Вы находитесь на странице: 1из 150

SURFACE MODIFICATION OF SILICA NANOPARTICLES

A Dissertation
Presented to
The Graduate Faculty of the University of Akron

In Partial Fulfillment
of the Requirements for the Degree
Doctor of Philosophy

Rajesh Ranjan
May, 2008

SURFACE MODIFICATION OF SILICA NANOPARTICLES

Rajesh Ranjan
Dissertation

Approved:

Accepted:

__________________________
Advisor

________________________
Department Chair

Dr. Roderic P. Quirk

Dr. Mark D. Foster

__________________________
Committee Member

________________________
Dean of the College

Dr. Scott Collins

Dr. Stephen Z. D. Cheng

__________________________
Committee Member

________________________
Dean of the Graduate School

Dr. Ali Dhinojwala

Dr. George R. Newkome

__________________________
Committee Member

________________________
Date

Dr. Coleen Pugh

________________________
Committee Member
Dr. Bi-min Zhang Newby
ii

ABSTRACT
Surface modification of nanosized silica particles by polymer grafting is gaining
attention. This can be attributed to the fact that it provides a unique opportunity to
engineer the interfacial properties of these modified particles; at the same time the
mechanical and thermal properties of the polymers can be improved. Controlled free
radical polymerization is a versatile technique which affords control over molecular
weight, molecular weight distribution, architecture and functionalities of the resulting
polymer.

Three commonly used controlled free radical polymerizations include

nitroxide- mediated polymerization (NMP), atom transfer radical polymerization (ATRP)


and reversible addition fragmentation transfer (RAFT) polymerization. ATRP and RAFT
polymerization were explored in order to modify the silica surface with well-defined
polymer brushes.
A novel click-functionalized RAFT chain transfer agent (RAFT CTA) was
synthesized which opened up the possibility of using RAFT polymerization and click
chemistry together in surface modification. Using this RAFT CTA, the surface of silica
nanoparticles was modified with polystyrene and polyacrylamide brushes via the
grafting to approach. Both tethered polystyrene and polyacrylamide chains were found
in the brush regime. The combination of ATRP and click chemistry was also explored
for surface modification.

iii

A combination of RAFT polymerization and click chemistry was also studied to


modify the surface via the grafting from approach. Our strategy included the (1)
grafting from approach for brush formation (2) facile click reaction to immobilize the
RAFT agent (3) synthesis of R-supported chain transfer agent and (4) use of the more
active trithiocarbonate RAFT agent.

Grafting density obtained by this method was

significantly higher than reported values in the literature.


Polystyrene (PS) grafted silica nanoparticles were also prepared by a tandem
process that simultaneously employs reversible addition fragmentation transfer (RAFT)
polymerization and click chemistry. The click reaction doesnt interfere with RAFT
polymerization. With a suitable choice of a Cu(I) catalyst, it is possible to perform both
RAFT polymerization and click chemistry together. In a single pot procedure, azidemodified silica, an alkyne-functionalized RAFT agent and styrene were combined to
produce the desired product. As deduced by thermal gravimetric and elemental analysis,
the grafting density of PS on the silica in the tandem process was intermediate between
the grafting to and grafting from techniques. Relative rates of RAFT polymerization
and click reaction were altered to control grafting density.
ATRP was also used to modify the surface of silica nanoparticles via the grafting
from approach. The surfaces of silica with homopolymers and diblock copolymers
brushes were modified using surface initiated ATRP. The polymer grafted silica particles
were characterized by FT-IR, TGA, XPS and elemental analysis.

iv

DEDICATION
To my loving parent, siblings and supportive fiance

ACKNOWLEDGEMENTS
First, I would like to thank my research supervisors, Dr. William J. Brittain and
Dr. Roderic P. Quirk. Especially, Dr. Brittain I could not have imagined having a better
advisor and mentor for my PhD, and without your excellent guidance and continuous
support, knowledge and perceptiveness I would never have finished. I learnt so much
from you that I will always be indebted for all the research skills you have cultivated in
me. Thanks Dr. Quirk for your time and guidance you have provided me within the last
one year.
My special thanks to Dr. Scott Collins for agreeing to become chair of my
committee in a very short notice. I also like to thank him for his invaluable idea of
controlling the rate of the click reaction vs RAFT polymerization independently through
variation of Cu (I) concentration.

This concept is not only very useful in surface

modification but will also enhance the scope of this work in other fields. I also want to
thank my committee members Dr. Ali Dhinojwala, Dr. Coleen Pugh and Dr. Bi-min
Zhang Newby for their valuable suggestions and time.
There are many people to thank for their support and encouragement, without
which this milestone would be an unattainable dream. I want to thank my research
group, my friends in the department and my room mates who have made the last few
years an exciting and memorable experience.

vi

I express my deep thanks to my parents, my brother Rajeev and my sister Rashmi


for all their love and support. Thanks bro for all the encouragement and guidance you
have provided through out my academic career and for standing besides me when ever I
needed a backing. I also want to thank my fiance whom I met while I was in graduate
school. Thanks Garima for being my best friend, you have always motivated and inspired
me to work hard.

vii

TABLE OF CONTENTS
Page
LIST OF TABLES .....xi
LIST OF FIGURES...........xii
LIST OF SCHEMES.........xv
CHAPTER
I. INTRODUCTION....1
II. HISTORICAL BACKGROUND....................7
2.1. Silica nanoparticles.....................................................................................7
2.2. Surface modification of silica nanoparticles......8
2.2.1. Covalent attachment via grafting from approach.........................10
2.2.2. Surface modification using conventional radical polymerization ..11
2.2.3. Surface modification using controlled radical polymerization ...15
2.2.4. Ring-opening polymerization (ROP)...30
2.2.5. Ring-opening metathesis polymerization (ROMP) ....34
2.2.6. Anionic polymerization ..36
2.2.7. Cationic polymerization ..36
2.3. Characterization....38
2.3.1. Characterization of modified particles.38
2.3.2. Characterization of the degrafted polymer......40
viii

2.4. Applications of modified silica nanoparticles ..41


2.5. Surface modification of silica nanoparticles with polyacrylamide.......................42
2.6. RAFT polymerization of acrylamide....................................................................43
2.7. Click chemistry.............................................................................................44
2.8. Combination of click chemistry and living radical polymerization .47
2.8.1. Combination of ATRP and click chemistry 47
2.8.2. Combination of RAFT polymerization and click chemistry....48
III. EXPERIMENTAL....50
3.1. Materials.......50
3.2. Measurements...51
3.3. Combination of RAFT polymerization and click chemistry for surface
modification..... 52
3.3.1. Deposition of 3-bromopropyltrichlorosilane on silica nanoparticles...52
3.3.2. Synthesis of azide-modified silica nanoparticles.52
3.3.3. Esterification reaction of S-1-dodecyl-S-(,-dimethyl-acetic acid)trithiocarbonate .....53
3.3.4. Surface modification via grafting to approach.....54
3.3.5. Surface modification via grafting from approach ...56
3.3.6. Surface modification via novel tandem approach....58
3.4. Combination of ATRP and click chemistry for surface modification via
grafting to approach....59
3.5. Surface modification of silica nanoparticles using surface initiated
polymerization (grafting from approach)..........61
3.6. Surface-initiated ATRP of styrene on silica nanoparticles...64
3.7. Surface-initiated ATRP of MA using PS modified silica nanoparticles...64
ix

IV. RESULTS AND DISCUSSION ......65


4.1. Silica Nanoparticles..65
4.2. Combination of living radical polymerization and click chemistry for
surface modification...............66
4.3. Synthesis of click-functionalized RAFT chain transfer agent..67
4.4. Application of click-functionalized (alkyne terminated) RAFT chain
transfer agent.....70
4.5. Surface modification via grafting to approach..70
4.6. Combination of ATRP and click chemistry for surface modification via
grafting to approach82
4.7. Surface modification via grafting from approach.85
4.8. Tandem approach of surface modification...92
4.9. Surface modification of silica nanoparticles by surface-initiated atom
transfer radical polymerization.........99
V. SUMMARY AND CONCLUSIONS 108
REFRENCES...110
APPENDIX .117

LIST OF TABLES
Table

Page

2.1. Silane coupling agents used for grafting polymer chains on silica.........15
4.1. Elemental analysis of modified silica nanoparticles...79
4.2. Grafting densities of modified silica nanoparticles....92
4.3. Grafting density of PS on silica nanoparticles as prepared by different grafting
methods...97
4.4. Effect of Cu(I) concentration on tandem approach for surface modification.....99

xi

LIST OF FIGURES
Figure

Page

2.1. TEM image of silica nanoparticles..................7


2.2. Grafting to approach of surface modification................10
2.3. Grafting from approach for surface modification......11
2.4. Free radical initiators immobilized on nanosilica particles .......13
2.5. Mechanism for controlled free radical polymerization..........16
2.6. Metal catalyzed atom transfer radical polymerization. ..17
2.7. Structure of monosiloxane initiator used for immobilization on the surface. 19
2.8. Silane-functionalized asymmetric difunctional initiator for sequential
polymerization of PS and PtBA..................................................................................23
2.9. Reversible chain transfer in RAFT polymerization...26
2.10. RAFT-silane coupling agent. ...27
2.11. Structure of DEPN....29
2.12. Transition metal immobilized on silica particles......35
2.13. ROMP initiator used for immobilization on silica particles and the 1st generation
Grubbs catalyst (Grubbs I)....35
2.14. Schematic model for polyacrylamide-grafted silica.....43
4.1. TEM image of silica nanoparticles provided by Nissan Chemical.65
4.2. 1H NMR spectrum of click-functionalized RAFT CTA.69
4.3. 13C NMR spectrum of click-functionalized RAFT CTA........69
xii

4.4. 13C NMR spectrum of polyacrylamide...74


4.5. 1H NMR spectrum of polyacrylamide....74
4.6. MALDI-TOF-mass spectrum of polyacrylamide...75
4.7. FT-IR spectra of (a) PAAm, (b) silica-PAAm nanoparticles, (c) silica-azide
nanoparticles and (d) silica-bromide nanoparticles....77
4.8. Thermogravimetric analysis (TGA) traces of (a) bare silica nanoparticles,
(b) silica-bromide nanoparticles, (c) silica-azide nanoparticles and (d)
silica-PAAm nanoparticles.....78
4.9. Thermogravimetric analysis traces of (a) bare silica nanoparticles, (b)
silica-bromide, (c) silica-azide and (d) silica-PS-b-PMA (by RAFT)....81
4.10. FT-IR spectra of (a) bare silica nanoparticles, (b) silica-bromide, (c)
silica-azide and (d) silica-PS-b-PMA (by RAFT).......82
4.11. FT-IR spectra of (a) bare silica nanoparticles, (b) silica-bromide, (c)
silica-azide, (d) silica-PS (by ATRP) and (e) silica-PMA(by ATRP)..84
4.12. Thermogravimetric analysis (TGA) traces of (a) bare silica nanoparticles,
(b) silica-bromide, (c) silica-azide, (d) silica-PS (by ATRP) and (e) silicaPMA(by ATRP)........................................................................................................84
4.13. FT-IR spectra of (a) bare silica nanoparticle, (b) silica-bromide, (c) silicaazide, (d) silica-RAFT CTA, (e) silica-PS, and (f) silica-PS-b-PMA..90
4.14. TGA analysis traces of (a) bare silica nanoparticle, (b) silica-bromide,
(c) silica-azide, (d) silica-RAFT CTA, (e) silica-PS (Mn=6,000 g/mol)
and (f) silica-PS-b-PMA (Mn=34,000 g/mol).......91
4.15. TGA analysis traces of (a) bare silica nanoparticle, (b) silica-bromide,
(c) silica-azide, (d) silica-PS (Mn =9000 g/mol) by grafting to approach,
(e) silica-PS (Mn =7000 g/mol) by tandem approach and (f) silica-PS
(Mn =12000 g/mol) by grafting from approach.96
4.16. FT-IR spectra of (a) bare silica nanoparticle, (b) silica-bromide, (c)
silica-azide, silica-PS by (d) grafting to approach, (e) tandem approach
and (f) grafting from approach..97
4.17. Structure of BITES initiator....100
4.18. 1H NMR spectrum of BITES initiator101
4.19. 13C NMR spectrum of BITES initiator...102
xiii

4.20. XPS spectrum of BITES initiator modified silica nanoparticles....103


4.21. FT-IR spectra of modified silica nanoparticles...106
4.22. TGA traces of modified silica nanoparticles..106
4.23. TEM images of PS-coated silica particles..107
4.24. TEM images of PMA-coated silica particles..107

xiv

LIST OF SCHEMES
Scheme

Page

2.1. Synthesis of Stber silica particles...8


2.2. Stabilization of the silica particles............8
2.3. Surface modification of silica nanoparticles.....9
2.4. Synthesis of covalently attached polystyrene on silica using surface-immobilized
azo initiators................12
2.5. Synthesis of polymer grafted silica by ATRP.....18
2.6. Self-condensing vinyl polymerization (SCVP) of an AB* inimer from a
functionalized silica particle .......22
2.7. General scheme for RAFT polymerization.26
2.8. General scheme for nitroxide mediated polymerization.28
2.9. General scheme for ROP of cyclic monomers from the surface of silica
nanoparticles. ............................................................................................................31
2.10. Mechanism of ring opening polymerization of caprolactone (CLn)...32
2.11. Synthetic strategy for ROP of -caprolactone and L-lactide from GPS-grafted
silica with Al. Y, or Sn alkoxide as the catalyst...33
2.12. Modification of silica surface with polyisobutylene....37
2.13. Model click reaction.....45
2.14. Click coupling reactions using telechelic polymers prepared by ATRP. 48
2.15. Synthesis of click-functionalized RAFT CTA..49
4.1. Synthesis of click-functionalized RAFT CTA....67
4.2. Mechanism of carboxyl-terminated RAFT CTA synthesis....68
xv

4.3. Synthesis of alkyne-functionalized polyacryalamide.....71


4.4. Modification of silica nanoparticles via grafting to approach.....76
4.5. Surface modification using ATRP and click chemistry..83
4.6. Modification of silica nanoparticles via grafting from approach....88
4.7. Surface modification of silica nanoparticles using tandem approach.94
4.8. Synthesis of BITES ATRP initiator..101
4.9. Surface modification of silica particles via surface-initiated ATRP........102

xvi

CHAPTER I
INTRODUCTION

Surface functionalization of silica nanoparticles with polymer brushes is very


important as the polymer coating alters the interfacial properties.

Also, the

mechanical and thermal properties of the matrix polymers in hybrid systems can be
altered by the compatibility of the nanoparticles with the matrix. Nanoparticle
physical properties are governed by both the size and shape of the silica core and the
surrounding organic layer. Hybrid nanoparticles serve as an interesting example of
spherical brushes in which chains extend away from the colloidal core to minimize
steric crowding.1 These hybrid particles exhibit improved surface wettability and
lubricity and superior dispersion. The size of the particles range from 1 nm to several
m. These particles find a wide range of applications in optics, electronics,
engineering and biosciences.2
Generally the surfaces of inorganic materials are functionalized with polymer
chains either chemically (through covalent bonding) or physically (by physisorption).3
A drawback of physisorbed polymer is that they are thermally and solvolytically
unstable due to the relatively weak van der Waals forces or hydrogen bonding that
anchors them to the surface.4 Covalent grafting techniques are preferred to maximize
a stable interfacial compatibility between the two phases.

Covalent grafting

techniques involve either the grafting to or grafting from methods. In the

grafting to method, pre-formed, end-functionalized chains are reacted with a


chemically activated substrate to form a tethered polymer layer.5

Though

experimentally simple, the grafting to method has a number of limitations; most


notably the grafting densities are low because of steric crowding of reactive sites by
previously attached polymers.6 The limitations of physisorption and grafting to can
be circumvented by the grafting from technique. In the grafting from method,
surface immobilized initiators are used to grow polymers in situ to generate a polymer
brush. The grafting from technique results in significantly higher grafting density
because the steric barrier imposed by the in situ grafted chains does not limit the
access of smaller monomer molecules to the active initiation sites.7
Using the right system and techniques, one can control the functionality,
density and thickness of the polymer brushes. Preparation of polymer brushes on
silica can be accomplished by conventional free radical, controlled free radical,
cationic, anionic and ring-opening metathesis polymerization techniques.8
Among these techniques, controlled free radical polymerization is a versatile
method to engineer the silica particles surface. Controlled free radical polymerization
can produce polymers with controlled molecular weight, composition and functionality.9
Recent reports describe the use of controlled polymerization techniques, particularly
atom transfer radical polymerization (ATRP), to grow polymer chains from the surface of
silica nanoparticles.10 In the present study the surfaces of silica nanoparticles were
modified with diblock copolymer brushes using surface-initiated ATRP. It was proposed
that grafting polymer chains onto the silica surface will improve the compatibility
between different matrixes and a particle surface. Depending on the matrix environment

for the particles, it was predicted that the diblock copolymer brushes would
spontaneously rearrange to maximize interaction between the particles and matrix. Thus
it was speculated that diblock brushes modified particles could become universal fillers
by alternating their surface chemistry in response to different environments.
Reversible addition fragmentation techniques (RAFT) polymerization is among
the most versatile of the controlled/living radical polymerization techniques. RAFT
polymerization is compatible with most vinyl monomers, requires mild reaction
conditions and is free from metal contamination.11 Use of RAFT polymerization for
surface modification of silica particles is not common. Low grafting density and
difficulty in the synthesis of surface-bound RAFT chain transfer agent (CTA) are two
major problems that hinder the use of RAFT polymerization for surface modification.
Although there are a few reports on the application of RAFT techniques for silica
nanoparticle modification, 12-16 high grafting density remains challenging. For ATRP, the
grafting density of the surface bound initiator was between 1.8 to 3.7 groups/nm2,

10

which is three times higher than the highest grafting density reported for RAFT.15 The
polymer chain density is directly related to important material properties such as
hydrolytic stability and fundamental dimensions such as brush height.17

By taking

advantage of the properties of high density polymer brushes, one can create novel
materials with smart or responsive surfaces. There is a need to increase the grafting
density of polymer brushes and simplify the preparation of RAFT agents.
Sharpless and co-workers18 popularized the 1,3-dipolar cycloaddition of azides
and terminal alkynes, catalyzed by copper(I), in organic synthesis. These reaction have a
high thermodynamic driving force, which makes them one of the most efficient reactions

available. Such reactions are very practical, because they can be performed in high yield,
in multiple solvents (including water), and in the presence of many other functional
groups. Also, the 1,2,3-triazole product is chemically very stable. Due to their efficiency
and simplicity, these cycloadditions were classified as click reactions. Click
functionalized polymers can be used for functionalization and modification of a variety of
substrates.
Combining the chain-end functionality control of living free radical
polymerization and the efficiency and diversity of click chemistry is desirable. The
utility of the click reaction has been demonstrated in ATRP.19 Among controlled free
radical polymerizations, RAFT has arguably the most important commercial
significance.20 Little work has been done on combination of RAFT polymerization and
the click reaction. Hawker, Wooley and co-workers21 used RAFT polymerization to
synthesize alkyne functionalized block copolymers.

Sumerlin and co-workers22

synthesized functional telechelic polymers using this approach. Davis, Barner-Kowollik


and co-workers23 synthesized block copolymers via RAFT polymerization and click
chemistry. The possibility of combining RAFT polymerization and click chemistry as a
synthetic strategy for surface modification was explored. A novel click-functionalized
RAFT CTA was synthesized which opens up the possibility of using RAFT
polymerization and click chemistry together in surface modification. Combination of the
click reaction and RAFT polymerization was used for both the grafting to and grafting
from surface modification techniques.
In the grafting to approach, it was illustrated that a one step process to make
alkyne end-functional polymers via RAFT polymerization can be used in a multitude of

post-polymerization processes. These alkyne-functionalized polymers can participate in


different kinds of end functionalization reactions using click chemistry. Using the RAFT
CTA, the surfaces of silica nanoparticles were modified with homopolymer and diblock
copolymer brushes via the grafting to approach. A click reaction was used to attach
polymers on the surface, which ensured better grafting density. Tethered polymer chains
were found in the brush regime. Acrylamide (AAm) polymers are an important class of
materials because of applications such as coatings, flocculants, paper making, mining,
electrophoresis and biology. The unification of the sorption ability of a colloidal particle
surface and the binding properties of PAAm grafted silica produces a strong flocculative
capability.24 Although the use of PAAm has increased, there have not been many reports
on the preparation of silica-PAAm hybrid nanoparticles.24,25 The solubility of PAAm is
limited and presents an experimental challenge in the preparation of silica-PAAm hybrid
nanoparticles. Exploiting the versatility of RAFT polymerization and the efficiency of
click reaction, silica-PAAm hybrid nanoparticles were synthesized. An alkyne terminated
ATRP initiator was also employed to synthesize alkyne end functionalized polymer and
attach it to a silica surface via the click reaction.
We prepared high density, well-defined polymer brushes on silica nanoparticles
via the grafting from approach. Our strategy included (1) the grafting from approach
for brush formation, (2) a facile click reaction to immobilize the RAFT agent, (3)
synthesis of R-supported chain transfer agent (R group of RAFT CTA (R-S-(C=S)-Z)
attached to the surface) over Z-supported chain transfer agent (Z group of RAFT CTA
attached to the surface) and (4) the use of a more active trithiocarbonate RAFT agent
over the conventional dithioester type RAFT agent. Grafting density obtained by this

method was significantly higher than corresponding reported values for RAFT polymers
in the literature. Homopolymer and block copolymer grafted silica nanoparticles were
prepared and characterized.
The option of conducting tandem RAFT polymerization and click chemistry in
one pot was investigated. A tandem process simplifies this dual chemistry and can
potentially save time and cost. This tandem method was used as a novel method of
surface modification. We applied this strategy to the surface modification of silica
nanoparticles and controlling the grafting density of tethered polymer chains.
Simultaneous RAFT polymerization and click chemistry were performed in a flask
containing styrene, azide-modified silica, alkyne-terminated RAFT CTA, CuBr/
N,N,N,N,N-pentamethyldiethylenetriamine (in a catalytic amount) and toluene at 90
0

C under a inert atmosphere for 18 h. The click (alkyne) functionalized CTA participated

in RAFT polymerization as well as the click coupling reaction with azide-modified silica
particles. To the best of our knowledge, this is first report where RAFT polymerization
and click coupling reactions are conducted in tandem, thus demonstrating mutual
compatibility. We applied this tandem approach to modify the surface of silica particles
with polystyrene. As deduced by thermal gravimetry and elemental analysis, the grafting
density of PS brushes on silica obtained by the tandem method (0.49 groups/nm2) was in
between the analogous grafting to method (0.27 groups/nm2) and the grafting from
method (0.65 groups/nm2). We also altered the relative rate of both reactions to control
grafting density.

CHAPTER II
HISTORICAL BACKGROUND

2.1. Silica nanoparticle


Among the numerous inorganic/organic hybrid materials, silica-polymer
hybrid materials are one of the most commonly reported in the literature. This may
be attributed to their wide use and the ease of particle synthesis. Silica nanoparticles
have been used as fillers in the manufacture of paints, rubber products, and plastic
binders.25 Silica particles coated with organic modifiers are used in applications that
include stationary chromatography phases,26 heterogeneous supported catalysts,27 and
in the automotive, electronics, appliance, consumer goods,28 aerospace and sensor29
industries.
Colloidal silica is of particular interest due to the ease of synthesis and precise
control of the size and distribution of the particles.31

Stber and co-workers32

reported a simple synthesis of monodisperse spherical silica particles. The synthesis


proceeds with the hydrolysis and condensation of tetraethyl orthosilicate (TEOS) in a
mixture of alcohol, water and ammonia (catalyst). In general, the hydrolysis reaction
gives the slightly hydrolyzed TEOS monomer (Scheme 2.1, Equation 1). This
hydrolyzed intermediate undergoes condensation to eventually form silica according
to Scheme 2.1 (Equation 2). The resultant particles are stabilized by electrostatic
repulsion due to the ions in the ammonia solution (Scheme 2.2).

(RO)3Si(OH) +

Si(OR)4 + H2O
(RO)3Si(OH) + H2O

Equation 1

ROH

Equation 2

SiO2 + 3ROH

Scheme 2.1. Synthesis of Stber spherical silica particles.

Si(OCH2CH3)4

HO

NH3, H2O

Silica

CH3CH2OH

TEOS

O-NH4+
OH

O-NH4+

HO

OCH2CH3

Scheme 2.2. Stabilization of the silica particles.

2.2. Surface modification of silica nanoparticles


The chemical properties of the silica surface are mainly determined by the
various silanol and siloxane groups that are present on the external as well as the
internal structure. The hydroxy groups on the surface of silica particles can be easily
tailored with organic compounds or polymers.

Silanol groups can be easily

functionalized by different chemical procedures. The most convenient technique for


silica surface functionalization is the use of the reaction of silanol groups with

OOH
O-

R Solvent
X

Si
X

Y
-O
OO
-O
O
SiO
2
Si
HO
O
-O
O
O
O-O
Y
-O

HO
HO

X = -OC2H5 , -OCH3, -Cl

O-

OH

HO
OH
HO
OH
HO SiO2
OH
OH
HO
OH
HO

HO

OH
OH
OH

suitable silane reagents (Scheme 2.3).33

Scheme 2.3. Surface modification of silica nanoparticles.

Y = H or Si
CH2

CH2

Generally the surfaces of inorganic materials are functionalized with polymer


chains either chemically (through covalent bonding) or physically (by physisorption).3
Physisorption involves adsorption of polymers with sticky segments.

The non-

covalent adsorption makes the adsorption reversible, especially during processing,


and is not a favored technique.4

Covalent grafting techniques are preferred to

maximize a stable interfacial compatibility between the two phases. There are two
methods for chemically attaching polymer chains to a solid substrate: the grafting to
and grafting from techniques. Polymer films anchored in this manner are inherently
more thermally and solvolytically stable. In the grafting to method, pre-formed,
end-functionalized polymer chains are reacted with a chemically activated substrate
(Figure 2.1).5

One advantage of this method is that polymer chains can be

characterized before being attached to the substrate. The drawback, however, is that
only relatively low grafting densities are obtained due to steric crowding of already
attached chains on the surface, which hinder diffusion of additional chains to reactive
sites.6 Reactive end groups must diffuse through a barrier of existing polymer film to
couple with functional groups on the surface and this diffusion barrier becomes more
pronounced as more chains are attached. This effect was noticed by Yoshikawa and
Tsubokawa34 when it was discovered that the number of grafted chains decreased as
the molecular weight of the chains increased. Though brushes have been prepared via
the grafting to method, more densely grafted brushes may be prepared by using the
grafting from method to circumvent the diffusion barrier problem.

B
A

A A

A
A

A
A

A A

Figure 2.1. Grafting to approach of surface modification.

2.2.1. Covalent attachment via grafting from approach


The grafting from method involves formation of an initiator layer (I) on the
surface of the silica followed by polymerization of monomer (M) (Figure 2.2).
Different polymerization techniques can be used to modify the surface of silica.
Thick brushes with a high graft density can be formed because monomer can easily
diffuse to reactive sites of the growing polymer chains. In this method, the steric
barrier to incoming polymers imposed by the in situ grafted chains does not limit the
access of smaller monomer molecules to the active initiation sites.7

This

polymerization technique is also commonly referred to as surface-initiated


polymerization. Preparation of polymer brushes via the grafting from technique on
silica nanoparticles can be accomplished by conventional free radical, controlled free
radical, cationic, anionic and ring-opening metathesis polymerization techniques.8

10

I
I

*
M

I
I

I
M

M
M

Figure 2.2. Grafting from approach for surface modification.

2.2.2. Surface modification using conventional radical polymerization


Conventional free radical polymerization is one of the most studied systems.
Radical processes are more tolerant of functional groups and impurities and are well
suited for polar monomers. It is a leading industrial method to produce polymers.16
The grafting from method has long been used for the preparation of covalently
attached polymers by free radical techniques. This approach was first reported by
Prucker and Rhe.35,36 It consisted of grafting an azo initiator onto a particle or flat
surface followed by polymerization. A self-assembled monolayer (SAM) of the azo
initiator was grafted on the surface of silica and this was used for the radical chain
polymerization of styrene (Scheme 2.4).

11

Scheme 2.4. Synthesis of covalently attached polystyrene on silica using surfaceimmobilized azo initiators.

High molecular weight polymer brushes with high graft density were
successfully produced. Kinetic investigation of the polymerization process revealed
that the initiation and propagation of the polymer at low conversion were similar to
that of solution polymerization.

However, different kinetics of termination were

observed. Spontaneous bimolecular termination was reduced for the surface-initiated


processes because free chains had to diffuse against a concentration gradient into the
film to achieve termination.

12

Recently, Ueda and co-workers37 investigated the radical polymerization of


vinyl monomers on a surface. Two different methods were used to introduce the
initiator on the surface of the silica. In one case the azo group was introduced by the
reaction of surface amino groups (introduced by pretreatment with 3-aminopropyltriethoxysilane) with 4,4-azobis(4-cyanopentanoyl chloride).

In the second case,

surface immobilization was achieved by reaction of surface amino groups with tbutylperoxy-2-methacryloxyethylcarbonate through a Michael addition. Thus silica
nanoparticles coated with an azo initiator (Figure 2.3A) and peroxy initiator were
prepared seperately (Figure 2.3B). The radical chain polymerization was initiated by
thermal decomposition.

Higher grafting density (compared to Prucker and Rhe

approach) was achieved and the formation of free polymer was considerably reduced.

CH3

R NH C (CH2)2 C N
CN

CH3
N C (CH2)2 COOH
CN

(A) Silica- Azo

CH3
R NH CH2 CH C O CH2 CH2 O C O OtBu
O

(B) Silica-Peroxy

Figure 2.3. Nanosilica particles immobilized with free radical initiators.

13

Molecularly imprinted polymers (MIPs) are highly crosslinked polymers with


recognition towards a target molecule or class of molecules.

Sellegren and co-

workers38 reported the grafting of films of MIPs using surface-bounded free radical
initiators on a 10 nm silica particle.

The grafting resulted in tightly packed

agglomerates interconnected by the polymer.


A different approach to MIPs to achieve graft polymers on silica surfaces
using radical polymerization involved the immobilization of a silane coupling agent
bearing a polymerizable double bond. This was then used as a comonomer for the
free radical polymerization of vinyl monomers in an emulsion or dispersion medium.
The monomers studied and the commonly used coupling agents are in Table 2.1.39-43
Well-defined, core-shell structures comprised of silica-core and the polymeric shell
were attained. The density of the coupling agent as well as the size of silica particle
have a strong effect on the particle morphology.
In conventional free radical polymerizations, propagation is a pseudo first
order reaction while termination is second order in chain end radical concentration.9
Thus, the proportion of termination to propagation increases substantially with
increasing free radical concentrations. Chain transfer and termination are impossible
to control in classical free radical polymerization.

These polymerizations are

generally characterized by broad molecular weight distributions, poor control of


molecular weight and chain end functionality, and the inability to synthesize welldefined block copolymers. However, many of these drawbacks can be overcome by
using controlled free radical polymerization.

14

Table 2.1. Coupling agents used for grafting polymer chains on silica.39-43
No

Coupling reagent

1.

Polymer
OCH3

CH2

CH2 CH2 CH2 Si OCH3

OCH3

CH3

PS, PMAA,
PS-coPBuA,

3-methacryloxypropyl trimethoxysilane
(MPS)
2

OC2H5

CH2
O

PS

CH2 Si OC2H5
OC2H5

CH3

Methacryloxy methyltriethoxysilane (MMS)


3

H3C(H2C)7

(CH2)7COOH

PS

Oleic acid

2.2.3. Surface modification using controlled radical polymerization


The use of controlled/living radical polymerization has proven to be a versatile
approach for incorporating different types of organic polymers with varied
architectures on silica surfaces. By using this technique, one can manipulate the
structure of the resultant polymer through changes in grafting density, composition
and molar mass.

Surface grafting polymerization via living free radical

polymerization route is preferred for at least two reasons.44 First, the solid surface

15

onto which initiating groups are placed confers a mobility barrier for termination by
coupling. Second, a limited number of initiating groups need be attached to the
surface to promote property changes, thereby reducing the concentration of free
radicals present in the system (controlled free radical polymerization works by
maintaining a low concentration of radicals present at any one time during the course
of the reaction).
activation
Active polymer radical

Dormant polymer
deactivation

Monomer
(Propagation)

Capping
group

Figure 2.4. Reversible chain transfer in controlled free radical polymerization.

In controlled radical polymerization, an initiator generates radicals, which


combine with monomer molecules to form chains. The radicals on the ends of the
polymeric chains are reversibly deactivated by the presence of capping groups
(Figure 2.4). Reaction of a radical polymer chain with a capping group eliminates the
radical, temporarily stopping polymerization. The loss of the capping group allows
propagation to continue. The life-time of the growing radicals is controlled, resulting
in the synthesis of predefined molar mass polymers with low polydispersity,
controlled composition, and functionality. A typical product has a polydispersity less
than 1.3.

16

In general, controlled/living polymerization can be achieved by stable freeradical polymerization, e.g. nitroxide-mediated processes (NMP), metal-catalyzed
atom transfer radical polymerization (ATRP) and degenerative transfer, e.g. reversible
addition-fragmentation chain transfer (RAFT) polymerization.9

2.2.3.1. Atom transfer radical polymerization (ATRP)


Matyjaszewski45 and Sawamoto46 simultaneously reported ATRP in the mid
1990s. It is an extension of the Kharasch addition47 reaction commonly referred to
as atom transfer radical addition (ATRA). Since then it has been a useful tool for the
synthesis of polymers of different architecture and morphology.

P-X + Cu(I)Lm

kact
kdea

P.

+M

kt
kp

X-Cu(II)Lm

P.

P-P

Figure 2.5. Copper-catalyzed atom transfer radical polymerization.

In ATRP, dormant polymer chains are capped with a halogen. The addition of
a metal catalyst and a ligand (L) removes the halogen from the chain, generating an
alkyl radical, activating the chain, and oxidizing the catalyst (Figure 2.5).

This

process is reversible; the oxidized catalyst may combine with the chain and recap it
with the halogen.

Propagation occurs when active chains react with monomer

molecules. Termination occurs when two active chains combine and is minimized by

17

adjusting the reactant concentration so that the majority of chains are in the
deactivated state at any given time.
The ATRP initiator (typically an -halo ester or an -benzyl moiety) is activated
in the presence of metal salts such as those of Cu, Ru, Fe, and others. The solubility and
activity of the metal salts are enhanced by ligation with aliphatic or aromatic amines.
ATRP is tolerant of functional groups and impurities and hence a wide range of
monomers can be polymerized in organic as well as aqueous phases.

This facile

polymerization and less stringent experimental conditions promoted the application to


brush growth on nanoparticles, especially silica.48
Patten and von Werne49 reported the functionalization of silica nanoparticles
with polymer brushes using ATRP. Their approach consisted of immobilization of
the initiator molecule on the silica particle followed by surface-initiated ATRP
(Scheme 2.5).9

R= Initiator
THF, 80oC

R
R

R
R

R
R
R

R
R
R
R
R
R

[M]0, CuX / 2L
90-110oC

Scheme 2.5. Synthesis of polymer grafted silica by ATRP.


The silane coupling agents [2-(4-chloromethylphenyl)ethyl]dimethylethoxysilane (CPTS), [3-(2-bromo-isobutyryl)propyl]dimethylethoxysilane (BPDS), and [3(2-bromopropionyl)propyl]dimethylethoxysilane (BIDS) were used for initiator
immobilization on the silica surface (Figure 2.6).

18

C2H5O Si CH2 CH2


CPTS

CH2Cl

C2H5O Si

O
Br

BPDS

C2H5O Si

Br

BIDS

Figure 2.6. Structures of monosiloxane initiators used for immobilization on the surface
of silica.9,47

The initiator-grafted silica particles were then used to perform the ATRP of
styrene and methyl methacrylate (MMA) under different conditions. They showed
that polymerization from smaller particles with a 75 nm diameter exhibited a higher
degree of control, which was lacking when larger particles (300 nm) were used.
However, the addition of deactivator or a small amount of free initiator helped in
controlling the polymerization on larger particles. Their study also showed that the
larger particles with their small quantities of initiation sites exhibited the kinetic and
molecular weight evolution as that from the flat substrates.50
Surface-initiated ATRP has been exploited to control the interparticle
aggregation of nanoparticles by grafting polystyrene chains onto particles.51 The
silica particles were maintained in an organic solution throughout the surface
modification; irreversible aggregation is often observed if the particles are redispersed
after solvent removal.

Matyjaszewski and co-workers52 reported similar work to

Patten and von Werne49 where they used Cu(II)Br2 as the deactivating transition metal
species. Styrene, n-butyl acrylate and MMA monomers were used for this study and
various core-shell colloids containing tethered AB diblock copolymers were
synthesized.

19

Armes and co-workers53 observed that several hydrophilic monomers can be


polymerized via ATRP under mild conditions in aqueous media. They extended this
process to perform surface modification on silica particles.54

Several hydrophilic

methacrylate monomers were used to explore the feasibility of using aqueous ATRP at
ambient

temperature.

glycol)methacrylate]
(PMEMA).

The

polymers

(POEGMA)

and

investigated

were

poly[oligo(ethylene

poly[2-(N-morpholino)ethyl

methacrylate]

The colloidal stability of these sterically stabilized particles was

investigated; several thermoresponsive systems were reported. POEGMAsilica particles


were stable from 20-65 oC because these polymer chains remain hydrophilic over this
temperature range.

However PMEMA modified particles displayed the onset of

aggregation at 34 oC, which corresponds to the lower critical solution temperature


(LCST) of this polymer coating.
Polyelectrolyte-grafted silica particles were also used in studies related to the
colloidal stability of the nanoparticles.55 Monodisperse silica particles of around 300
nm were surface-grafted with suitable ATRP initiators, followed by surface initiated
ATRP of ionic monomers such as sodium 4-styrenesulfonate, sodium 4-vinylbenzoate, 2-(dimethylamino)ethyl methacrylate (DMAEMA) and 2-(diethylamino)ethyl methacrylate (DEAEMA) in protic solvents.

In principle, polyelectrolyte

brushes should enhance the colloidal stability in aqueous medium due to the
steric/electrostatic repulsive forces. Investigation into the pH-responsive colloidal
stability demonstrated that cationic polyelectrolyte-grafted silica particles were
colloidally stable at low or neutral pH, but became aggregated at higher pH. However
a reverse effect was found for anionic polyelectrolyte-grafted silica particles. Hence

20

the stability of these particles was determined not only by the polyelectrolyte brushes,
but was also highly pH dependent, as might be expected based on the pH-dependence
of ionization.
Vairon and co-workers56 studied the growth of poly(n-butyl acrylate) (nBA)
brushes from the surface of 12 nm silica particles. The use of nBA produced a
particle with a hard silica interior and an elastomeric shell, which produced films with
enhanced mechanical properties.
High density PMMA brushes with molecular weights near 500,000 g/mol have
been successfully grafted onto the silica nanoparticles using surface-initiated ATRP.57
(2-Bromo-2-methyl)propionyloxyhexyltriethoxysilane was used as an ATRP initiator
to achieve well-defined, high density PMMA brushes. These particles were found to
have an exceptionally good dispersibility in organic solvents. Observation of the
monolayers by transmission electron microscopy (TEM) and atomic force microscopy
(AFM) showed that they formed an ordered 2-dimensional lattice at the air-water
interface throughout the monolayer. The controlled nature of the polymerization was
confirmed by the similar molecular weights and low polydispersity of both the free
polymer from the solution and the polymer cleaved from the surface of the silica
using HF. Boetcher and co-workers58 focused on the same ATRP procedure and
reported on the graft density of a first and a second generation (block) of polystyrene
grafted from silica. They showed that approximately 10-15% of the first generation
of the grafts were not active for reinitiation of the second monomer feed.
Mller

and

co-workers59

grafted

hyperbranched

polymer

nanoparticles by self-condensing ATRP from a silica surface.

21

on

silica

An -bromo type

ATRP initiator was first attached to silica particles (0.8-1.6 initiator molecules/nm2)
followed by a self-condensing vinyl polymerization (SCVP) of an AB* inimer
(Scheme 2.6),which had both a polymerizable acrylic group and an initiating group in
the same molecule. Well-defined polymer chains were grown from the surface to
yield hybrid nanoparticles comprised of silica cores and hyperbranched polymer
shells having multifunctional bromoester end groups. The relatively high weight
fraction and the surface coverage of the grafted polymers were observed for SCVP.
Self

condensing-vinyl

copolymerization

(SCVCP)

in

the

presence

of

the

functionalized silica particles was also used to prepare functional branched polymersilica hybrid nanoparticles, in which one could tune the architecture, chemical and
physical properties, and particle morphology, by the choice of comonomers and their
composition in the feed. Using SCVCP of inimer and tert.-butyl acrylate followed by
hydrolysis, branched poly-(acrylic acid)-silica hybrid nanoparticles were created.
A b

A B*

B*

B*

B*

A B*
SiO2

A*

A B*
b

SiO2

A B*

A B*

A*

B*

B*

*B
A*B*
A*B*

A*
A*B*

A*

A B*

kA

SiO2
b

A B*

A B*

kB

SiO2

A*B*

A B*

B*

A*

A*B*

A*B*
A*B*

A*

A B*

A*B* A*B*

A*
A*B*

B*

A*

B*

A*

A B*

A B*
SiO2

A*B*
B*

kB

A*B*

A*B*

A*

B*
O
SiO2

Br

O Si

Br

SiO2

B*

AB*

Scheme 2.6. SCVP of an AB* inimer from a functionalized silica particle.

22

A novel approach to attain an environmentally responsive hairy nanoparticles


was achieved by Zhao and co-workers60 using a combination of surface initiated
ATRP and NMP. An asymmetric difunctional initiator having structure shown in
Figure 2.7 was used to sequentially synthesize mixed poly(t-butyl acrylate)
(PtBuA)/PS brushes. PtBuA was converted to poly(acrylic acid) (PAA) by acidic
hydrolysis. Kinetic and molecular weight analysis confirmed the controlled nature of
the polymerization. Chain reorganization of the mixed brushes was studied in blockselective solvents using

H NMR spectroscopy.

H NMR spectroscopy was

performed in DMF-d7, a good solvent for both PS and PAA, and showed
characteristic peaks at 6.0-7.4 ppm corresponding to styrene and 2.45 ppm
corresponding to PAA. However the peak at 2.45 ppm decreased upon addition of
CDCl3 suggesting the collapse of PAA chains and the mobility of PS chains. A
similar decrease was observed for the aromatic PS peaks on addition of CD3OD.
These studies suggested a reorganization of the brushes in response to the
environment. The particles also formed stable dispersions in CHCl3, (a selective
solvent for PS) and in methanol, a selective solvent for PAA.

H3C
Br
O C
CH3
C
O
Cl

Si

O
N

Figure 2.7. Silane-functionalized asymmetric difunctional initiator for sequential


polymerization of PS and PtBA.
23

2.2.3.2. Reverse ATRP


Reverse ATRP is the polymerization using a traditional radical initiator, e.g.
2,2-azobisisobutyronitrile (AIBN), in the presence of a copper/ligand complex. It
differs from conventional ATRP in the initiation process.

The primary radical

abstracts a halogen atom from the catalyst/ligand complex and forms a dormant halide
species and the reduced transition metal species activator. Peroxide groups were
introduced on the surface of silica particles to perform reverse ATRP of MMA.61
Controlled polymerization was achieved in cyclohexanone as solvent at 70 oC.
Detailed GPC study of the PMMA polymer by dissolution of the silica by treatment
with HF showed the controlled nature of the polymerization.

2.2.3.3. Reversible addition-fragmentation chain transfer (RAFT) polymerization


RAFT was reported in 1998 by Rizzardo and co-workers.20 The mechanism of
RAFT polymerization involves a sequence of additionfragmentation equilibria as shown
in Scheme 2.7 and Figure 2.8. In RAFT polymerization, a propagating chain is capped
when the radical attacks the carbon-sulfur double bond of the RAFT agent (Scheme
2.7, compound 1). RAFT CTA possesses two important areas of functionality, known as
the R group and the Z group. Each group plays an important role in the RAFT process.
The R group has two primary purposes; first it must be a good free radical leaving
group, second, it must reinitiate polymerization readily. Like the R group, the Z group
also has two primary purposes, these are to activate the C=S bond for radical addition and
(most importantly) to stabilize the intermediate radical adduct. RAFT polymerization
requires selection of a suitable chain transfer agent, which has a very high transfer

24

constant in radical polymerization of the desired monomers under appropriate reaction


conditions. Initiation and radicalradical termination occur as in conventional radical
polymerizations.

After initiation, addition of a propagating radical (Pn) to the

dithiocarbonyl compound (Scheme 2.7, compound 1) occurs.

This is followed by

selective fragmentation of the intermediate radical resulting in a polymeric


thiocarbonylthio compound (Scheme 2.7, compound 3) and a new radical (R). Reaction
of the radical (R) with monomer forms a new propagating radical (Pm).

Rapid

equilibration between the active propagating radicals (Pn and Pm) and the dormant
polymeric thiocarbonylthio compounds (Scheme 2.7, compound 3) provides an equal
probability for all chains to grow, and allows for the production of polymers with narrow
polydispersities.

At the end of the polymerization, most chains retain the

thiocarbonylthiolate end group.


Compared with ATRP and NMP, RAFT has some advantages, including: (1) it
can be successfully carried out under a wide range of reaction conditions (bulk,
solution, suspension, emulsion) and give polymers with a narrow molecular weight
distribution; (2) it is applicable to a wide range of functionality in monomer types
(e.g., OH, -COOH, -CONR2, -NR2, -SO3Na); and (3) the majority of chains in the
product polymer possess the S=C(Z)S- group and thus polymerization can be
continued in the presence of a second monomer resulting in a block copolymer.

25

Initiation and Propagation


Initiator

Pn

Reversible Chain transfer


Pn
M

S R

+ S

Pn

S R

Pn

k-

k-add

kp

kadd

Z
2

Reinitiation
M
k1

R M

Pm

Chain equilibration and Propagation


Pm

S Pn

+ S

kp

S Pn

Pm S
Z

Pm S

S
Z

Termination
Pn + Pm

kt

Dead polymer

Scheme 2.7. General scheme for RAFT polymerization.

Polymer radical

RAFT CTA

Figure 2.8. Reversible chain transfer in RAFT polymerization.

26

Pn
M

kp

Unlike flat surfaces and gold nanoparticles, there are very few reports on
RAFT polymerization to synthesize polymer brushes on silica nanoparticles. A few
papers have been published to date. A RAFT-mediated polystyrene brush growth on
silica was reported by Tsuji and co-workers.14 An ATRP initiator was anchored onto
a silica particle and polystyrene brushes were initially grown. The halogen end group
from ATRP polymerization was converted to the dithiobenzyl group for the RAFT
technique reagent by reacting 1-phenylethyldithiobenzoate in the presence of
CuBr/4,4-di-n-heptyl-2,2-bipyridine (dHbipy) complex.

The resulting silica

particles with anchored PS-RAFT agent were used to perform polymerization of


styrene in bulk in the presence of added free RAFT agent to effectively control the
polymerization. This multi-step procedure was totally avoided by Benicewicz and coworkers.15 They reported the first synthesis of a RAFT-silane coupling agent (Figure
2.9), which was used directly to prepare brushes on silica nanoparticles.
S

H O

CH3

C S C C O (CH2)3 Si OCH3
CH3
CH3

Figure 2.9. RAFT-silane coupling agent.


A grafting density of 0.15-0.68 RAFT agents/nm2 was achieved upon
immobilization of the RAFT-silane coupling agent on the silica particle.
Polymerization was then conducted at low conversions to avoid gelation or
interparticle radical coupling. Homopolymer brushes of PS-silica, and PnBuA-silica
were synthesized and the block copolymer of PS-b-PnBuA-silica was successfully
synthesized.

27

2.2.3.4. Nitroxide-mediated polymerization (NMP)


NMP, also referred to as stable free radical polymerization (SFRP), is based on
the concept of an activation-deactivation equilibrium between dormant species and a
small fraction of propagating macroradicals.

The reactions are done at elevated

temperatures where the initiation is rapid and all the chains are formed at the same
time. The initiated polymer chains are reversibly capped by a stable free radical to
give a dormant polymer chain (Scheme 2.8).
M
kp
R1
CH2 CH O N

R2

R1

ka
kd

CH2 CH

+ O N

R2

Scheme 2.8. General scheme for nitroxide-mediated polymerization.

SFRP is usually mediated by nitroxides as outlined by Georges62 and Hawker


and Malmstrm63 The nitroxides can reversibly react with the growing chain but do
not initiate polymerization. Initiators are usually peroxides, azo compounds, and
redox systems commonly used in conventional free radical polymerizations. SFRP
was initially restricted to the high-temperature polymerization of styrene.
SFRP has certain advantages over the other controlled polymerization
techniques; it does not suffer from the gel effect and it is clearly suitable for
industrial-scale production.64 Bartholome and co-workers65 extensively studied the
NMP of styrene on functionalized silica particles of 13 nm diameter.

28

EtO

P O
OEt

Figure 2.10. Structure of DEPN.


In

this

work,

N-tert-butyl-N-(1-diethylphosphono-2,2-dimethylpropyl)-

nitroxide (DEPN), was used as the NMP initiator (Figure 2.10). DPEN functionalized
with a silane group was grafted onto a silica particle under an argon atmosphere at 60
o

C for 24 h. A relatively low grafting density of 0.95 mol/m2 was reported. This

low density was attributed to the steric crowding by the bulky initiator. The NMP of
the styrene was performed in toluene; free nitroxide initiator was added to ensure
good control over the polymerization. Molecular weight and polydispersity (PDI) of
the free and grafted polymers were agreed well; a low PDI was consistent with a
controlled polymerization.
In a later study the same research group investigated two different routes for
grafting

alkoxyamine

initiator

onto

silica.66

In

one

route,

(acryloxypropyl)trimethoxysilane (APTMS) was grafted onto silica and DEPN was


trapped by the acryloxy radical formed in the presence of AIBN. In a second route an
alkoxyamine grafted silica particle was obtained by simultaneous reaction of DEPN,
AIBN and APTMS. Though these procedures eliminated the multi-step synthesis of
the functional alkoxyamine, no increase in the grafting density was observed.
Polystyrene brushes with a Mn = 60,000 g/mol were successfully synthesized on the
silica particles.

29

A polymerization compounding approach to graft DEPN radical was taken by


both Kasseh and co-workers67 and Parvole and co-workers.68 A free radical azo or a
peroxide initiator was initially immobilized on the silica surface. The NMP process
was then initiated from these initiators. Styrene and acrylate brushes were
successfully synthesized using these procedures. Blomberg and co-workers69 used the
concept of surface-initiated living polymerization to synthesize hollow nanoparticles.
Trichlorosilane-functionalized alkoxyamine initiator was immobilized onto silica
particles. Polystyrene brushes were grown using NMP. The polymer shell was then
cross-linked thermally or chemically. Crosslinking was achieved by incorporation of
benzocyclobutene group or chemically by the reaction of maleic anhydride repeat
units with a diamino crosslinker. The dissolution of the silica with HF created hollow
polymeric nanocapsules.

2.2.4. Ring-opening polymerization (ROP)


Radical polymerizations are limited to vinyl monomers. However for medical
applications,70 there is a need for biocompatible and biodegradable polymers such as
polylactides and polylactones. These polyesters are synthesized by the ring-opening
polymerization of cyclic monomers catalyzed or initiated by organometallic
derivatives.71 Many of these processes are living/controlled methods suitable for the
controlled grafting of polymers from surfaces.

Surface-initiated, ring-opening

polymerization is generally performed after the formation of a self-assembled


monolayer (SAM) that contains a polymerization initiator. In the case of aliphatic

30

polyesters, usually a hydroxyl- or amine terminated SAM is immobilized on the


surface to initiate ring-opening polymerization (Scheme 2.9).

OH

OMe
X (CH2)n Si OMe
OMe

O Si (CH2)n X
X = OH, -NH2

O Si (CH2)n X

1. AlEt3

O Si (CH2)n X

2. n CL
3. H3O+

Scheme 2.9. General scheme for ROP of cyclic monomer caprolactam (CL) from the
surface of silica nanoparticles.

Polyester grafting by ring-opening polymerization of lactones (-caprolactone)


was reported by Carrot and co-workers.72 The covalent grafting of the polyester chain
was achieved by the surface functionalization of the silica surface with a
trimethoxysilane agent containing amine functionality.

The initiation process

occurred selectively from the surface by activation of the amine groups by an


aluminum alkoxide derivative.

The polymerization proceeds by a coordination-

insertion mechanism, involving the coordination of the carbonyl group of the caprolactone by triethylaluminum, followed by the nucleophilic addition of the amine
(Scheme 2.10). A controlled polymerization was achieved and the initial monomerto-amine ratio determined the molecular weight of the resultant polyester.

31

The

polyester-grafted silica particles were easily dispersed in organic solvents such as


THF and CHCl3.

NH2

Et
Et

NH2

AlEt3

Et
Al

NH2

O
(-C2H6)

H
N

C O

(CH2)5 O

AlEt2

(-C2H6)

AlEt2 O

AlEt2 O
N

CLn

AlEt3

(CH2)5 O

AlEt2

n CLn

(CH2)5 O

AlEt2

H3O+
O
H
N

(CH2)5 O

Scheme 2.10. Mechanism of ring opening polymerization of caprolatone (CLn).

Choi and co-workers73 investigated the use of tin(II) octanoate to produce


poly(p-dioxanone) (PPDX) and poly(1,5-dioxepan-2-one) (PDXO), a biodegradable,
aliphatic poly(ether-ester). These were prepared by the ring-opening polymerization
of p-dioxanone and 1,5-dioxepan-2-one.

For the PPDX brush, an oligo(ethylene

glycol) terminated SAM was used to initiate the polymerization.

Field-emission

scanning electron microscopic (FE-SEM) images showed an increase in particle size


with the polymerization that confirmed the surface-initiated polymerization. In a later

32

report a different approach was taken; the silanol groups of the silica particles were
used to initiate polymerization. The FE-SEM images of the PDXO-silica particles
were spherical with no aggregation.

A brush thickness of around 140 nm was

achieved, which was confirmed by thermogravimetric analysis (TGA) and infrared


(IR) spectroscopy. A detailed study exploring the ROP of -caprolactone and Llactide with aluminium, yttrium and tin alkoxides on silica nanoparticles has been
reported by Joubert and co-workers74 (Scheme 2.10).

OH

Prehydrolyzed
GPS

O Si (CH2)3 O CH2 CH OH
CH2 OH

2M(OR)3
O Si (CH2)3 O CH2 CH O M
CH2 O M

OR
OR
OR

+ 2ROH

OR

Scheme 2.11. Synthetic strategy for ROP of -caprolactone and L-lactide from GPSgrafted silica with Al, Y, or Sn alkoxide as the catalyst.

According to their study, hydroxyl groups were introduced on the surface by


grafting of pre-hydrolyzed 3-glycidoxypropyl trimethoxysilane (GPS) at 100 oC for 1
h. A high grafting density of 3.4 mol/m2 was achieved. Catalytic polymerization of
the cyclic monomers was initiated in the presence of the hydroxyl group by the
different metal alkoxides. Free and grafted polymer chains were formed, in amounts

33

governed by the proportion of the metal atoms and the alcohol groups. The metal
oxides were highly influential in the grafting process. Tin(II) octanoate [Sn(Oct)2]
provided a high grafting density while negligible amounts of polymer were grafted in
the presence of the highly efficient Yttrium catalyst. This observation was attributed
to the slow alcohol-alkoxide interchange reaction due to the reduced mobility of the
initiator immobilized on the silica surface. They extended this strategy to introduce
poly(ethylene oxide) (PEO) onto silica nanoparticles by in situ ROP of oxirane
monomers in the presence of aluminum isopropoxide [Al(OiPr)3] as the initiator.
High [alcohol]/[metal] ratios used in this study resulted in low conversion and low
molecular weights of the grafted polymer; 40 wt% of PEO was grafted onto the silica
nanoparticles based on TGA measurements.

2.2.5. Ring-opening metathesis polymerization (ROMP)


ROMP is the transition-metal-catalyzed ring-opening polymerization of
strained cyclic olefins.

Mingotaud and co-workers75 immobilized a ruthenium

catalyst on 200 nm silica nanoparticles (Figure 2.11). This was achieved by reacting
the methathesis catalyst bearing a hydroxyl group with an acyl chloridefunctionalized silica particle. The grafting density was around 7 mol m-2. This was
then used to perform ROMP of norbornene. The authors reported that 30% of the
catalyst initiated polymerization.

TEM characterization showed a core shell

morphology that suggested the catalyst presence on the silica surface.

34

O
NH

(CH2)8

O (CH2)10 PCy2
Ru
Ph
PCy2R

Figure 2.11. Transition metal catalyst immobilized silica particle.


Seery and Jordi17 also reported the synthesis of polynorbornene brushes
using ROMP on silica nanoparticles with diameters ranging from 15-30 nm. They
immobilized

ROMP

initiators

5-norbornene-2-yl(ethyl)ethoxydimethylsilane

(NEEDS) and 5-(bicycloheptenyl)triethoxysilane (BCH), as shown in the Figure 2.12.


The initiator-immobilized particles were then reacted with a stoichiometric amounts
of Grubbs 1st generation catalyst, then followed by polymerization. Depending on
the coupling agent, the density of polynorbornene groups on the surface was reported
to be in the range of 1.82.6 groups/nm2.

NEEDS

Grubbs 1
CH3
Si

OCH2CH3

Cl2L2Ru

CH3
BCH

L = PCy3 or PPh3

Si(OCH2CH3)3

Figure 2.12. ROMP initiator used for immobilization on silica particles and the 1st
generation Grubbs catalyst (Grubbs 1).

35

2.2.6. Anionic polymerization


Anionic polymerization is a powerful method to make well-defined
architectures of hydrocarbon-based polymers.

Due to the stringent conditions

required for this type of polymerization, difficulties often arise due to the presence of
moisture and other impurities.

Despite these difficulties, there have been a few

reports on anionic polymerization on silica particles. Oosterling and co-workers76


reported the grafting of polystyrene and polystyrene-b-polyisoprene using an anionic
technique. Anionic polymerization was initiated in the presence of t-butyllithium by
immobilizing a ractive double bond on the silica surface, by reacting vinyl benzyl
trichlorosilane with silica nanoparticles.

However, high molecular weight and

molecular weight distribution were obtained due to slow and inefficent initiation by tBuLi. In contrast to this work, Advincula and co-workers77 reported a controlled
anionic polymerization of styrene from silica nanoparticles. The initiator precursor
1,1-diphenylethylene (DPE) was functionalized with alkyldimethylchlorosilane and
grafted onto silica particle surfaces. n-BuLi was used to activate the DPE, which
allowed the anionic polymerization of the styrene monomer to proceed in benzene
solution. However, the reported reproducibility of the polymerization was poor. This
was attributed to the free initiator and aggregation of the particles.

2.2.7. Cationic polymerization


Little work has been done on the application of cationic polymerization to the
synthesis of polymer brushes. Silica nanoparticles modified with polyisobutylene
have advantageous chemical and physical properties. Silica nanoparticles modified

36

with different chain-lengths of polyisobutylene (PIB) show different rheological


(solid, rubber-like material), thermal (decomposition temperature much higher than
unbound polyisobutylene) and chemical characteristics compared to unmodified PIB.
The living cationic polymerization is a useful tool to bind polymeric chains onto
nanoparticles with a controlled chain length and grafting density of the attached
polymer brushes.
Zirbs and co-workers78 investigated the grafting-from polymerization of
isobutylene from silica nanoparticles (Scheme 2.11). Living cationic polymerization
methods were used to grow polymer on the silica nanoparticles. Silanol groups on the
surface of the nanoparticles were modified with the initiator chloro{2-[3-(2chloropropane-2-yl)phenyl]propyl}dimethylsilane

for

the

carbocationic

polymerization. Polymerization of isobutylene was catalyzed by itanium tetrachloride


and di-t-butylpyridine (DTBP).

Modified nanoparticles with a well-defined PIB-

polymeric shell were synthesized.

Cl
Si
Cl
OH

Si

DTBP

Cl

CH2Cl2:n-Hex = 1:1
TiCl
DTBP

Scheme 2.12. Modification of silica surface with polyisobutylene.

37

2.3. Characterization
Polymer-modified silica particles have been the subject of extensive research.
The characterization studies improve the understanding of the mechanisms involved,
the final properties of the resultant composites, and help determine the types of
application. There are a number of literature reports of various analytical techniques
used to characterize these particles at each stage of modification and also to analyze
the organic shell after de-grafting from the surface.
2.3.1. Characterization of modified particles
There are number of particle analysis techniques. Gravimetric techniques are
used to determine the changes in weight before and after modification of the surface.
Thermal analysis, such as thermogravimetric analysis (TGA) and differential scanning
calorimetry (DSC) provide a bulk compositional property of the hybrid material.
Matyjaszewski and co-workers1 used DSC to study the effect of tethering and chain
mobilization on the glass transition temperature of the polymer shell-polystyrene.
The glass transition temperature of polystyrene immobilized on the silica surface was
higher due to a decrease in chain mobility, which is affected by tethering and chain
confinement. With increasing molecular weight, the Tg values of the tethered chains
approached that of bulk polymers.
Microscopy has been used to determine the size and nature of the silica
particles, before and after each stage of modification. Various techniques such as
atomic force microscopy (AFM), scanning electron microscopy (SEM), transmission
electron microscopy (TEM), and brewster angle microscopy (BAM) have proven
useful. These methods provide a visual image of the formation of the polymer chains

38

and the aggregation properties. Fukuda and co-workers57 reported the analysis of
high density PMMA brushes on silica particles using TEM and AFM techniques. A
surface film of PMMA-silica was formed by depositing a drop of 10 wt % suspension
of PMMA-silica in toluene onto the surface of purified water. The TEM images
showed the silica core as dark circles uniformly dispersed throughout the film, while
the PMMA chains formed fringes around the core. The AFM images showed the
surface structure of the film as protrusions composed of a silica core and the polymer
layer was found at a constant spacing on the substrate. BAM is used to obtain in situ
information about the long-range structure of a silica particulate layer formed at the
water surface in a film balance.79 These studies showed that the dispersibilty of the
particle changed inversely with the particle hydrophobicities at the water-air interface
due to attractive interparticle forces.
Electrophoresis studies were used to assess the effect of the grafted
polyelectrolyte on the colloidal stability of the silica particles.

Armes and co-

workers54 combined aqueous electrophoresis and dynamic light scattering (DLS)


studies to show that the colloidal stabilities of the polyelectrolyte-grafted silica
particles were dictated by the nature of the grafted polyelectrolyte and were highly
pH-dependent.
Spectroscopic techniques such as NMR, IR and X-ray photoelectron
spectroscopy (XPS) are widely used to characterize modified silica particles. Diffuse
reflectance infrared Fourier transform (DRIFT) spectroscopy is used to observe the
characteristic peaks for the initiator and polymer bound to the silica surface. The
diffuse reflectance technique is sufficiently sensitive and appropriate for the relatively

39

low coverage of immobilized groups on the silica surface. The carbonyl absorption is
particularly useful in the characterization of the bonded phase.

13

C and

29

Si solid-

state CP-MAS NMR spectroscopy are widely used to study the immobilization of the
initiator layers on the silica particle. XPS in particular is useful for monitoring the
presence of different oxidization states of atoms, relative abundance of atomic species
and the silica core itself. Armes and co-workers54 investigated the elemental and
chemical surface composition of the initial silica particle and the final polymer
grafted silica particles by XPS. XPS studies confirmed the presence of the grafted
polymer layer.
The size and state of dispersion of the modified and unmodified nanoparticles
is followed using various light scattering techniques. El Harrak and co-workers80
followed dispersion at each functionalization stage using small angle neutron
scattering (SANS). SANS helped to highlight the role played by the polymer in
preventing aggregation.

The wettability properties of polymer-grafted silica are

estimated by the penetration rate of water through a column packed with polymer.81
Untreated silica is hydrophilic in nature and the wettability of the silica particles can
be controlled by grafting hydrophilic or hydrophobic polymers on the surface of
silica.82

2.3.2. Characterization of the degrafted polymer


Silica nanoparticles have a high surface to volume ratio. They provide the
advantage of producing significant amounts of grafted polymer. These polymers can
be degrafted from the surface and subjected to various characterization studies. These

40

polymers can be degrafted from the surface either by direct cleavage or by dissolution
of the particle.83 A widely used procedure to recover the polymer is by dissolution of
the silica particle with HF. The isolated polymer can be analyzed to determine its
molecular weight, molecular weight distribution and composition using conventional
analytical techniques.
2.4. Applications of modified silica nanoparticles
The main challenge in dispersing silica particles lies in controlling the interparticle aggregation. It is necessary to stabilize these particles to prevent aggregation,
which is done by grafting polymer chains using the different methods mentioned
above. These long polymer chains control the nanoparticle aggregation by steric
repulsions. Another advantage of grafted polymer chains is that they can improve the
compatibility of the matrix with the particle surface.

It has been reported that

controlling the aggregation improves the mechanical properties of the composite.84


The formation of stable dispersions in solvents or polymer matrices is an important
variable affecting the final properties of formulations and coatings. Smart surfaces
that respond to external stimuli are important for changing wettability and adhesion
properties.
Brushes of molecularly imprinted polymers on silica particles have been used
for chromatographic supports.85

They have been used as a stationary phases for

separation of triazinic herbicides. Synthesis of hollow polymeric nanocapsules using


silica nanoparticles were developed by Blomberg and co-workers69 These materials
have potential applications for the encapsulation of active substances such as drugs
and dyes.

Silica/polymer hybrids are important candidates for environmental

41

pollution control processes. The application of silica/polymer hybrid systems for


removal of heavy metal ions such as cobalt and copper salts from waste water has
been studied by Meyer and co-workers86 The large surface area, well-defined pore
size and pore shape of the silica particles, and the specific binding sites of the
polymer help removal of these metals.

2.5. Surface modification of silica nanoparticles with polyacrylamide


Surface modification of silica nanoparticles has attracted considerable attention in
recent years. Acrylamide (AAm) polymers are an important class of materials because of
their use in many industrial applications such as coatings, flocculants, paper making,
mining, electrophoresis and biological applications.87 Not much work has been done to
study silica-polyacrylamide hybrid nanoparticles.
Jang and Park24 used the sol-gel process to make polyacrylamide-silica
nanocomposites. The nanocomposites were monolithic and formed a thin film. Fourier
transform IR spectroscopy (FTIR), AFM, and SEM were utilized to prove the existence
of an interaction between the two phases and to examine the morphological features of
the hybrid materials. The morphology of the hybrid materials was discussed in terms of
the particlematrix morphology. At low tetraethylorthosilicate (TEOS) concentrations,
the particles were well dispersed. These particles aggregated to form clusters with
increasing TEOS contents. AFM studies of fracture surfaces showed that the hybrid
materials appeared to form particulates. This is due to sufficient water concentration.
With an increase of TEOS contents, the roughness of the fracture surface and the
aggregation of silica particles were increased.

42

Demchenko and co-workers88 performed a radical graft polymerization of


acrylamide onto silica nanoparticles using ceric ions. They also compared grafted PAAm
and free PAAm. Grafted PAAm was found to have a more homogeneous structure, less
rigidity, and denser packing in comparison with free PAAm. A cooperative system of Hbonds exists between the silica surface and PAAm which forms the dense polymer shell.
The unification of the sorption ability of the colloidal particle surface and binding
properties of long polymer chains in PAAm-grafted silica produces its highly flocculative
capability. Graft polymerization of PAAm on silica nanoparticles results in a polymercolloid complex where the polymer chains are covalently bond to silica and there is an
attraction of the grafted chains with the particle surface through H-bonding (Figure 2.13).

Figure 2.13. Schematic model for polyacrylamide grafted silica.

2.6. RAFT polymerization of acrylamide


McCormick and co-workers89 studied the kinetics and molecular weight control of
the polymerization of acrylamide via RAFT.

Higher rates of polymerization were

achieved with CTAs having higher rates of fragmentation, initiators with lower
43

decomposition temperatures, and higher initiator to CTA concentrations.

High

conversions, however, required a constant supply of initiator derived radicals. Molecular


weight deviations for polymers prepared in water were greater than those synthesized in
DMSO.

Adjustment of pH to values below 5.0 using appropriate buffers greatly

diminished hydrolysis and aminolysis of the dithioesters; however, for increasingly


extended reaction times (retarded rates), loss of the macro CTA became an issue.
Polymers produced with lower CTA to initiator concentrations exhibited greater
deviations, while polymers prepared with CTAs with faster rates of fragmentation had
molecular weights closer to those predicted by theory. For RAFT polymerizations of
acrylamide in DMSO mediated by the trithiocarbonate type CTA (2-(1-Carboxy-1methyl-ethylsulfanylthiocarbonylsulfanyl)-2-methyl-propionic acid), reaction rates were
significantly faster, and near quantitative conversions could be reached with proper
initiator choice. Trithiocarbonate RAFT agents gave better control of molecular weight
and polydispersity of polyacrylamide. They also afforded much faster polymerization
rates and the polydispersity values were always less than 1.2.

2.7. Click chemistry


Click chemistry is a concept introduced by K. Barry Sharpless and co-workers18 in
2001. Click reactions are high yield reactions that follow the following criteria:
1.

The reaction is modular and wide in scope.

2.

There is a large thermodynamic driving force (> 84 kJ/mol) for a fast reaction
with a single reaction product.

3.

High chemical yield.

44

4.

Selective.

5.

Simple reaction conditions.

6.

No solvent involved or a benign solvent (preferably water).

7.

Simple product isolation.

8.

No offensive by-products.

9.

Readily available starting materials and reagents.

The following classes of chemical transformations are the most common examples of
click reactions:
1.

Cycloaddition reactions, particularly the Huisgen 1,3-dipolar cycloaddition


reaction between alkynes and azides.

2.

Nucleophilic substitution especially for small strained rings like epoxy and
aziridine compounds.

3.

Carbonyl chemistry involving formation of ureas and amides.

4.

Epoxidation and dihydroxylation of carbon carbon double bonds.

R1

N +

R1

R2

Cu(I)
N
2

Scheme 2.13. Model click reaction.

The Huisgen 1,3-dipolar cycloaddition of azides and terminal alkynes, catalyzed


by copper(I) is the most common click reaction (Scheme 2.12). This reaction has a high
thermodynamic driving force usually greater than 84 kJ/mol.18 The reactions can be

45

performed in high yield, in multiple solvents (including water), and in the presence of
many other functional groups. The 1,2,3-triazole product is chemically very stable. This
reaction is useful because azides and alkynes can be easily introduced into a molecule
and these chemical moieties are stable under a variety of conditions. Azides and alkynes
are inert to most biological and organic conditions, including highly functionalized
biological molecules, molecular oxygen, water, and the majority of common reaction
conditions in organic synthesis. In particular, despite the thermodynamic favorability of
azide decomposition, kinetic factors allow aliphatic azides to remain nearly invisible until
presented with a good dipolarophile. The kinetic stability of alkynes and azides is
directly responsible for their slow cycloaddition, which generally requires elevated
temperatures and long reaction times.

Cu(I) catalysis dramatically improves

regioselectivity to afford the 1,4-regioisomer exclusively and increases the reaction rate
up to 107 times, eliminating the need for elevated temperatures.90
Although the thermal dipolar cycloaddition of azides and alkynes occurs through
a concerted mechanism, density functional theory (DFT) calculations on monomeric
copper acetylide complexes indicate that the concerted mechanism is strongly disfavored
relative to a stepwise mechanism.91 Stepwise cycloaddition catalyzed by a monomeric
Cu(I) species lowers the activation barrier relative to the uncatalyzed process by as much
as 11 kcal/mol, which is sufficient to explain the rate enhancement observed under Cu(I)
catalysis.

46

2.8. Combination of click chemistry and living radical polymerization


Combining the chain-end functionality control of living free radical
polymerization and the efficiency of click chemistry is attractive. Post-polymerization
modification remains a viable means of incorporating functionality into a polymer that is
potentially incompatible with synthetic, characterization, or processing conditions.92 A
drawback of this method of functionalization is relatively low yields and side reactions
with other groups within the polymer. Efficient click reactions are suitable for successful
post-polymerization modification.

2.8.1. Combination of ATRP and click chemistry


Lutz and co-workers19a used click chemistry to prepare an end-functionalized
ATRP polymer. The bromine chain ends of the polymer were first transformed into an
azide end group and subsequently reacted with terminal alkynes to create different
functional end groups. van Hest and Opsteen19b demonstrated a modular approach for the
synthesis of block copolymers via a Huisgen 1,3-dipolar cycloaddtion of terminal alkynes
and azides. Terminal alkyne and azide functionalities were introduced via ATRP using
functionalized initiators, or via a post-polymerization end group modification. SEC
measurements of the subsequent cycloaddition reactions demonstrated that block
copolymers were formed quantitatively and that residual polymer precursors could be
removed successfully from the reaction mixture.

Matyjaszewski and co-workers19c

synthesized an alkyne-terminated ATRP initiator to polymerize well defined -alkyne-bromo-terminated polystyrene (PS). These polymers prepared by ATRP were coupled
via a step growth mechanism using click coupling to yield PS containing triazole linkages

47

in the repeat units (Scheme 2.14). Matyjaszewski and co-workers92,93 also synthesized
,-dihydroxypolystyrene and star polymers via ATRP and click chemistry.
Macrocyclic polymer, neoglycopolymer, and 1st generation dendritic copolymers have
also been synthesized by combining ATRP and click chemistry.19d,94,95

Scheme 2.14. Click coupling reactions using telechelic polymers prepared by ATRP.

2.8.2. Combination of RAFT polymerization and click chemistry


RAFT polymerization has received increasing attention in recent years. Among
available controlled free radical polymerization techniques, RAFT has arguably the most
important commercial significance because it works with the greatest range of vinyl
monomers and under a wide variety of experimental conditions. Little work has been
done on the combination of RAFT polymerization and the click reaction. Davis, BarnerKowollik and co-workers23 synthesized well-defined block copolymers PS-b-PVAc by
combining RAFT polymerization and click chemistry. In the first step, new clickable
RAFT agents were designed and used to produce homopolymers of PS and PVAc under
controlled/living conditions (Scheme 2.15). In the second step, click coupling reactions
were performed between PS and PVAc to afford various block copolymers with variable
block ratios.
Hawker, Wooley and co-workers21 used alkene functionalized RAFT CTA
directly for the RAFT-based sequential polymerization of tetrahydropyran acrylate and

48

styrene, followed by selective cleavage of the tetrahydropyran esters to give alkenefunctionalized block copolymers.

These click-functionalized polymers, with the

functionality located at the hydrophilic polymer termini, were then self-assembled using a
mixed-micelle methodology to afford surface-functionalized clickable micelles in
aqueous solutions.

Scheme 2.15. Synthesis of click functionalized RAFT CTA.

Sumerlin and co-workers22 synthesized two novel azido-functionalized chain


transfer agents and subsequently employed them to mediate the RAFT polymerizations of
styrene and N,N-dimethylacrylamide under a variety of conditions.

The resulting

homopolymers retained end group functionality, as evidenced by the successful


formation of block copolymers.

The -azido terminal polymers and the azido-

functionalized CTAs were coupled to various alkynes (propargyl acrylate, propargyl


methacrylate, and propargyl alcohol) using the high efficiency of click chemistry in the
presence of a Cu(I) catalyst, demonstrating the ability to prepare a range of functional
telechelics and CTAs.

49

CHAPTER III
EXPERIMENTAL

3.1. Materials
Colloidal silica (D = 70-100 nm) as a 30 wt % dispersion in isopropanol was
provided by Nissan Chemical.

N-(3-Dimethylaminopropyl)-N-ethylcarbodiimide

hydrochloride (EDC, 98%), 4-N,N-dimethylaminopyridine (DMAP, 99%), 4,4'-azobis(4cyanovaleric

acid)

(VA

501,

98%),

azobisisobutyronitrile

(AIBN,

98%),

dimethylsulfoxide (DMSO), dimethylformamide (DMF), tetrahydrofuran (THF),


acrylamide (electrophoresis grade >99%), methanol, sodium azide (99.5%), 2bromoisobutyryl bromide (98%), triethoxysilane (95%), triethylamine (99.5%), allyl
alcohol (98.5%), Aliquat 336 (tricaprylylmethylammonium chloride), copper sulfate
(99%) and sodium ascorbate (98%) were purchased from Aldrich. Styrene (99%), methyl
acrylate

(99%),

ethyl

2-bromoisobutyrate

(98%),

N,N,N,N,N-

pentamethyldiethylenetriamine (PMDETA, 99%) were purchased from Aldrich Chemical


and were purified (removed inhibitor) by passage through a column of activated basic
alumina (Aldrich 150 mesh). CuBr (Aldrich, 98%) was purified as described in the
literature.96 3-Bromopropyltrichlorosilane (95%) was purchased from Gelest Inc. Unless
otherwise specified all the chemicals were used as received.

50

3.2. Measurements
Fourier transform infrared (FT-IR) spectra were obtained on a Bruker Tensor 27
FT-IR spectrometer using a diffuse reflection apparatus (Cricket, Harrick Scientific).
Thermogravimetric analysis (TGA) was performed on a TA 2950 instrument at a scan
rate of 20 oC/min under nitrogen atmosphere. The molecular weights of polymers were
measured in THF (at 35 oC with a flow rate of 1 mL/min) by gel permeation
chromatography (GPC) using a Waters 501 pump, Waters HR4 and HR2 styragel
columns, a Waters 410 differential refractometer, and a differential viscometer (Viscotek
100) and a laser light scattering detector (Wyatt Technology, DAWN EOS, = 670 nm).
Matrix-assisted laser desorption/ionization (MALDI) spectra were acquired by Prof.
Chrys Wesdemiotis group using a Bruker Reflex III MALDI-TOF mass spectrometer
equipped with a nitrogen laser (337 nm), a single-stage pulsed ion extraction ion source, a
two-stage grid-less reflector, and the two dual microchannel plate detectors for detection
in linear and reflectron mode. MS spectra were measured with reflectron mode, with the
ion source and reflector lens potentials kept at 20 keV and 22.5 keV, respectively.
Solutions

of

-cyano-4-hydroxycinnamic

acid,

polymer

sample,

and

sodium

trifluoroacetate were prepared in water. These solutions were mixed in the ratio of
matrix:cationizing salt:polymer (10:1:2), and 0.5 L of the mixture was applied to the
MALDI sample target and allowed to dry. Elemental analyses were obtained from
Galbraith Laboratories in Knoxville, TN. X-ray photoelectron spectroscopy (XPS) was
performed on a Perkin Elmer instrument at the MATNET Surface Analysis Centre at
Case Western Reserve University.

51

3.3. Combination of RAFT polymerization and click chemistry for surface modification
Alkyne-functionalized polymer and azide-functionalized silica nanoparticles were
separately synthesized.

3.3.1. Deposition of 3-bromopropyltrichlorosilane on silica nanoparticles


Into a Schlenk flask, 6 g of dried silica and 45 mL of anhydrous toluene were
added under an inert atmosphere. This flask was sonicated for 30 min and then placed
into an oil bath preset at 80 oC. 3-Bromopropyltrichlorosilane (5.5 mL) in 15 mL of
toluene was added dropwise and the solution was maintained at 80 oC for 18 h. The
particles were recovered by centrifugation at 3000 rpm for 30 min. These particles were
redispersed in toluene and centrifuged. This cycle was repeated 6 times to obtain the
final product.

After drying, these modified nanoparticles were characterized by IR

spectroscopy, TGA and elemental analysis. Elemental analysis confirmed the presence
of bromine element (1.53 %) on modified silica nanoparticles (calculation shown in
Appendix A1).

3.3.2. Synthesis of azide modified silica nanoparticles


3-Bromopropyl modified silica nanoparticles (5 g) and 2 g of NaN3 in 100 mL of
DMF were combined under an inert atmosphere. This solution was stirred at 80 oC for 18
h. The particles were recovered by centrifugation at 3000 rpm for 30 min. The cycle of
centrifugation and redispersion in water was repeated 3 times. The isolated particles
were characterized by IR spectroscopy, TGA and elemental analysis. Elemental analysis

52

of nitrogen (0.76 %) indicated that the modified nanoparticles contained, on average, 3.45
azide groups per nm2 (calculation shown in Appendix A2).

3.3.3.

Esterification

reaction

of

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

acid)trithiocarbonate
The synthesis of S-1-dodecyl-S'-(,'-dimethyl--acetic acid) trithiocarbonate
chain transfer agent (trithiocarbonate CTA) was carried out according to a previously
reported method in the literature.97

Trithiocarbonate CTA (1 g, 2.7 mmole), N-(3-

Dimethylaminopropyl)-N-ethylcarbodiimide hydrochloride (EDC) (0.78 g, 4.1 mmole),


4-N,N-dimethylaminopyridine (DMAP) (0.5 g, 4.1 mmole) and 10 mL of
dichloromethane were added to a round bottomed flask and stirred for few minutes under
inert atmosphere. Propargyl alcohol (0.5 mL, 8.9 mmole) was added and the mixture was
stirred overnight at room temperature. The product was washed with acidic water (0.05
N HCl), water and brine several times. The organic layer containing was removed, dried
under reduced pressure. Product (RAFT CTA 1) was collected as a dark yellow viscous
liquid (at room temperature). The yield (0.8 g) was 80%.
1

H NMR (CDCl3, 300 MHz): = 0.87 (t, -CH2CH3, 3H), 1.2-1.4(m, -CH2CH2, 20H), 1.65

(s, -C(CH3)-CH3, 6H), 2.5(s, -CHC, 1H), 3.25 (t, -CH2S, 2H), 4.7 (s, -O-CH2C, 2H).
13

C NMR (CDCl3): = 17 (CH2-CH3), 22 (-CH2-CH3), 24.5 (2C, CH3-C), 26-32 (9C,

CH2(CH2)-CH2), 37 (CH2(CH2)-S), 53 (C(CH3)-CH3), 55 (CH2(C)-O), 76 (CHC), 77


(C(CH2)CH, 173 (C=O), 222 (C=S).

53

3.3.4. Surface modification via grafting to approach


RAFT polymerization of styrene and acrylamide were performed and obtained
polymers were coupled to silica nanoparticles using click reaction.

3.3.4.1. RAFT polymerization of acrylamide


The alkyne terminated RAFT CTA 1 (0.4 g, 0.96 mmole), 8.0 g (0.112 mole) of
acrylamide, 70 mL of DMSO and 0.01 g (0.04 mmole) of azo intiator (VA501) were
combined in a Schlenk flask. The flask was subjected to 3 freezepumpthaw cycles to
remove oxygen. The mixture was heated to 70 oC for 4 h under an inert atmosphere. The
polymerization was stopped by quenching the mixture in ice water and PAAm was
obtained after precipitation into cold methanol. Two independent analytical techniques
were used to characterize the low molecular weight: PAAm, MALDITOF mass
spectrometry and 1H NMR spectroscopy (calculation shown in Appendix A3). Good
agreement was observed between calculated molecular weight (Mn = 4,500 g/mole) and
experimental molecular weight by both techniques.

13

C NMR was used to confirm the

presence of the alkyne end group.

3.3.4.2. Click coupling between polyacrylamide and silica nanoparticle


PAAm (3 g, 0.66 mmole) and 15 mL of DMSO were combined in a round
bottomed flask and stirred until the PAAm dissolved; 1 g of azide-modified silica
nanoparticles was then added and the mixture was sonicated for 30 min. A solution of
0.005 g (0.03 mmole) of CuSO4 in 1 mL of water was added to the mixture followed by
addition of a freshly prepared solution of 0.015 g (0.75 mmole) of sodium ascorbate in 1

54

mL of water. The mixture was heated in an oil bath at 50 oC overnight. The particles
were recovered by centrifugation at 3000 rpm for 30 min. The particles were redispersed
in water and centrifuged; this cycle was repeated 4 times. Finally, the particles were
placed in a soxhlet extractor and extracted with water for 18 h. The particles were dried
under vacuum and analyzed by FT-IR, TGA and elemental analysis.

3.3.4.3. RAFT polymerization of styrene


The alkyne-terminated RAFT CTA 1 (0.4 g, 1.0 mmole), 24 mL of (0.21 mole)
styrene, 25 mL of toluene and 0.01 g (0.06 mmole) of AIBN were combined in a Schlenk
flask. The flask was subjected to 3 freezepumpthaw cycles to remove oxygen and the
mixture was heated at 90 oC for 18 h under an inert atmosphere. The polymerization was
stopped by quenching the flask in ice water and polystyrene was obtained after
precipitation into cold methanol. GPC was used to determine molecular weight (Mn =
9,000 g/mol) and polydispersity (1.1). Monomer conversion was determined to be 41%
(model calculation shown in Appendix A4).

13

C NMR spectroscopy was used to confirm

the alkyne functionality in PS.

3.3.4.4. Synthesis of alkyne-terminated diblock copolymer (PS-b-PMA) by RAFT


polymerization
Alkyne terminated RAFT CTA 1 (1.2 g, 2.9 mmole), 24 mL (0.21 mole) of
styrene, 25 mL of toluene and 0.03 g (0.2 mmole) of AIBN were combined in a Schlenk
flask. The solution was degassed and the polymerization was carried out at 90 oC for 18
h under an inert atmosphere. The polymerization was stopped by quenching the mixture

55

in ice water and polystyrene was obtained after precipitation into cold methanol. GPC
was used to determine molecular weight (Mn = 3,000 g/mol) and polydispersity (1.1).
Monomer conversion was determined to be 39%.

Then 6.6 g (2.2 mmole) of this

polystyrene, 12 mL (0.11 mole) of methyl acrylate, 12 mL of toluene and 0.018 g (0.10


mmole) AIBN were combined in a Schlenk flask and polymerization was carried out at
80 0C for 3 h in similar way to obtain diblock copolymer, PS-b-PMA (Mn = 6,000 g/mol,
polydispersity =1.1).

3.3.4.5. Click coupling between alkyne-terminated PS-b-PMA and silica nanoparticles


PS-b-PMA (3 g, 0.5 mmole), 3.4 g of azide-modified silica and 40 mL of DMF
were combined in a flask and sonicated to produce a uniform suspension. Solutions of
0.005 g (0.03 mmole) of CuSO4 in 0.5 mL of water and 0.022 g (0.11 mmole) of sodium
ascorbate in 1 mL of water were added and the mixture was heated at 70 oC for 18 h.
Silica particles were separated after centrifugation of the mixture.

The silica

nanoparticles were redispersed in toluene and centrifuged. This cycle was repeated 3
times with toluene, 2 times with water and 1 time in THF to remove physisorbed polymer
and Cu catalyst. The modified silica particles were isolated after drying at reduced
pressure. Presence of block copolymers on silica nanoparticles were confirmed by IR
spectroscopy. TGA and elemental analysis was also used to characterize particles.

3.3.5. Surface modification via grafting from approach


RAFT CTA were first immobilized on silica nanoparticles and then surface mediated
RAFT polymerization were performed.

56

3.3.5.1. Synthesis of RAFT CTA immobilized silica nanoparticle


Azide modified silica (1.5 g), 0.25 g (0.62 mmole) of alkyne-terminated RAFT
CTA 1 and 10 mL of anhydrous DMF were combined in a flask under an inert
atmosphere. This mixture was subjected to sonication for 30 min. Then 0.005 g (0.031
mmole) of CuSO4 in 1 mL of water and 0.012 g (0.06 mmmole) of sodium ascorbate in 1
mL of water were added into the flask. The mixture was stirred for 15 h at 50 oC. The
modified silica particles were isolated by centrifugation.

These particles were

redispersed and centrifuged to remove any physisorbed RAFT CTA 1. This cycle was
performed two times with toluene, one time with water and one time with toluene again.
Modified silica particles were dried under reduced pressure. Obtained particles were
characterized by IR, TGA, and elemental analysis.

3.3.5.2. Typical surface mediated RAFT polymerization of styrene


RAFT CTA 1 modified silica nanoparticles (1.1 g), 20 mL of styrene, 20 mL of
toluene,

0.15

(0.4

mmole)

of

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

acid)

trithiocarbonate (carboxyl-terminated RAFT CTA) and 0.0067 g (0.04 mmole) of AIBN


were added into a flask under inert atmosphere. Polymerization was typically done at 90
o

C for 18 h. Silica particles were separated after centrifugation of the mixture. Alkyne-

terminated free polystyrene was obtained by precipitation into cold methanol. The silica
nanoparticles were redispersed in toluene and centrifuged. This cycle was repeated 5
times to remove physisorbed polymer. The collected silica particles were dried under
reduced pressure to obtain PS modified silica particles. IR, TGA, and elemental analysis
were used to confirm the presence of attached PS on these particles.

57

3.3.5.3. Typical surface mediated RAFT polymerization of MA over PS modified silica


nanoparticles
PS modified silica particles (0.8 g), 0.15 g (0.4 mmole) of carboxyl-terminated
RAFT CTA, 0.005 g (0.03 mmole) of AIBN, 12 mL (0.11 mole) of MA and 15 mL of
toluene were combined in a flask under nitrogen. The polymerization was allowed to
proceed at 80 oC for 8 h. Modified silica particles were collected and characterized using
the previously described procedure (in section 3.3.5.2).

3.3.5.4. General procedure for cleavage of polymers from silica nanoparticles


Polymer silica hybrid nanoparticles (0.25 g) were dispersed in 5 mL of toluene.
Then 75 mg of tricaprylmethylammonium chloride (Aliquat 336) was added as a phase
transfer catalyst. A 5% aqueous HF solution (5 ml) was added and the mixture was
stirred overnight. The organic layer was removed and the polymer was isolated by
precipitation into methanol.

Volatiles were removed from the collected polymer in

vacuo.

3.3.6. Surface modification via novel tandem approach


Tandem RAFT polymerization and click chemistry were performed to modify the surface
of silica nanoparticles. In this process RAFT polymerization and click chemistry were
performed in a single pot. Alkyne terminated polystyrene were also obtained due to
presence of excess amount of RAFT CTA.

58

3.3.6.1. Surface modification of silica nanoparticles using tandem RAFT polymerization


and click chemistry
Azide-modified silica nanoparticles (1 g), 0.3 g (0.718 mmole) of alkyne
terminated RAFT CTA 1, 30 mL (0.26 mole) of styrene, 30 mL of toluene, 0.004 g
(0.028 mmole) of CuBr, 0.010 mL (0.048 mmole) of PMDETA and 0.01 g (0.06 mmole)
of AIBN were combined in a Schlenk flask. The flask was subjected to three freeze
pumpthaw cycles to remove oxygen. The mixture was heated at 90 oC for 18 h under an
inert atmosphere. The polymerization was stopped by cooling the flask in ice water.
Silica particles were separated after centrifugation of the mixture. Alkyne-terminated PS
was obtained from the solution by precipitation into cold methanol. The silica particles
were redispersed in toluene and centrifuged. This cycle was repeated 5 times to remove
physisorbed polymer. The silica particles were dried under reduced pressure to afford
PS-modified silica particles. These nanoparticles were characterized by IR, TGA, and
elemental analysis.

3.3.6.2. Modification of silica nanoparticles using the grafting to approach


Free polystyrene (3 g, 0.33 mmole) obtained from a tandem experiment, 1.3 g of
azide-modified silica and 40 mL of DMF were combined in a round bottomed flask and
stirred for several min. The mixture was sonicated for 30 min and solutions of 0.005 g
(0.031 mmole) of CuSO4 in 0.5 mL water and 0.022 g (0.11 mmole) of sodium ascorbate
in 1 mL of water were added to that flask. This mixture was heated at 70 oC for 18 h.
The silica nanoparticles were isolated and characterized as described previously in
section 3.3.4.5.

59

3.4. Combination of ATRP and click chemistry for surface modification via grafting to
approach
ATRP of styrene and MA were performed using alkyne-terminated ATRP initiator.
Obtained polymers were click coupled to silica nanoparticles.

3.4.1. ATRP of styrene


The

synthesis

of

alkyne-terminated

ATRP

initiator,

propargyl

2-

bromoisobutyrate, was carried out according to a previously reported method in the


literature.98 Propargyl 2-bromoisobutyrate (0.3 mg, 1.5 mmole), 0.225 g (1.56 mmole) of
CuBr , 0.6 mL (3.1 mmole) of PMDETA, 22 mL (0.19 mole) of styrene and 19 mL of
toluene were combined in a Schlenk flask. The mixture was subjected to 3 freezepump
thaw cycles and polymerization was performed under inert atmosphere at 110 oC for 9 h.
The polymerization was stopped by quenching the mixture into ice water. The polymer
solution was passed though an alumina column 3 times to remove any traces of copper.
The purified solution was precipitated into cold methanol to afford the polymer. The
polystyrene was characterized by GPC, 1H and

13

C NMR spectroscopy.

Molecular

weight of polystyrene was determined to be 6000 g/mol (PDI = 1.1) with 47% monomer
conversion.

3.4.2. ATRP of MA
Propargyl 2-bromoisobutyrate (0.3 mg, 1.47 mmole), 0.225 g (1.56 mmole) of
CuBr , 0.6 mL (3.12 mmole) of PMDETA, 17 g (0.20 mole) of MA and 22 mL of toluene
were combined in a Schlenk flask and the mixture was deoxygenated (three freeze /

60

vacuum / N2 refill cycles). The polymerization was performed under an inert atmosphere
at 90 oC for 6 h. The polymerization was stopped by quenching the mixture into ice
water. The polymer solution was passed though an alumina column 3 times to remove
any traces of copper. The purified solution was precipitated into cold methanol to afford
PMA (Mn = 8,000 g/mol, PDI = 1.14) with 70% monomer conversion.

3.4.3. Click coupling between polystyrene and silica nanoparticles


Polystyrene (3 g, 0.33 mmole), 1.3 g of azide-modified silica and 40 mL of DMF
were combined in a round bottomed flask and the mixture was stirred for 5 min.
Following a brief sonication, solutions of 0.005 g (0.031 mmole) of CuSO4 in 0.5 mL
water and 0.022 g (0.11 mmole) of sodium ascorbate in 1 mL of water were added and
the mixture was heated at 70 oC for 18 h. The particles were recovered by centrifugation
at 3000 rpm for 30 min. These particles were redispersed in solvent and centrifuged; this
cycle was repeated 4 times in toluene, 1 time in water and then 1 time in toluene. The
modified silica particles were obtained after drying at reduced pressure. Obtained
particles were characterized by IR, TGA, and elemental analysis.

3.5. Surface modification of silica nanoparticles using surface-initiated polymerization


(grafting from approach)
ATRP initiator was synthesized in two steps and attached to silica nanoparticles.
Surface initiated polymerizations were performed to modify the surface of silica
nanoparticles.

61

3.5.1. Synthesis of allyl 2-bromoisobutyrate


A 1 L round bottomed flask was charged with 400 mL of anhydrous
dichloromethane, 7.58 g (130 mmole) of allyl alcohol, and 13.25 g (130 mmole) of
triethylamine. The flask was fitted with a 250 mL addition funnel which was charged
with 31.5 g (274 mmole) of 2-bromoisobutyryl bromide. The mixture was cooled to 0 C
in an ice/water bath and the 2-bromoisobutyryl bromide added dropwise over a 30 min
period. After addition, the reaction mixture was stirred for an additional 18 h at ambient
temperature.
The reaction mixture was transferred to an l L separatory funnel. The salts and
unreacted acid halide were extracted from the organics by sequential washings with 250
mL of 1N HCl, 3 x 250 mL of aqueous saturated sodium bicarbonate, and 2 x 250 mL of
deionized water. The organic phase was isolated and dried over anhydrous magnesium
sulfate. The solvent was removed by rotary evaporation and the product isolated as a
slightly yellow liquid by distillation at 53-54 C. 1H NMR (CDCl3): = 5.75.9 (m, 1H,
=CH-), 5.155.33 (m, 2H, =CH2-), 4.58 (d, 2H,-CH2O-), 1.94 (s, 6H, -CH3).

13

C NMR

(CDCl3):= 170 (CO), 130 (2C, CH), 117 (2C, CH2), 68 (CH2O), 56 (Cquaternary), 29 (2C,
CH3-C-Br).

3.5.2 Synthesis of [3-(2-bromoisobutyryl)propyl]triethoxysilane (BITES)


In a 100 mL round bottom flask, 23.5 g (113.5 mmole) of allyl 2bromoisobutyrate, 25 mL of anhydrous toluene and 35.5 g (216 mmole) of
triethoxysilane were mixed.

Karstedts catalyst [Platinum(0)-1,3-divinyl-1,1,3,3-

tetramethyldisiloxane complex solution in xylene, Pt ~2%] (0.11 mL, 0.01 mmole) was

62

added and the apparatus was fitted with a reflux condenser and rubber septum. The
mixture was heated to reflux (110 oC) overnight. The excess triethoxysilane and toluene
were removed under reduced pressure, yielding a light yellow product.

The crude

product was purified by column chromatography using 15:1 (v/v) hexanes/ethyl acetate
as eluent to afford BITES (60% yield).

H NMR (CDCl3): = 3.9 (m, 6H, CH2(OSi)-

CH3), 4.58 (d, 2H, -CH2O-), 1.94 (s, 6H, CH3-C-Br), 1.85 (m, 2H, CH2(CH2)-CH2), 1.1
(t, 9H, CH3-CH2), 0.84 (t, 2H, CH2(Si)-CH2).

13

C NMR (CDCl3): = 170 (CO), 68

(CH2(OC)-CH3), 59 (2C, CH2(OSi)-CH3) 56 (Cter), 29 (2C, CH3-C-Br), 22 (3C, CH3CH2), 19 (CH2(CH2)-CH2), 8 (CH2(Si)-CH2).

3.5.3. Deposition of ATRP initiator onto silica nanoparticles


Silica nanoparticles (10 g) were suspended in 125 mL of toluene for 30 min under
sonication. The toluene suspension of silica nanoparticles was heated at 80 oC for 30
min. An excess (3 g, 8.1 mmole) of BITES initiator was added dropwise, and the mixture
was heated at reflux for 18 h. The particles were isolated via centrifugation. The
particles were washed free of adsorbed initiator using 5 cycles of centrifugation and
resuspension in toluene, and then volatiles were removed under vacuum. These modified
nanoparticles were characterized by IR, TGA and elemental analysis. Using elemental
analysis of bromine, the grafting density of ATRP initiator was 4.2 groups/nm2 (model
calculation shown in Appendix A1).

63

3.6. Surface-initiated ATRP of styrene on silica nanoparticles


Initiator modified silica (9 g), CuBr (0.6 g, 4.2 mmole), PMDETA (1.47 g, 8.4
mmole), and ethyl 2-bromoisobutyrate (0.6 g, 3.06 mmole) were combined in a Schlenk
flask containing a stir bar. Styrene (27.0 g, 259.6 mmole) and toluene (30 mL) were
added via syringe, and the mixture was deoxygenated (three freeze / vacuum / N2 refill
cycles). The mixture was stirred and sonicated to get a uniform suspension and the
mixture was heated at 110 oC for 5 h. Silica particles were separated after centrifugation
of the mixture. Free polystyrene was obtained by precipitation into cold methanol with a
molecular weight (Mn) of 6000 g/mol (80% conversion) and PDI = 1.1. The silica
nanoparticles were redispersed in toluene and centrifuged. This cycle was repeated 3
times with toluene, 2 times with water and 1 time with THF to remove physisorbed
polymer and Cu catalyst

3.7. Surface initiated ATRP of MA using PS-modified silica nanoparticles


PS modified silica particles (6 g), 0.3 g (0.2 mmole) of CuBr, 0.70 g (0.40
mmole) of PMDETA, 0.3 g (1.53 mmole) of ethyl 2-bromoisobutyrate, 8 g (95 mmole) of
MA, and 16 mL of anisole were combined in a flask under nitrogen. The mixture was
sonicated for 30 min to obtain a uniform suspension. The polymerization was allowed to
proceed at 90 oC for 5 h. Modified silica particles and free PMA were collected using the
previously described procedure. Modified silica nanoparticles were characterized by
TGA, FT-IR, XPS and elemental analysis. Free PMA was found to have molecular
weight (Mn) of 3000 g/mol (55% conversion) and PDI = 1.1.

64

CHAPTER IV
RESULTS AND DISCUSSION

4.1. Silica Nanoparticles


Silica nanoparticles are commercially available and they can also be synthesized
in the laboratory. Silica particles were synthesized using modified Stber process, where
50 mL of ethanol was added to 3 mL of 28% aqueous ammonia (NH4OH).99 The
tetraethylorthosilicate (TEOS) was added dropwise and the particle was allowed to grow
for 24 h.

The modified Stber process provides good control of particle size and

distribution. Uniform particles with a size range of 75-100 nm (determined by TEM)


were obtained.

Figure 4.1. TEM image of silica nanoparticles provided by Nissan Chemical.

- 65 -

In these studies, commercially available silica nanoparticles [OrganosilicasolTM,


IP-ST-ZL from Nissan Chemical America Corporation, Houston, (TX)] were mostly
used. Silica particles were obtained as 30% dispersion in isopropyl alcohol. The solvent
was evaporated and the particles were dried overnight in a vacuum oven at 80 oC before
use. These nanoparticles had particle sizes of 70-100 nm and a specific surface area 32
g/m2 (provided by Nissan Chemical America Corporation). TEM studies of these
nanoparticles showed a narrow particle size distribution (Figure 4.1).

4.2. Combination of living radical polymerization and click chemistry for surface
modification
The Cu(I) catalyzed variant of the Huisgen 1,3-dipolar cycloaddition of azides
and alkynes is one of the most efficient reactions available.18 Combining the chain-end
functionality control of living free radical polymerization and the efficiency and diversity
of click chemistry was desirable.

Among controlled free radical polymerizations,

reversible addition fragmentation transfer (RAFT) has arguably the most important
commercial significance because it works with the greatest range of vinyl monomers.20
Hence in this research, the possibility of combining RAFT polymerization and click
chemistry as a synthetic strategy for surface modification was explored. A novel click
functionalized RAFT chain transfer agent (CTA) was synthesized and this opened up the
possibility of using RAFT polymerization and click chemistry together in surface
modification.

66

4.3. Synthesis of click functionalized (alkyne-terminated) RAFT chain transfer agent


A carboxyl-terminated trithiocarbonate RAFT CTA was prepared by a previously
reported one step procedure (Scheme 4.1 and 4.2).97 This CTA is readily used to obtain
carboxyl-functionalized polymers. For the application of this type of CTA in click
chemistry, The terminal carboxyl group of CTA was converted to an alkyne ester group
via esterification with propargyl alcohol. The esterification reaction was done in the
presence

of

EDC

coupling

agent

and

DMAP

base.

EDC

[1-ethyl-3-(3-

dimethylaminopropyl) carbodiimide hydrochloride] is a water soluble carbodiimide


which is employed to increase the efficiency of the esterification reactions at lower
temperature.

O
SH

+ CS2 + CHCl3 + NaOH +

Step 1
S
OH
S

O
Carboxyl terminated RAFT Chain Transfer Agent (CTA)
HO

Carbodiimide Mediated
Esterification

Step 2

S
O
S

S
O

Alkyne terminated RAFT CTA

Scheme 4.1. Synthesis of click functionalized RAFT CTA.

67

Na+ CCl3

NaOH

CHCl3

H2 O

CCl3

: CCl2
O

Cl

S
Cl
NaOH
C12H25

SH

C12H25

S
OH

C12H25

NaOH

Cl
C12H25

Scheme 4.2. Mechanism of carboxyl-terminated RAFT CTA synthesis.

The product structure was confirmed by 1H and 13C NMR spectroscopy. The 1H
NMR spectrum displayed resonance at 4.7 ppm (CH2(O)-CCH) and 2.5 ppm (CHC)
which confirmed the attachment of propargyl group to CTA (Figure 4.2). The 13C NMR
spectrum also showed alkyne resonances at 75.4 and 76.8 ppm (Figure 4.3).

68

Figure 4.2. 1H NMR spectrum of click functionalized RAFT CTA.

Figure 4.3. 13C NMR of click-functionalized RAFT CTA.


69

4.4. Application of click-functionalized (alkyne-terminated) RAFT chain transfer agent


The combined use of click and RAFT chemistry can be used for surface
modification by both grafting to and grafting from approaches. In the first part of this
work, the modification of silica nanoparticles via the grafting to approach was studied.
In this approach alkyne-terminated CTA was used to prepare alkyne-terminated
polymers. Then a click reaction was used to attach the functionalized polymers on azidefunctionalized silica nanoparticles which ensured better grafting density. In the second
part, the combination of RAFT polymerization and click chemistry was studied to modify
the surface via the grafting from approach. In this approach an alkyne-terminated CTA
was directly coupled with azide modified silica nanoparticles. These CTA immobilized
silica particles were modified using surface mediated RAFT polymerization. In the third
part, the option of conducting tandem RAFT polymerization and click chemistry in one
pot was explored. A tandem process simplified this dual chemistry and could potentially
save time and money. This tandem method was developed as a novel method of surface
modification.

4.5. Surface modification via grafting to approach


In the grafting to method, end-functionalized chains react with substrates to
form covalently attached polymer brushes, making them more robust to solvent and
thermal treatment. Controlled polymerization techniques have been used to synthesize
chains of desired molecular weight and narrow molecular weight distribution that can
later be grafted to substrates.

The click functionalized RAFT CTA was used to

synthesize well-defined, end-functionalized polymers.

70

4.5.1. Synthesis of alkyne terminated polystyrene


The CC bond is susceptible to addition reactions as well as polymerization. The
trimethylsilyl group has been used as a protecting group for alkyne functionality during
ATRP19b,94 and anionic polymerization.100 Opsteen and van Hest19b used a protecting
group to circumvent complexation of the alkyne group with the copper catalyst during
polymerization. In anionic polymerization, the proton of the alkyne bond is acidic and
can react with the initiator. Matyjaszewski and co-workers19c had polymerized styrene at
90 oC using an alkyne-terminated ATRP initiator without using any protecting group.
With milder conditions and absence of copper ions in RAFT polymerization, we expect
little interference from the alkyne group was expected.
RAFT polymerization of styrene using the alkyne-terminated CTA was performed
at 90 oC (6% initiator) for 18 h. Well-defined PS with polydispersity of 1.1 and Mn =
9,000 g/mol was obtained (Figure A1). The presence of alkyne group was confirmed by
13

C NMR which showed alkyne resonances at 75 and 77 ppm.


S
O
S

O
NH2

S
O

Acrylamide

Alkyne terminated RAFT CTA


CN

NC
HO

OH

NN

700 C

3-5 h

Azo initiator

NH2
O

S
S

DMSO

O
O

Alkyne terminated PAAm

Scheme 4.3. Synthesis of alkyne functionalized polyacryalamide.


71

4.5.2. Synthesis of alkyne-terminated polyacrylamide


Well-controlled RAFT polymerization of acrylamide has been difficult due to
CTA degradation as a result of hydrolysis of the acrylamide monomers.101 Even a small
amount of monomer hydrolysis could produce enough ammonia to convert all
dithioesters in solution to thiols. Therefore appropriate choice of CTA and reaction
conditions are very important for achieving well-controlled RAFT polymerization of
acrylamide. McCormick and co-workers89 studied the kinetics and molecular weight
control of the RAFT polymerization of acrylamide in order to determine reaction
conditions that would provide excellent control of polymerization.

Due to the less

stabilized intermediate radicals, trithiocarbonate CTAs afford much faster polymerization


rates which are necessary for controlled polymerization of acrylamide. McCormick and
co-workers89 established that RAFT polymerizations of acrylamide in DMSO mediated
by the trithiocarbonate CTA in the presence of VA 501 free initiator at 70 oC provides
good control of molecular weight and the polydispersity for PAAm.
Exactly

the

same

experimental

conditions

were

employed

trithiocarbonate CTA to polymerize acrylamide (Scheme 4.3).

with

our

In 3.5-4 h of

polymerization, 50-60 % conversion was achieved, which indicated the fast rate of
polymerization. The goal was to prepare low molecular weight PAAm to facilitate
characterization by end group analysis. PAAm synthesized with low molecular weight
(Mn = 4000-5000 g/mol) and with a terminal alkyne group. The

13

C NMR spectrum

(Figure 4.4) revealed peaks at 75.4 and 76.8 ppm, which correspond to an alkyne. The
1

H NMR spectrum (Figure A2) displayed resonance at 4.7 ppm (OCH2-CCH).

72

The molecular weight analysis of PAAm was determined by

H NMR

spectroscopy (Figure 4.5) and MALDI-TOF-MS (Figure 4.6). By integrating the peak
corresponding to terminal a terminal proton (CH3-CH2, 0.87 ppm) with those protons of
repeating units (1.8 and 2.4 ppm), the Mn was calculated to be Mn = 4,500 g/mol
(Appendix A3).

4.5.3. MALDI-TOF-MS study of polyacrylamide


MALDI-TOF mass spectrometric analysis of the polyacrylamide was performed using cyano-4-hydroxycinnamic acid as the matrix. The molecular weight difference between
two consecutive peaks in the MALDI spectrum was 71 (Figure 4.6), which corresponds
to a single acrylamide monomer unit. It is generally believed that the C-S bond of the
chain end (close to the thiocarbonyl function) may fragment inside the spectrometer.102
MALDI-TOF-MS results were inconclusive for the investigation of polymer structure.

4.5.4. Synthesis of azide modified silica nanoparticles


A two step synthetic route was used to modify the surface of silica nanoparticle
with azide functionalities (Scheme 4.4).103

In the first step, 3-bromopropyl

trichlorosilane was used to introduce a bromide group on the surface of a silica


nanoparticle. IR analysis of the bromide-modified silica nanoparticle (Figure 4.7(d))
revealed an absorption at 2900 cm-1 corresponding to C-H stretching. Elemental analysis
confirmed the presence of bromine (1.53 %) which corresponds to a surface grafting
density of 3.65 groups/nm2 (calculation shown in Appendix A1).

73

Figure 4.4. 13C NMR spectrum of polyacrylamide.

Figure 4.5. 1H NMR spectrum of polyacrylamide.

74

Figure 4.6. MALDI-ToF-MS of polyacrylamide.

This bromide was substituted with azide by reaction with sodium azide. An azide
group was verified by an IR absorption at 2110 cm-1 (Figure 4.7(c)). Elemental analysis
also showed the presence of nitrogen (0.76 %) which corresponds to a surface grafting
density of 3.47 groups/nm2 (calculation shown in Appendix A2). Using TGA, weight
loss of 1.9% was observed [Figure 4.8(b)] for the bromide-modified nanoparticle and 1.6
% weight loss for the azide-modified nanoparticle [Figure 4.8(c)], which is consistent
with the difference in atomic masses of the functional groups.

75

OH

HO

HO

HO

Cl

OH

SiO2

HO

Y = H or Si

Y
Br

Cl Si

OH
OH

Y
O

80 C
Y
Silica--Bromide

Si

Br

Y = H or Si
O

SiO2

SiO2

Cl

Silica

Toluene

NaN3

N N N

Si
O

Y
Silica--Azide
Na Ascorbate
Click Reaction

CuSO4 , 5 H2O

DMSO , 500 C

NH2

O
O

Si
Y

Y = H or Si

Y
SiO2

N N

S
NH2

Silica--PAAm

Scheme 4.4. Modification of silica nanoparticle via grafting to approach.

4.5.5. Click coupling between azide-modified silica nanoparticle and alkyne-terminated


PAAm
The click reaction commonly corresponds to the Cu1-catalyzed Huisgen 1,3dipolar cycloaddition of azides and alkynes to afford 1,2,3-triazoles (Scheme 3).104 This
coupling reaction was first tried with alkyne-terminated PS and azide-modified silica
particles. PS-coated silica nanoparticles with a relatively high grafting density of 0.37
groups/nm2 were obtained which corresponds to a distance between grafting sites of 1.85
nm (model calculation shown in Appendix A5).105 The radius of gyration for PS (Mn =

76

9,000 g/mol) was calculated as 6 nm (model calculation shown in Appendix A6).106,107


Thus the diameter of the coils was larger than the distance between grafting points.
Based on the conventional interpretation of polymer brushes, these tethered PS chains
exist in the brush regime. These PS-modified particles were characterized by IR, TGA
and elemental analysis.
This 1,3-dipolar cycloaddition reactions are often performed in water for faster
reaction times and to alleviate the requirement for additional base. Aqueous solutions of
polyacrylamide are usually viscous.

Therefore, we utilized a mixture of water and

DMSO (1:7.5, v/v) as a medium for the click reaction between alkyne-terminated PAAm
and azide-modified silica nanoparticles (Scheme 4.4).

Figure 4.7.FT-IR spectra of (a) PAAm (b) silica-PAAm nanoparticles (c) silica-azide
nanoparticles (d) silica-bromide nanoparticles.

77

Figure 4.8. Thermogravimetric analysis (TGA) traces of (a) bare silica nanoparticles, (b)
silica-bromide nanoparticles, (c) silica-azide nanoparticles and (d) silica-PAAm
nanoparticles.

Grafting of polyacrylamide onto silica nanoparticles was characterized by IR


(Figure 4.7(b)) where a reduction of the azide resonance at 2110 cm-1 was observed, a
concomitant with an increase in the alkyl resonance at 2900-3000 cm-1 and appearance of
a carbonyl resonance at 1735 cm-1 that corresponds to the amide group.

The

polyacrylamide was degrafted from the silica by cleavage in HF; the 1H NMR spectrum
indicated the disappearance of the alkyne group and the presence of the triazole group.
TGA also confirmed the presence of polyacrylamide on the silica nanoparticle as
evidenced by an additional 8% weight loss corresponding to the grafted polyacrylamide
[Figure 4.8(d)]. Using elemental analysis, we confirmed the presence of polyacrylamide

78

on the surface of silica nanoparticle.

The increased carbon and nitrogen weight

percentage on silica particle after click coupling (Table 4.1). The ratio of carbon (3.84
%) and nitrogen (1.52 %) (2.52:1) on silica-PAAm matched with relative atomic mass of
carbon (36 amu) and nitrogen (14 amu) in repeating units.
Elemental analysis results were used to calculate the surface grafting density. We
calculated that 0.31 polyacrylamide chains were present on 1 nm2 surface area of silica
which corresponds to a 2 nm distance between grafting sites.105 The calculated diameter
of polyacrylamide coils (Mn = 4,500 g/mol, Rg = 2.6 nm) is larger than the distance
between grafting sites (model calculation shown in Appendix A5).

Thus, grafted

polyacrylamide was also in the brush regime. The grafting density of polyacrylamide
indicates that roughly 10% of the azide groups were converted during the click reaction
with the alkyne-functionalized polyacrylamide. This grafting density is lower than the
polystyrene coupling (0.37 chains/nm2).
Table 4.1. Elemental analysis of modified silica nanoparticles.

Result
(%)

Surface
grafting
densitya
(groups / nm2)

Bromine

1.53

3.68

Carbon

0.98

Nitrogen

0.76

Silica

Carbon

3.84

PAAm

Nitrogen

1.52

SilicaPS

Carbon

12

Sample

Elemental
Analysis

Silica
bromide
Silicaazide

3.45

0.31

Appendix A1, A2 and A5


79

0.37

There have been only few reports on the modification of silica nanoparticles by a
RAFT polymer.

Benicewicz and co-workers15 deposited RAFT CTA on silica

nanoparticles with grafting densities of 0.15-0.54 RAFT agents/nm2.

This was the

surface density of RAFT CTA prior to surface mediated polymerization. Guo and coworkers16 immobilized a lactose-containing polymer onto silica gel particles with grafting
densities 0.035-0.178 groups/nm2. In contrast, the surface grafting density obtained with
an ATRP initiator was 2-5 groups/nm2.10 Compared to the grafting density of the ATRP
initiator, the grafting densities were lower for the RAFT CTA. This is likely due to the
attachment of a more bulky RAFT CTA onto silica using the silanization reaction.
The grafting to approach has advantages and disadvantages. This approach is
experimentally simple and provides better control of polymerization, although it usually
suffers from a lower grafting density. In the grafting to approach reported here, the
grafting density was relatively high and even comparable to that of the grafting from
approach. We speculate that this is attributable to the facile nature of click chemistry.
Thus, we predict that the use of click chemistry for grafting-to modification of
nanoparticles will play an increasingly important role.
This click reaction was preferred in the presence of excess polymer. After the
click reaction, silica particles were easily collected by centrifugation.

The excess

polymer was precipitated into cold methanol to recover the alkyne-terminated polymer,
which can be used in subsequent nanoparticle modifications.
This approach was also used to attach alkyne-terminated diblock copolymer to the
surface of silica nanoparticles. Using alkyne-terminated RAFT CTA, PS-b-PMA block
copolymers were synthesized through two-step sequential RAFT polymerizations. RAFT

80

polymerization of styrene was performed at 90 oC using AIBN as a radical initiator to


synthesize well-defined PS [Mn = 3,000 g/mol, PDI = 1.1, (Figure A3)]. PS was then
employed as a macro RAFT CTA for the subsequent growth of PMA to form the final
diblock copolymer PS-b-PMA [Mn = 6,000 g/mol, PDI = 1.1, (Figures A4 and A5)].
These alkyne-terminated PS-b-PMAs were grafted onto azide-modified silica
nanoparticles using the click coupling reaction. Diblock brush-modified particles act as
responsive materials since they are able to alter their surface chemistry in response to
different environments. Diblock copolymer modified nanoparticles were characterized
by TGA (Figure 4.9) and IR (Figure 4.10).

Figure 4.9. Thermogravimetric analysis of (a) bare silica nanoparticles, (b) silicabromide, (c) silica-azide and (d) silica-PS-b-PMA (by RAFT).

81

Figure 4.10. FT-IR spectra of (a) bare silica nanoparticles, (b) silica-bromide, (c) silicaazide and (d) silica-PS-b-PMA (by RAFT).

4.6. Combination of ATRP and click chemistry for surface modification via grafting to
approach
ATRP and click chemistry were combined to modify the surface of silica
nanoparticles with well-defined diblock copolymer brushes (Scheme 4.5).

Alkyne

terminated ATRP initiator propargyl 2-bromoisobutyrate was prepared by reacting 2bromoisobutyryl bromide with propargyl alcohol in the presence of triethylamine.98
Using this functionalized initiator, ATRP of styrene was conducted at 90 0C for 9 hours
to obtain alkyne-terminated PS (Mn = 6,000 g/mol, PDI = 1.1) (Figures A6). The

13

NMR spectrum revealed peaks at 75.4 and 76.8 ppm, which correspond to an alkyne.

82

This alkyne-functionalized PS was coupled to azide-modified silica nanoparticles in a


DMF/water mixture to create PS silica hybrid nanoparticles.

These modified silica

nanoparticles were thoroughly washed to remove any physisorbed polymers.

Dried

nanoparticles were characterized by IR spectroscopy and TGA. IR spectroscopic analysis


of PS-modified particles showed new aromatic resonances at 1450-1500 cm-1 and 30003100 cm-1 which confirmed the presence of polystyrene on the surface of the particles
[Figure 4.11(d)]. TGA showed an additional weight loss that corresponds to attached
polystyrene [Figure 4.12(d)]. Methyl acrylate (MA) was also polymerized using the
propargyl 2-bromoisobutyrate ATRP initiator to synthesize alkyne-terminated PMA (Mn
= 6,000 g/mol, PDI = 1.1) (Figures A7) which was then click coupled to azide modified
silica nanoparticles to make PMA silica hybrid nanoparticles (Figures 4.11 and 4.12).
Thus silica particles were also modified with ATRP polymers via the grafting to
approach.
O
OH

Br

Br

Br

ATRP

O
SiO2

+
N N

Si

Br
O

Styrene, Toluene

Click
Reaction

Azide modified silica

CuBr / PMDETA

CuSO4/Na Ascorbate
DMF , 70 oC
Br

O
SiO2

O Si
O

N
N

Polystyrene modified silica nanoparticle

Scheme 4.5. Surface modification using ATRP and click chemistry.


83

Figure 4.11. IR spectra of (a) bare silica nanoparticles, (b) silica-bromide, (c) silica-azide,
(d) silica-PS (by ATRP) and (e) silica-PMA(by ATRP).

Figure 4.12. Thermogravimetric analysis (TGA) of (a) bare silica nanoparticles, (b)
silica-bromide, (c) silica-azide, (d) silica-PS (by ATRP) and (e) silica-PMA(by ATRP).
84

4.7. Surface modification via grafting from approach


The grafting from approach has become the method of choice to produce thick,
densely grafted, and covalently attached polymer brushes. A layer of polymerization
initiator/RAFT CTA is first covalently attached to a substrate, followed by an in situ
surface-initiated polymerization to produce the desired brush.
In this part, we report the successful preparation of high density, well-defined
polymer brushes by RAFT polymerization on silica nanoparticles. Our strategy includes
(1) grafting from approach for brush formation, (2) facile click reaction to immobilize
the RAFT agent, (3) synthesis of R-supported chain transfer agent over Z-supported
chain transfer agent14,15 and (4) use of a more active trithiocarbonate RAFT agent over
conventional dithioester type RAFT agent. With the advantage of the properties of high
density polymer brushes, these diblock copolymer modified silica nanoarticles open the
route to the elaboration of a new family of nanocomposite colloids with potential
applications in responsive or smart systems.
The grafting from approach for RAFT polymerization can be achieved in two
different ways. Free radical initiator can be anchored to the surface, followed by surfaceinitiated RAFT polymerization in presence of a RAFT CTA. Brittain and Baum12 used
this approach to modify silicate surfaces. However, the sensitive nature of the initiator
group only allows mild reaction conditions, which limits the grafting density. The
second approach involves surface immobilization of RAFT CTA followed by surfacemediated RAFT polymerization in the presence of free initiator. This approach was used
in this work as it gives more flexibility in reaction conditions.

85

The selection of RAFT CTA is crucial to synthesize well-defined polymer


brushes. RAFT CTA possesses two important areas of functionality, known as the R
group and the Z group. Each group plays an important role in the RAFT process. In
general, RAFT polymerization based on solid supports can be performed using both (a)
the R-group approach where the CTA is attached to the backbone via the leaving and
reinitiating R group and (b) the Z-group approach where the CTA is attached to the
backbone via the stabilizing Z group, and both methods have advantages and limitations.
14,15

In the R-group approach, where the solid support is part of the leaving R group,

higher molecular weight of grafted polymers and grafting density can be achieved, but
the molecular weight distribution may be broadened by the possible chain coupling. In
the Z-group approach, where the backbone is part of Z group, the RAFT process involves
the reaction of linear radical chains with the functional backbone, leading to a
monomodal molecular weight distribution and a better-defined grafted polymer, but the
grafting density is liable to decrease due to the shielding effect.108 We used the R group
approach to exploit its inherent ability to synthesize polymer brushes. This approach
affords both higher molecular weight and grafting density of the attached polymer. A
highly reactive trithiocarbonate type RAFT CTA was used to further increase the grafting
density.
Generally, a silyl condensation reaction is used to immobilize a RAFT CTA onto
silica nanoparticles. We used the more efficient click reaction to attach the RAFT CTA
onto silica nanoparticles. We synthesized an alkyne-functionalized RAFT CTA and
azide-functionalized silica nanoparticle in order to facilitate this coupling.

86

4.7.1. Synthesis of RAFT CTA anchored silica nanoparticles


The alkyne-terminated RAFT CTA 1 was prepared by an easy two step synthetic
procedure (Scheme 4.6). Azide-modified silica nanoparticles were also synthesized in
two steps (Scheme 4.6).

Silica nanoparticles were modified with 3-bromopropyl

trichlorosilane, and than the bromine group was converted into an azide group by a
nucleophilic substitution with sodium azide.
A click reaction was used to anchor the RAFT CTA onto the silica nanoparticle.
The reaction was performed in presence of Cu(I) for 18 h at 50 oC. The particles were
exhaustively washed with toluene to remove any ungrafted RAFT CTA. Figure 4.13d
shows the FT-IR spectra of RAFT-modified particles. It displays new peaks at 1735 cm-1
(C=O) and at 2900 cm-1 (C-H) confirming the presence of CTA group. It also shows a
reduction of azide absorption peak at 2098 cm-1. The residual azide absorption peak
indicates an incomplete conversion. TGA analysis (Figure 4.14d) revealed an additional
2.5% weight loss compared to the azide-modified particles.
Elemental analysis was used to calculate the loading of active sites (Table 4.2).
Percent sulfur was used to determine the amount of RAFT CTA on the surface of the
silica nanoparticles. It revealed that the grafting density of RAFT CTA was 1.2-1.3
groups/nm2 which is higher than the literature grafting density of RAFT CTA (0.15 to
0.68 groups/nm2).11 The RAFT CTA group is sterically large so direct immobilization is
inefficient.

After initial silation reaction with 3-bromopropyltricholorsilane (3.67

groups/nm2) followed by nearly quantitative azide displacement of the bromide, click


chemistry was used to introduce a RAFT CTA with a much higher grafting density.

87

Silica
nanoparticle

NaOH

Cl
Toluene , 80

+ CS2 + CHCl3 + CH3COCH3

SH

0C

Silica--bromide

Br

Cl Si
Cl

S
OH
S

O
O Si
O

O
Br

Carboxyl terminated RAFT CTA


HO

70 0C, 18 h

Esterification

NaN3 , DMF

S
Silica--azide

O
O Si
O

N N

O
Alkyne terminated RAFT CTA

Click Coupling

Silica--RAFT CTA

O
O Si
O

N N

Styrene, Toluene

S
S

RAFT Polymerization

AIBN
90 0C, 18 h

Silica--PS

O
O Si
O

S
O

N N

S
S

MA, Toluene
RAFT Polymerization

AIBN
80

Silica--PS-b-PMA

O
O Si
O

0C,

8h
O
S

N N

O S
O
CH3

Scheme 4.6. Modification of silica nanoparticle via grafting from approach.

88

4.7.2. Homopolymerization
For high grafting density of the polymer it is important to introduce RAFT CTA
on the surface with a sufficiently high grafting density. But not all of the RAFT CTA
sites on nanoparticle surface may participate in the growth of polymer chains because the
growing chains can sterically block access of the adjacent CTA sites on the nanoparticle
surface. Thus it is important to have an efficient surface-immobilized CTA to grow a
polymer chain. Therefore, the reactive trithiocarbonate CTA was immobilized in order to
achieve better efficiency. This CTA was found to have high chain transfer efficiency and
control over the radical polymerization because the carbon attached to the labile sulfur
atom is tertiary and bears a radical stabilizing carboxyl group.
Surface-mediated RAFT polymerization of silica nanoparticles was performed in
the presence of free RAFT CTA in order to better control the polymerization.
Polymerization was conducted in toluene at 90 oC for 18 h in the presence of AIBN (6
mol % of RAFT agent). Since the trithiocarbonate-based CTA is reactive, a small
amount of AIBN is adequate for the polymerization. The polymerization was conducted
at low conversion to avoid possible interparticle polymeric radical coupling. After 18 h
of polymerization, the Mn of free PS was 6,000 g/mol and the polydispersity was 1.2
(Figure A8). This is consistent with a slow and controlled RAFT polymerization. HF
was used to cleave the grafted polymer from silica the hybrid nanoparticles. The grafted
polymer had slightly lower molecular weight (Mn = 5,000 g/mol) and broader
polydispersity (1.3) than free polymer (Figure A9). Similar behavior was also observed
and explained for surface-initiated ATRP.10 The presence of free RAFT CTA slows
down the propagation of free chain in solution and favors the addition-fragmentation

89

reactions on the solid support. Formation of free polymer is also useful for assessment of
grafted polymer chains.
Either alkyne-terminated CTA or carboxyl-terminated CTA could be used for the
free RAFT CTA. This was another advantage of this approach as there was no need to
make any separate surface-bound CTA and free CTA. Usually surface active CTA
contains a trichlorosilane group making these CTAs unsuitable for use a free CTA. In
this work carboxyl-terminated RAFT CTA was used as free CTA because it was easier to
make.

Figure 4.13. IR spectra of (a) bare silica nanoparticles, (b) silica-bromide, (c) silica-azide,
(d) silica-RAFT CTA, (e) silica-PS and (f) silica-PS-b-PMA.

90

Figure 4.14. TGA analysis of (a) bare silica nanoparticles, (b) silica-bromide, (c) silicaazide, (d) silica-RAFT CTA, (e) silica-PS (Mn=6,000 g/mol) and (f) silica-PS-b-PMA
(Mn=34,000 g/mol).

The FT-IR spectrum showed evidence for the aromatic group at 1400 cm-1 and 3200
cm-1, which are characteristic absorbtion peaks for PS (Figure 4.13e). TGA analysis also
confirmed the presence of PS brush on the silica nanoparticles (Figure 4.14). There was
an additional 14% weight loss corresponding to the PS brushes. Elemental analysis
confirmed the presence of PS and was used to calculate the grafting density of PS brushes
(Table 4.2).

For 6000 molecular weight polystyrene surface grafting density was

calculated to be 0.65 chains/nm2. This grafting density correspond to a 1.4 nm distance

91

between grafting sites.105 The diameter of polystyrene coils (Mn = 6,000 g/mol, Rg = 5.1
nm) is larger than the distance between grafting sites (model calculation shown in
Appendix A5). Thus grafted polystyrene was in the brush regime.
Table 4.2. Grafting densities of modified silica nanoparticle.
Sample

Elemental

Result (%)

analysis

Surface grafting density on


silica nanoparticle
(groups/nm2)a

Silica - bromide

Bromine

1.53

3.68

Carbon

0.98

Nitrogen

0.76

Silica RAFT

Carbon

3.84

CTA

Sulfur

1.52

1.20

Carbon

30.02

0.65

Carbon

14.58

0.60

Carbon

14.14

0.22

Silica - azide
3.45

Silica PS
(Mn=12,000 g/mol)
Silica PS
(Mn=6,000 g/mol)
Silica-PS-b-PMA
(Mn=34,000 g/mol)
a

Appendix A1, A2 and A5

4.7.3. Block copolymerization


One of the advantages of controlled free radical polymerization is the potential to
synthesize diblock copolymers. The PS silica hybrid nanoparticle was used as a macro
RAFT CTA to perform surface-mediated RAFT polymerization of methyl acrylate (MA).

92

Surface-mediated RAFT polymerization was performed in the presence of free CTA.


Free polymer in solution was used to monitor the molecular weight of grafted block
polymer on the silica nanoparticle. Free PMA was found to have a Mn of 28,000 g/mol
and a polydispersity of 1.10 (Figure A10).
Figure 4.13f shows the IR spectra of the block copolymer-modified silica
nanoparticle. The characteristic absorption bands of PS are still present in addition to the
new adsorption at 1727 cm-1 (carbonyl peak) which is characteristic of PMA. In the TGA
analysis, these hybrid particles underwent an additional 24 % weight loss due to second
block.

Grafting density of second block was also calculated and indicated a 40%

reinitiation efficiency.

Similar reinitiation efficiency was also observed in surface

initiated ATRP.10
In this study, the successful preparation of high density, well defined polymer
brushes on silica nanoparticles via RAFT polymerization was reported. A four pronged
strategy was used to maximize the grafting density.

4.8. Tandem approach to surface modification


Traditionally there are two main routes for surface modification. Route I is the
grafting to approach where a preformed, end functionalized polymer couples to an
activated surface. Route II is the grafting from approach that typically proceeds via
propagation from a surface-immobilized initiator. Because route II affords a higher
grafting density, this route is generally preferred.

In both routes, the surface

immobilization reaction and polymerization reactions are performed sequentially. In


route 1 polymerization is done first then those polymers are attached to the surface via

93

surface immobilization reaction. While in route II, these two reactions are done in
reverse order.
Herein a novel approach for surface modification is reported in which
polymerization and surface immobilization reactions proceed simultaneously.

Both

RAFT polymerization and click coupling reactions were conducted in tandem, thus
demonstrating mutual compatibility. This tandem approach was applied to modify the
surface of silica particles with polystyrene. The relative rates of both reactions were
altered to control grafting density.

Silica
nanoparticle

SH +
NaOH

Cl
Toluene , 80 0C

Cl Si

Br

OH

Cl
Silica--bromide

70 0C, 18 h

Silica--azide

O
O Si
O

- +
N

O
S
N

O
Alkyne terminated RAFT CTA

RAFT Polymerization

CuBr / PMDETA

Click Coupling

900 C, 18h
O
O Si
O

Esterification

HO

NaN3 , DMF
O
O Si
O

Carboxyl terminated RAFT CTA

Br

Styrene, Toluene

Silica--PS

CS2 + CHCl3 + CH3COCH3

O
N

S
n
O

N N

S
S

Scheme 4.7. Surface modification of silica nanoparticles using the tandem approach.

94

For tandem RAFT polymerization and click chemistry, alkyne-terminated RAFT


chain transfer agent (CTA) and azide-modified silica nanoparticles were separately
synthesized (Scheme 4.7).

Traditionally click chemistry is performed using

CuSO4/sodium ascorbate [as a source of Cu(I)] catalyst in water or a mixture of water


and polar solvent. Simultaneous RAFT polymerization and click chemistry were initially
attempted in the presence of CuSO4/sodium ascorbate in dimethylformamide. But good
control over the polymerization was not observed. It is speculated that the ascorbic acid
may be acting as a radical scavenger.109 Another source of Cu(I) catalyst that has been
used

in

click

chemistry

of

polymers

is

the

CuBr/PMDETA

(pentamethyldiethylenetriamine) complex.110 The CuBr/PMDETA complex is widely


used in ATRP where it undergoes a reversible chain transfer reaction with alkyl halide.
In the absence of an alkyl halide group, it was expected that this copper complex would
be innocuous during RAFT polymerization.
Simultaneous RAFT polymerization and click chemistry were performed in a
flask containing styrene, azide-modified silica, alkyne-terminated RAFT CTA,
CuBr/PMDETA (in a catalytic amount) and toluene at 90 oC under an inert atmosphere
for 18 h. The click (alkyne)-functionalized CTA participated in RAFT polymerization as
well as the click coupling reaction with azide-modified silica particles. At any particular
reaction time, the CTA would be attached to the silica nanoparticles and present in
solution. Polystyrene (PS) chains were growing via RAFT polymerization whether the
process is occurring on the silica or in solution. At the reaction end, there will be PS
attached to silica particles and free PS in solution. PS-modified silica nanoparticles were
separated by centrifugation and characterized by thermogravimetric analysis (TGA)

95

(Figure 4.15), Fourier transform infrared (FT-IR) spectroscopy (Figure 4.16) and
elemental analysis (Table 4.3).
Table 4.3. Grafting density of PS on silica nanoparticles as prepared by different grafting
methods.

Synthetic
approach

Grafted
PS (Mn)
g/mol

Elemental
analysis of
Carbon
(weight %)

Grafting
density by
elemental
analysis
(groups/nm2)a

Grafting
density by TGA
(groups/nm2)a

Grafting to

9000

11.74

0.30

0.32

Tandem

7000

15.39

0.51

0.54

Grafting
from

12000

30.02

0.65

0.68

Appendix A5

Figure 4.15. TGA analysis of (a) bare silica nanoparticles, (b) silica-bromide, (c) silicaazide, (d) silica-PS (Mn =9000 g/mol) by grafting to approach, (e) silica-PS (Mn =7000
g/mol) by tandem approach and (f) silica-PS (Mn =12000 g/mol) by grafting from
approach.
96

Figure 4.16. IR spectra of (a) bare silica nanoparticles, (b) silica-bromide, (c) silica-azide,
silica-PS by (d) grafting to approach, (e) tandem approach and (f) grafting from
approach.

Free PS was precipitated from methanol and found to have a molecular weight Mn
= 9,000 g/mol and polydispersity of 1.1 (Figure A12). The presence of an alkyne group
was confirmed by

13

C NMR spectroscopy.

These results are consistent with the

controlled nature of RAFT polymerization. PS chains attached to silica were cleaved


using HF to afford PS with Mn = 7,000 g/mol and polydispersity of 1.3 (Figure A13).
The molecular weight was similar for PS immobilized on silica compared to PS formed
in solution.

Similar behavior has been reported when surface-initiated ATRP was

performed in presence of free ATRP initiator.10


This tandem approach was compared with grafting to and grafting from in
terms of grafting density. For the grafting to method, the isolated free PS with an
97

alkyne group was reacted with azide-modified silica. For the grafting from method,
alkyne-terminated RAFT CTA was coupled with azide-modified silica followed by
surface-ediated RAFT polymerization in the presence of free CTA. For each method, the
grafting density was calculated using elemental analysis and TGA.
The grafting density obtained by the tandem method was between the grafting
densities obtained by the grafting to and the grafting from approaches (Table 4.3). In
the grafting to method, polymer chains must diffuse to the silica particles, while in the
grafting from approach, only low molecular weight monomer diffuses to the silica
surface. Consistent with the literature, the lower grafting densities are for grafting
to.111 The tandem process of RAFT and click chemistry represents an intermediate case
in terms of sterics consistent with the intermediate grafting density.
By changing the relative rates of the click reaction and RAFT polymerization, this
tandem method can be shifted between grafting to and grafting from methods. For
example, with a faster click reaction, there would be a higher probability that only RAFT
CTA or RAFT CTA having a few monomers would attach to silica nanoparticles and it
would be closer to the grafting from approach. Recent kinetic studies indicate that the
rate of the catalytic process in the click reaction is generally second order in copper.112
The Cu(I) concentration was used to change the rate of the click reaction and this variable
was studied to determine its effect on grafting density and molecular weight (Table 4.4).
When 10-fold more Cu(I) was used in the tandem process, grafting density increased
significantly (from 0.51 groups/nm2 to 0.70 groups/nm2) and became closer to the
grafting from approach.

But there were no effects on molecular weight and

polydispersity of the RAFT polymers. Therefore, it is possible to control the length of a

98

polymer brush (RAFT polymerization) and grafting density (click reaction) of the
tethered polymer brushes separately.
Table 4.4. Effect of Cu(I) concentration on tandem approach of surface modification.

Concentration
of CuBr
(in ratio)

1
10

Elemental
analysis

Free PS

Sulfur
(%)

Carbon
(%)

Mn
(g/mole)

Polydispersity

0.29
0.52

15.39
20.76

9000
9000

1.11
1.10

Grafting
density by
elemental
analysis
(groups/
nm2)a
0.51
0.72

Grafting
density
By TGA
(groups/
nm2)a
0.54
0.70

Appendix A5
This tandem approach definitely has clear advantages over the other two methods

in terms of labor, time and cost. It definitely saves the step of surface immobilization on
surface.

This step requires a toxic solvent and the washing of modified silica

nanoparticles. The tandem approach provides grafting density in the brush regime. In a
typical grafting from approach, there is a need to separately synthesize surface-bound
RAFT CTA (mostly silane group terminated RAFT CTA) and free CTA. But in this
method there is only one kind of CTA which performs both as a free CTA and surface
bound CTA.

4.9. Surface modification of silica nanoparticles by surface-initiated atom transfer radical


polymerization
The synthesis of polymer brushes on nanosized silica particles by surface-initiated
polymerization (SIP) techniques is being increasingly reported in the literature.8,111
Polymer brushes by SIP on silica can be prepared using free radical, cationic, ringopening methathesis, atom transfer radical polymerization (ATRP) and reverse addition
fragmentation transfer polymerization. Among these various methods, controlled ATRP
99

has emerged as a useful technique to covalently bond polymers to a surface due to its
versatility, simplicity, and tolerance to adventitious impurities such as water.

The

strategy involves pretreating inorganic nanoparticles with compounds having initiating


groups. In the present study the surfaces of silica were modified with diblock copolymer
brushes. It is proposed that grafting polymer chains onto the silica surface will improve
the compatibility of the particles in different media.

4.9.1. Surface-bound ATRP initiator


An ATRP initiator commonly used with copper halide catalyst systems is an
bromoisobutyrate,This type of initiator has been successfully immobilized on surfaces to
prepare a wide range of polymer brushes.10,48 In this study, -(2-bromoisobutyryl)propyl
triethoxysilane (BITES) was used for immobilization onto the silica surface. Due to the
proven versatility and facile attachment chemistry, this initiator was selected for surfaceinitiated polymerizations.

OEt
EtO

Si

O CH3
(CH2)3 O C C Br
CH3

OEt

Figure 4.17. Structure of BITES initiator.

BITES initiator was synthesized by a two step process (Scheme 4.8).10 BITES
was prepared via the hydrosilation of allyl ester group-containing ATRP initiators with
triethoxysilane. The use of a triethoxysilane group provides three sites for attachment and
100

improves the surface coverage. The initiator was characterized by 1H NMR and

13

NMR spectroscopy (Figure 4.18 and 4.19).

O CH3
OH +

Br

C C

triethylamine
Br

CH3
allyl alcohol

C2H5O

Si

OC2H5

triethoxysilane

CH3

CH3

OC2H5

CH3

Br

Br

+ HBr

allyl 2-bromoisobutyrate

2-bromoisobutyryl bromide

OC2H5

CH3

"Pt"
Hydrosilation

C2H5O

Si

(CH2)3

OC2H5

allyl 2-bromoisobutyrate

CH3

(3-[2 bromoisobutyryl]propyl)
triethoxysilane

Scheme 4.8. Synthesis of BITES ATRP initiator.

Figure 4.18. 1H NMR spectrum of BITES initiator.

101

CH3

Br

C2 H5 O

Si

Y = H or Si

Toluene
Br

80
1100 C0C

-O
HO

OC2H5

Y
-O
OO
-O
O
HO SiO 2
Si
O
-O
O
O
O
-O
Y

Br

Silica

O-

HO
HO

OC2H5

OH

HO
OH
HO
OH
SiO
2
HO
OH
OH
HO
O
H
HO

OOH
O-

OH
OH
OH

Figure 4.19. 13C NMR spectrum of BITES initiator

Silica-ATRP int.
Toluene
CuBr
PMDETA

SiO2

Styrene
Ethyl 2-bromoisobutyrate

Toluene, MA, 90 0C
SiO 2

CuBr / PMDETA
Ethyl 2-bromoisobutyrate

Silica--PS

Silica PS-b-PMA

Scheme 4.9. Surface modification of silica particles via surface-initiated ATRP.

102

4.9.2. Immobilization of initiator on the surface of silica nanoparticles


The initiator BITES was deposited onto silica at 110 oC in anhydrous toluene
(Scheme 4.9). XPS studies (Figure.4.20) were performed on the initiator-immobilized
silica particles which clearly indicated the presence of bromine. FT-IR spectroscopy was
performed on the initiator-immobilized silica and the presence of an ester absorbance
peak at 1730 cm-1 (Figure 4.21) confirmed the presence of the surface-immobilized
initiator. The attached polymer density was calculated using elemental analysis and TGA
(Figure 4.22) and corresponded to 3.5-4 initiator groups/nm2 by both methods.

Figure 4.20. XPS spectrum of BITES initiator modified silica nanoparticles.

103

4.9.3. ATRP on the initiator-modified nanoparticles


Surface initiated polymerizations were performed to synthesize homopolymer and
diblock copolymer brushes.

4.9.3.1. Synthesis of homopolymer brushes


ATRP was performed on the initiator-modified silica particles using two different
monomers: styrene and methyl acrylate. Ethyl 2-bromoisobutyrate (EBiB) was added as
a free initiator in solution. The addition of EBiB has two advantages. First, molecular
weight data of the free polymer has been shown to correlate well with that of the surfacebound

polymer9,

affording

important

information

about

the

ATRP-initiated

polymerization. A second and more important benefit of added free initiator is that it
provides a high enough concentration of the deactivating Cu(II) species to control
polymerization both at the surface and throughout the rest of solution. Styrene and MA
polymerizations were successfully performed using surface-immobilized initiators BITES
in toluene and anisole solutions respectively.
The molecular weights of free polymer and cleaved polymer were determined
using GPC. It was possible to make homopolymers of PS and PMA having molecular
weights of 3,000-15,000 g/mol. The polydispersities (1.05-1.2) were relatively narrow
for both the polymers indicating that polymerization occurred in a controlled manner.

104

4.9.3.2. Synthesis of block copolymers brushes.


The homopolymer-silica particle was used as an initiator for the synthesis of the
second block. Two block copolymers-modified silica nanoparticles (silica-PS-b-PMA
and silica-PMA-b-PS) were synthesized.

The addition of free initiator, 2-ethyl

bromoisobutyrate, gave free polymer which was used to determine the molecular weight
of the second block. IR (Figure 4.21), TGA (Figure 4.22), and elemental analyses were
performed to confirm the presence of the second block. Polymer modified particles were
also analyzed by TEM (Figure 4.23 and 4.24)
The reinitiation efficiency for the second block was low. In a typical TGA
measurement for silica-PS-b-PMA, bare silica had a weight loss of 2.35 % which is
attributed to adsorbed water (Figure 4.22). Immobilization of the initiator increased the
weight loss to 3.77 %. After polymerization of styrene, the silica-PS [PS (Figure A13);
Mn = 7000 g/mol, PDI =1.1] had a weight loss of 18.8 %, and for the silica-PS-b-P MA
[free PMA (Figure A14); Mn = 3000 g/mol, PDI = 1.4] the weight loss increased by just
3.15 % for a residual weight of 77.93 %. TGA was used to estimate the grafting density
of attached homopolymers and block polymers. The grafting densities of first block (PS,
Mn = 7000 g/mol) and second block (PMA, Mn = 3000 g/mol) were 0.55 groups/nm2 and
0.20 groups/nm2 respectively (model calculation shown in A5). These results suggest
poor re-initiation efficiency (36%) of the second block which is consistent with the
literature.10 Inaccessibility of living chain ends of the bottom PS block to the second
monomer and unavoidable termination events during styrene polymerization (resulting in
dead polymer chains) are two main culprits for this poor re-initiation efficiency.

105

Figure 4.21. IR spectra of modified silica nanoparticles.

106

Figure 4.22. TGA of modified silica nanoparticles.

Figure 4.23. TEM images of PS-coated silica particles.

Figure 4.24. TEM images of PMA coated silica particles.

107

CHAPTER V
SUMMARY AND CONCLUSIONS

The purpose of this research was to utilize a combination of controlled free


radical polymerization and click chemistry for surface modifications. In this research,
reversible addition fragmentation transfer (RAFT) polymerization was combined with
Huisgen 1,3-dipolar cycloaddition reactions to modify the surface of silica nanoparticles
with different homopolymer and diblock copolymer brushes via the grafting to, the
grafting from and a novel tandem approach. Combination of atom transfer radical
polymerization (ATRP) and click chemistry was also explored for surface modification
of silica nanoparticles.

Surface-initiated ATRP was also used to attach diblock

copolymer brushes onto the surface.


A novel alkyne-terminated trithiocarbonate chain transfer agent (CTA) was
synthesized, which permitted the combination of RAFT polymerization and click
chemistry.

Using this CTA, alkyne-terminated RAFT homopolymers and diblock

copolymers were synthesized. Click chemistry was then used to graft polymer brushes
onto azide-modified silica nanoparticles. To the best of our knowledge, the highest
reported surface grafting density (0.6-0.7 groups/nm2) was obtained for RAFT polymer
attachment via the grafting to approach. Using this approach, we reported the first
synthesis of polyacrylamide-silica hybrid nanoparticles via any living free radical
polymerization method. It was determined that the attached polymers exist in the brush

108

regime. ATRP and click chemistry were also combined to make polymer-modified silica
nanoparticles via the grafting to approach.
The combined use of RAFT polymerization and click chemistry enabled the
preparation of high density polymer brushes on silica nanoparticle via the grafting from
approach also.

In this approach, the click reaction was used to immobilize a

trithiocarbonate RAFT agent onto silica nanoparticles. The click reaction resulted in a
grafting density of 1.2-1.3 RAFT agents/nm2, a value much higher than previously
reported in the literature. The high reactivity of a trithiocarbonate CTA was further
utilized to produce a high grafting density (0.6-0.7 groups/nm2) of polymer brushes. The
living nature of RAFT polymerization was also exploited to make diblock copolymer
modified silica nanoarticles.
In a third study, a novel surface modification was developed which
simultaneously employed RAFT polymerization and click chemistry. This approach has
advantages of time and cost as the polymerization and surface immobilization reactions
were performed simultaneously in a tandem manner. Grafting density obtained by this
tandem method was in between grafting densities observed for the grafting to and
grafting from methods. The grafting density of tethered chains was varied by changing
relative rates of RAFT polymerization and click chemistry by adjusting the level of the
Cu(I) catalyst for the click reaction.

109

REFERENCES

1.

Savin, D. A.; Pyun, J.; Patterson, G. D.; Kowalewski, T.; Matyjaszewski, K. J.


Polym. Sci. Part B: Polym. Phys. 2002, 40, 2667.

2.

Garca, M.; van Zyl, W. E.; ten Cate, M. G. J.; Stouwdam, J. W.; Verweij, H.;
Pimplapure, M. S.; Weickert, G. Ind. Eng. Chem. Res. 2003, 42, 3750.

3.

Zhao, B.; Brittain, W. J. Prog. Polym. Sci. 2000, 25, 677.

4.

Belder, G. F.; ten Brinke, G.; Hadziioannou, G. Langmuir 1997, 13, 4102.

5.

Mansky, P.; Liu, Y.; Huang, E.; Russell, T. P.; Hawker, C. Science 1997, 275,
1458.

6.

Lyatskaya, Y.; Balazs, A. C. Macromolecules 1998, 31, 6676.

7.

Pruker, O.; Rhe, J. Mater. Res. Soc. Symp. Proc. 1993, 304, 1675.

8.

Radhakrishnan, B.; Ranjan, R.; Brittain, W. J. Soft Mater 2006, 2, 386.

9.

Matyjaszewski, K.; Davis, T. P. Eds. Handbook of Radical Polymerization;


Wiley: Hoboken, NJ, 2002.

10.

von Werne, T.; Patten, T. E. J. Am. Chem. Soc. 2001, 123, 7497.

11.

Chiefari, J.; Chong, Y. K.; Erocle, 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.

12.

Ranjan, R.; Brittain, W. J. Macromolecules 2007, 40(17), 6217.

13.

Baum, M.; Brittain, W. J. Macromolecules 2002, 35, 610.

14.

Tsujii, Y.; Ejaz, M.; Sato, K.; Goto, A.; Fukuda, T. Macromolecules 2001, 34,
8872.

15.

(a) Li, C.; Han, J.; Ryu, C. Y.; Benicewicz, B. C. Macromolecules 2006, 39(9),
3175. (b) Li, C.; Benicewicz, B. C. Macromolecules 2005, 38(14), 5929.
110

16.

Guo, T. Y.; Liu, P.; Zhu, J. W.; Song, M. D.; Zhang, B. H. Biomacromolecules
2006, 7, 1196.

17.

Jordi, M. A.; Seery, T. A. P. J. Am. Chem. Soc. 2005, 127, 4416.

18.

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

19.

(a) Lutz, J. F.; Borner, H. G.; Weichenha, K. Macromol. Rapid Commun. 2005,
26, 514. (b) Opsteen, J. A.; van Hest, J. C. M. Chem. Commun. 2005, 1, 57. (c)
Tsarevsky, N. V.; Sumerlin, B. S.; Matyjaszewski, K. Macromolecules 2005, 38,
3558. (d) Laurent, B. A.; Grayson, S. M. J. Am. Chem. Soc. 2006, 128, 4238. (e)
Gao, H.; Matyjaszewski, K. J. Am. Chem. Soc. 2007, 129, 6633.

20.

Moad, G.; Rizzardo, E.; Thang, S. H. Aust. J. Chem. 2005, 58, 379. (b) Moad, G.;
Rizzardo, E.; Thang, S. H. Aust. J. Chem. 2006, 59, 669.

21.

O'Reilly, R. K.; Joralemon, M. J.; Lui, W.; Hawker, C. J.; Wooley, K. L. J.


Polym. Sci. Part A: Polym. Chem. 2006, 44, 5203.

22.

Gondi, S. R.; Vogt, A. P.; Sumerlin, B. S. Macromolecules 2007, 40, 474.

23.

Qumener, D.; Davis, T. P.; Barner-Kowollik, C.; Stenzel, M. H. Chem. Commun.


2006, 48, 5051.

24.

Jang, J.; Park, H. J. App. Polym. Sci. 2002, 83, 1817.

25.

Hommel, H.; Touhami,A.; Legrand, A. P. Makromol. Chem. 1993, 194 879.

26.

Wei, Y.; Fan, L. M.; Chen, L. R. Chromatographia 1997, 46, 637.

27.

Mathew, J. P.; Srinivasan, M. Polym. Int. 1992, 29, 179.

28.

(a) Hideyo, M.; Yutaka, S.; Sumio, T.; Yasubumi, T. Polypropylene resin
composition for automobile component parts. Japanese Patent JP 9165478 A. (b)
Mizutani, Y.; Nago, S. J. Appl. Polym. Sci. 1999, 72, 1489.

29.

Mittal, K. L., Ed. Adhesion Aspects of Polymeric Coatings; Plenum: New York,
1983.

30.

Chang, S. M.; Lee, M.; Kim, W.S. J. Colloid Interface Sci. 2005, 286, 536.

31.

Yu, Y.; Chen, W. Mat. Chem. Phys. 2003, 82, 388.

32.

Stber, W.; Fink, A.; Bohn, E. J. J. Colloid Interface Sci. 1968, 26, 62.

33.

Spange, S. Prog. Polym. Sci. 2000, 25, 781.


111

34.

Yoshikawa, S.; Tsubokawa, N. Polym. J. 1996, 28, 317.

35.

Pruker, O. ; Rhe, J. Macromolecules 1998, 31, 592.

36.

Pruker, O.; Rhe, J. Macromolecules, 1998, 31, 602.

37.

Ueda, J.; Sato, S.; Tsunokawa, A.; Yamauchi, T. ; Tsubokawa, N. Eur. Polym. J.
2005, 41, 193.

38.

Sulitzky, C. ; Rckert, B.; Hall, A. J.; Lanza, F.; Unger, K.; Sellergren, B.
Macromolecules 2002, 35, 79.

39.

Zhang, K.; Chen, H.; Chen, X.; Chen, Z.; Cui, Z.; Yang, B. Macromol. Mater.
Eng. 2003, 288, 380.

40.

Gu, S.; Kondo, T.; Konno, M. J. Colloid Interface Sci. 2004, 272, 314.

41.

Zheng, Z.; Yu, J.; Guo, Z. X. Macromol. Chem. Phys. 2004, 205, 2197.

42.

Reculusa, S.; Mingotaud, C.; Bourgeat-Lami, E.; Duguet, E.; Ravaine, S.


Nanoletters 2004, 4, 1677.

43.

Ding, X.; Zhao, J.; Liu, Y.; Zhang, H.; Wang, Z. Materials Lett. 2004, 58, 3126.

44.

Hawker, C. J. J. Am. Chem. Soc. 1994, 116, 1185.

45.

Wang, J. S.; Matyjaszewski, K. J. Am .Chem. Soc. 1995, 117, 5614.

46.

Kato, M.; Kamigaito, M.; Sawamoto, M.; Higashimura, T. Macromolecules 1995,


28, 1721.

47.

a) Kharasch, M. S.; Engelmann, H.; Mayo, F. R. J. Org. Chem.1938, 2, 288. (b)


Kharasch, M. S.; Jensen, E. V.; Urry, W. H. Science 1945, 102, 169. (c)
Kharasch, M. S.; Elwood, E. V.; Urry, W. H. J. Am.Chem. Soc. 1947, 69, 1100.
(d) Kharasch, M. S.; Friedlander, H. N. J.Org. Chem. 1949, 14, 239. (g) Walling,
C.; Huyser, E. S. Org. React. 1963, 13, 91.

48.

(a) Matyjaszewski, K.; Xia, J. Chem. Rev. 2001, 101, 2921. (b) Patten, T.;
Matyjaszewski, K. Adv.Mater. 1998, 10, 901. (c) Lutz, J. F.; Jahed, N.;
Matyjaszewski, K. J. Polym. Sci., Part A: Polym. Chem. 2004, 42, 1939.

49.

von Werne, T.; Patten, T. E. J. Am. Chem. Soc. 1999, 121, 7409.

50.

(a) Matyjaszewski, K.; Miller, P.; Shukla, N.; Immaraporn, B.; Gelman, A.;
Luokala, B.; Siclovan, T.; Kickelbick, G.; Vallant, T.; Hoffmann, H.; Pakula, T.
Macromolecules 1999, 32, 8716. (b) Kim, N. Y.; Jeon, N. L.; Choi, I. S.; Takami,
112

S.; Harada, Y.; Finnie, K. R.; Girolami, G. S.; Nuzzo, R. G.; Whitesides, G. M.;
Labinis, P. E. Macromolecules 2000, 33, 2793.
51.

El Harrak, A.; Carrot, G.; Oberdisse, J.; Eychenne-Baron, C.; Bou, F.


Macromolecules 2004, 37, 6376.

52.

Pyun, J.; Jia, S.; Kowalewski, T.; Patterson, G. D.; Matyjaszewski, K.


Macromolecules 2003, 36, 5094.

53.

(a) Wang, X. S.; Lascelles, S. F.; Jackson, R. A.; Armes, S. P. Chem. Commun.
1999, 18, 1817. (b) Wang, X. S.; Armes, S. P. Macromolecules 2000, 33, 6640.

54.

Perruchot, C.; Khan, M. A.; Kamitsi, A.; Armes, S. P.; von Werne, T.; Patten, T.
E. Langmuir 2001, 17, 4479.

55.

Chen, X.; Randall, D. P.; Perruchot, C.; Watts, J. F.; Patten, T. E.; von Werne, T.;
Armes, S. P. J. Colloid Interface Sci. 2003, 257, 56.

56.

Carrot, G.; Diamanti, S.; Manuszak, M.; Charleux, B.; Vairon, J. P. J. Polym. Sci.,
Part A: Polym. Chem. 2001, 39, 4294.

57.

Ohno, K.; Morinaga, T.; Koh, K.; Tsujii, Y.; Fukuda, T. Macromolecules 2005,
38, 2137.

58.

Boettcher, H.; Hallensleben, M. L.; Nub, S.; Wurm, H. Polym. Bull.(Berlin) 2000,
44, 223.

59.

Mori, H.; Seng, D. C.; Zhang, M.; Mller, A. H. E. Langmuir 2002, 18, 3682.

60.

Li, D.; Sheng, X.; Zhao, B. J. Am. Chem. Soc. 2005, 127, 6248.

61.

(a) Wang, Y.; Pei, X.; He, X.; Yuan, K. Eur. Polym. J. 2005, 41, 1326. (b) Wang,
Y.; Pei, X.; Yuan, K. Mater. Lett. 2004, 59, 520.

62.

Georges, M.; Veregin, R.; Kazamaier, P.; Hamer, G. G. Macromolecules 1993,


26, 2987.

63.

Malmstrm, E. E.; Hawker, C. J. Macromol. Chem. Phys. 1998, 199, 923.

64.

Saban, M. D.; Georges, M. K.; Veregin, R. P. N.; Hamer, G. K.; Kazmaier, P. M.


Macromolecules 1995, 28, 7032.

65.

Bartholome, C.; Beyou, E.; Bourgeat-Lami, E.; Chaumont, P.; Zydowicz, N.;
Macromolecules 2003, 36, 7946.

66.

Bartholome, C.; Beyou, E.; Bourgeat-Lami, E.; Chaumont, P.; Lefebvre, F.;
Zydowicz, N. Macromolecules 2005, 38, 1099.
113

67.

Kasseh, A.; Ait-Kadi, A.; Riedl, B.; Pierson, J. F. Polymer 2003, 44, 1367.

68.

(a) Parvole, J.; Billon, L.; Montfort, J. P. Polym. Int. 2002, 51, 1111. (b) Parvole,
J.; Laruelle, G.; Khoukh, A.; Billon, L. Macromol. Chem. Phys. 2005, 206, 372.

69.

Blomberg, S.; Ostberg, S.; Harth, E.; Bosman, A. W.; van Horn, B.; Hawker, C. J.
Polym. Sci., Part A: Polym. Chem. 2002, 40, 1309.

70.

Nouvel, C.; Ydens, I.; Degee, P.; Dubois, P.; Dellacherie, E.; Six, J. L. Macromol.
Symp. 2001, 175, 33.

71.

Mecerreyes, D.; Jrme, R.; Dubois, P. Adv. Polym. Sci. 1999, 147, 1.

72.

Carrot, G.; Rutot-Houz, D.; Pottier, A.; Dege, P.; Hilborn, J.; Dubois, P.;
Macromolecules 2002, 35, 8400.

73.

(a) Yoon, K. R.; Koh, Y. J.; Choi, I. S. Macromol. Rapid Commun. 2003, 24, 207.
(b) Yoon, K. R.; Lee, Y. J.; Lee, J. K.; Choi, I. S. Macromol. Rapid Commun.
2004, 25, 1510.

74.

(a) Joubert, M.; Delaite, C.; Bourgeat-Lami, E.; Dumas, P. J. Polym. Sci., Part A:
Polym. Chem. 2004, 42, 1976. (b) Joubert, M.; Delaite, C.; Bourgeat-Lami, E.;
Dumas, P. Macromol. Rapid Commun. 2005, 26, 602.

75.

Mingotaud, A. F.; Stephane, R.; Mingotaud, C.; Keller, P.; Sykes, C.; Duguet, E.;
Serge, R. J. Mater. Chem. 2003, 13, 1920.

76.

Oosterling, M. L. C. M.; Sein, A.; Schouten, A. J. Polymer 1992, 33, 4394.

77.

Zhou, Q.; Wang, S.; Fan, X.; Advincula, R. Langmuir 2002, 18, 3324.

78.

Zirbs, R.; Binder, W.; Gahleitner, M.; Mach, D. Macromol. Symp. 2007, 254, 93.

79.

Tolnai, Gy.; Csempesz, F.; Kabai-Faix, M.; Klmn, F.; Keresztes, Zs.; Kovca,
A. L.; Ramsden, J. J.; Hrvlgyi, Z. Langmuir 2001, 17, 2683.

80.

El Harrak, A.; Carrot, G.; Oberdisse, J.; Bou, F. Polymer 2005, 46, 1095.

81.

Takeuchi, Y.; Fujiki, K.; Tsubokawa, N. Polym. Bull. 1998, 41, 85.

82.

Hayashi, S.; Fujiki, K.; Tsubokawa, N. React. Funct. Polym. 2000, 46, 193.

83.

Advincula, R. C., Ed. Polymer Brushes: Synthesis,


Applications; John Wiley & Sons: New York, 2004.

84.

Oberdisse, J.; Dem, B. Macromolecules 2002, 35, 4397.


114

Characterization,

85.

Tamayo, F. G.; Titirici, M. M.; Martin-Esteban, A.; Sellergren, B. Anal. Chim.


Acta 2005, 542, 38.

86.

Meyer, T.; Prause, S.; Spange, S.; Friedrich, M. J. Colloid Interface Sci. 2001,
236, 335.

87.

Kroschwitz, J. I. Encyclopedia of Polymer Science and Engineering, 2nd ed.;


Wiley; New York, 1990.

88.

Demchenko, O.; Zheltonozhskaya, T.; Filipchenko, S.; Syromyatnikov, V.


Macromol. Symp. 2005, 222, 103.

89.

Thomas, D. B.; Convertine, A. J.; Myrick, L. J.; Scales, C. W.; Smith, A. E.;
Lowe, A. B.; Vasilieva, Y. A.; Ayres, N.; McCormick, C. L. Macromolecules
2004, 37, 8941.

90.

Appukkuttan, P.; Dehaen, W.; Fokin, V. V.; van der Eycken, E. Org. Lett. 2004,
6, 4223.

91.

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

92.

Gao, H.; Louche, G.; Sumerlin, B. S.; Jahed, N.; Golas, P.; Matyaszewski, K.
Macromolecules 2005, 38, 8979.

93.

Gao, H.; Matyjaszewski, K. Macromolecules 2005, 39, 4960.

94.

Ladmiral, V.; Mantovani, G.; Clarkson, G. J.; Cauet, S.; Irwin, J. L.; Haddleton,
D. M. J. Am. Chem. Soc. 2006, 128, 4823.

95.

Johnson, J. A.; Lewis, D. R.; Diaz, D. D.; Finn, M. G.; Koberstein, J. T.; Turro,
N. J. J. Am. Chem. Soc. 2006, 128, 6564.

96.

Keller, R. N.; Wycoff, H. D. Inorg. Synth. 1946, 2, 1.

97.

Lai, J. T.; Filla, D.; Shea, R. Macromolecules 2002, 35, 6754.

98.

Luedtke, A. E.; Timberlake, J. W. J. Org. Chem. 1985, 50, 268.

99.

Nozawa, K.; Gailhanou, H.; Raison, L.; Panizza, P.; Ushiki, H.; Sellier, E.;
Delville, J.P.; Delville, M.H. Langmuir 2005, 21, 1516.

100.

Tsuda, K.; Ishizone, T.; Hirao, A.; Nakahama, S.; Kakuchi, T.; Yokota, K.
Macromolecules 1993, 26, 6985.

101.

Thomas, D. B.; Sumerlin, B. S.; Lowe, A. B.; MaCormick, C. L. Macromolecules


2003, 36, 1436.
115

102.

F. Arnaud, L. Catherine; C. Marie-Thrse, P. Christian Macromolecules 2004,


37, 2026.

103.

Lummerstorfer, T.; Hoffmann, H. J. Phys. Chem. B, 2004, 108 (13), 3963.

104.

Huisgen, R. In 1,3-Dipolar Cycloaddition Chemisrty; Padwa, A.; Ed.; Wiley:


New York, 1984.

105.

Voronov, A.; Shafranska, O. Langmuir 2002, 18, 4471.

106.

Fetters, J. L.; Lohse, D. J.; Milner, S. T. Macromolecules 1999, 32, 6847.

107.

Fetters, L. J.; Lohse, D. J.; Richter, D.; Witten, T. A.; Zirkel, A. Macromolecules
1994, 27, 4639.

108.

Perrier, S.; Takolpuckdee, P.; Mars, C. A. Macromolecules 2005, 38, 6770.

109.

Erdem, G.; Oner, C.; nal, A. M.; Kisakrek, D.; gs, A. Y. J. Biosci. 1994, 19,
9.

110.

(a) Mantovani, G.; Ladmiral, V.; Tao, L.; Haddelton, D. M. Chem. Commun.
2005, 16, 2089. (b) Dirks, A. J.; van Berkel, S. S.; Hatzakis, N. S.; Opsteen, J. A.;
van Delft, F. L.; Cornelissen, J. J. L. M.; Rowan, A. E.; van Hest, J. C. M.; Rutjes,
F. P. J. T.; Nolte, R. J. M. Chem. Commun. 2005, 33, 4172.

111.

Edmondson, S.; Osborne, V.L.; Huck, W.T.S. Chem Soc. Rev. 2004, 33, 14.

112.

Rodinov, V. O.; Fokin, V. V.; Finn, M. G. Angew. Chem. Int. Ed. Engl. 2005, 44,
2210.

116

APPENDIX

A1. Model calculation of grafting density


(Table 4.1. Silica-bromide)
Surface grafting density of bromopropyltrichlorosilane = n groups/nm2.
Specific surface area of silica nanoparticles = 32 g/m2 (provided by Nissan Chemical
America Corporation)
1 g of silica nanoparticles contain 32 x n x 1018 / 6.023x 1023 moles of bromopropyltrichloro-silane
= n x 5.313 x 10-5 moles
= n x 5.313 x 10-5 x 80 g of bromine element
n x 5.313 x 10-5 x 80
Amount of bromine present on silica particles = -------------------------------------- x 100 %
1+ n x 5.313 x 10-5 x 122
Molecular weight of bromopropyl group = 122
From elemental analysis, amount of bromine present in modified silica particles = 1.53 %
Hence,

1.53 =

n x 5.313 x 10-5 x 80
------------------------------------------ x 100 %
1+ n x 5.313 x 10-5 x 122

Surface grafting density of bromopropyl group, n = 3.68 groups/nm2.

117

A2. Model calculation of grafting density


(Table 4.1. Silica-azide)
Surface grafting density of azide = n groups/nm2.
Specific surface area of silica nanoparticles = 32 g/m2 (provided by Nissan Chemical
America Corporation)
1 g of silica nanoparticles contain 32 x n x 1018 / 6.023x 1023 moles of azide group
= n x 5.313 x 10-5 moles
= n x 5.313 x 10-5 x 42 g of nitrogen element
n x 5.313 x 10-5 x 42
Amount of bromine present on silica particles = -------------------------------------- x 100 %
1+ n x 5.313 x 10-5 x 122
Molecular weight of azidopropyl group = 84
From elemental analysis, amount of nitrogen present in modified silica particles = 0.76 %
Hence,

0.76 =

n x 5.313 x 10-5 x 42
------------------------------------------ x 100 %
1+ n x 5.313 x 10-5 x 84

Surface grafting density of bromopropylazide groups, n = 3.45 groups/nm2.

A3. Molecular weight calculation by 1H NMR spectroscopy


Total integration values for peaks at 1.8 and 2.4 ppm (3 H of repeating units) = 32.22 +
56.96 = 89.18
Integration value for peak at 0.87 ppm (3H, CH3-CH2) = 1.45
Number of repeating unit in polyacrylamide = 89.18/1.45 61
118

Molecular weight of RAFT CTA = 402


Molecular weight of Polyacrylamide = 61 x 71 + 400 = 4731

A4. Model calculation of monomer conversion


(3.3.4.3. RAFT polymerization of styrene)
Target molecular weight = Amount of monomer (g) / moles of RAFT CTA (initiator for
ATRP)
= 21.84 g / 0.001
= 21840
Percent conversion = (actual molecular weight x 100) / target molecular weight
=

(9000 x 100) / 21840

41 %

A5. Model calculation for surface grafting density and distance between grafting sites of
attached polymers
(Table 4.3. Grafting to approach)
Surface grafting density of polystyrene (PS) = n groups/nm2.
Specific surface area of silica nanoparticles = 32 g/m2 (provided by Nissan Chemical
America Corporation)
1 g of silica nanoparticles contain 32 x n x 1018 / 6.023x 1023 moles of PS
= n x 5.313 x 10-5 moles of PS
= n x 5.313 x 10-5 x 9000 g of PS
119

Weight fraction of carbon in styrene repeating unit = 96/104


n x 5.313 x 10-5 x 9000 x (96/104)
Amount of carbon present on silica particles = -------------------------------------- x 100 %
1+ n x 5.313 x 10-5 x 9000
From elemental analysis, amount of carbon present in PS modified silica particles =
11.74 %
From elemental analysis amount of carbon present in azide modified silica nanoparticles
= 0.98 %

Hence,

(11.74 0.98) % =

n x 5.313 x 10-5 x 9000 x (96/104)


------------------------------------------ x 100 %
1+ n x 5.313 x 10-5 x 9000

Surface grafting density of polystyrene, n = 0.28 groups/nm2.


Distance between grafting sites (nm) = (4 / n )1/2 , From Reference 105
= 2.1 nm

A6. Model calculation of Rg (radius of gyration) for polystyrene


For polyacrylamide, From Reference 106-107
(Rg2 / Mn) = 0.152 2 mol/g
Hence Rg at Mn = 4500 = 2.6 nm
For polystyrene, From Reference 106-107
(Rg2 / Mn) = 0.434 2 mol/g
Hence Rg for polystyrene (Mn = 9000) = 6 nm

120

Figure A1. GPC analysis of PS prepared by RAFT polymerization.

121

Figure A2. 1H NMR spectra of polyacrylamide.

122

Figure A3. GPC analysis of PS prepared by RAFT polymerization.

123

Figure A4. GPC analysis of PS-b-PMA prepared by RAFT polymerization.

124

Figure A5. GPC chromatogram of PS-b-PMA prepared by RAFT polymerization.

125

Figure A6. GPC analysis of PS prepared by ATRP.

126

Figure A7. GPC analysis of PMA prepared by ATRP.

127

Figure A8. GPC analysis of PS prepared by grafting from approach.

128

Figure A9. GPC analysis of cleaved PS prepared by grafting from approach.

129

Figure A10. GPC analysis of second block PMA prepared by surface-mediated RAFT
polymerization.

130

Figure A11. GPC analysis of free PS obtained in tandem approach.

131

Figure A12. GPC analysis of cleaved PS prepared by tandem approach.

132

Figure A13. GPC analysis of free PS (first block) prepared by ATRP.

133

Figure A14. GPC analysis of free PMA (second block) prepared by ATRP.
134

Вам также может понравиться