selective precipitation and then purify by
recrystallization from toluene. The EFA monomer accounts 10% of the
total weight of outer birch bark. In the present work an average yield of
60% of pure EFA monomer was obtained after extraction. The purity of
the monomer was evaluated by 1H NMR were it is also possible to see
that the epoxide remains intact during the whole process (see Figure 4).

Figure 4. 1H NMR spectrum of 9,10-epoxy-18-hydroxyoctadecanoic acid (EFA)1.

4.2. Synthesis of telechelic oligomers by


enzyme catalyzed esterification

One of the main advantages of enzyme catalyzed esterification is the


possibility of tuning the degree of polymerization by adjusting the

18
Results and discussion

stoichiometric ratios of the reactants and introducing an end-capper.


The reactants, usually diacids or diesters, react with each other in an
alternated way until they are consumed. In that moment the end-capper,
usually an alcohol, finishes the reaction by reacting at the end of the
chain. In a general manner, the synthesis of the telechelic oligomers
follows the reaction scheme shown in Scheme 5.

Scheme 5. Synthesis of epoxy-derived ,ω-telechelic polyesters by enzyme


catalyzed esterification. R-OH represents an alcohol used as end-capper.

However, it is also possible to perform the reaction using a methyl ester


that will act both as reactant and end capper through a
transesterification reaction (see Scheme 6).

Scheme 6. Enzyme catalyzed synthesis of epoxy-derived ,ω-telechelic polyesters


through transesterification reaction.

Two approaches have been considered for the synthesis of


multifunctional telechelic oligomers derived from the EFA monomer:
synthesis of oligomers with two functional groups with different
reactivities or with similar reactivities.
For the synthesis of oligomers containing functional groups with
differentiated reactivity, a methacrylate was introduced as end group of
the epoxy-functional polyester. Epoxides exhibit high reactivity towards
ring opening and methacrylates are known for their ease to polymerize
radically. Both reactions can be triggered by UV-light, nevertheless the
activated species are different, i.e. a cation or a radical, and therefore is
possible to tune the reactivity of the oligomer.
In the case of the multifunctional oligomers containing two groups with
similar reactivities, it was necessary to introduce as end functionality
another group capable to undergo cationic polymerization in a similar

19
Results and discussion

way as the epoxide does. It was also necessary to consider that the
intrinsic epoxide of the EFA oligomer is a secondary epoxide and its
reactivity differs to the reactivity of primary epoxides. In these terms, an
oxetane was introduced as suitable end groups. Although the epoxide is
slightly more ring strained than the oxetane and will ring open faster
when attacked by a nucleophile, the oxetane is slightly more basic which
will make it more reactive towards cationic polymerization and therefore
these two groups are considered equivalent. In Figure 5 are depicted all
the oligomers used in this work.

Figure 5. List of telechelic oligomers used in this thesis

All oligomers were characterized by 1H NMR, MALDI-TOF MS and SEC.

20
Results and discussion

1
4.2.1. H NMR

In order to evaluate the success of the polymerization some factors were


taken into consideration: permanence of the epoxy group during the
whole process and the appearance/ disappearance of new peaks
corresponding to the formation of new ester bonds. In Figure 6 are
represented the 1H NMR spectrum of two of the prepared oligomers
together with the spectrum of the oligomer.

Figure 6. 1H NMR spectra of the EFA monomer (lower spectrum), EODP3


oligomer (mid spectrum) and EFA-EGMA oligomer (upper spectrum).

1H NMR could also be used to estimate the degrees of polymerization of


the oligomers.
Oligomer DP NMR
EODP3 2.2
EODP6 5.1
EDP3 2.7
ODP3 2.3
EFA-EGMA 4.4

Table 1. Degrees of polymerization of all oligomers calculated from their 1H NMR


spectrum. More information about the 1H NMR spectra and the calculations of
the DP can be found in Paper II.

21
Results and discussion

Due to volatility issues, all the epoxide/ oxetane oligomers possess a DP


slightly lower than the one aimed for. All products followed the aimed
trend with respect to size although some deviations were found. This is
proposed to be due to factors such as slight errors in stoichiometry as
well as possible evaporation of reactants.

4.2.2. MALDI-TOF MS

The structure of the oligomers was also determined by MALDI-TOF


mass spectroscopy. As exemplified in Figure 7 for the EODP3, an
average degree of polymerization close to the targeted was obtained in
majority. As can be seen, a small fraction of monosubstituted oligomer is
also obtained. This does not significantly interfere with the controlled
formation of the final polymer networks although it must be considered
if it appears to a larger extent.

Figure 7. MALDI-TOF mass spectrum of the EODP3 oligomer. “n” represents the
degrees of polymerization of the disubstituted oligomer and “m” the degrees of
polymerization of the monosubstitued56.

22
Results and discussion

4.2.3. SEC

As can be seen in Figure 8 for the cases of the epoxy-derived oligomers


with DP3 end-capped with methacrylate (EFA-EGMA) and with oxetane
(EODP3), a broad distribution of molecular weights was obtained for
these enzymatic polymerizations. This type of distributions is common
in synthesis of all oligomers in the present study.

Figure 8. SEC traces of EFA-EGMA oligomer (left) and the EODP3 oligomer
(right).

4.3. Thermoset formation

The presence of two or more functional groups in the same polymer


chain, gives us the possibility of tailoring the final properties of the
network by controlling the way this groups react with each other. The
possibilities of network formation will increase with the number of
different functional groups. For the type of oligomers used in this work,
i.e. with two different functional groups, four different networks can be
formed. As shown schematically in Scheme 7, this networks can be
formed by homopolymerization reaction of the groups with themselves
separately (Networks A and B) or at the same time (Network C), or by
copolymerization reaction of the two groups with each other (Network
D).

23
Results and discussion

Scheme 7. Schematic representation of the possible networks formed starting


from a ,ω- end functional oligomer with 3 functional groups in the backbone

The different ways the networks are formed will have direct
repercussions on their mechanical properties as will be explained in
point 4.3.2. Mechanical properties of the formed networks.

In order to study the reactivity


of the different functional
groups towards network
formation, FTIR and RT-FTIR
studies were performed.
The real time measurements
were performed by recording
spectrum continuously while
irradiating the sample put on
Scheme 8. RT-FTIR setup for the in
the ATR crystal. A schematic situ photopolymerization analysis.
representation of the
experimental setup can be
found in Scheme 8.

24
Results and discussion

4.3.1. Polymerization performance. Fourier Transformed Infrared


and Photo-Real Time Infrared Analysis

Epoxide/ methacrylate systems. Cationic polymerization vs


radical polymerization.

For the epoxide/ methacrylate


oligomers, two possibilities of
network formation yielding in
three different types of networks
were contemplated: either by
reaction of the groups
independently or in parallel by
using one of the photoiniators at
the time or together.
The success of the radical
polymerization could be Figure 9. FTIR spectrum of the
methacrylate/epoxy resin measured by
monitored by complete
RT-FTIR during the course of the
disappearance of the peak at radical polymerization1.
1638 cm -1 as well as the
conjugated ester carbonyl at 1710
cm-1. The polymerization proceeds rapidly with a full conversion
obtained after 2.5 min (see Figure 9).
In the case of the thermosets crosslinked by cationic polymerization, it is
possible to observe a variation of the peaks in the region between 600
and 1100 cm-1 (see Figure 10). However, due to peak overlapping the
assignment of the peaks corresponding to the vibrations of the oxirane
ring was challenging. Those vibrations appeared in the same region as
peaks originating from the methacrylate moiety (around 845 cm -1).
Nevertheless, the obtainment of free standing films in an “open air
system” together with the variations in the mechanical properties of the
network were considered as sufficient evidences of the cationic
polymerization. Similar behaviors were observed for the dual cured
networks. Whilst the peaks used to assess the reaction between
methacrylates can clearly be differentiated, the ones corresponding to
the epoxides appear overlapping in the aforementioned region. As will
be explained in point 4.3.2.Mechanical properties of the formed

25
Results and discussion

networks, the formation of the network by reaction of both groups will


have a direct repercussion in the properties of the network as a much
“tighter” network is formed when both groups react. The last was taken
as trustworthy evidence of the dual curing procedure

Figure 10. FTIR spectra of the methacrylate/ epoxide oligomer (lower) and the
three networks formed from it through radical polymerization, cationic
polymerization and dual curing (from bottom to top).

Epoxide/ oxetane systems. Homopolymerization vs copolymerization


in cationic polymerization

When two functional groups with similar reactivity are present in the
oligomer, the possibilities for network formation increase, since these
two groups may also be able to copolymerize with each other. This is the
case of the oligomers containing an epoxide and an oxetane; both
compounds are cyclic ethers and therefore suitable monomers for
cationic polymerization57.
The oligomers were then subjected to cationic polymerization monitored
by FTIR-RT In order to be able to differentiate the peaks corresponding
to each group, analogous oligomers only containing epoxides or only
containing oxetanes were synthesized (EDP3 and ODP3).
It was observed, that for the oligomers only containing one of the
functional groups, the cationic ring opening was faster for epoxides than

26
Results and discussion

for oxetanes. This was proven by monitoring the appearance of new


peaks in the region between 1120-1000 cm-1 and tracking the changes
every 10 minutes during one hour (see Figure 11). Homopolymerization
of secondary epoxides gives rise to a new absorption band at 1073 cm -1
due to the formation of a substituted polyethylene glycol structure
(Figure 10 left), while homopolymerization of the oxetane yields in a new
band at 1062 cm-1 corresponding to a polypropylene glycol structure
(Figure 10 right).

Figure 11. Spectra of EDP3 (left) and ODP3 (right) oligomers taken every 10 min
with focus in the area between 1120-1000 cm-1 56

In order to understand the reactivity towards cationic polymerization of


these two groups when present in the same oligomer, similar oligomers
with different degrees of polymerization, 3 and 6, were synthesized.
Different degrees of polymerization will also mean different oxetane/
epoxy ratios and so, for the oligomer with DP=3 the oxetane/ epoxy ratio
will be 2/3 and for DP=6, 1/3.
As can be seen in Figure 12, already after 10 minutes of exposure to the
UV-light of the epoxy/ oxetane oligomer with DP3 (EODP3), there is a
large increase in absorbance in the region between 1120-1000 cm-1 as
consequence of the formation of new ether bonds. When the absorbance
of the peaks corresponding to the epoxide (1073 cm -1) and the oxetane
(1062 cm-1) are plotted against time (Figure 13), it is possible to see that
the absorbance increase faster for the EODP3 oligomer than for the
EDP3 or ODP3 oligomers. This is due to the preferred copolymerization
reaction between epoxide and oxetane. As already proven by Sasaki 58
and Crivello59, 60, when a small amount of epoxide is present during the

27
Results and discussion

cationic polymerization of a specie containing oxetane, the reaction rate


increases dramatically in comparison with the polymerization without
epoxide. This is due to the fast generation of an oxonium ion from the
epoxy ring that gives a “kick-start” to the cationic polymerization of the
oxetane61. This and other studies mainly dealing with primary epoxides
and oxetane monomers clearly shown that epoxy and oxirane monomers
readily copolymerize with each other to give polyether polymers 62, 63. In
the present study we have extended the concept of copolymerization
between the two groups present in the same polymer chain, to its use as
tool for tailoring the final properties of the thermoset.

Figure 12. Spectra of EODP3 (left) and EODP6 (right) oligomers taken every 10
min with focus in the area between 1120-1000 cm-1 56

Figure 13. Absorbance vs time of the peak resulting from the ring opening of
epoxides (1073 cm-1, left graph) and the peak resulting from the ring opening of
oxetanes (1062 cm-1, right graph)56

As can be seen in Figure 12, the increase in absorbance is not as


remarkable for the epoxy/ oxetane oligomers with DP6 as for the

28
Results and discussion

oligomers with DP3. This is because the oxetane/ epoxy ratio is smaller
for the oligomers with DP6 and the copolymerization is less probable.

4.3.2. Mechanical properties of the formed networks

The main factors affecting the mechanical performance of a thermoset


are: rigidity of the backbone, polarity, crosslink density and the presence
of loose chain ends. All the present resins are aliphatic polyesters and
therefore possess similar polarity and rigidity of the backbone. The two
distinctive parameters will therefore be the crosslink density and the
presence of dangling ends.
As can be seen in Figure 14 and summarized in Table 2, the lowest glass
transition temperature was obtained for the epoxy/methacrylate resin
only crosslinked radically. In this case, the network is formed by
covalent binding of the end groups yielding in a rather loose network
(network Type B, according to Scheme 7). On the other hand, the highest
Tg value is obtained for the dual cured network (network type C). This
network is formed by covalent reaction of two functional groups
simultaneously, the epoxide and the methacrylate, which produces a
tighter network.
Although similar, the networks formed by cationic polymerization of the
epoxy/ methacrylate resin and the EDP3 resin present different Tg
values: 24 and 5, respectively. In both cases the initial oligomer present
an average of 3 epoxy groups capable to react through cationic
polymerization (network Type A). However, the EDP3 oligomer possess
long carbon chains as end groups which confers more flexibility to the
network and therefore decreases the Tg.
The networks formed starting from oligomers EODP3 and EODP6
present Tg values different to those mentioned before. This is due to the
fact that the networks are built through copolymerization reaction
between both groups (network Type D). The difference in Tg values
between these two are related to the difference in oxetane/ epoxy ratio.

29
Results and discussion

Figure 14. Tan vs temperature representations of the cured resins

The homogeneity of the network is furthermore reflected in the width of


the Tg transition64. In this work, the two types of polymerization used,
radical polymerization and cationic polymerization, follow a chain wise
mechanism. Chain wise mechanisms usually present broader glass
transition intervals, since they are less homogeneous processes than the
step wise mechanisms. The last is evident in all cases except of the
networks formed by copolymerization reaction, EODP3 and EODP6. The
narrow transitions are seen as prove of the high homogeneity of this
networks, especially evident in the EODP3 network, where the higher
oxetane to epoxide ratio enhances the possibilities of copolymerization
reaction and therefore the homogeneity of the network. Exact details on
the reaction mechanism for the copolymerization, is not included in the
present study but the results indicate that this is an area worth to further
study. An increased homogeneity with increased crosslink density
indicates a change in mechanism.
The increase in crosslink density is also seen as an increase in the
rubbery modulus (see Figure 15). This parameter should only be taken
into consideration for the epoxide/ oxetane systems. In the case of the
epoxide/ methacrylate oligomers the films were too fragile to be able of
making conclusions in this region.

30
Results and discussion

Figure 15. Storage modulus vs temperature representations of the cured resins.

E’ E’ rp Tg Tan
Code DP Type of network
(GPa) (GPa) (oC) (E’/E’’)

Cationic
3 3 - 24 0.26
polym.

Radical
3 3 - 3 0.53
polym.

Cationic
+ 3 3 - 40 0.31
Radical

EDP3 3 2 0.3 5 0.51

EODP3 3 2 1 15 1.03

EODP6 6 2 0.2 12 0.79

Table 2. Summary of the mechanical properties of the cured resins (E’: storage
modulus at -60oC; E’rp: storage modulus in the rubbery plateau).

31
Conclusions

5. Conclusions
In the continuous search of natural based raw materials the utilization of
a side product from the wood industry to prepare polymer networks is
not only of interest from an academic perspective but also from an
industrial point of view. At the same time, the possibility of tuning the
mechanical properties of the polymer network opens up a broad range of
applications in terms of material design. In the present work a series of
polymer networks based on a renewable monomer have been prepared.
In order to be able of tuning the properties of the network, different end
groups and different polymerization methods were used. In a first term,
methacrylates were used as end cappers of the epoxy-derived oligomers.
In this case, the approach used for the formation of the networks was to
react the two functional groups either separately or at the same time
using two different polymerization methods, i.e. radical and/or cationic
polymerization. The second approach was to end cap the oligomers with
functional groups with similar reactivity as the epoxides. For that
purpose, oxetanes were selected as end capping groups. The overall aim
was to understand if in this case the two functional groups would again
form networks by reacting with themselves or if they would react with
each other. What we observed in this case was the clear preference of
these two groups to react with each other which not only yielded in new
mechanical properties but also in an increase in the reaction rate.
We therefore conclude that the main achievements in this work have
been:
1 monomer. Extraction and purification of a monomer that
proceeds from a “zero value source” as the outer bark of the
birch tree. We herein demonstrate the possibility of obtaining
a bio-based compound, pure enough to be polymerized, with
a straightforward process.
2 functionalities. The synthetic route presented in this
thesis, gives us the possibility of preparing telechelic
polymers with a large variety of functional groups in the main
chain. The control of the degree of polymerization and end-
functionalization while the epoxide stays intact throughout
the whole process, are achieved by enzyme catalyzed
polymerization.
3 polymerization methods. All oligomers prepared
possess 2 different functionalities: the intrinsic of the epoxy
monomer and the telechelic functionality. In order to
selectively react the different functional groups, 3

32
Conclusions

polymerization methods were used: radical polymerization,


cationic polymerization and a combination of both (dual
curing).
4 combinations. As shown in Scheme 7, for the oligomers
prepared in this work containing 2 different functional
groups in the main chain (present in different proportions), 4
different combinations for network formation are possible:
reaction of the end groups with themselves, reaction of the
epoxides with themselves, reaction of the two groups at the
same time and reaction of the two groups with each other.
The different ways these groups are “connected” will
determine the properties of the networks.

33
Future work

6. Future work
The knowledge gained during the present work encourages us to
continue our research towards other natural based monomers and other
polymerization methods that will allow the formation of polymer
thermosets. The preparation of polymeric materials derived from
natural resources is an emerging field and in continue development and
therefore invite us to challenge ourselves in this search.
In the present work we have demonstrated how polymer networks can
be easily synthesized by combining an epoxy-functional monomer with
different end-functional groups. The properties of the network can then
be tailored by choosing how these groups will react with each other. In
this context, the only variation of the monomer or the end-capper will
result in the formation of new networks with, inevitably, different
properties.
The use of enzymes in the synthesis of polymers is of great interest and
therefore a field we still want to explore. Enzymatic catalysis has been
proven to be a useful tool for the synthesis of polymers that with other
“traditional” routes cannot be prepared, e.g. polyesters prepared by
eROP of macrolactones65-67. This opens many new possibilities also in
terms of material design.
As already proven by us and others, the selectivity of the enzymes allows
the synthesis of polymers while keeping intact other groups intrinsic of
the monomer. This allows e.g. the possibility of post-functionalizing the
surface of the formed network, which will increase its range of
applications.
The polymers and oligomers prepared in this work are rather fragile and
therefore far away of having an industrial application. However, there
are some options for strengthening the networks as for example using
more bulky monomers.

34
Acknowledgements

7. Acknowledgements
In a first term, I would like to thank the People Programme (Marie Curie
Actions) of the European Union’s Seventh Framework Programme
FP7/2007-2013/ under REA grants agreement no [289253] which have
funded the research leading to these results and the REFINE project.
With this, I would also like to thank all the partners of the REFINE
project, ESRs and ERs, for good discussion about science and not
science, collaborations and great meetings around Europe.
My greatest thank is to my main supervisor Mats Johansson for always
being supportive and helpful and for having given me the opportunity of
performing my PhD-studies.
I would also like to thank my main co-author, Stefan Semlitsch and my
co-supervisor, Mats Martinelle, for a great work together and for always
being open for new ideas and projects. In this context I would also like to
thank some of the other members of the BioCat group, Karl Hult and
Maja Finnveden, for great collaborations and meetings.
Eva Malmström is also thanked for having given me the opportunity to
learn more about “Telechelic polymers”.
Sandra Garcia Gallego is thanked for all the “technical” help while
writing this thesis.
All my current and former colleagues (and friends) at the Division of
Coating Technology are thanked for creating the absolute best working
environment possible, because “there are never stupid questions” and
for always being warm and welcoming to new people.
Last, but definitely not least, I would like to thank my family for always
being a great support and for helping me to see the positive side of the
things.

All this people are thanked because in larger or smaller way, more
professional or more personal, have contributed to this first half of my
PhD studies. My greatest thank to all of you!

35
References

8. References
1. S. Torron, S. Semlitsch, M. Martinelle and M. Johansson,
Macromol. Chem. Phys., 2014, 215, 2198-2206.
2. R. Ekman, Holzforschung, 1983, 37, 205-211.
3. WO2010093320A1, 2010.
4. R. Jerome, M. Henrioulle-Granville, B. Boutevin and J. J. Robin,
Prog. Polym. Sci., 1991, 16, 837-906.
5. C. Hedfors, E. Oestmark, E. Malmstroem, K. Hult and M.
Martinelle, Macromolecules, 2005, 38, 647-649.
6. A. Rüdiger, P. Hendil-Forssell, C. Hedfors, M. Martinelle, S. Trey
and M. Johansson, Journal of Renewable Materials, 2013, 1,
124-140.
7. L. Montero de Espinosa and M. A. R. Meier, European Polymer
Journal, 2011, 47, 837-852.
8. M. N. Belgacem, A. Gandini and Editors, Monomers, Polymers
and Composites from Renewable Resources, Elsevier Ltd., 2008.
9. M. A. R. Meier, Macromol. Chem. Phys., 2009, 210, 1073-1079.
10. S. A. Miller, ACS Macro Lett., 2013, 2, 550-554.
11. P. A. Krasutsky, Nat. Prod. Rep., 2006, 23, 919-942.
12. A. Gandini, C. Pascoal Neto and A. J. D. Silvestre, Progress in
Polymer Science, 2006, 31, 878-892.
13. M. A. Bernards, Canadian Journal of Botany, 2002, 80, 227-240.
14. P. C. R. O. Pinto, A. F. Sousa, A. J. D. Silvestre, C. P. Neto, A.
Gandini, C. Eckerman and B. Holmbom, Industrial Crops and
Products, 2009, 29, 126-132.
15. H. Kim, J. V. Olsson, J. L. Hedrick and R. M. Waymouth, ACS
Macro Lett., 2012, 1, 845-847.
16. R. A. Gross, B. Kalra and A. Kumar, Appl. Microbiol. Biotechnol.,
2001, 55, 655-660.
17. S. Kobayashi and A. Makino, Chemical Reviews, 2009, 109,
5288-5353.
18. S. Kobayashi, H. Uyama and S. Kimura, Chem. Rev.
(Washington, D. C.), 2001, 101, 3793-3818.
19. R. A. Gross, A. Kumar and B. Kalra, Chem. Rev. (Washington, D.
C.), 2001, 101, 2097-2124.
20. I. K. Varma, A.-C. Albertsson, R. Rajkhowa and R. K. Srivastava,
Prog. Polym. Sci., 2005, 30, 949-981.

36
References

21. J.-i. Kadokawa and S. Kobayashi, Current Opinion in Chemical


Biology, 2010, 14, 145-153.
22. A. Olsson, M. Lindström and T. Iversen, Biomacromolecules,
2007, 8, 757-760.
23. S. Okumura, M. Iwai and Y. Tominaga, Agric. Biol. Chem., 1984,
48, 2805-2808.
24. A. Ajima, T. Yoshimoto, K. Takahashi, Y. Tamaura, Y. Saito and Y.
Inada, Biotechnol. Lett., 1985, 7, 303-306.
25. A. L. Margolin, J. Y. Crenne and A. M. Klibanov, Tetrahedron
Lett., 1987, 28, 1607-1610.
26. J. S. Wallace and C. J. Morrow, J. Polym. Sci., Part A: Polym.
Chem., 1989, 27, 3271-3284.
27. S. Kobayashi, Macromol. Rapid Commun., 2009, 30, 237-266.
28. B. Yeniad, H. Naik and A. Heise, Adv. Biochem. Eng./Biotechnol.,
2011, 125, 69-95.
29. S. Kobayashi, H. Uyama and S. Namekawa, Polymer
Degradation and Stability, 1998, 59, 195-201.
30. O. X. o.-l. c. v. h. g. i. o.-c. Compiled by A. D. McNaught and A.
Wilkinson. Blackwell Scientific Publications, in IUPAC.
Compendium of Chemical Terminology, 2nd ed. (the "Gold
Book").IUPAC. Compendium of Chemical Terminology, 2nd ed.
(the "Gold Book").
31. D. Knani, A. L. Gutman and D. H. Kohn, J. Polym. Sci., Part A:
Polym. Chem., 1993, 31, 1221-1232.
32. H. Uyama and S. Kobayashi, J. Mol. Catal. B: Enzym., 2002, 19-
20, 117-127.
33. H. Uyama and S. Kobayashi, Chem. Lett., 1993, 1149-1150.
34. M. Takwa, K. Hult and M. Martinelle, Macromolecules
(Washington, DC, U. S.), 2008, 41, 5230-5236.
35. M. Takwa, N. Simpson, E. Malmstroem, K. Hult and M.
Martinelle, Macromol. Rapid Commun., 2006, 27, 1932-1936.
36. A.-C. Albertsson and R. K. Srivastava, Advanced Drug Delivery
Reviews, 2008, 60, 1077-1093.
37. A. Cordova, T. Iversen and K. Hult, Polymer, 1999, 40, 6709-
6721.
38. M. Eriksson, L. Fogelstroem, K. Hult, E. Malmstroem, M.
Johansson, S. Trey and M. Martinelle, Biomacromolecules,
2009, 10, 3108-3113.

37
References

39. M. Eriksson, A. Boyer, L. Sinigoi, M. Johansson, E. Malmstroem,


K. Hult, S. Trey and M. Martinelle, J. Polym. Sci., Part A: Polym.
Chem., 2010, 48, 5289-5297.
40. M. Eriksson, K. Hult, E. Malmstroem, M. Johansson, S. M. Trey
and M. Martinelle, Polym. Chem., 2011, 2, 714-719.
41. Anon, Ceram.-Silik., 2008, 52, 299.
42. H. A. Pohl, J. Am. Chem. Soc., 1951, 73, 5660-5661.
43. C. Decker, Macromol. Rapid Commun., 2002, 23, 1067-1093.
44. Y. Yagci, S. Jockusch and N. J. Turro, Macromolecules, 2010, 43,
6245-6260.
45. C. Decker, Progress in Polymer Science, 1996, 21, 593-650.
46. J. P. Fouassier, X. Allonas and D. Burget, Progress in Organic
Coatings, 2003, 47, 16-36.
47. C. E. Hoyle and C. N. Bowman, Angew. Chem., Int. Ed., 2010, 49,
1540-1573.
48. K. Studer, C. Decker, E. Beck and R. Schwalm, Progress in
Organic Coatings, 2003, 48, 92-100.
49. T. Y. Lee, C. A. Guymon, E. S. Jönsson and C. E. Hoyle, Polymer,
2004, 45, 6155-6162.
50. T. Y. Lee, T. M. Roper, E. S. Jonsson, I. Kudyakov, K.
Viswanathan, C. Nason, C. A. Guymon and C. E. Hoyle, Polymer,
2003, 44, 2859-2865.
51. K. S. Anseth, C. M. Wang and C. N. Bowman, Polymer, 1994, 35,
3243-3250.
52. M. Sangermano, N. Razza and J. V. Crivello, Macromol. Mater.
Eng., 2014, 299, 775-793.
53. K. Horie, M. Barón, R. Fox, J. He, M. Hess, J. Kahovec, T.
Kitayama, P. Kubisa, E. Maréchal and W. Mormann, Pure and
Applied Chemistry, 2004, 76, 889-906.
54. J. V. Crivello and J. L. Lee, J. Polym. Sci., Part A: Polym. Chem.,
1989, 27, 3951-3968.
55. S. R. Akhtar, J. V. Crivello and J. L. Lee, J. Org. Chem., 1990, 55,
4222-4225.
56. S. Torron and M. Johansson, "Oxetane terminated telechelic
epoxyfunctional-polyesters as cationically polymerizable
thermoset resins – tuning the reactivity with structural design".
Submitted manuscript, 2015.
57. S. Penczek, M. Cypryk, A. Duda, P. Kubisa and S. Słomkowski,
Progress in Polymer Science, 2007, 32, 247-282.

38
References

58. H. Sasaki, J. M. Rudzinski and T. Kakuchi, J. Polym. Sci., Part A:


Polym. Chem., 1995, 33, 1807-1816.
59. U. Bulut and J. V. Crivello, Journal of Polymer Science Part A:
Polymer Chemistry, 2005, 43, 3205-3220.
60. J. V. Crivello, B. Falk and M. R. Zonca, Jr., J. Polym. Sci., Part A:
Polym. Chem., 2004, 42, 1630-1646.
61. J. V. Crivello, J. Polym. Sci., Part A: Polym. Chem., 2014, 52,
2934-2946.
62. S.-C. Yang, J.-H. Jin, S.-Y. Kwak and B.-S. Bae, J. Appl. Polym. Sci.,
2012, 126, E380-E386.
63. US20060041032A1, 2006.
64. M. Chanda and Editor, Introduction to Polymer Science and
Chemistry: A Problem Solving Approach, CRC Press, 2006.
65. H. Uyama, H. Kikuchi, K. Takeya and S. Kobayashi, Acta Polym.,
1996, 47, 357-360.
66. H. Uyama, K. Takeya, N. Hoshi and S. Kobayashi,
Macromolecules, 1995, 28, 7046-7050.
67. L. Van der Mee, F. Helmich, R. De Bruijn, J. A. J. M. Vekemans,
A. R. A. Palmans and E. W. Meijer, Macromolecules, 2006, 39,
5021-5027.

39

Das könnte Ihnen auch gefallen