through surface-attached monomers

248
Grafting of Polymers onto SiO 2 Surfaces through Surface-attached Monomers genehmigte Dissertation zur Erlangung des Doktorgrades Doktor der Naturwissenschaften der Fakultät für Angewandte Wissenschaften der Albert-Ludwigs-Universität Freiburg im Breisgau von Diplom-Chemiker Daniel Mädge geboren am 01.01.1974 in Oldenburg (Oldb.) Freiburg im Breisgau 2007

Upload: others

Post on 18-Dec-2021

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: through Surface-attached Monomers

GGrraaffttiinngg ooff PPoollyymmeerrss oonnttoo SSiiOO22 SSuurrffaacceess

tthhrroouugghh SSuurrffaaccee--aattttaacchheedd MMoonnoommeerrss

genehmigte

Dissertation

zur Erlangung des Doktorgrades

Doktor der Naturwissenschaften

der Fakultät für Angewandte Wissenschaften

der Albert-Ludwigs-Universität Freiburg im Breisgau

von

Diplom-Chemiker

Daniel Mädge

geboren am 01.01.1974 in Oldenburg (Oldb.)

Freiburg im Breisgau 2007

Page 2: through Surface-attached Monomers

The present work was carried out from August 2000 to March 2005 at the Institute for

Microsystem Technology (IMTEK), Albert-Ludwigs-University of Freiburg, Germany,

at the department “Chemistry and Physics of Interfaces”, under the supervision of

Prof. Dr. Jürgen Rühe.

Page 3: through Surface-attached Monomers

Dekan: Prof. Dr. Bernhard Nebel (FAW, Freiburg)

Vorsitzender: Prof. Dr. Eicke R. Weber (Fraunhofer ISE, Freiburg)

Erstgutachter: Prof. Dr. Jürgen Rühe (IMTEK, Freiburg)

Zweitgutachter: PD Dr. Thomas Hanemann (IMTEK, Freiburg & FZK, Karlsruhe)

Beisitzende: Prof. Dr. Margit Zacharias (IMTEK, Freiburg)

Datum der mündlichen Prüfung: 02. 08. 2007

Page 4: through Surface-attached Monomers

The beginning of knowledge is the discovery of something we do not understand.

Frank Herbert (Dune series)

Page 5: through Surface-attached Monomers

Table of Contents i

Table of Contents

1 Introduction....................................................................................................................................1

1.1 Modified Silicon Dioxide Substrates ..................................................................................1

1.1.1 Chromatographic Material...........................................................................................1

1.1.2 Immobilization of Biomolecules..................................................................................2

1.1.3 Catalyst Immobilization ...............................................................................................4

1.2 Fabrication of Silica Particles & Silicon ..............................................................................5

1.3 Surface Decoration of Silica, Glass Beads & Silicon .........................................................7

1.3.1 Functional Silanes & Silane Anchors ..........................................................................7

1.3.2 Polymer Layers by “Grafting-to” & “Grafting-from”..............................................9

1.3.3 Polymerizing in Presence of Surface-attached Monomers ....................................10

1.3.4 Polymer Micropatterning...........................................................................................13

1.4 Radical polymerization.......................................................................................................14

2 Goals & Strategy of the Work ....................................................................................................21

2.1 Goals......................................................................................................................................21

2.2 Strategy .................................................................................................................................21

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers ..............26

3.1 Silane Monomers .................................................................................................................26

3.2 Homopolymers on Silica ....................................................................................................31

3.3 Mechanism of the Grafting of Polystyrene onto Silica Surfaces through Surface-attached Monomers ..............................................................................................35

3.3.1 Structure Variation of the Immobilized Monomer.................................................35

3.3.2 Variation of the Polymerization Time ......................................................................37

3.3.3 Influence of the Silica Gel Concentration.................................................................40

3.3.4 Influence of Monomer Concentration in Solution ..................................................42

Page 6: through Surface-attached Monomers

ii Table of Contents

3.3.5 Influence of Initiator Concentration......................................................................... 44

3.3.6 Variation of the Polymerization Temperature........................................................ 46

3.3.7 Surface Concentration of Polymerizable Silanes & its Influence on Building Polymer Monolayers .................................................................................................. 50

3.3.8 Polymerization in Two Steps .................................................................................... 56

3.4 Grafting onto Various Porous Substrates........................................................................ 57

3.4.1 Influence of Substrates and Surface-attached Monomer Silanes ......................... 57

3.4.2 Influence of the Polymerization Time for Aerosil300............................................ 59

3.4.3 Influence of the Polymerization Time for Differently Modified Glass Beads.... 62

3.4.4 Influence of Initiator Concentration for Different Substrates............................... 63

3.4.5 Influence of the Monomer Concentration for Glass Beads ................................... 66

3.4.6 Influence of the Glass Bead Concentration ............................................................. 68

3.5 Discussion............................................................................................................................ 70

3.6 Conclusion ........................................................................................................................... 76

3.7 Grafting of Poly(N,N-dimethyl acrylamide) onto Silica Surfaces through Surface-attached Monomers ............................................................................................. 79

3.7.1 Polymerization Time Influence................................................................................. 79

3.7.2 Variation of Initiator Concentration......................................................................... 82

3.7.3 Influence of Monomer Concentration...................................................................... 86

3.7.4 Discussion & Conclusion for P(DMAA) on Silica.................................................. 88

4 Grafting of Functional Copolymers onto Silica Surfaces ...................................................... 90

4.1 Polystyrene Copolymers on Silica Particles.................................................................... 92

4.1.1 Poly(styrene-co-N-acryloyl-N-methyl-propyl phthalimide) & Poly(styrene-co-N-acryloyl-N-methyl-propyl amine)........................................... 92

4.1.2 Poly(styrene-co-N-methacryloyl-β-alanine succinimide ester).......................... 100

4.1.3 Two-step Grafting Processes................................................................................... 103

4.2 Poly(N,N-dimethyl acrylamide) copolymers on silica particles................................ 107

Page 7: through Surface-attached Monomers

Table of Contents iii

4.2.1 Poly(N,N-dimethyl acrylamide-co-N-acryloyl-N-methyl-propyl phthalimide) & Poly(N,N-dimethyl acrylamide-co-N-acryloyl-N-methyl-propyl amine) ........107

4.2.2 Poly(N,N-dimethyl acrylamide-co-N-methacryloyl-β-alanine succinimide ester) ....................................................................................................................................108

4.3 Discussion & Conclusion Copolymers...........................................................................111

5 Applications for Functional Copolymers on Silica Substrates............................................113

5.1 Enzyme Immobilization ...................................................................................................113

5.1.1 Glucose Oxidase Assay.............................................................................................114

5.1.2 Stability against Leaching ........................................................................................116

5.1.3 Influence of the Immobilization Time on the Conversion Rate of the Enzyme Modified Glass Beads ...............................................................................................118

5.1.4 Influence of the Immobilization Concentration on the Conversion Rate of the Enzyme Modified Glass Beads ................................................................................119

5.2 Biomolecule Immobilization............................................................................................122

5.2.1 Biotin Labeling...........................................................................................................122

5.2.2 DNA Immobilization & Hybridization ..................................................................125

5.3 Catalyst Immobilization ...................................................................................................127

5.4 Discussion & Conclusions Applications ........................................................................130

6 Polymer Systems on Silicon Wafers........................................................................................133

6.1 Grafting of Poly(N,N-dimethyl acrylamide) Layers onto Silicon Wafers through Surface-attached Monomers ............................................................................................133

6.1.1 Influence of Monomer Concentration ....................................................................133

6.1.2 Polymerization Time Influence ...............................................................................136

6.1.3 Variation of Initiator Concentration .......................................................................138

6.1.4 Discussion & Conclusion P(DMAA) on Wafers....................................................140

6.2 Microstructured Surfaces .................................................................................................143

6.2.1 Photoablation of Surface-attached Monomers Followed by Polymerization ...143

6.2.2 Patterning through Polymer Network Formation................................................145

Page 8: through Surface-attached Monomers

iv Table of Contents

6.2.3 Discussion & Conclusion Micropatterning........................................................... 160

7 Outlook....................................................................................................................................... 162

8 Experimental I – Materials & Methods.................................................................................. 163

8.1 Materials ............................................................................................................................ 163

8.1.1 Material List............................................................................................................... 163

8.1.2 Material Preparation ................................................................................................ 164

8.2 Methods ............................................................................................................................. 166

8.2.1 Elemental Analysis ................................................................................................... 166

8.2.2 Infrared Spectroscopy .............................................................................................. 166

8.2.3 Gel Permeation Chromatography (GPC) .............................................................. 166

8.2.4 Ellipsometry & Imaging Ellipsometry................................................................... 166

8.2.5 X-ray Photoelectron Spectroscopy ......................................................................... 167

8.2.6 Nuclear Magnetic Resonance Spectroscopy ......................................................... 167

8.2.7 UV/vis Spectroscopy ................................................................................................ 167

8.2.8 Fluorescence Microscopy & Reading..................................................................... 167

8.2.9 Profilometry............................................................................................................... 168

8.2.10 Titration...................................................................................................................... 168

9 Experimental II – Polymerizations ......................................................................................... 169

9.1 Immobilization of Silanes ................................................................................................ 169

9.1.1 Immobilization of Trimethoxy Silanes onto Silica Gel ........................................ 169

9.1.2 Immobilization of Trimethoxy Silanes onto Glass Beads ................................... 171

9.1.3 Co-Immobilization of MPS and OS onto Silica Gel ............................................. 172

9.2 Homopolymerizations ..................................................................................................... 173

9.2.1 Synthesis of Poly(3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10-heptadecafluorodecyl acrylate)...................................................................................................................... 173

9.2.2 Synthesis of Polystyrene & Poly(methyl methacrylate) ...................................... 173

Page 9: through Surface-attached Monomers

Table of Contents v

9.2.3 Synthesis of Poly(N-isopropyl acrylamide)...........................................................174

9.2.4 Synthesis Poly(N,N-dimethyl acrylamide)............................................................175

9.2.5 Synthesis of Poly(methacrylic acid) ........................................................................175

9.3 Polystyrene – Polymerization Parameters .....................................................................176

9.3.1 Influence of Reaction Time.......................................................................................176

9.3.2 Variation of the Overall Concentration of the Solid.............................................177

9.3.3 Monomer Concentration Variation.........................................................................178

9.3.4 Initiator Concentration Variation............................................................................178

9.3.5 Variation of the Polymerization Temperature ......................................................179

9.3.6 Influence of MPS Surface Concentration on the Polystyrene Graft Density.....180

9.3.7 Variation of the Overall Concentration of Glass Beads .......................................181

9.3.8 Two-step Grafting .....................................................................................................181

9.3.9 Polymerization Reactions without Surface-attached Monomers (blind tests) .181

9.4 Poly(N,N-dimethyl acrylamide) – Polymerization parameters..................................182

9.4.1 Influence of Reaction Time.......................................................................................182

9.4.2 Monomer Concentration Variation.........................................................................183

9.4.3 Initiator Concentration Variation............................................................................183

9.5 Copolymerizations ............................................................................................................185

9.5.1 Synthesis of Poly(styrene-co-N-acryloyl-N-methyl-propyl phthalimide), P(S-co-AC3pht) ...........................................................................................................185

9.5.2 Synthesis of Poly(styrene-co-N-acryloyl-N-methyl-propyl amine), P(S-co-AC3amine) ......................................................................................................187

9.5.3 Synthesis of Poly(styrene-co-N-methacryloyl-β-alanine succinimide ester), P(S-co-MAC2ae) .........................................................................................................189

9.5.4 Synthesis of Poly(N,N-dimethyl acrylamide-co-N-acryloyl-N-methyl-propyl phthalimide), P(DMAA-co-AC3pht) .......................................................................191

9.5.5 Synthesis of Poly(N,N-dimethyl acrylamide-co-N-acryloyl-N-methyl-propyl amine), P(DMAA-co-AC3amine).............................................................................192

Page 10: through Surface-attached Monomers

vi Table of Contents

9.5.6 Synthesis of Poly(DMAA-co-N-methacryloyl-β-alanine succinimide ester), P(DMAA-co-MAC2ae).............................................................................................. 193

10 Experimental III – Reactions with Functional Groups ........................................................ 195

10.1 Two Step Grafting on Functionalities ............................................................................ 195

10.1.1 Immobilization of Acryloyl Chloride as Monomer ............................................. 195

10.1.2 Immobilization of 2-Brom-Propionic Acid Bromine as Starter for ATRP ........ 195

10.1.3 Polymerization of Styrene with Polymer-attached Monomers.......................... 196

10.1.4 Atom Transfer Radical Polymerization of Styrene with Polymer-attached Initiators ..................................................................................................................... 197

10.2 Coupling of Dyes to the Functionalities of the Copolymer ........................................ 198

10.2.1 Coupling of Fluorescein Isothiocyanate ................................................................ 198

10.2.2 Coupling of DY-635-NH2......................................................................................... 199

10.3 Enzyme Immobilization .................................................................................................. 199

10.3.1 Activity Assay ........................................................................................................... 199

10.3.2 Washing Cycles......................................................................................................... 200

10.3.3 Varied Immobilization Time................................................................................... 200

10.3.4 Varied Initial Immobilization Concentration ....................................................... 201

10.3.5 Immobilization of Dyed Enzymes.......................................................................... 201

10.3.6 Hydrolysis of Immobilized Enzymes .................................................................... 201

10.4 Oligo Nucleotide Immobilization .................................................................................. 202

10.4.1 Immobilization Reaction.......................................................................................... 203

10.4.2 Hybridization ............................................................................................................ 203

10.5 Biotin assay........................................................................................................................ 204

10.6 Catalyst Immobilization .................................................................................................. 204

11 Experimental IV – Silicon Wafers ........................................................................................... 205

11.1 Silane Immobilization ...................................................................................................... 205

11.2 P(DMAA) Layers .............................................................................................................. 205

Page 11: through Surface-attached Monomers

Table of Contents vii

11.2.1 Influence of the Polymerization Time ....................................................................206

11.2.2 Influence of the Monomer Concentration..............................................................206

11.2.3 Influence of the Initiator Concentration.................................................................206

11.3 Microstructured Layers ....................................................................................................206

11.3.1 UV Photoablation of Surface-attached Monomers ...............................................207

11.3.2 Grafting of Polystyrene.............................................................................................207

11.4 Combined Functional Layers & Networks ....................................................................208

11.4.1 Preparation of Copolymers for Network Formation............................................208

11.4.2 Coating & UV-cross-linking.....................................................................................210

11.4.3 Functional layers .......................................................................................................211

12 Summary.....................................................................................................................................212

13 Zusammenfassung ....................................................................................................................216

14 References ...................................................................................................................................220

Page 12: through Surface-attached Monomers

viii Abbreviations & Symbols

Abbreviations & Symbols monomer

N-acryloyl-N-methyl-propyl phthalimide AC3pht

N,N-dimethyl acrylamide DMAA

3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10-heptadecylfluordecyl acrylate

HDFDA

Methacrylic acid MAA

N-methacryloyl-β-alanine succinimide ester

MAC2ae

Methacryloyl-4-oxy-benzophenone MABP

Methyl methacrylate MMA

N-isopropyl acrylamide NIPAM

polymer

Poly(N,N-dimethyl acrylamide) P(DMAA)

Poly(3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10-heptadecylfluordecyl acrylate)

P(HDFDA)

Poly(Methacrylic acid) P(MAA)

Poly(Methyl methacrylate) P(MMA)

Poly(N-isopropyl acrylamide) P(NIPAAm)

Polystyrene PS

silane surface monomer

modification

3-methacryloylpropyl trimethoxysilane MPS -ME

N-methacryloyl-N-methyl-propyl trimethoxysilane

MNPS -MA

3-acryloylpropyl trimethoxysilane APS -AE

N-acryloyl-N-methyl-propyl trimethoxysilane

ANPS -AA

octyl trimethoxysilane OS -Oct

substrate

LiChrospher Si60 LC700

Aerosil 300 AR300

Glass beads 200 Å GB250

Glass beads 500 Å GB80

Glass beads 1000 Å GB40

Page 13: through Surface-attached Monomers

Abbreviations & Symbols ix

Ai, Ap, At Arrhenius constants for initiation, propagation and termination Asp, Amod specific surface area before/after monomer immobilization ABTS 2,2’-Azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) AIBN Azobis(isobutyronitrile)

ATRP Atom transfer radical polymerization

c concentration

CM, CS transfer constant, transfer to monomer and solvent respectively

δ, δmod amount of grafted silane or polymer δ chemical shift (nuclear magnetic resonance spectroscopy)

d distance or diameter

DCC N,Nʹ-dicyclohexyl-carbodiimide

DMF N,N-dimethyl formamide

DMSO Dimethylsulfoxide

DMT [1,1-bis(4-methoxyphenyl)-1-phenyl]methyl

DNA Deoxyribonucleic acid

DRIFT Diffuse reflection infrared Fourier transformation spectroscopy

DSC Differential scanning calorimetry

ελ extinction coefficient at wavelength λ

Eλ extinction at wavelength λ

Ei, Ep, Et activation energy for initiation, propagation and termination EA Elemental analysis

FITC Fluorescein-5-isothiocyanate

Freon® 113 1,1,2-trichloro-trifluoro ethane

FT Fourier transformation

Γ graft density of silane or polymer

GC Gas chromatography

GOD Glucose oxidase

GPC Gel permeation chromatography

HPLC High-performance liquid chromatography

I, I0 intensity of transmitted light through sample or blank cuvette

IR Infrared (spectroscopy)

[I] initiator concentration

ϕ percentage of functional units inside a copolymer

ki, kp, kt rate constants for initiation, propagation and termination

km, ks transfer rate constant to monomer and solvent respectively

KPS Potassium peroxodisulfate

Page 14: through Surface-attached Monomers

x Abbreviations & Symbols

LC liquid chromatography

m mass

MC atomic weight of carbon

MF atomic weight of a molecule fragment

Mj dead polymer chain with j repeat units •jM growing polymer chain with j repeat units

Mn number average of the molecular weight distribution

Mw weight average of the molecular weight distribution

[M] monomer concentration

n amount of substance

ν kinetic chain length

NMR nuclear magnetic resonance (spectroscopy)

PD polydispersity

PEBr 1-Phenylethylbromine

PMDETA N,N,N’,N’,N”-Pentamethyl-diethylenetriamine

ppm parts per million

R gas constant

RCM Ring closing metathesis •R radical initiator fragment

rpm rounds per minute

[S] solvent concentration

SEC Size exclusion chromatography

SEM Surface electron microscopy

t time

T temperature

TEA Triethylamine

TEM Transmission electron microscopy

UV Ultraviolet

UV/vis Ultraviolet/visible light spectroscopy

v enzyme conversion rate

V volume

THF Tetrahydrofurane

nX degree of polymerization

XPS x-ray photoelectron spectroscopy

Z number of atoms in a molecule fragment

Page 15: through Surface-attached Monomers

1 Introduction 1

1 Introduction

1.1 Modified Silicon Dioxide Substrates In the last decades there has been remarkable theoretical, experimental and practical

interest in the surface decoration of solid SiO2 substrates with molecules and polymers for a

broad range of applications such as analytics[1], chromatography for separation and

purification[2-4] or catalysis [5]. One of the most common substrates for immobilization of

complex molecules on a solid support are silica gels and glass beads. One general advantage

why these systems are widely used is the easy handling of such materials. Pouring them into

a chromatography column or reaction vessel and separate them after use from a liquid or

gaseous phase is quite simple. A description of applications and a discussion of advantages

of these materials compared to other materials such as variable size of the particles and

surface topology will be addressed in the following subchapters. There are many reports on

silica solids which have been physically or chemically surface modified to enhance their

properties towards the selected application and its needs.

1.1.1 Chromatographic Material

Separation or purification via chromatography is based on differences in partitioning

behavior of substances between a mobile and a stationary phase. The stationary phase is

always a solid and SiO2 substrates play an important role in such applications. The analyte

(together with solvent) travels as mobile phase through the stationary phase and interaction

with both phases results in specific retention of the analyte in the system. The interaction of

the analyte with the stationary phase can be based on relative solubility, adsorption or

interaction with charges and determines the retention of the analyte. Therefore it is

important to properly control the surface properties of materials used in chromatography.

Chromatography, especially the very common liquid chromatography (LC) has been of great

interest for industrial applications. In fact, many patents dealing with liquid chromatography

have been filed using silanes[6-9] or polymers[10-13] as surface modification. For example

for reversed phase liquid chromatography oxidic support material has to be “switched”

regarding its surface polarity. Yang et al. have bound octadecyl silane to silica (ODS-SiO2)

and polybutadiene on zirconia (PBD-ZrO2)[14]. Amino-silane modified silica is a very

versatile material because the reactivity of the amino groups allows many surface reactions.

Page 16: through Surface-attached Monomers

2 1 Introduction

Based on such material Ascah et al. generated amide-functionalized reversed phase column

material[15], Jackson et al. used such modified silica in continuous flow reactors[16] and

Bruckner et al. separated amino acid enantiomers with column material prepared from

amino-modified silica[17]. Urea-functionalized materials were prepared by Silva et al. for

LC[18] and by Chen et al. for chiral LC[19]. Ruffing et al. basically changed the surface

properties of silica particles and capillaries by immobilizing polysiloxane using a platinum

hydrosilylation catalyst. Parallel, hydrosilylation was used for mounting chiral species to the

stationary phase to finally perform LC in organic liquids or supercritical fluids[20]. Based on

silica immobilized carbohydrate biopolymers Felix et al., Okamoto et al. and Wang et al.

performed chiral high-performance liquid chromatography (HPLC)[21-25]. The separation of

racemic mixtures into pure enantiomers is of great interest for pharmaceutical industries.

1.1.2 Immobilization of Biomolecules

Immobilization of biomolecules onto the surfaces of solid substrates is an area of great

interest for science and technology because the obtained biomaterials have numerous

applications in the fields of biotechnology, immunosensing and biomedical applications. At

the example of protein immobilization, especially enzyme immobilization, and

DNA/oligonucleotide immobilization the large application field of biomolecule

immobilizations is introduced.

A number of different pathways have been described which allow the immobilization of

proteins. They can be adsorbed tightly to the surface of insoluble materials[26, 27], covalently

bound onto carriers[28-30], entrapped within gels[31] or microcapsules[32, 33], or

crosslinked with bi- or multifunctional reagents[34]. Enzymes are proteins, which contain at

least one active center which catalyzes biotransformations. Immobilizing enzymes gives the

advantages of facile separation from the reaction mixture, minimized enzyme loss,

reusability and rapid termination of reactions. But it is crucial that enzymes do not lose their

activity after the immobilization process due to deformation of their structure or blocking the

active center. Proper immobilization techniques, adjusted to the particular reaction

conditions, have to be chosen for enzyme applications to maintain enzyme activities and get

the advantages of enzyme immobilization. That is the reason why enzyme engineering is a

vital and fast-growing application in pharmaceutical and food additive markets.

Among the many enzymes that have been immobilized in the past (Figure 1-1) are

dehydrogenase[35], isomerase[28], lipase[27], amylase[36], urease[29], oxidoreductase[37]

Page 17: through Surface-attached Monomers

1 Introduction 3

and oxidase[30] just to name few that have been immobilized to silica supports. Recently

Kim et al. immobilized IgG-peroxidase species on polymer layers covalently bound to glass

substrates using a “grafting-to” method via polymers containing surface reactive amino

groups and a “grafting-from” method[38]. Nakamura et al. immobilized IgG-monoclonal

antibodies or glucose oxidase to silica via silane linker molecules[39]. Glucose oxidase is a

well known and intensively investigated system and therefore a good model compound for

immobilization as experiments in the past have shown[30, 40-42]. Biofunctionalization for a

wide variety of applications can be achieved by coating silica surfaces with biomolecules

such as lipids, proteins or enzymes. However, to decorate surfaces with such biomolecules a

surface optimization of the inorganic SiO2 with a silane or polymer coating is favorable in

many respects, e.g. to create a layer being capable to covalently bind the desired

molecules[43], to enhance adsorption specificity of the surface[44], or as support layers for

lipid membranes[45]. The modified materials may then be used for high-throughput

screening[46].

++ --

a) b) c) d)

e) f)

g)

Figure 1-1. Immobilization techniques: carrier binding: adsorption (a) [26, 27], binding via

electrostatic forces (b)[47], covalent binding (c) [28-30], biospecific binding (d); entrapment: lattices

(e)[31], microcapsules (f) [32, 33]; direct cross-linking (g) [34].

In biotechnology and clinical medicine modified beads and also planar systems are of

great interest for the generation of analytic chips and biosensors[48, 49]. Here especially

biotin- and DNA- or oligonucleotide immobilization[50-53] come into play. Composites of

Page 18: through Surface-attached Monomers

4 1 Introduction

glass beads and polymers with immobilized biomolecules were developed to improve the

function of these supports[54-56]. In addition it was found that a polymer cushion of

functional hydrogels gives shear protection to biomolecules that were covalently bound by

activated ester groups within the hydrogel[31].

Various methods using polymer layers on planar substrates like plasma borne

polymers[57], dendrimers[58] or latex based polymer film[59] have been developed for

biomolecule immobilization. Patterned surfaces on planar systems with immobilized

biomolecules can be used in an even better way as analytical chips, e.g. DNA microarrays as

Freidank et al. and Oh et al. have shown[60, 61]. Protein chips for proteomics are a quite new

development. Snyder et al. examined the use of protein chips that have been generated by

just using densely packed wells[62]. Melnyk et al. immobilized peptides in a structured

fashion by using linker molecules[63]. Of course also combinations of the previously

mentioned polymer layer systems with common patterning methods are possible. Another

combination is the use of patterned planar systems with functional particles to immobilize

biomolecules[64, 65] as realized by the Frauenhofer Gesellschaft and Dickinson et al. The first

approach uses polymeric layers with reactive groups to bind the particles onto the surface,

the second is simpler by using pre-fabricated wells or well plates.

1.1.3 Catalyst Immobilization

As various industries use immobilized enzymes as biocatalysts there is also a large market

for immobilizing classic catalysts for chemical synthesis. Already 1975 rhodium catalyst

species for hydroformylation reactions were immobilized by a quite complicated method to

phosphine ligands contained within a polystyrene coating on silica gel by Arai et al. In

heterogeneous catalysis these substrates showed high efficiency[66]. Several approaches

within the last decades have been made on the immobilization of rhodium compounds and

use them to catalyze olefin hydroformylation[66-68] which is a very common industrial

reaction. Various other reactions have been performed by the use of functional silica

immobilized catalyst species, e.g. oxidation of alkanes and cyclohexanes with iron and

copper by Kurusu et al.[69, 70], or Heck re[71]action with palladium by Pan et al.[72]

Immobilization of catalytic compounds were also applied to the pyrogenic silica Aerosil.

Hultman et al. fixed the catalyst directly to the surface[73, 74] whereas Sarkany et al. used a

combination of polymer and [71]palladium[71].

Page 19: through Surface-attached Monomers

1 Introduction 5

In the last years metathesis reactions grew considerably and techniques to immobilize

adequate catalyst species to solid supports evolved. Kingsbury et al. immobilized and used

olefin metathesis catalyst on particles[75], Buchmeiser et al. did the same on particles

equipped with a surface polymer layer[76]. Ring opening metathesis polymerization (ROMP)

with immobilized catalyst was performed by Mingotaud et al. to grow polymer layers on

silica nanoparticles constructing core-shell architecture[77]. Besides ROMP Krause et al. also

performed ring closing metathesis (RCM) reactions with immobilized Grubbs-Hermann type

catalysts[78].

1.2 Fabrication of Silica Particles & Silicon Silica is a material that offers a great variety of particle sizes, porosities and surface

topologies. Some variants exhibit high surface areas within a small amount of solid mass like

infusorial earth which is a naturally structured form of silica built by sedimentation of

diatoms and infusorians[79]. The detailed structure of synthetic silica depends on the

production process. The way it is generated determines particle size, particle agglomeration

and ageing respectively sintering. An important method of silica production is the sol-gel

process[80-82] where soluble silicates are converted into silicic acid that then condenses

building a large network of interconnected microscopic pores (Scheme 1-1a). Using alkoxy

silanes instead of sodium silicate, e.g. tetramethoxysilane or tetraethoxysilane, gives highly

pure and metal free silica (Scheme 1-1b).

a) Na2SiO3 + 2 H+ → SiO2 + H2O + 2 Na+

b) Si(OR)4 + 2 H2O → SiO2 + 4 ROH

Scheme 1-1. Reaction of sodium silicate (a) or alkoxy silanes (b) to silica by the sol-gel process[83, 84].

First a stable colloidal dispersion, the sol, is generated. During the condensation reaction

chain growth and network formation or particle growth can be influenced by parameters like

pH-value or salt concentration[85]. The viscosity rises until the gel point is reached. The gel

point is defined as the degree of conversion where a material is formed which exhibits an

elastic response when mechanical stress is delivered to the system[86]. The obtained

hydrogels are aged by moderate drying. Upon this evaporation process the pores collapse

resulting in socalled xerogels (Greek: ξερός, xerós = dried). In order to retain the original

pore structure and take advantage of the high specific surface of a silica network,

Page 20: through Surface-attached Monomers

6 1 Introduction

supercritical drying is applied to the hydrogels[87] and the received material is called

aerogel (Greek: αέρας, aeros = air). These aerogels were developed as storage material for

other substances. Another advancement in the generation of silica materials is the Stöber

process[88] where large, well defined, monodisperse silica particles are obtained through

hydrolysis of alkoxysilanes.

Another major method for silica generation is hydrolysis of halide silanes within an

oxyhydrogen flame (Scheme 1-2). These pyrogenic silicic acids form very pure and compact

silica particles without pores. The primary particles may be sintered to larger agglomerates

creating pores between the primary particles. These aggregates have the advantage of a very

open structure with mainly macropores.

Scheme 1-2. Synthesis of pyrogenic silica: different stages of particle formation in an oxyhydrogen

flame (T > 1200 °C)[89].

An advantage of the sol-gel process is that already within the process hybrid materials are

accessible. “Hybrid materials” means a material consisting of clear regions where either

inorganic or organic structures dominate. Using silanes with organic substituents, instead of

pure silicates with four oxygen substituents, give particles with the chosen organic moieties.

Such particles with a low compressive modulus known as Tospearl have been used in

cosmetic formulations[90]. Other particles with a high compressive modulus and mesopores

have been developed by Cheng and Fisk for utilization in reverse-phase high-performance

liquid chromatography[91]. Another commercial species are “MQ resins”, particles whose

cores can be prepared by using tetraethoxy silane[92] or sodium silicate[93] from Q units

analog to the sol-gel process. The shell is then generated by using (functional) M units

(Scheme 1-3). The largest application for MQ resins is in pressure-sensitive adhesives

(PSA)[94, 95].

Page 21: through Surface-attached Monomers

1 Introduction 7

Scheme 1-3. Nomenclature of siloxane units bearing different numbers of organic substituents (= R).

The combination of single letters in the name of a resin indicates the which kind units are used to

prepare the resin.

Another important substrate we use are silicon wafers. These planar substrates have a

silicon dioxide layer in the nanometer range and therefore adequate silanes may be attached

to the surface similarly as for the silica and glass substrates. Silicon wafers are mainly used in

applications of the semiconductor industry. But in the recent years miniaturization has

become an important aspect in many fields of technology. Especially the construction of

micro devices is interdisciplinary and combines amongst others semiconductor technology,

microelectronics, microoptics as well as surface chemistry[96, 97]. For example Jeyaprakash

et al. managed to reduce the surface conductivity of a gas sensor device by tailoring the

surface with a fluoropolymer film[98].

1.3 Surface Decoration of Silica, Glass Beads & Silicon

1.3.1 Functional Silanes & Silane Anchors

Silica, glass and silicon have in common that the surface of these substrates consists of a

silicon dioxide layer where reactive silanes can be attached to. As reactive silanes mostly

chloro- or alkoxy silanes are used. Already in the 70ies silanes were used to influence the

surface of solid substrates to get more specified interactions for liquid chromatography

(LC)[99-102]. Because of the great interest in such materials many basic investigations have

been carried out. Velasquez et al. gathered information about the surface free energy of

silane covered materials[103] whereas Tripp et al. had a closer view on the surface

attachment of chloro silanes[104, 105] and Kurth et al. investigated surface reactions of

surface-attached silanes[106]. Because of the good reactivity of amino groups special interest

has been set on amino-silanes and different measurements have been carried out concerning

Page 22: through Surface-attached Monomers

8 1 Introduction

their surface adsorption and immobilization like calorimetry[107, 108], ellipsometry[109] or

x-ray photoelectron spectroscopy and magic angle spinning-NMR[110]. But also silanes with

other reactive groups have been immobilized, e.g. epoxy groups[111]. Another large field of

interest is the hydrophobization of silica substrates with long alkyl chain silanes like

octadecyl silane (ODS)[112] for chromatography purposes[14, 101]. This short overview

shows that the surface-attachment of various silanes with divers functionalities on solid

substrates has been intensively studied in the past.

The grafted amounts δ (in mol·g-1) of monomer in the composite material can be

determined from elemental analysis according to equations derived from calculations of

Berendsen and de Galan[113]:

ZM

g

C

Cmod ⋅

=δ equation 1-1

δmod is here the grafted amount of silane per gram modified silica (weight of silica

including the silane) and gC is the carbon content as determined by elemental analysis. Z

represents the number of carbon atoms in the immobilized unit and MC is the atomic mass of

carbon (12.01 g·mol-1). MF is the molecular mass of the part of the silane which is surface

attached (note: the methoxy groups are lost to surface anchoring or hydrolysis and are,

therefore, disregarded in this calculation). For a more detailed comparison in some cases it is

desirable to describe the grafted amount in relation to the pure silica (equation 1-2). However

to do so it has to be taken into account that the specific surface is altered by the surface

attachment (equation 1-3).

( ) 1F

Fmod

Fmod MM1

M −⋅

⋅−

⋅=

δδδ equation 1-2

( )( ) modunFmodmod AM1A ⋅⋅−= δ equation 1-3

δ is the grafted amount of silane per gram pure silica (weight of silica without the silane).

Aunmod and Amod are the specific surfaces (in m²·g-1) of the solid before and after silane

immobilization.

Page 23: through Surface-attached Monomers

1 Introduction 9

1.3.2 Polymer Layers by “Grafting-to” & “Grafting-from”

The decoration of surfaces with thin layers of polymers is a wide spread approach for the

tailoring of surface properties of materials[114-117]. A huge number of different techniques

have been developed ranging from simple painting or solution casting processes to

sophisticated chemical approaches which allow for the generation of films of covalently

grafted polymers. The latter typically exhibit a outstanding stability against harsh

environments including treatments with solvents that would usually dissolve the coating.

Typical chemical strategies range from systems in which polymers with reactive groups

(either at the chain end or along the backbone) are attached to suitable groups at the

surfaces[118-121] (“grafting-to”; cf. Figure 1-2a) to more recent approaches that utilize

monolayers of initiator molecules to “grow” the chains at the surfaces[122-127] where also

living polymerization methods are available[128-131] (“surface-initiated polymerization”,

“grafting-from”, cf. Figure 1-2b). These two techniques have been thoroughly studied over

the past twenty years and books and reviews are available that describe the particularities of

each approach[132, 133]. As we will see in 1.3.3 some theoretical considerations and

calculations play an important role for understanding the behavior of the third grafting

technique using surface attached monomers.

Surface modifications on the substrates used in this thesis have been well studied. In the

70ies Negievich et al. already performed different graft polymerization techniques on Aerosil

material[134, 135]. A few years later Ivanchev et al. used surface-attached

methacryloyloxymethyl methyldiethoxysilane units for a graft polymerization of methyl

methacrylate on Aerosil[136] and in a “grafting-from” like approach polystyrene was grown

by initiation through the redox decay of immobilized hydroperoxides[137]. Grafted

polymethacrylate layers on Aerosil were investigated by Yushkova et al.[138]

Different types of LiChrospher, sol-gel borne particles, have been used for surface sensitive

applications in the past. Silica particles with a polyacrylate layer (LiChrospher Diol) showed

good separation properties in size exclusion chromatography (SEC)[139, 140] and in high-

performance liquid chromatography (HPLC) of raw sugars and polyols[141] as well as the

commercially available reversed phase material LiChrospher RP[142-144]. Guillaume et al.

coated LiChrospher successfully with a cross linked polymer film[145] and Kurganov et al.

built a thin polyvinylpyrrolidone film on a LiChrospher species[146] with pores of 30 to 50 nm

comparable to the glass beads used here. As we see the history of these solid supports and

the many surface modifications that have been delivered to them to get the desired

Page 24: through Surface-attached Monomers

10 1 Introduction

properties, our choice of silica solids seems to be reasonable to explore the behavior of the

approach using surface-attached monomers.

1.3.3 Polymerizing in Presence of Surface-attached Monomers

Surprisingly, the third approach for the generation of monolayers at solid surfaces has not

received equal attention in the scientific community so far, even though it is frequently used

in industry as a so-called primer system for the modification of surfaces of inorganic

components of acrylate based formulations. As already mentioned above in this approach

(cf. Figure 1-2c) functionalized monomers are attached to the surfaces of the solid substrates

and the resulting surface-attached groups of the materials are then used as a “comonomer”

during a free radical polymerization. During this process all chains are initiated in solution

but some of the growing chains are captured by surface-attached monomers. From there they

can continue to grow away from the surface until a chain transfer or a termination reaction

ends growth of the individual chain.

Consequently the chains on the surface are not tethered to the substrate by one end but

with one or several anchor points somewhere along the chain. Among the first to study such

approaches were Hamann and Laible who used silane based systems on silica particles[147]

and Trachenko et al. who directly coupled acrylic acids to various oxide surfaces[148]. In

case of the silane based system the monomer is bound on the silica surface via siloxane

bridging links. Both systems were used for the copolymerization of the surface-attached

monomers with monomers in solution. Cohen et al have published the results of

experimental and theoretical work that to some extent describe the overall kinetics of the

process[149], but up to now there is no comprehensive study that elucidates the special

kinetic and mechanistic features of this approach. Delivering this information is the main

task of this thesis and the important targets as well as the strategy to reach them is described

in chapter 2. Furthermore it is important to establish our grafting technique for a wide

variety of silica because with the choice monomer, functionality and substrate it will be

applicable for many different purposes in a simple and successful way.

Page 25: through Surface-attached Monomers

1 Introduction 11

Figure 1-2. Schematic depiction of three different techniques for the generation of surface-attached

polymer monolayers; attachment of functionalized polymers, “grafting-to” (a), surface-initiated

polymerization from immobilized initiators, “grafting-from” (b) and incorporation of surface-attached

monomer units into a growing chain during a solution-borne polymerization (c).

Before moving the focus towards applications we tackle the behavior of this grafting

system by comparing it (cf. chapter 3.3) with theoretical description of “grafting-to”

approaches of Chakrabarti et al.[150] and Binder et al.[151] which use Monte Carlo

calculations to explain kinetic details of the irreversible adsorption of polymer chains which

are uniform in length onto a surface with reactive sites. End-functionalized polymer coils

approach the surface because a growing polymer chain only attaches to the surface monomer

by its radical moiety at the growing end of the chain. Furthermore the polymer adsorption

process is irreversible because the growing polymer chain forms a chemical bond with the

surface attached monomer. One may draw parallels to a system of homopolymer adsorption

that is also discussed by Zajac and Chakrabarti[152]. There a polymer chain has several

Page 26: through Surface-attached Monomers

12 1 Introduction

anchor points on a surface. This system may be realized such that the radical center moves

from surface monomer to surface monomer either building trains of directly connected

surface monomers or building loops with interconnected monomer units from solution. But

multiple attachment is not very favorable because even close to the surface the number of

free monomers that surround the active center of growing chains is much higher than the

number of surface attached monomers in close proximity[153]. Another important factor that

is taken into account by Binder et al. and used in the calculations of Chakrabarti et al. is

described by Ligoure and Leibler[154]. The early formation of a polymer layer attached to

the surface exhibits steric repulsion of polymer chains that approach the coated surface.

There are two successive regimes: First diffusion in solution governs the adsorption kinetics.

Later on adsorbed chains begin to overlap and penetration of this barrier controls the

kinetics. When assuming that an active center of a growing chain rapidly connects to a

surface attached monomer the grafting kinetics are controlled by diffusion in solution and

penetration of chains through the polymer layer protecting the surface. As this layer gets

thicker it becomes harder for polymer chains to permeate it. Shorter chains have then steric

and entropic advantages to penetrate the polymer layer and grow onto the surface. This

model, depicted in Figure 1-3, is important for the grafting system we describe because in

standard radical solution polymerization there are always growing chains of different

lengths as well as newly initiated chains present throughout the polymerization.

Page 27: through Surface-attached Monomers

1 Introduction 13

Figure 1-3. The kinetic barrier built up by already surface attached polymer coils hinders larger

polymer radicals that approach the layer at a later point of time to reach the surface and get connected

to it (a) more effectively than smaller polymer chains that may have a better chance to get through to

an unreacted surface monomer and get linked to it (b).

1.3.4 Polymer Micropatterning

In the last decades different patterning techniques such as embossing[155], photoresist

technology, photolithography[156, 157], microcontact printing[158, 159] or ink-jet

printing[160, 161] have been used to give a surface a desired structure and also to deposit

various substances in a structured manner. Low molecular weight compounds that attach to

surfaces of glass and silicon like silanes or to surfaces of noble metals like thiols have been

used to tailor the chemical composition of the solid substrates with spatial resolution.

Page 28: through Surface-attached Monomers

14 1 Introduction

Additionally there have been recent efforts to expand patterning to polymers because they

can deliver thicker layers and higher densities of functional groups. Various approaches with

destructive patterning of a homogeneous polymer layer like masking coupled plasma

treatment[162], UV-ablation, electron-beam lithography[163, 164] or even enzymatic

degradation[165] have been investigated. Different constructive approaches are possible

when using surface attached initiators to perform “grafting-from” polymerization like

polymerization of a pre-patterned initiator layer[166], photopolymerization of a

homogeneous initiator layer through a mask[166, 167], or even using a sequence of

photopolymerization and thermal polymerization to grow different polymer brushes from

the same initiator layer[168]. Another constructive approach is patterning by forming and

attaching a polymer network, also in multiple layers[60], through mask guided UV-cross-

linking of benzophenone units within a polymer. Direct photopolymerization of

poly(ethylene glycol) macromers with cross-linking compounds like di-, tri- or tetraacrylates

through a mask formed patterned hydrogels attached to a methacryloyl propyl

trichlorosilane modified substrate[169]. With direct thermal polymerization using pre-

patterned methacryloyl propyl trimethoxysilane modified substrates and UV-cross-linking of

benzophenone containing copolymers through a mask onto substrates with a homogeneous

MPS-layer we like to introduce a technique that offers robustness during the preparation

process as well as for the resulting product and the versatility of the previously mentioned

patterned polymer systems.

1.4 Radical polymerization Because radical polymerization is the main reaction in this work a short introduction to

this reaction collected from different textbook sources[170-172] will be given in the

following. Generally, in radical polymerization monomers containing a double-bond are

added to the radical end of a growing chain. The polymerization process can be divided into

three different stages: Initiation, propagation and termination. Before starting a growing

chain free radicals have to be formed. Special starter molecules, called initiators, are

decomposed by thermal or photochemical energy or generated by a redox process. In our

experiments we use azoisobutyronitrile (AIBN) as thermal initiator that decomposes into

two cyanopropyl radicals and nitrogen. The starting reaction is the addition of the first

monomer to the start radical. The addition of further monomers then is called propagation

reaction and it has a rate of around one monomer per millisecond although the rate may

vary dependent on the the used monomer. The growth is terminated through a termination

Page 29: through Surface-attached Monomers

1 Introduction 15

reaction which has two different possibilities: Recombination is the addition of two radical

ends of growing chains and disproportionation is when a hydrogen is transferred from the

end of one radical chain to the end of another radical chain.

The description of the kinetics of a radical polymerization can be divided into the three

stages mentioned above. Equation 1-4 shows the decomposition of an initiator molecule I

into two radical fragments R• with an initiation rate constant ki followed by the starting

reaction where a monomer M adds to the radical fragment (equation 1-5).

equation 1-4 •→ R2I ik

equation 1-5 •• →+ 1k M-RMR s

The initiation, where the primary radicals are formed, is typically slower than the starting

reaction where the first monomer is added. Because of that the initiation is the rate

determining step and the reaction rate is defined as the decrease of initiator concentration [I]

with time (equation 1-6).

[ ] [ ]Ikdt

Idi ⋅=− equation 1-6

The propagation reaction is when a monomer molecule M adds to a growing chain Mj•

where j is the number of already added repeating units. The propagation rate constant is kp

(equation 1-7).

equation 1-7 •+

• →+ 1jk

j M-RMM p

Now the polymerization rate can be calculated as shown in equation 1-8 where the

decrease of the monomer concentration [M] with time is a product of the propagation rate

constant, monomer concentration and the sum of all growing chains with various lengths.

[ ] [ ] [ ]∑ •⋅⋅=− jp MMkdtMd

equation 1-8

After initiation the growing chain adds monomer units within a short time period

(milliseconds to seconds) as long as monomers are present and unless any termination

reaction takes place.

Page 30: through Surface-attached Monomers

16 1 Introduction

Termination reactions are recombination (equation 1-9) and disproportionation

(equation 1-10) with the corresponding termination rate constants tc and td.

equation 1-9 •+

•• →+ kjk

kj MMM tc

equation 1-10 kjk

kj MMMM td +→+ ••

One termination reaction eliminates two undistinguishable radical chains. Now adding

both specific termination constants to one overall termination rate constant kt leads to

equation 1-11.

[ ] [ ]( 2

jtj Mk2

dtMd

∑ ••

=− ) equation 1-11

To simplify the kinetic model one presumes that during the reaction there is a steady state

where as many radicals are built as are consumed:

[ ] [ ]

dtMd

dtRd j

••

= equation 1-12

Using equation 1-11 and a combination of equation 1-6 and equation 1-4 we obtain for the

steady state

[ ] [ ]t

ij k

IkM =∑ • equation 1-13

Together with equation 1-8 the polymerization rate is calculated as

[ ] [ ] [ ] [ ] [ ]2

1

21

21

IMk

kkMMk

dtMd

t

ipjp ⋅⋅

⋅=⋅⋅=− ∑ • equation 1-14

If we define the kinetic chain length ν as the quotient of the propagation rate divided by

the initiation rate, we obtain for the steady state the following equation:

[ ]

[ ]( )21

Ikk2

Mkν

ti

p

⋅⋅

⋅= equation 1-15

Page 31: through Surface-attached Monomers

1 Introduction 17

With the termination factor a = 1 for disproportionation and 2 for recombination we get the

degree of polymerization nX .

νXn ⋅= a equation 1-16

The number average of the molecular weight distribution Mn is the degree of

polymerization multiplied with the monomer weight M0:

0nn MXM ⋅= equation 1-17

Combining equations 1-15 to 1-17 and unifying all constant values to a constant b results in

[ ][ ]2

1

IMM n ⋅= b equation 1-18

Here we see that the number average molecular weight for a standard radical

polymerization is proportional to the monomer concentration and reciprocal to the square

root of the initiator concentration.

The polydispersity PD is defined as the quotient of the weight average molecular weight

divided by the number average molecular weight.

n

w

MMPD = equation 1-19

For the different termination modes the polydispersity differs. In case of pure

disproportionation the theoretical polydispersity (derived from Schulz-Flory distribution) is

PD = 2 because all chains keep their length upon termination. For a termination by a pure

recombination mode the theoretical polydispersity is PD = 1.5 because the recombination

process statistically reduces the number of different chain lengths and the molecular weight

distribution is narrowed.

The temperature influence on a radical polymerization can be derived from the Arrhenius

equations for the initiation, propagation and termination velocity constants. The general

equation for a reaction rate constant k is

⋅−

⋅= TRE

eAk equation 1-20

Page 32: through Surface-attached Monomers

18 1 Introduction

where A is the frequency factor, E the activation energy of the reaction and R the gas

constant ( = 8.31451 J·K-1·mol-1). Using equation 1-20 with equation 1-15 and 1-16 delivers the

term for the temperature dependence of the degree of polymerization:

[ ][ ] 2

12

12

1IM

TR

E2E

2E

expAA2

AX

pti

ti

pn ⋅

−+⋅

⋅⋅⋅= a equation 1-21

Multiplied with the molecular weight M0 of the monomer (cf. equation 1-17) we get the

temperature dependence of the number average of the molecular weight:

[ ][ ] 0

pti

ti

pn M

IM

TR

E2E

2E

expAA2

AM

21

21

21 ⋅⋅

−+⋅

⋅⋅⋅= a equation 1-22

For batch reactions where defined amounts of initiator and monomer in the start reaction

mixture are given there is a time dependence of the molecular weight. Significant changes of

the reaction conditions, e.g. change of monomer/initiator concentration or viscosity, are the

reason for reaction stages a polymerization reaction runs through. Each stage may exhibit a

different distribution of the molecular weight of the polymer. Here conversion comes into

play as an important factor advancing with polymerization time.

Page 33: through Surface-attached Monomers

1 Introduction 19

Figure 1-4. Different types of progress in conversion with advancing time are depicted. Curve a)

shows a typical conversion diagram for a polymerization in the bulk. In a relatively early stage the gel-

effect causes an auto-acceleration of the polymerization. The final conversion lies below 1 because the

glass-effect “freezes” the mixture, polymer segments lose mobility and monomers cannot diffuse to

growing chain ends. Curve b) shows a conversion diagram for a polymerization with a medium

monomer concentration. Auto-acceleration takes place but mobility of the ingredients is retained by a

certain amount of a solvent. Curve c) shows a conversion diagramm where no gel-effect occurs and the

conversion rate decreases with time because the monomer concentration decreases as well. In the latter

case full conversion can be reached in a longer period of time.

With ongoing conversion the polymer concentration increases in the polymerization

mixture and therefore the viscosity of the mixture rises. The higher the initial monomer

concentration and the higher the molecular weight of the formed chains the more

pronounced this effect is. The influence of this diffusion controlled process is reflected in

higher viscous polymerization systems show stronger gel-effects. Above a certain polymer

concentration the polymer coils begin to overlap and entangle. This impedes the diffusion of

polymer radicals and the termination rate decreases whereas the advancing initiator

decomposition generates constantly new radicals leading to an increase of radical

concentration and growing chains (of any length). Consequently the polymerization rate

increases (Figure 1-4).

Page 34: through Surface-attached Monomers

20 1 Introduction

The decrease of the termination rate during the gel-effect causes an increase in kinetic

chain length. Therefore the degree of polymerization jumps up when the gel-effect starts

(arrow in Figure 1-5a). Polymer built during a gel-effect possesses a different molecular

weight distribution than the polymer built during the steady state. The distribution of the

degree of polymerization becomes broader.

Figure 1-5. Development of the average molecular weight with polymerization time in a

polymerization solution where the monomer content is high enough to lead to a gel-effect (a). The

arrow marks the beginning of this effect. In dilute polymerization solutions the average molecular

weight decreases parallel with the monomer concentration (b) as monomers are used up by polymer

formation. The grey area marks a region of linear molecular weight decrease where conversion increase

and monomer decrease are constant because of a steady state.

When the polymerization solution is as diluted as there is no gel-effect and the monomer

concentration decreases continuously with time because ongoing polymer growth consumes

monomer. When the initiation rate is held constant the average molecular weight will

decrease according to the monomer content within the solution (Figure 1-5b). As the

polymerization rate and conversion are constant in the steady state the molecular weight

decreases linearly (cf. Figure 1-5b). If significant amounts of initiator are consumed during

polymerization the decrease of the molecular weight may be compensated by a slight

increase of molecular weight due to lower initiator concentrations.

Page 35: through Surface-attached Monomers

2 Goals & Strategy of the Work 21

2 Goals & Strategy of the Work

2.1 Goals Surface-attached monomers are well established and often used in polymerization systems

but there is only little comprehension of such systems in regard to mechanistic details.

Monomer-modified substrates enable us to graft polymer onto them by radical

polymerization in solution. Such a reaction will occasionally attack a surface-attached

monomer leading to surface-attachment of the whole chain. We like to investigate this

method of surface modification using surface-attached monomers. First goal of this work is

to improve the understanding of the mechanism of a graft polymerization in the presence of

surface-attached monomers. This is attempted to reach by varying important polymerization

parameters which influence the kinetic chain length and in consequence the amount of

grafted polymer and graft density of the attached chains. Further goals emanate from the

knowledge we obtain by reaching the first goal. One is to explore the versatility of this

method by using different monomers on the surface as well as in the surrounding media.

Furthermore we investigate if silica porosity and size influence the result of the investigated

graft polymerization as these substrate properties are important for the use in different

applications. With the gathered knowledge we will be able to tailor polymer grafted surfaces

in a desired way. This leads to an investigation plot where feasibility studies for possible

applications for such modified substrates are performed.

2.2 Strategy Good polymerizability has to be a premised property of the surface-attached monomer

because the grafting step where the radical end of a growing chain connects to a monomer

on the surface has to be a fast and irreversible step to ensure a good yield of surface-attached

polymer chains. 3-methacryloyl-propyl trimethoxysilane (MPS) is a commercially available

and widely used compound with a silane anchor group. Because the methacrylic moiety is

linked to the spacer via an ester group that may be hydrolyzed under certain conditions it is

attempted to synthesize a similar compound where the methacrylate forms an amide group

which is more inert to hydrolysis. Amides can be hydrolyzed or transamidated only under

drastic conditions. To complete the picture also the acrylic versions of the ester linked and

amide linked compound are used in the experiments.

Page 36: through Surface-attached Monomers

22 2 Goals & Strategy of the Work

To get a deeper insight to graft mechanism of a polymerization using surface-attached

monomers, the influence of important polymerization parameters such as polymerization

time and temperature, concentration of monomer in solution and on the surface as well as

initiator concentration on the resulting material is investigated. The focus at this is on the

determination of grafted polymer amount via elemental analysis (EA), molecular weight

distribution of the built free polymer by gel permeation chromatography (GPC). From these

analytical results the graft density is calculated. The influence of the polymerization

parameters on both values, grafted polymer amount and graft density, is then evaluated.

Laible and Hamann already discovered differences in the grafted polystyrene amount on

MPS-modified Aerosil[147]. They found out that with higher styrene concentration during

polymerization the amount of the grafted PS increases. Similar graft amount determinations

were made by Browne, Chaimberg and Cohen to investigate the influence of polymerization

time on polyvinyl acetate grafting on vinyl triethoxysilane modified silica spheres[173].

Cohen et al. developed a complex kinetic model for free radical graft-polymerization based

on a conservational polymerization and molecular weight distribution numerical algorithm

to predict graft yield for polyvinylpyrrolidone (PVP) on material with polymerizable surface

sites[149]. These “kinetic modeling” considerations do not provide a clear picture or answers

to questions like what happens if the surface-attached monomer has a different reactivity

compared to the monomers in solution or how does the porosity of the substrate or the

viscosity of the surrounding media influence the polymer monolayer buildup. As already

mentioned in 1.3.3 the formation of a graft polymer layer limits the grafting of further

polymer in a certain way. It is important to know in what manner the polymerization

parameter influence the surface polymer layer formation and in consequence the control of

this process.

Besides the understanding of the mechanism of this method we investigate its versatility

by varying the polarity of the used monomers from a hydrophobic perfluoro monomer to a

hydrophilic monomer like methacrylic acid. We will investigate if with surface-attached

monomers these different monomers in solution form a chemically bound polymer layer on

the modified substrates. To underline and expand this versatility the introduction of

functional groups by copolymerization in solution with special tailor-made monomers will

be a suitable way. These monomers bear amino or activated ester groups which allow many

further reactions, e.g. with catalyst species or biomolecules. If successfully applied to our

system this may lead to valuable applications for the silica-polymer hybrid systems. As

examples we will carry out copolymerizations with the aforementioned monomers in

Page 37: through Surface-attached Monomers

2 Goals & Strategy of the Work 23

solution in combination with styrene to form a hydrophobic layer, or in combination with

N,N-dimethyl acrylamide to form a hydrophilic layer on different silica substrates.

According to the strategy of this work, substrates with different particle and pore sizes are

used which are commercially available and used in a broad range of applications as we have

introduced in the previous chapter. Here the specific surface of the used substrates ranges

from 700 m2 per gram for the highly microporous LiChrospher Si60 down to 40 m2 per gram

for macroporous glass beads with 100 nm pores. All values were determined by the producer

via BET method[174].

LiChrospher Si60 is a silica gel consisting of monolithic particles with spherical shape, a size

of around 25 µm and micropores with a diameter of around 2 nm (Figure 2-1).

Figure 2-1. SEM images of LiChrospher Si60 before (a) and after 24 hours (b) of polymerization of

styrene.

Page 38: through Surface-attached Monomers

24 2 Goals & Strategy of the Work

Aerosil300 is a pyrogenic silica gel and has a specific surface of 300 m2·g-1. Its body consists

of small sintered silica beads building up a three-dimensional network with concave pore

geometry (Figure 2-2).

Figure 2-2. TEM images showing Aerosil300 in a pure state (a) and after 24 h PS polymerization (b).

The glass beads involved in our studies have pore sizes of 20 nm, 50 nm and 100 nm

(“mesoporous” up to “macroporous”). Their specific surfaces cover with 250 m2, 80 m2 and

40 m2 per gram the regions of medium and lower specific surfaces (Figure 2-3).

Figure 2-3. A CCD image of GB80 before used for polymerization (a); TEM reveals mesopores on the

surface of the glass beads (b) that stay intact after been in polymerization solution for 24 h.

Furthermore we investigate in how far chemical microstructures can be formed by using a

system with surface-attached monomers. Flat substrates such as silicon wafers will be

Page 39: through Surface-attached Monomers

2 Goals & Strategy of the Work 25

equipped with surface-attached monomers and chemically microstructured in two different

ways. The first method will use a mask on the surface and surface monomers will be

destroyed in the illuminated areas by UV ablation. For the second method a polymer layer

with incorporated benzophenone units will be deposited and subsequently cross-linked by

UV irradiation. In the latter method C-H-bonds of the propyl spacer of the surface-attached

monomers will be used as anchor points for the benzophenone. Additionally a second

polymerization using the graft process will be possible afterwards to form a polymer layer

with different properties which might be an interesting feature.

Page 40: through Surface-attached Monomers

26 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-

attached Monomers

3.1 Silane Monomers For the system described here the first step towards the deposition of the polymer

monolayers is the modification of the silica surfaces with appropriate monomer units

(Scheme 3-1). To allow a deeper insight into the subsequent polymerization reaction we have

looked at monomers that differ somewhat in their structure. The silanes we used were

acrylates and methacrylates. All of them carried a trimethoxysilane anchor group separated

from the (meth)acrylic acid residue by a propyl spacer. This trimethoxysilylpropyl moiety

was linked to the methacrylic acid via a ester or amide linkage as it can be expected that the

amide might be less prone to hydrolysis compared to the esters in chemically harsh

environments. On the other side, the nature of the polymerizable group i.e. acrylic and

methacrylic tails, was varied to study the influence of the reactivity of the monomer, i.e. their

polymerization behavior, of these two groups during the subsequent free radical

polymerization process for the deposition of the polymers on the silica particles. The name of

the compounds are 3-methacryloyl-propyl trimethoxysilane (MPS), N-methacryloyl-N-

methyl-propyl trimethoxysilane (MNPS), 3-acryloyl-propyl trimethoxysilane (APS) and N-

acryloyl-N-methyl-propyl trimethoxysilane (ANPS). It is well-known that trifunctional

silanes like the ones used in this study tend to form not ideal monolayers but surface

attached networks with varying structure due to hard to control pre-condensation reactions

in solution prior to deposition on the surface[175-177]. An factor which is here especially

important is the presence of water. As we used a porous silica gel such processes could,

under certain reaction conditions, lead to situations in which the pores become clogged with

a rather thick disordered siloxane layer. In order to avoid such complications we decided to

deposit the monomer silanes under strictly inert conditions, meaning that all solvents and

reagents were rigorously dried by appropriate means and all deposition reactions were

carried out in an atmosphere of dry nitrogen.

Page 41: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 27

Scheme 3-1. Reaction scheme for the preparation of SiO2 substrates with surface-attached monomers.

Substrates modified with polymerizable groups are generated by immobilization of a trimethoxy silane

carrying a (meth)acrylic group (1). These modified substrates can then be used in a radical

polymerization to form ultrathin (co)polymer monolayers on the substrate surface (2).

The resulting layers were characterized qualitatively using diffuse reflectance FT-IR

spectroscopy (DRIFT). Due to the fairly similar chemical nature of the silanes, very similar

spectra were obtained for all monolayers. As an example, Figure 3-1 shows the spectrum of a

MPS layer in comparison to a spectrum obtained from the as-obtained, i.e. unmodified, silica

gel. Signals as the C-H stretching vibrations below 3000 cm-1, a strong carbonyl absorption

band at 1700 cm-1, and the C=C stretching band at 1636 cm-1 are a clear indication for the

presence of an MPS layer on the silica surface.

Page 42: through Surface-attached Monomers

28 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

4000 3000 2000 1000

HH( C=C )ν

ν (5H Aryl)

ν

ν

ν

νν

(CH)

(SiOH)

(SiOSi)

(C=C)(C=O)

ν

(Aryl-H)

tran

smitt

ance

(a.u

.)

wavenumber (cm-1)

Figure 3-1. Diffuse reflectance FT-IR (DRIFT) spectra obtained from unmodified silica gel (1), SiO2-

MPS (2), and a silica gel carrying a polystyrene monolayer (3).

In order to obtain a more quantitative picture the materials were investigated by elemental

analysis which allows for the determination of the grafting density of the silanes on the

LiChrospher surfaces in first place. The carbon contents of all materials was determined by

elemental analysis. The values were all rather close to each other and ranged from 6-8% (cf.

Table 3-1). The other substrates such as Aerosil and the different glass beads come into play

later in this work. Surface attachment of the monomer silanes is fundamental for all

investigations so these results are the first to be discussed. From the carbon content values

the immobilized amount δ can be calculated according to equations 1-1 and 1-3. Using the

immobilized amount and the specific surface area of the LiChrospher (700 m2·g-1 according to

the BET method[174]) the surface concentration Γ0 of the polymerizable groups can be

calculated.

Page 43: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 29

Table 3-1. Carbon contents and graft densities on different SiO2 surfaces with different immobilized

monomers.

substrate silica material

+ specific surface (m2·g-1)

carbon*

(%)

Γ0

(µmol·m-2)

LC700-ME (MPS) LiChrospher Si60 (700 m2) 6.9 ±0.9 0.9 ±0.1

LC700-MA (MNPS) LiChrospher Si60 (700 m2) 7.9 ±0.8 0.9 ±0.1

LC700-AE (APS) LiChrospher Si60 (700 m2) 6.0 0.9

LC700-AA (ANPS) LiChrospher Si60 (700 m2) 6.4 0.9

LC700-Si-octyl† (OS) LiChrospher Si60 (700 m2) 5.5 ±1.3 0.7 ±0.2

AR300-ME (MPS) Aerosil300 (300 m2) 6.4 2.0

GB250-ME (MPS) glass beads (250 m2) 4.8 1.8

GB250-AA (ANPS) glass beads (250 m2) 5.2 2.1

GB80-ME (MPS) glass beads (80 m2) 1.96 ±0.02 2.27 ±0.03

GB80-MA (MNPS) glass beads (80 m2) 2.13 ±0.09 2.47 ±0.11

GB80-AE (APS) glass beads (80 m2) 1.72 ±0.06 2.24 ±0.08

GB80-AA (ANPS) glass beads (80 m2) 1.97 ±0.09 2.56 ±0.11

GB40-ME (MPS) glass beads (40 m2) 1.3 3.1

GB40-MA (MNPS) glass beads (40 m2) 1.9 4.4

* obtained from elemental analysis, † reference sample with surface-attached octylsilane monolayer

It can be seen that monomer graft densities around 0.9 µmol·m-2 are found for LiChrospher

systems. These values appear to be somewhat lower than that of other trifunctional silanes

on comparable surfaces. However, it was not our premier goal to deposit as many

polymerizable groups as possible, but we were solely interested in finding conditions that

allow for a preparation of these layers in a very reproducible fashion and to avoid extensive

network formation which might cause problems in the polymerization reactions.

Appropriate monomer silane units are immobilized on the other substrates as well. The

standard MPS was used for all substrates, LiChrospher, Aerosil & different glass beads. In

special cases of interest also MNPS, APS and ANPS were attached to the surface silanol

Page 44: through Surface-attached Monomers

30 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

groups. For MPS the carbon amount after immobilization is proportional to the specific

surface of the substrates with pore sizes between 5 and 50 nm (AR300-ME, GB250-ME, GB80-

ME) and a resulting monomer graft density of around 2 µmol·m-2 (Table 3-1). LC700-ME

possess a lower graft density (Γ = 0.9 µmol·m-2) probably due to inaccessibility of its smallest

pores. GB40-ME shows a higher graft density (Γ = 3.1 µmol·m-2) because the larger sized

100 nm pores allows larger monomer silane clusters bind to the surface. Different monomers,

which have the same reactivity towards silica surfaces because of the same head group

architecture build monomer layers of similar graft densities on the same glass beads (GB80-

ME, GB80-AE & GB80-AA).

Page 45: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 31

3.2 Homopolymers on Silica Different polymers have been attached to the surface during formation and incorporation

of the surface-attached monomers. Regardless of their polarity all polymers formed a well

defined layer on the silica surface (Figure 3-2).

Figure 3-2. The pathways from a silane modified silica substrate to different immobilized polymers

layers are summarized.

Page 46: through Surface-attached Monomers

32 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

The results of a series of different polymerization reactions are summarized in Table 3-2.

The P(HDFDA) achieved the highest polymer load for the given parameters, but molecular

weight and graft density could not be determined because the perfluoro polymer was not

soluble in any solvent which could be used for GPC. P(MMA) showed also loading values

for mid sized polymer coils above those of comparable PS graft polymerizations. NIPAAm

and MAA with high propagation rate constants formed homopolymers that exceeded the

measuring range of the GPC columns so that the molecular weight was estimated to be

higher than the highest calibration standard for the SUPREMA column (> 1,300,000 g·mol-1).

Probably these large polymer coils clog easily the micropores of the LiChrospher. Therefore

the difference in polymer loading between silica gel with high specific surface area and glass

beads with medium specific surface area is less pronounced and meets at a level of around

0.4 g polymer per gram silica. PS as standard system was well described in recent

publications and its loading is higher and the graft density is lower for the LiChrospher case

in comparison to the glass beads. The polymer load with P(DMAA) was for both substrates

quite low because the chosen monomer concentration was low to avoid gel formation and

lead to low molecular weight polymer (~18,000 g·mol-1). Higher monomer concentrations

caused gelation in early stages of the polymerization. It was shown that the polymer load can

be driven to much higher values in a controlled manner despite to a gel effect occurring at

high conversion[178]. In general the high surface LiChrospher silica allowed a higher polymer

load compared to the mesoporous glass beads whereas the glass beads show in general

higher graft densities.

Page 47: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 33

Table 3-2. Polymer loads and graft densities from polymerization reactions of different monomers on

different silica substrates.

product mʹpoly per g2SiO

*

(g)

nM #

(g·mol-1)

Γ0

(µmol·m-2)

LC700-AE-P(HDFDA) 1.04 - -

GB80-AA-P(HDFDA) 0.59 - -

LC700-ME-PS 0.46 36,000 0.018

GB80-ME-PS 0.14 32,500 0.054

LC700-ME-P(MMA) 0.78 76,500 0.015

GB80-ME-P(MMA) 0.36 80,000 0.056

LC700-ME-P(NIPAAm) 0.40 > 1,300,000‡

GB80-ME-P(NIPAAm) 0.32 > 1,300,000‡

LC700-ME-P(DMAA) 0.12 18,000 0.009

GB80-AE-P(DMAA) 0.09 18,000 0.061

LC700-ME-P(MAA) 0.46 > 1,300,000‡

GB80-ME-P(MAA) 0.42 > 1,300,000‡

* determined from elemental analysis, # determined from GPC analysis of the free polymer, ‡ GPC column exclusion limit reached

The various surface-attached homopolymer monolayers were first characterized by DRIFT

spectroscopy. P(HDFDA) showed adsorption bands of C=O stretching vibrations at 1738 cm-1

and C-F stretching vibrations at 1193 cm-1 although the latter band overlapped with the

broad Si-O-Si scaffold vibrations (1200-1000 cm-1). Bands due to aromatic C-H stretching

vibrations (3100-3000 cm-1, several bands ) were observed in case of the PS covered solids. In

the P(MMA) case there was a typical signal for C=O stretching vibration at 1730 cm-1.

P(NIPAAm) and P(DMAA) showed both C=O stretching of the amide group (amide I) which

lied for the monosubstituted monomer P(NIPAAm) at 1650 cm-1; a few wavenumbers higher

than the disubstituted case P(DMAA) at 1641 cm-1. For P(NIPAAm) as a monosubstituted

amide there was an additional band due to a N-H stretching vibration (amide II) at 1530 cm-1.

Page 48: through Surface-attached Monomers

34 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

A characteristic C=O stretching vibration band at 1710 cm-1 was observed for P(MMA) on the

solids.

For those homopolymer carrying substrates that were measured with x-ray photoelectron

spectroscopy (cf. Figure 3-3) the measured elemental composition of the surface-attached

polymer monolayer corresponds to values one expects from elemental analysis of free

polymer and silica combined (cf. Table 3-3). For the P(DMAA) monolayer the silicon signals

are quite intense and are reduced for the polystyrene layer on silica compared to the

P(DMAA) layer. Because in XPS experiments the depth from that photoelectrons can escape

from the material through the surface to be detected and analyzed is limited to a few

nanometers. Due to this fact stronger silicon signals indicate a thinner polymer layer. We

conclude that for the used polymerization conditions the P(DMAA) layer is thinner than the

PS layer. In case of the perfluoro polymer both silicon signals have almost vanished and we

got the here the thickest layer what is in agreement with the results of elemental analysis.

1000 800 600 400 200 00

50

100

150

200

c)

b)

a)

O KLLF KLL

F 2s

F 1s

Si 2pSi 2s

C 1s

N 1s

O 1s

coun

ts (1

000/

s) (a

.u.)

binding energy (eV)

Figure 3-3. X-ray photoelectron spectra show the surface composition of glass beads with different

homopolymer monolayers: (a) polystyrene, (b) poly(heptadecafluordecyl acrylate) and (c)

poly(dimethyl acrylamide).

Page 49: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 35

Table 3-3. Ratios of different elements on the surface of silica particles modified by various polymers.

product C*

(%)

N*

(%)

F*

(%)

O*

(%)

Si*

(%)

GB80-AA-P(HDFDA) 33.0 - 52.2 - 4.7

GB80-AA-PS 45.9 - - 35.6 18.5

GB80-AE-P(DMAA) 31.9 5.0 - 43.3 19.9

LC700-ME-PS 58.1 - - 27.0 14.9

* obtained from XPS and standardized

3.3 Mechanism of the Grafting of Polystyrene onto Silica

Surfaces through Surface-attached Monomers As substrate for the following detailed investigations the highly porous LiChrospher Si60

(LC700) was used because its large specific surface grants that significant amounts of

polymer bind to the silane monomers which had been attached in the first step. The behavior

of other substrates, especially various glass beads, upon monomer immobilization and

subsequent polymerization is described in 3.7.

3.3.1 Structure Variation of the Immobilized Monomer

To study the polymerization behavior of various surface-attached acrylate and

methacrylate monomers, styrene was used as the “free” monomer and polymerizations were

triggered using azobis(isobutyronitrile) (AIBN) as an initiator. The following conditions were

employed: cstyrene = 2.9 mol·L-1, cAIBN = 9 mmol·L-1, csilica = 10 g·L-1, temperature T = 60°C. The

polymerization time was set to 24 h. In two separate reference experiments unmodified silica

gel and silica gel treated with an alkyl silane (octyl trimethoxysilane) were used as substrates

for the polymerization mixture.

All gels were extensively washed after polymerization to remove any adsorbed, “free”

polymer from the surfaces (for details see 9.1.1 and 9.3). The so obtained solids were

characterized by FTIR spectroscopy (DRIFT) and elemental analysis. Figure 3-1 shows a

spectrum obtained from an MPS-SiO2 sample after it was exposed to a polymerization

reaction. The attachment of polystyrene is proven by the presence of adsorption bands

Page 50: through Surface-attached Monomers

36 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

typical for polystyrene, especially in the region just above 3000 cm-1 which is indicative for

the presence of aromatic C,H groups in the sample.

The amount of polymer attached to the surfaces was determined by elemental analysis. For

the two reference samples which contained no immobilized monomer at the surface only

very small carbon contents of 0.5 – 3% (Γ0 < 1 nmol) were determined. Such values are often

found even on the surfaces of untreated silica gels due to airborne carbon containing

contaminants and are therefore not indicative for any attachment of polymers. For all other

silanes much higher carbon values of 28 – 33% (Γ0 = 15 - 18 nmol·m-2) were found except for

SiO2-MNPS where much lower polymer loadings (i.e. only one forth or one fifth of those of

the other monomers) were observed. All values are summarized in Table 3-4 together with

the molecular weights of the “free” polymers recovered from the respective polymerization

mixture.

Table 3-4. Carbon contents and graft densities of surfaces prepared from different immobilized

monomers and surface-attached polystyrene monolayers on LiChrospher Si60.

carbon*

(%)

nM #

(g·mol-1)

Γ0

(µmol·m-2)

substrate

monomer polystyrene monomer polystyrene

LC700-ME (MPS) 8.1 32.9 36,000 1.3 0.018

LC700-ME (MNPS) 8.2 15.0 36,000 1.2 0.004

LC700-ME (APS) 6.0 28.1 33,000 1.0 0.018

LC700-ME (ANPS) 6.5 31.5 36,000 1.0 0.017

LC700-Si-octyl† (OS) 6.5 7.0 37,000 0.9 <0.001

LC700 (unmodified) - 3.0 40,000 - 0.001

* obtained from elemental analysis, # determined from GPC analysis of the free polymer, † reference sample with surface-attached octylsilane monolayer

In order to obtain values for the graft density of the polystyrene on the various surfaces we

assumed that the molecular weights of the free and the attached polymers are similar. This

assumption seems to be reasonable because the molecular weight of a polymer prepared by

free radical polymerization is determined by the ratio of the rate constant of propagation to

that of the termination reaction. It is known from the corresponding kinetics of solution

Page 51: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 37

polymerization that a significant change of the termination reaction only occurs at

conversions of 50% or above. Such segment densities, however, do not occur in the grafted

layers.

As we compare the graft densities of the polymers on the silica gels to the graft density of

the polymerizable groups we find that only about 3-5% of all surface-bound vinyl groups

were actually “caught” by a growing chain and thereby incorporated into the surface-

attached polymer, if we assume that each chain gets only attached by the incorporation of

one monomer anchored to the surface.

3.3.2 Variation of the Polymerization Time

In order to further elucidate this process for the modification of silica particles with

polymers and to explore the limits of the method in terms of available graft densities or

grafted amounts we performed a series of polymerizations in which all conditions were kept

constant and only the polymerization time was varied from 2 –72 h. The grafted amount and

the graft density of the resulting materials were derived from elemental analysis data and

from the molecular weights of the “free” polymers in the same way as described above

(Figure 3-4). The data depicted as a triangle in Figure 3-4 Figure 3-5 and Figure 3-6 were

obtained from a sample for which the polymerization mixture was renewed after a initial

polymerization time of 48 h. Details are explained in the text.

Page 52: through Surface-attached Monomers

38 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

0 10 20 30 40 50 60 7005

101520253035404550

reaction time (h)

Mn (

103 g

/mol

)

Figure 3-4 Influence of the polymerization time on the number average of the molecular weight

distribution as obtained by GPC analysis of the free polymers

It can be seen that the molecular weight of the polymers stays constant for the entire range

of polymerization times as it is expected for a typical free radical polymerization.

Furthermore, the value of 36,000 g·mol-1 agrees well with the expected molecular weight

derived from the Mayo equation (equation 3-1) for the given set of polymerization

parameters[179, 180].

[ ][ ] 0

SMn X

1MSCC

X1

+⋅+= equation 3-1

With this equation the molecular weight can be calculated from the degree of

polymerization nX according to equation 1-17. The constant CM (= 1,67·10-4) is the quotient of

the propagation rate constant kp divided by the transfer rate constant km for a transfer

reaction from a growing chain to a monomer. The constant CS (= 8,3·10-4) is the quotient of kp

as dividend and ks the transfer rate constant for a transfer from a growing chain to a solvent

molecule as divisor. For the given polymerization mixture the styrene concentration [M] is

2.9 mol·L-1 and the toluene concentration [S] is 6.3 mol·L-1.The degree of polymerization in

absence of a solvent 0X (= 1920) we get from the experiments carried out by Bialk et al.[153].

The used initiator concentrations are same in the compared cases. All values put into

Page 53: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 39

equation 3-1 give a calculated number average of the molecular weight of 41,200 g·mol-1.

More, in this simplified equation unconsidered, transfer reactions may reduce the molecular

weight further. Additionally the GPC measurement method implies an error of up to 10%.

Taking these two factors into account the theoretical value is close to the practical one.

0 10 20 30 40 50 60 700.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

graf

ted

poly

mer

(g P

S / g

SiO

x)

reaction time (h)

Figure 3-5. Influence of the polymerization time on the grafted amount of polystyrene onto a silica

surface.

The grafted amount in terms of g PS per g SiOx is initially a function of the polymerization

time but after about 25-30 h of polymerization no further polymer chains (or only minor

amounts) become attached to the particles. Accordingly, also the graft density initially

increases as the polymerization proceeds but then levels off after about 25-30 h at around

0.5 g PS·g-1 SiOx. This corresponds to an theoretical average layer thickness of 0.7 nm when

the total surface of the LiChrospher material is covered equally. If the molecular weight of the

polymer is taken into account that is a graft density of 20 nmol·m-2. It should be noted, that

this behavior is not caused by exhaustive conversion of initiator or monomer. This is

demonstrated in a control reaction in which styrene was firstly polymerized in the presence

of SiO2-MPS for 48 h, then the silica gel was recovered, washed and subjected again to a

polymerization solution for another 24 h corresponding to a total polymerization time of

again 72 h. The total grafted amount of 0.52 g PS / g SiO2 determined from this sample is very

similar to the value found for the sample which was directly polymerized for 72 h (cf. Figure

Page 54: through Surface-attached Monomers

40 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

3-5 and Figure 3-6). The error of elemental analysis measurements (~1%) is too small to be

displayed as error bars in the diagrams that show grafted polymer amounts.

0 10 20 30 40 50 60 700.000

0.005

0.010

0.015

0.020

0.025

0.030

graf

t den

sity

(µm

ol/m

²)

reaction time (h)

Figure 3-6. Influence of the polymerization time on the graft density of the polystyrene chains on the

silica.

3.3.3 Influence of the Silica Gel Concentration

In another set of experiments we explored the practical limits of the described approach

and varied the overall amount of modified silica for a given set of polymerization conditions

(styrene/toluene 1/2 v/v, 15 mL, cAIBN = 9 mmol·L-1, T = 60°C, t = 24 h). The amount of

modified silica (SiO2-MPS) was varied from 0.15 g to 4.8 g, i.e. by a factor of 32. The highest

silica concentration used in this set of polymerizations poses an upper practical limit as the

solids are barely wetted by the monomer solution. Again, the resulting materials were

characterized in terms of the grafted amount, the molecular weight of the polymers and the

graft density as described above and the results are plotted as a function of the concentration

of the solid in Figure 3-7 and Figure 3-8. This plot demonstrates that the amount of solid

does not influence the overall procedure and graft densities of 19±1 nmol/m² were obtained

for all cases. However, it was found that the amount of “free” polymer which is formed

during the reaction diminishes if higher amounts of solid are added to the mixture.

Comparing the amount of isolated “free” PS from the 0.15 g and the 4.8 g sample the weight

Page 55: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 41

is halved from 1.3 g to 0.6 g for 15 mL of polymerization solution (10 mL toluene, 5 mL

styrene) after a conversion of approximately 30 – 40%.

0 50 100 150 200 250 300 3500.0

0.1

0.2

0.3

0.4

0.5

0.6

graf

ted

poly

mer

(g P

S / g

SiO

x)

silica gel concentration (g/L)

Figure 3-7. Grafted amount of polystyrene as determined by elemental analysis for varied

concentrations of the monomer modified silica gel inside a polymerization dispersion.

0 50 100 150 200 250 300 3500.000

0.005

0.010

0.015

0.020

0.025

0.030

graf

t den

sity

(µm

ol/m

²)

silica gel concentration (g/L)

Figure 3-8. Graft density of polystyrene for varied concentrations of the monomer modified silica gel

inside a polymerization dispersion.

Page 56: through Surface-attached Monomers

42 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

3.3.4 Influence of Monomer Concentration in Solution

The monomer concentration is an important parameter for free radical polymerizations

and how it effects the built-up of the polymer monolayer onto silica is explored in this set of

experiments. A change of the monomer concentration in solution by nearly a factor of ten

causes countable differences in polystyrene load. The higher the styrene concentration the

higher the final polymer load on LiChrospher. For polymerization in pure styrene (8.7 mol·L-1)

the grafted polymer achieved 0.49 g PS per gram silica gel as maximum. Diluting the styrene

with toluene down to 0.97 mol·L-1 (v/v 1/8) the polystyrene content drops down to 0.15 g PS

per gram of silica (Figure 3-9). The factor of dilution (x8) is higher than the decrease in

polymer load (x3).

0 1 2 3 4 5 6 7 8 90.0

0.1

0.2

0.3

0.4

0.5(1/0)

(1/1)(1/2)

(1/4)

(1/8)

graf

ted

poly

mer

(g P

S / g

SiO

x)

c(M) (mol/L)

Figure 3-9. Influence of the monomer concentration on the amount of grafted polystyrene as

determined by elemental analysis; values in brackets: ratio styrene/toluene (v/v)

The average molecular weight for the formed free polystyrene increases proportionally

(Figure 3-10). For a styrene concentration of 0.97 mol·L-1 the number average of the molecular

weight is 21000 g·mol-1. Up to a styrene concentration of 4.35 mol·L-1 and a number average

of the molecular weight of 114000 g·mol-1 the graft density drops linearly from 10 down to

4 nmol·m-2 (Figure 3-11).

Page 57: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 43

0 1 2 3 4 5 6 7 8 90

100

200

300

400 (1/0)

(1/1)

(1/2)(1/4)

(1/8)

Mn (

103 g

/mol

)

c(M) (mol/L)

Figure 3-10. Correlation between molecular weight of the free polymer and monomer concentration in

solution; higher concentration leads to higher molecular weights.

0 1 2 3 4 5 6 7 8 90.000

0.002

0.004

0.006

0.008

0.010

0.012

(1/0)

(1/1)

(1/2)

(1/4)

(1/8)

graf

t den

sity

(µm

ol/m

²)

c(M) (mol/L)

Figure 3-11. Graft density as a function of monomer concentration calculated from elemental analysis

and GPC results.

The molecular weight of the polystyrene created during the polymerization in pure

styrene does not follow this trend. In this case the molecular weight is higher than expected.

During polymerization this mixture got highly viscous. Caging effects of growing radical

chains due to reduced mobility of the polymer coils lead to higher molecular weights. This

Page 58: through Surface-attached Monomers

44 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

effect occurs during polymerization at higher conversion and is known as Trommsdorf effect

[178]. So it is due to the long polymerization time of 48 hours for this sample. The molecular

weight was in this case 400000 g·mol-1 and a factor of 20 higher than for the lowest monomer

concentration. Therefore the graft density leaves the linear trend and is settled at 2 nmol·m-2.

3.3.5 Influence of Initiator Concentration

Molecular weight of radical polymerization reactions can be easily controlled by variation

of the initiator concentration. To elucidate how this parameter influences the polymer

monolayer while the monomer concentration is kept constant, the initiator concentrations

were varied in a range from 10-4 to 10-1 mol·g-1.

1E-4 1E-3 0.01 0.10.0

0.1

0.2

0.3

0.4

0.5

graf

ted

poly

mer

(g P

S / g

SiO

x)

c(I) (mol/L)

Figure 3-12. Influence of initiator concentration on the amount of deposited polystyrene as determined

by elemental analysis.

Page 59: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 45

0 20 40 60 80 1000

50

100

150

200

250

Mn (

103 (g

/mol

))

(c(I))-1/2 (L1/2/mol1/2)

Figure 3-13. The molecular weight as a function of the reciprocal square root of the initiator

concentration according to the dependence in equation 1-18.

1E-4 1E-3 0.01 0.10.00

0.02

0.04

0.06

0.08

0.10

0.12

graf

t den

sity

(µm

ol/m

²)

c(I) (mol/L)

Figure 3-14. Influence of the initiator concentration on the graft density as calculated from grafted

amount and average molecular weight.

With increasing initiator concentration the polystyrene load increases from 0.02 to 0.45 g

per gram silica (Figure 3-12). The molecular weight decreases from 225000 down to

6000 g·mol-1. As shown in Figure 3-13 the molecular weight increases linearly with the

reciprocal square root of the initiator concentration. For initiator concentrations above

Page 60: through Surface-attached Monomers

46 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

1 mmol·L-1 the dependence of the molecular weight from the initiator concentration derived

from equation 1-18 is valid. Both effects combined result in a strongly raised graft density

which increases by a factor of 1000 from 0.1 nmol·m-2 to 100 nmol·m-2 (Figure 3-14).

3.3.6 Variation of the Polymerization Temperature

In several experiments the reaction temperature for the polymerization was varied in a

range from 60 to 90 °C. Two sets of polymerizations were performed: In the first set of

experiments polymerization reactions were carried out using polymerization times that

correspond to one half-life time of the initiator for each temperature. In the second set of

experiments the polymerization time was set to 21.5 h for each run regardless of the

temperature chosen (cf. Table 9-6).

330 340 350 360 370

0.1

0.2

0.3

0.4

0.5

21.50 h

16.05 h10.85 h

4.00 h

1.05 h

0.4 h

polymerization times

graf

ted

poly

mer

(g P

S / g

SiO

x)

T (K)

Figure 3-15. Temperature dependence of the grafted amount of polystyrene to a silica surface. All

samples were polymerized up to an initiator conversion of 50%. The polymerization time for each

sample is given next to the respective data point.

Page 61: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 47

330 340 350 36005

101520253035

Mn (

103 g

/mol

)

T (K)

Figure 3-16. Temperature dependence of the number averaged molecular weight of the free polystyrene

during polymerization of a monomer modified silica (LC700-ME), initiator conversion 50%.

330 340 350 3600.000

0.005

0.010

0.015

0.020

graf

t den

sity

(µm

ol/m

²)

T (K)

Figure 3-17. Temperature dependence of the graft density of polystyrene on a silica surface after an

initiator conversion of 50%.

The results obtained from the polymerizations in which the reaction times were set to one

half-life time of initiator decomposition are shown in Figure 3-15 to Figure 3-17. Both, the

Page 62: through Surface-attached Monomers

48 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

grafted amounts and the molecular weight decrease with increasing temperature. The

dependence of the molecular weight of the obtained polymer on the reaction temperature

shows a behavior as it is expected on the basis of the standard description of the mechanism

of free radical polymerizations that predicts a decrease of the molecular weight with

increasing temperature. As we calculate the graft density of the resulting polymer

monolayers we find that this value does not depend on the polymerization temperature but

stays constant for the full range of temperatures under investigation. This result

demonstrates that the decrease of the grafted amount with increasing temperature is solely

caused by the decreasing molecular weight for this set of polymerizations.

A different picture was obtained from the experiments in which the polymerization time

was set to 21.5 h for all temperatures. The results of these experiments are shown in Figure

3-18 to Figure 3-20.

330 340 350 3600.0

0.1

0.2

0.3

0.4

0.5

0.6

graf

ted

poly

mer

(g P

S / g

SiO

x)

T (K)

Figure 3-18. Grafted amount of polystyrene obtained from polymerizations at different temperatures

for a constant polymerization time of 21.5 h.

Page 63: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 49

330 340 350 3600

10

20

30

40

50

60 Mn Mw

mol

ecul

ar w

eigh

t (10

3 g/m

ol)

T (K)

Figure 3-19. Polymerization at different temperatures for a constant polymerization time of 21.5 h

gives for polystyrene the molecular weights as shown in Figure 3-18.

330 340 350 3600.000

0.005

0.010

0.015

0.020

0.025

0.030

0.035

graf

t den

sity

(µm

ol/m

²)

T (K)

Figure 3-20. Polystyrene graft densities from polymerizations where the polymerization temperature

was varied and the polymerization time was kept constant at 21.5 h.

Both the grafted amount and the graft density firstly decrease as the samples are prepared

at higher temperatures but the minimum is reached for a polymerization temperature of

65°C after which higher amounts of polymer can be grafted to the surfaces if higher

polymerization temperatures are employed. Also, a different trend for the molecular weight

Page 64: through Surface-attached Monomers

50 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

as a function of temperature is observed. Initially, Mn decreases with increasing

polymerization temperature as it was observed for the first set of experiments. At moderate

temperatures up to 65°C the graft process and the polymer formation follow the expected

kinetics and the molecular weight distribution is similar to that of the experiment with

constant initiator conversion. Also the graft density for the lower temperature range can be

considered constant within the error margins. However, for temperatures of 70°C and higher

the value for Mn settles at a certain value while Mw increases sharply with increasing

polymerization temperature. At higher polymerization temperatures the rapid consumption

of initiator causes a significant drop of its concentration and formation of longer polymer

chains occurs at later polymerization stages. Therefore an increase for the polydispersity

index from initially 1.7 to values more than 2 is observed which is largely due to a more and

more bimodal distribution of the molecular weight distribution as evidenced in the GPC

traces. This is probably caused by the change of polymerization conditions at high monomer

conversions with low initiator concentrations. For higher polymerization temperatures the

polymer chains become shorter and the polymer formation is faster. So the yield of surface-

attached polymer coils becomes better again with higher polymerization temperatures when

polymerization time is kept constant.

3.3.7 Surface Concentration of Polymerizable Silanes & its

Influence on Building Polymer Monolayers

3.3.7.1 Co-Immobilization of MPS & OS

The initial step for attaching polymer monolayers to surfaces through the strategy

described here is the immobilization of a monomer unit on the silica. To get a better

understanding how the grafting process is influenced by the density of monomer molecules

on the surface, the concentration of the monomer MPS is varied by “diluting” the

immobilization mixture with octyl silane (Figure 3-21). Although it has a very similar

reactivity towards silica surfaces the latter trialkoxy silane cannot be incorporated into a

polymer.

Page 65: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 51

Figure 3-21. Co-immobilization of MPS and OS (a) to control double bond density on the silica gel

surface (b) and subsequent polymerization (c).

The overall concentration of the two silanes was the same for all immobilization solutions.

Only the ratio between MPS (monomer unit) and OS (aliphatic unit) was varied. Elemental

analysis of the dried LiChrospher samples after the immobilization process shows similar

levels of surface coverage with the two silanes, independent of the starting composition. The

samples had between 0.65 and 0.75 mmol silanes per gram of silica gel. The overall graft

density for all samples was around 1.0 µmol·m-2 ( Figure 3-22). Only for the immobilization

of pure OS a slightly lesser density of 0.8 µmol·m-2 is observed.

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0

1.2

graf

t den

sity

(µm

ol/m

²)

molar part MPS on surface

Figure 3-22. Comparison of overall graft densities (MPS+OS) from the silane co-immobilization with

the effective contingent of the polymerizable MPS after analysis and calculation of elemental analysis

and FT-IR results. The circles are overall and the triangles particular MPS graft densities.

Page 66: through Surface-attached Monomers

52 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

The infrared spectra in Figure 3-23 show clearly a change in transmittance bands with

changing surface composition of the two silanes, which are the result of the different

compositions of the immobilization solutions. The difference in the intensity of the C=O-

vibration band is huge, whereas the band of the C-H2-vibration reveals less but significant

changes. The intensity of carbonyl vibration bands increases with the MPS content on the

silica gel surface and the intensity increase of methylene vibration bands proves a higher

amount of OS on the surface.

4000 3000 2000 1000

100% OS

5% MPS, 95% OS

20% MPS, 80% OS

50% MPS, 50% OS

80% MPS, 20% OS

100% MPS

tran

smitt

ance

(a.u

.)

wavenumber (cm-1)

Figure 3-23. DRIFT spectra of LiChrospher with different mixtures of immobilized MPS and OS on

the surface

Of these two characteristic IR-bands the extinction ratio was calculated as the quotient of

the appropriate integrals ( Figure 3-24). The calibration curve results as the same calculation

was done for the liquid silane mixtures consisting of known volumes of MPS and OS

( Figure 3-24).

Page 67: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 53

0.0 0.2 0.4 0.6 0.8 1.005

101520253035

extin

ctio

n ra

tio (

ν(C

H2)

/ ν(C

=O) )

partial volume MPS0.0 0.2 0.4 0.6 0.8 1.005

101520253035

Figure 3-24. A quantitative IR-analysis of the surface-attached silane-mixture compared to the

calibration curve calculated by measured extinction coefficients with exponential fit functions

A comparison of the DRIFT results of the modified LiChrospher with the calibration curve

shows that MPS is at higher concentration at the surface than what to expect from the ratio in

the initial immobilization solutions. This shift is depicted in Figure 3-25 as a co-

immobilization diagram. It shows that more MPS than OS is attached to the surface as it can

be derived from the silane ratios at the start. Now we can compare the overall surface

densities ( ) in Figure 3-22 with the effective surface density of MPS only ( ). Although the

molar amount of MPS in the immobilization solutions was varied by a factor of 20 the

effective variation of the surface density is within a factor of 10.

Page 68: through Surface-attached Monomers

54 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

0.0 0.2 0.4 0.6 0.8 1.00.0

0.2

0.4

0.6

0.8

1.0M

PS o

n su

rfac

e (m

olar

par

t)

MPS in mixture (molar part)

1.0 0.8 0.6 0.4 0.2 0.0

1.0

0.8

0.6

0.4

0.2

0.0

OS

on s

urfa

ce (m

olar

par

t)

OS in mixture (molar part)

Figure 3-25. Co-immobilization diagram of MPS vs. OS

3.3.7.2 Influence of Monomer Surface Density on PS Monolayer

Formation

The polystyrene load after polymerization shows no significant variation when the MPS

surface density on the silica gel surface changes by more than one magnitude of 10 from

1.07 µmol·m-2 down to 0.08 µmol·m-2 (Figure 3-26). The calculated polystyrene graft densities

settle constantly around 0.02 µmol·m-2 (Figure 3-27).

Page 69: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 55

0.0 0.2 0.4 0.6 0.8 1.0 1.20.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

graft density MPS (µmol/m²)

graf

ted

poly

mer

(g P

S / g

SiO

x)

Figure 3-26. Relation between amount of polystyrene and graft density of MPS; only at very low

MPS densities the polymer gets a bit less

0.0 0.2 0.4 0.6 0.8 1.0 1.20.000

0.005

0.010

0.015

0.020

0.025

graf

t den

sity

(µm

ol/m

²)

graft density MPS (µmol/m²)

Figure 3-27. Graft densities dependence between the immobilized monomer and polystyrene

Page 70: through Surface-attached Monomers

56 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

3.3.8 Polymerization in Two Steps

For this set of experiments polystyrene grafted samples come into play. The polystyrene

attached to these silica was grafted with our herein described technique. The carbon content

gets higher while the length of the immobilized polymer chains gets smaller (Figure 3-28 and

Figure 3-13). These different substrates were placed under same polymerization conditions

for short polymer chains into a second polymerization batch for 24 h (Table 3-5). In order to

get relatively short polystyrene chains a high amount of AIBN (c = 50 mmol·L-1) and a

medium monomer concentration (c = 2.9 mol·L-1) was used. It turns out that after final

elemental analysis in all cases the amount of polystyrene is almost the same and lies between

0.50 and 0.55 g polystyrene per gram LiChrospher (Figure 3-28).

Table 3-5. Molecular weights and graft densities of the two-step experiments

sample nM PS (1st)

(g/mol)

nM PS (2nd)

(g/mol)

ΓPS (1st)

(nmol/m²)

ΓPS (2nd)

(nmol/m²)

PS1 230000 13000 < 1 51

PS2 100000 13500 2 38

PS3 60000 13000 7 29

PS4 32000 14000 17 15

PS5 6500 13500 98 9

Page 71: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 57

PS1 PS2 PS3 PS4 PS50.0

0.1

0.2

0.3

0.4

0.5

0.6

previous amount new amount

graf

ted

poly

mer

(g P

S / g

SiO

x)

sample

Figure 3-28. Grafting of short chains ends up with almost the same amount of polymer regardless of

the starting amount of polymer

3.4 Grafting onto Various Porous Substrates Transferring the simple grafting technique to various silica substrates with varying specific

surfaces and surface topologies is necessary to demonstrate the versatility of the system

towards applications.

3.4.1 Influence of Substrates and Surface-attached Monomer

Silanes

Several radical polymerization reactions with styrene as monomer were carried out and

the results of the polymerization under “standard conditions” (or mostly used conditions)

are summarized in Table 3-6. There is a trend to higher polymer loading with higher specific

surface. It is shown by the graft densities in Table 3-6 that the amount of surface-bound

polystyrene is not proportional to this surface when considering all silica topologies.

Looking at the different glass beads with similar molecular weight of the PS the polymer

loading is proportional to the surface and graft densities are in the same range. High surface

silica types have a bit more polymer load due to the higher specific surfaces. But only

equipped with smaller pores the graft densities are somewhat smaller than those of the glass

bead types.

Page 72: through Surface-attached Monomers

58 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

Very obvious is also a difference in copolymerization ability between the different

monomer silanes. Initially all monomer silanes have comparable graft densities after the

immobilization step (Table 3-1). From Table 3-6 one sees that under comparable

polymerization conditions not just specific surface makes the difference but with MNPS the

polymer loading and graft density are significantly smaller. In the first copolymer

experiments MNPS-modified substrates show the same behavior resulting in low polymer

loadings and graft densities. Therefore we discard these products in the further course of this

work. Further details on polymerization on substrates with different topologies are given in

the following paragraphs (3.4.2 to 3.4.4).

Table 3-6. Resulting carbon contents and graft densities after polymerization (cstyrene = 2.9 mol·L-1,

cAIBN = 9 mmol·L-1, T = 60°C, t = 24 h) with different immobilized silanes and according molecular

weights of the “free” polystyrene.

substrate & surface

modification

carbon*

(%)

Γ0

(µmol·m-2)

nM

(g·mol-1)

LC700-ME 32.9 0.018 36,000

LC700-MA 12.0 0.002 43,900

AR300-ME 29.8 0.038 35,300

GB250-ME 28.2 0.050 30,900

GB80-ME 12.8 0.054 32,500

GB80-AE 13.7 0.060 32,700

GB80-AA† 10.7 0.025 53,400

GB40-ME† 8.8 0.031 71,700

GB40-MA† 3.7 0.006 69,100

* obtained from elemental analysis, † cAIBN = 3 mmol

Page 73: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 59

3.4.2 Influence of the Polymerization Time for Aerosil300

To get a better picture of the influence of the pore topography on the overall process of

grafting polymer monolayers on silica surfaces we used Aerosil300 in a polymerization time

dependence setup and compared it to already won data of LiChrospher. The polystyrene

loading for both silica substrates is very similar in amount and increase with time (Figure

3-29).

0 5 10 15 20 25 300.0

0.1

0.2

0.3

0.4

0.5

graf

ted

poly

mer

(g P

S / g

SiO

x)

polymerization time (h)

Figure 3-29. Comparison between LiChrospher ( ) and Aerosil300 ( ); for same periods of

polymerization time and same molecular weights the amount of grafted polymer is independent from

the topography of the silica gel

The highest polymer load increase is in the first 10 hours. Then the increase levels down to

nearly zero for a polymerization time below 30 hours. The final polymer loadings are around

0.4 g PS per gram silica.

Page 74: through Surface-attached Monomers

60 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

0 5 10 15 20 25 300.00

0.01

0.02

0.03

0.04

0.05

graf

t den

sity

(µm

ol/m

²)

polymerization time (h)

Figure 3-30. The equal polymer amount results in graft density shift because of the specific surface:

LiChrospher = 700 m²/g ( ) & Aerosil300 = 300 m²/g ( )

The specific surface of Aerosil300 is lower than that of LiChrospher what means on same

polymer loading the final graft density of 38 nmol·m-2 is higher by a factor of two (Figure

3-30). Obviously the bigger concave pore structure of the Aerosil agglomerates, built from

primary particles, is more likely to build in polystyrene than the compact LiChrospher spheres

where the significant inside part of the surface area is only accessible through micropores.

Looking on how the Aerosil silica gel responds on two parallel changes of polymerization

parameters gives an interesting picture. On the one hand less initiator in solution leads to

longer polymer chains and the final polymer amount is reduced. On the other hand more

monomer in solution creates larger polymer coils and increases the polymer load analog to

the LiChrospher experiments. Both parameter changes combined then result in the same

grafted amount of polystyrene (Figure 3-31).

Page 75: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 61

0 5 10 15 20 25 300.0

0.1

0.2

0.3

0.4

0.5

graf

ted

poly

mer

(g P

S / g

SiO

x)

polymerization time (h)

Figure 3-31. Polystyrene deposition with time on Aerosil300 with two different molecular weights

( : 35000 g/mol, : 95000 g/mol) results in same amounts for each molecular weight

Because the molecular weight is driven to higher values by both parameter changes a

decrease of graft density is the result (Figure 3-32).

0 5 10 15 20 25 300.00

0.01

0.02

0.03

0.04

0.05

graf

t den

sity

(µm

ol/m

²)

polymerization time (h)

Figure 3-32. The graft densities of polystyrene differ due to two different molecular weights

( : 35000 g/mol, : 95000 g/mol) on Aerosil300

Page 76: through Surface-attached Monomers

62 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

3.4.3 Influence of the Polymerization Time for Differently

Modified Glass Beads

The main aspect of the polymerization time experiments on glass beads with differently

modified surfaces is to find out if the mesoporous glass beads show a different behavior

from the already used silica gel systems. Having a look at the polystyrene load (Figure 3-33)

and at the resulting graft densities (Figure 3-34) similarities with the high specific surface

substrates LiChrospher and Aerosil300 are discovered: There is an increase in loading over the

first 24 hours of polymerization up to 0.11 g PS per gram glass beads and a decreasing

deposition rate.

0 5 10 15 20 25

0.02

0.04

0.06

0.08

0.10

0.12

0.14

graf

ted

poly

mer

(g P

S / g

SiO

x)

polymerization time (h)

Figure 3-33. Increase of the polystyrene amounts with polymerization time for glass beads modified

with different silanes (GB80-ME , GB80-AE , GB80-AA ). During elapsing time for every

silane monomer on the GB80 glass beads the mass of grafted polymer is nearly the same.

Page 77: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 63

0 5 10 15 20 250.00

0.02

0.04

0.06

graf

t den

sity

(µm

ol/m

2 )

polymerization time (h)

Figure 3-34. Increase of the polystyrene graft densities with polymerization time for glass beads

modified with different silanes (GB80-ME , GB80-AE , GB80-AA ). Graft densities have a

similar increase with time but differ in the total values because molecular weights are influenced by

polymerization parameters (here initiator concentration) whereas the overall process is not.

The average values for Mn are 56000 g·mol-1 for GB80-AA-PS and 49500 g·mol-1 for GB80-

AE-PS as polymerizations carried out with 3 mM initiator and 26000 g·mol-1 for GB80-ME-PS

as polymerization executed with 9 mM initiator. The molecular weight stays at the same

level during the whole polymerization time, so the development of the graft densities with

time is similar for all polymerization sets and depends only on the chosen initiator

concentration. A closer look on the influence of AIBN concentration is depicted in the

following paragraph.

3.4.4 Influence of Initiator Concentration for Different Substrates The strong influence of the initiator concentration on the molecular weight of “free”

polymer is for all substrates similar (Figure 3-36) because the polymerization takes mainly

place in solution. Because the molecular weight values vary over a larger scale than the

specific surface areas of the substrates the calculation of the graft densities are mainly

influenced by them. So the graft densities exhibit an indifferent behavior regarding the

different substrates with the main trend of getting higher graft densities with more initiator

in polymerization solution (Figure 3-37).

Page 78: through Surface-attached Monomers

64 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

1E-4 1E-3 0.01 0.10.0

0.1

0.2

0.3

0.4

0.5

c(I) (mol/L)

graf

ted

poly

mer

(g P

S / g

SiO

x)

Figure 3-35. Polymer loading for different silica solids at varied initiator concentrations

(LC700-ME , AR300-ME , GB250-ME , GB80-ME , GB80-AE ).

1E-4 1E-3 0.01 0.10

50

100

150

200

250

300

Mn (

103 g

/mol

)

c(I) (mol/L)

Figure 3-36. Molecular weights for different silica solids at varied initiator concentration

(LC700-ME , AR300-ME , GB250-ME , GB80-ME , GB80-AE ).

Page 79: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 65

1E-4 1E-3 0.01 0.10.00

0.05

0.10

0.15

0.20

0.25

graf

t den

sity

(µm

ol/m

2 )

c(I) (mol/L)

Figure 3-37 Graft densities for different silica solids at varied initiator concentration (LC700-ME ,

AR300-ME , GB250-ME , GB80-ME , GB80-AE ). The development with increasing

initiator concentration of the molecular weight as major factor and the graft densities is similar for all

different silica solids. A significant influence of the substrate topography on the result is recognized

for the final polystyrene loading.

Interesting is a detailed picture of the polymer load on the different substrates for different

initiator concentrations as the volume ratio of styrene/toluene is 1/2 (Figure 3-35). Reverence

is the LC700-ME-PS system where the polymer loading varies strongly with initiator

concentration/molecular weight between 0.02 g and 0.45 g polystyrene per gram substrate.

The value for the AR300-ME silica gel generated under same monomer concentration is

equal to the corresponding LiChrospher polystyrene loading although the specific surface is

more than twice lower. The GB250-ME material with less specific surface than the AR300-ME

achieves under same conditions even a higher polymer loading than the two silica gel types.

Interestingly the increase of polystyrene load between the two chosen initiator

concentrations is much smaller than observed for LiChrospher. GB80-ME and GB80-AE

systems exhibit a relatively small increase in PS load with higher initiator concentration in

the polymerization solution as well. Here the absolute value is smaller due to the less specific

surface but in the case of 0.1 mM AIBN the polystyrene load on LC700-ME drops even below

the value for these glass beads. The mesoporous structure of the glass beads allows polymer

Page 80: through Surface-attached Monomers

66 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

coils of all sizes grown to access the whole surface. The small micropores of the used silica

gels are more easily clogged by formation of a surface-attached polymer layer than the

mesopores of the glass beads. Therefore the surface areas of the particle core are more

effectively blocked which reduces the overall number of growing radical chains diffusing in

to free surface sites.

3.4.5 Influence of the Monomer Concentration for Glass Beads

Three experiments with varying styrene concentrations in the polymerization solution

with GB80 glass beads complement the picture we already got from the experiments with

LiChrospher Si60. Variation of the PS loading by variation of the styrene concentration in the

polymerization process is also possible when glass beads are concerned. The higher the

initial styrene concentration the more PS will be bound covalently to the glass bead surface.

Interesting is that after polymerization in bulk styrene the PS loading of more than 0.5 g PS

per gram substrate is as high as it is on LiChrospher Si60 the case (Figure 3-38).

0 1 2 3 4 5 6 7 8 90.0

0.1

0.2

0.3

0.4

0.5

0.6(1/0)

(1/1)

(1/2)

graf

ted

poly

mer

(g P

S / g

SiO

x)

c(M) (mol/L)

Figure 3-38. Amount of grafted polystyrene on GB80-ME glass beads after polymerization with

different initial styrene concentrations in toluene; values in brackets: styrene/toluene (v/v).

The number average and the weight average of the molecular weight distribution follow

the expected route: Higher styrene concentrations lead to significantly higher molecular

weights of the polymer (Figure 3-39).

Page 81: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 67

0 1 2 3 4 5 6 7 8 90

100

200

300

400

500

600

(1/0)

(1/1)

(1/2)

mol

ecul

ar w

eigh

t (10

3 g/m

ol)

c(M) (mol/L)

Mn Mw

Figure 3-39. Averaged molecular weights for PS after polymerization with different monomer

concentrations.

The increase in molecular weight is more pronounced than the increase of the polymer

load with higher initial styrene concentrations in the polymerization solution. Therefore the

graft densities are apparently lower for styrene concentrations above 4 mol·L-1 (Figure 3-40).

We observe a similar tendency in monomer concentration influence for glass beads like we

do for silica gel. Graft density on LiChrospher drops from 6.6 nmol·m-2 down to 1.8 nmol·m-2.

The smaller graft density decrease factor for glass beads (~2) and if compared to that of

LiChrospher (~3) maybe a hint for the theory where smaller pores are more efficiently blocked

by the surface-attached polymer layer than mesopores.

Page 82: through Surface-attached Monomers

68 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

0 1 2 3 4 5 6 7 8 90.00

0.01

0.02

0.03

0.04

0.05

0.06

(1/0)(1/1)

(1/2)

graf

t den

sity

(µm

ol/m

2 )

c(M) (mol/L)

Figure 3-40. Graft densities of PS on glass beads (GB80-ME) after polymerization with varied

monomer concentrations.

3.4.6 Influence of the Glass Bead Concentration

When the practical limits of a technique are explored it is important to see if it is applicable

where a maximum output is needed. The overall amount of modified glass beads for a given

set of polymerization conditions (styrene/toluene 1/2 v/v, 12 mL, cAIBN = 3 mmol·L-1, T = 60°C,

t = 24 h) was varied within a region of 0.6 g to 4.8 g, i.e. by a factor of 8. For the highest solid

concentration used in this set the glass beads are barely wetted by the polymerization

solution and therefore a practical limit for a batch reaction is reached. Characterization of the

resulting solids (GB80-ME-PS) in terms of the grafted polystyrene load and graft densities

leads to the plotted graphs shown in Figure 3-41 and Figure 3-42.

Page 83: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 69

0 100 200 300 400 5000.000.020.040.060.080.100.120.140.16

glass bead concentration (g/L)

graf

ted

poly

mer

(g P

S / g

SiO

x)

Figure 3-41. Variation of the concentration of the modified glass beads (GB80-ME) during a grafting

polymerization reaction does not effect the grafted amount as determined by elemental analysis.

0 100 200 300 400 500

0.00

0.01

0.02

0.03

0.04

0.05

glass bead concentration (g/L)

graf

t den

sity

(µm

ol/m

2 )

Figure 3-42. Variation of the concentration of the modified glass beads (GB80-ME) during a grafting

polymerization reaction does not effect the graft densities.

The amount of glass beads does not influence the deposition of polystyrene and graft

densities of 12±1 nmol·m-2 are achieved. Parallel the forming of “free” polymer was

Page 84: through Surface-attached Monomers

70 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

dramatically reduced using higher glass bead concentrations. When 0.6 g GB80-ME glass

beads were used in 12 mL polymerization solution (styrene/toluene, 1/2, v/v) about 600 mg

of free PS had been separated after polymerization by precipitation of the solution in

methanol. For the case where 4.8 g GB80-ME glass beads were used in the same amount of

polymerization solution about 150 mg free PS had been separated.

3.5 Discussion The surface attachment of the polymers is based on a copolymerization of surface attached

monomers with free monomers in solution. As the number of monomer units at the surface

is almost negligible in comparison to the number of monomers in solution, the majority of all

chains formed during the process are simply “free” homopolymers. However, the described

experiments show that the polymers can be readily separated from the modified silica gel by

simple washing procedures with suitable solvents. Furthermore, the fraction of free polymer,

as compared to the amount of polymer on the substrate, can be reduced significantly as it is

possible to run the polymerization reactions at fairly high solid concentration and thus limits

volume for generating free polymer without influencing the results of the polymerization,

i.e. the graft density and the molecular weight of the polymers on the surface. For substrates

with lower specific surface and therefore less surface-attached monomer units the same

behavior is expected and confirmed by the results of the experiments with mesoporous glass

beads (GB80) where the polystyrene loads and the graft densities stay at the same level for

all substrate concentrations. So the most important factor for use in advanced applications, a

good cover with a good polymer loading, applies for silica gel with small particle sizes as

well as for glass beads with particle sizes of several hundreds of microns.

Elemental analysis of all polymer modified silica gels and glass beads revealed that

relatively large amounts of polymer can be grafted to the silica gel for most monomer moiety

carrying silanes. For the studied polystyrene systems materials, silica gel as well as glass

beads, can be obtained with an overall polymer content of around 30 weight-% for

LiChrospher Si60. This amount is remarkable especially when it is considered that the

densities of the two components are rather different so that the material itself consists of

about equal volumes of organic and inorganic components. Based on the assumption that the

whole surface (~ 700 m2·g-1) is covered by a PS layer with constant thickness the calculated

theoretical layer thickness is ultrathin with around 1 nm. Because experiments with MPS-

modified silicon wafers show under comparable polymerization conditions layer thicknesses

Page 85: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 71

up to 15 nm it is obvious that not the whole surface area of the silica is covered by a PS layer.

Especially surface areas in the small pores of the inner core of the silica particles may not be

reached because pores in the shell got clogged with polymer during the polymerization

process.

Reference experiments in which polymerizations were carried out in the presence of

unmodified silica or silica carrying an inert alkyl silane demonstrate that the reaction indeed

proceeds according to the scheme shown in Figure 1-2c. The silica gels recovered after these

polymerizations remained largely unaltered and within the experimental error no polymer

could be found on the surfaces after completion of the polymerization reactions.

It is shown that suitable silane monomer molecules can be delivered and attached to the

different surfaces we offered in a more than sufficient density. The graft densities of the

monomer units are not influenced by different surface topologies (microporous,

mesoporous) and are settled at 2±1 µmol·m-2 which is at least higher by a factor of 40 than

the expected polystyrene graft densities on the monomer modified silica substrates.

The results show that a fairly wide variety of different acrylate and methacrylate groups

can be successfully tethered to inorganic surfaces and used for the attachment of solution-

borne macroradicals. The stability of the resulting materials in acidic or basic environments

can be improved if the more stable amide derivatives are used instead of the esters. On the

other hand, the esters of these silanes are more readily (i.e. commercially) available which

might be interesting and advantageous for the use of such hybrid materials in less critical

applications.

A somewhat special case is the surface grafted monomer N-methacryloyl-N-methyl-propyl

trimethoxysilane (MNPS) which yields much lower grafted amounts than the other

monomers. No effort was made to elucidate this behavior in detail, but it is rather well

known that methacrylamides show only low propagation rates in free radical

polymerizations[181] and yield only low molecular weight polymer under the reaction

conditions employed here. This case, however, proofs that it is necessary to immobilize

molecules with decent polymerization properties for the described process and that the sole

presence of a double bond is not sufficient.

In various experiments it is shown that the diversity of different polymers grown by free

radical polymerization based on a huge variety of available monomers can be expanded to

the technique where surface attached monomers are included. As it is exemplarily shown for

Page 86: through Surface-attached Monomers

72 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

styrene the parameters and conditions of the radical polymerization can be tuned towards

the desired loading. For further monomers in radical polymerization different

polymerization behavior has to be taken into account influencing the details of the described

general. To get a detailed picture of how the response of the system (polymer load or layer

thickness) towards the polymerization conditions is, every group of chemically similar

monomers has to be observed separately.

In a number of experiments in which selected important polymerization parameters were

systematically varied it was found that the technique of grafting polymers via

copolymerization of free and surface-attached monomers is a well-controllable process. The

amount of grafted polymer, its molecular weight and the density of the grafted chains on the

surface can be controlled over a reasonable range by adjusting proper polymerization

conditions.

Varying the polymerization time, for example, the amount of polymer which is grafted on

highly porous LiChrospher Si60 can be adjusted in a broad range up to 0.5 g PS/g SiO2 in a

fairly simple way. As it is expected for this free radical polymerizations a variation of the

polymerization time does not influence the molecular weight of the generated polymers.

According the grafted amount of polymer increases at first and then slowly levels off. Using

Aerosil300 as highly porous silica gel, built from sintered smaller primary particles, shows a

behavior similar to LiChrospher on time dependence but the concave formed, larger pores

compensate the smaller specific surface and lead to same polymer loads for both silica gels.

As on the silica gels the amount of polymer that is grafted onto the glass beads can be

adjusted in a fairly simple way via polymerization time up to an amount of 0.5 g PS/g SiO2.

The time kinetics on glass beads reveal also a slow increase of deposited polystyrene

although the larger mesopores do not limit diffusion of growing chains into them. The

practical results support the theoretical calculations made by Chakrabarti et al., Binder et al.

and Leibler et al. [150, 151, 154] where we first have diffusion controlled regime and a

polymer monolayer forms on the surface and when the surface is covered by polymer the

approaching growing chains have to overcome this polymer barrier in order to connect to

one of the surface-attached monomers and be irreversibly bound into the polymer layer.

Having a look at the polymer deposition as a function of time (Figure 3-43) we see that in a

relatively short period of time the grafted amount of polymer increases to a value that is the

lion’s share of the overall grafted polymer amount (green area). This is the time range where

the growing chains can reach quite easy the surface with the monomer sites. Then there is a

Page 87: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 73

longer time period where this increase levels of so that we get nearly a plateau for the

amount of grafted polymer (orange area). Here we get less polymer grafted to the surface

because the polymer monolayer as steric barrier makes it more difficult for growing chains to

get through to the surface-attached monomers.

Figure 3-43. Exemplary time-polymer deposition curve for a radical polymerization in presence of

silica with surface-attached monomers.

Variation of monomer or initiator concentration in the polymerization solution shows that

by these two important polymerization parameters the polystyrene loading on LiChrospher

can be controlled as well as the graft density. The grafting process is influenced by the size of

the polymer coils that approach the surface and by the probability of an growing chain

linking to a surface monomer. In the case of monomer concentration these two factors act as

antagonists because with higher monomer concentrations in solution the probability of

attaching grows as well as the size of the polymer chains that than are able to block a larger

area on the substrate. The resulting cross interaction leads to higher loading with polymer

while the graft density decreases with higher monomer concentrations during

polymerization. For the variation of initiator concentration the attaching probability and the

coil size work hand in hand because with higher initiator concentrations in solution the

probability of getting connected to the surface increases because of a higher number of

growing chains as well as the polymer coils become smaller. The molecular weight is linear

to the reciprocal square root of the initiator concentration as to expect from a radical

Page 88: through Surface-attached Monomers

74 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

polymerization in solution. Both factors combined lead then to an increase in polymer load

on higher initiator concentrations and to a more pronounced increase of the graft density.

With the Aerosil300 experiments an important insight is achieved that equal polymer

loading with different molecular weights can be achieved by choosing the right ratio

between the influence of the monomer concentration and the initiator concentration. This

result strengthens the statement that both parameters influence the grafting process from

different directions. The combination of the initiation rate, building new growing chains and

the molecular weight of the surface approaching chains is important as polymer load is

gained with a high number of chains of lower molecular weight or a fewer chains with a

higher molecular weight that soon crowd the surface of the substrate.

If the initial initiator concentration is varied, glass beads reveal the same trend as

microporous silica but the influence is less pronounced. This means that in the same

concentration range of the styrene the variation of graft density is larger for silica gel than for

glass beads. Although the glass beads show higher increase of the absolute value for the final

PS load attached to the surface due to better accessibility of the surface regions in mesopores.

The variation of initiator concentration has not such an impact on the polystyrene loading of

mesoporous glass beads as on the polymer load of silica with smaller pores. On the different

glass beads the polymer load is mainly influenced by the specific surface and insensitive

towards molecular weight affected by initiator concentration. Comparing the amount of

polystyrene on LiChrospher to those on the glass beads of varying specific surface one can

estimate the “effective surface area” of LiChrospher for certain molecular weights, e.g. about

80 m2·g-1 for a Mn around 100000 g·mol-1 referring to the value for an initiator concentration of

1 mM (Figure 3-35). Under the given conditions the modified glass beads with 20 nm pores

(GB250-ME) are the most effective substrate concerning the specific polymer load.

Varying the polymerization temperature a special behavior is found. As long as the

initiator conversion stays more or less the same (i.e. as long as the polymerization time is

adjusted according to the temperature dependence of the initiator decomposition) the

molecular weight of the resulting polymers follows closely the expectations for a free radical

process for which a decrease of the molecular weight with increasing polymerization

temperature is predicted. Interestingly, the graft density remains constant for all

temperatures investigated so that the grafted amount is decreasing with increasing

polymerization temperature solely due to a decrease of the molecular weight. Choice of the

appropriate polymerization temperature therefore offers a handle for the preparation of

Page 89: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 75

silica-polymer hybrid materials with different molecular weights of the grafted

macromolecules but constant anchor density. At very high conversions of monomer (> 60%)

and initiator (> 70%) as it is the case for the polymerizations at temperatures above 70°C in

the second set of temperature experiments we get a different picture. For these

polymerization at higher temperatures there seems to be a phase where still shorter polymer

chains are built because there is still enough initiator left. These chains can penetrate through

the surface polymer layer and bind to a surface-attached monomer even after half of the

initiator is consumed (results see above) and before the initiator diminishes so that are larger

polymer coils are formed that will not get through the surface polymer layer and bind to a

surface-attached monomer. The polydispersity of the generated free polymer increases and

GPC analysis of the free polymer reveals a more and more pronounced bimodal distribution

where a second peak with higher molecular weights superimposes. This might because of

leaving the steady state kinetics far behind and polymerizing at the end with different

initiator and monomer concentrations compared to the start of the polymerization reaction.

A central role plays an experiment, where silica gel with different surface monomer

densities is used in the graft polymerization process to elucidate the influence of this factor

on the layer formation. In detail the controlled deposition of certain monomer densities on

the surface is an important point of interest. To handle different monomer concentrations

during the silane immobilization step we chose a co-immobilization technique (Figure 3-21).

Delivering the known MPS as monomer silane in the presence of the “inactive” aliphatic OS

to the silica surface in certain molar ratios works well as it can be seen by DRIFT

measurements (Figure 3-23). The increase/decrease of the vibrational bands for each silane

follows the trend given by the initial ratio. Both silanes are of nearly same molecular weight

and size. The reactivity towards silica surfaces is identical because of the trimethoxy anchor

architecture. Taking a closer look by quantitative analysis of the IR-extinction values, i.e. the

carbonyl/methylene ratio, and comparison with extinction ratios of pure silane mixtures of

known composition (Figure 3-24), reveals that it is more likely to immobilize MPS than OS

(Figure 3-25). This “co-immobilization diagram” is of significance because the silanes were

present in double excess in the immobilization solution in order not to run out of silanes

during this attachment step. Interestingly, the following polymerization step under standard

conditions (9.3.6) is not affected by a 10fold variation of the MPS surface density (Figure 3-26

& Figure 3-27). That is plausible because it has been shown that on a silica surface where

only MPS has been immobilized only 3-5% of these double bonds are used up under given

conditions. The rest can be regarded as being “active” beneath the PS monolayer.

Page 90: through Surface-attached Monomers

76 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

The experiment where silica material grafted with polymer of different molecular weight

in different previous polymerizations is used in a second polymerization generating low

molecular weight polymer allows insights into the layer limited graft process. Using MPS-

modified silica in two subsequent polymerizations with different polymerization solutions

the effect of smaller vs. larger polymer chains comes into play. As we have shown already

the number of attached polymer chains is much smaller than the number of available

monomer units on the surface. After a first polymerization step where polystyrene

monolayers with varying molecular weights have been delivered to the modified silica gel

(see samples from the initiator variation) there are still many “active” double bonds on the

surface. It has also been shown that applying polymer coils of the same molecular weight

(> 30000 g·mol-1) in a second polymerization step ends up in a minimal increase of the final

polymer load (see polymerization time experiments: 48 h + 24 h). When relatively short

polymer chains via high initiator concentration are generated, they are able to be built inside

the existing polymer monolayer. A high number of growing radical polymer chains makes

sure that there are enough chances to connect to surface sites and “fill up” the polymer

monolayer. In the end we have nearly same polystyrene loading for all 2-step samples that

indicates a saturation with polymer on the surface.

3.6 Conclusion The strategy of grafting polymers to solid surfaces by copolymerization of surface-attached

monomers with monomers in solution is a simple and versatile strategy for the covalent

anchoring of polymers to solid substrates. Using different surface attached monomers shows

that not every double bond is suitable for this radical polymerization process and one has to

ponder which surface monomer may be well copolymerized during the polymerization. The

nature of the covalently bound polymer, from nonpolar to polar, can be varied widely giving

the substrates surface the desired physicochemical properties. Well defined polymer

monolayers can be obtained without the need to synthesize polymers with specific (end-

)groups, which allow surface attachment. It might be considered as a draw-back that besides

the polymer monolayer a significant amount of free polymer is generated that needs to be

removed from the surface. However, a variety of applications can be envisaged where this

does not pose a significant problem. Actually the circumstance that the graft process can be

driven to high material output filling the polymerization solution completely with solids

without losing in polymer loading is a big bonus. In fact, such systems are used for a number

Page 91: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 77

of industrial applications in which silanes as the ones used in this study are used as

“primers” for filler materials in acrylic formulations.

Typical features of free radical polymerization reactions are still valid for this type of

surface capturing polymerization and can be used to control layer thickness, molecular

weight and graft density of the surface attached chains to some extent. Surprisingly, despite

the simplicity of polymer monolayer formation through surface-attached monomers, such an

approach has not attracted much attention in the scientific community and only little is

known about the exact mechanism and kinetics of layer formation. In several sets of

experiments we confirmed the simplicity and variability of controlling this method. The

polymer deposition is insensitive to a wide range of surface concentrations of the previously

immobilized monomer unit. Variations in the surface density that may occur in the initial

silane immobilization step do not hinder to get “standardized” polymer monolayers. In

consequence polymer load and graft density are independent of smaller fluctuations in

quality of the silane immobilization and the polymerization step fully controls the properties

of the polymer monolayer. Beneath polymerization time and temperature monomer and

initiator concentration have been approved to control polymer load (layer thickness) and

graft densities very well in different directions because size of polymer coils and number of

growing chains are factors that compete against each other for linkage to the surface

monomers. A special feature concerning the control polymer load via polymerization should

be mentioned: As shown in Figure 3-43 a control of the polymer load and of the layer

thickness is only possible at the early stage of the polymerization with small polymer

loadings. In the later polymerization stage where the increase is very low we get an almost

stable amount of grafted polymer and respectively polymer monolayer thickness over a large

period of time. After having crossed a certain point in time, the polymerization system is

almost insensitive to the duration of the polymerization regarding polymer load or layer

thickness. This robustness might be a big advantage in processing such modified substrates.

Stable process conditions can be found, which give highly reproducable materials. A

“grafting-to” system with end-functionalized polymer chains reaches a certain maximum

plateau for the polymer load (Figure 3-44a) according to the length (“monodisperse” case) or

the molecular weight distribution of the participating polymer chains quite fast. Our system

generates throughout the polymerization new growing polymer chains. These freshly

formed, shorter chains may diffuse easily through the polymer monolayer to attach to the

surface-bound monomer increasing the overall polymer load. This chance gets smaller when

the segment density of the polymer monolayer becomes higher with time and number of

Page 92: through Surface-attached Monomers

78 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

attached polymer chains but there might always be a newly formed polymer chain that gets

through to the surface unless all monomers are consumed. Therefore this method does not

reach a limit like the “grafting-to” process and is able to generate higher polymer loads and

thicker polymer monolayers (Figure 3-44b). Smaller polymer coils are able to penetrate the

existing polymer monolayer barrier to catch the immobilized monomers behind. These

“filled up” polymer monolayers have a discrete loading value that depends on the specific

surface and topography of the solid substrate and the whole method reminds of a “grafting-

through” process. This discussion, however, is directed towards rather subtle effects. For

practical polymerization reactions, the key feature is that a nearly constant polymer loading

is obtained rather easy for longer polymerization times although at any point in time of the

polymerization growing polymer chains that are small enough can diffuse through the

surface-attached polymer monolayer and increase the total amount of polymer that is

covalently bound to the surface.

Figure 3-44. The amount of grafted polymer on a surface as a function of time. For a genuine

“grafting-to” system with end-functionalized polymer chains the load reaches a maximum with a

stable plateau (a). For the radical polymerization system using surface-attached monomers the load

increase levels off but does not reach a maximum plateau (b).

The used grafting method works in a good controllable manner for a great variety of silica

substrates which differ in size, specific surface and surface topography. This underlines the

great versatility of this technique towards a wide field of applications where solids with a

special surface finish are involved. Based on our results and knowing the specifications of a

Page 93: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 79

silica substrate one is enabled to estimate and forecast the polymer loading when using this

grafting process.

After gaining these insights of a standard polymerization system like styrene the

knowledge has to be expanded to a counter player in polarity. With

N,N-dimethyl acrylamide as monomer we get hydrophilic polymers that may act as a

helpful base polymer in applications where swellability in polar fluids is a desired property.

The investigations on grafting this polar polymer using our approach are described in the

following chapter.

3.7 Grafting of Poly(N,N-dimethyl acrylamide) onto Silica

Surfaces through Surface-attached Monomers The set of model polymerizations with different monomers to generate homopolymer

monolayers on silica has shown that the polymerization properties of the monomers have a

strong influence on the build-up of the polymer layer even though the polymerization

conditions were the same. It is therefore necessary to investigate the relation between

polymerization conditions and obtained surface-attached monolayer for each monomer in

detail. Because the polystyrene based system so far investigated is basically hydrophobic, we

now focus on the poly(N,N-dimethyl acrylamide) system to have a hydrophilic complement.

3.7.1 Polymerization Time Influence

When the influence of the polymerization time on a GB250-ME-P(DMAA) system was

studied it was observed that already after 30 minutes the amount of grafted P(DMAA)

almost reached the final value of 0.4 g polymer per gram substrate (Figure 3-45). No further

P(DMAA) was covalently bound to the substrate during the following 24 hours.

Page 94: through Surface-attached Monomers

80 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

0 4 8 12 16 20 240.0

0.1

0.2

0.3

0.4

0.5

graf

ted

PDM

AA

(g/g

SiO

x)

polymerization time (h)

Figure 3-45. The amount of covalently bound P(DMAA) for different polymerization times.

The results of GPC measurements of the polymers obtained after polymerization times

between 0.5 and 24 hours show almost constant molecular weights (Figure 3-46). The

apparent slight increase in molecular weight at short polymerization times might be due to

thermal equilibration. The number average molecular weight is around 25000 g·mol-1 and the

weight average around 110000 g·mol-1. The system P(DMAA) in DMF gels very fast and all

reaction mixtures got very viscous with time at an initial DMAA concentration of 3.2 mol·L-1.

0 4 8 12 16 20 240

20

40

60

80

100

120

Mw Mn

mol

ecul

ar w

eigh

t (10

3 g/m

ol)

polymerization time (h)

Figure 3-46. The molecular weight averages for different polymerization times as obtained from GPC measurements.

Page 95: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 81

Shorter polymerization times than 30 minutes were not realized because on the one hand

thermal equilibration takes some time as we already see some effect on the molecular weight

of the P(DMAA) that increases the shorter the polymerization time gets. The average

temperature is lower for shorter polymerization times when heating up from ambient

temperature. On the other hand there it is unpractical to control the amount of deposited

polymer via polymerization time because handling is too difficult when polymerization

times are that short. The calculated graft densities are around 0.075 µmol·m-2 (Figure 3-47).

For polymerization times below 2 hours the graft densities are a bit lower because of higher

molecular weights of the P(DMAA) caused by insufficient thermal equilibration.

0 4 8 12 16 20 240.00

0.02

0.04

0.06

0.08

0.10

graf

t den

sity

(µm

ol/m

2 )

polymerization time (h)

Figure 3-47. The graft density of P(DMAA) on GB250-ME for different polymerization times.

Polymerization time experiments were not carried out in detail with the other substrates

but the same independence is true for glass beads with a lower specific surface (GB80) and

for highly porous silica gel (LC700) as a number of tests measurements have shown. The

immediate layer buildup is not influenced by differences in specific surface and porosity.

These properties themselves have an influence on the final amount of grafted polymer as it

has been shown and discussed in chapter 3.4.

Page 96: through Surface-attached Monomers

82 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

3.7.2 Variation of Initiator Concentration

Detailed experiments with different initiator concentrations in the polymerization

solutions were undertaken with GB80 type glass beads and LC700 silica particles.

Surprisingly neither the drastic change in initiator concentration over four decades nor the

change in molecular weight have an influence on the P(DMAA) loading. As it can be seen in

Figure 3-48 and Figure 3-49 the amount of grafted P(DMAA) is the same for the chosen

initiator concentration variation range for each substrate regardless of the surface topology.

1E-4 1E-3 0.01 0.10.00

0.05

0.10

0.15

0.20

graf

ted

PDM

AA (g

/g S

iOx)

c(I) (mol/L)

Figure 3-48. The amount of grafted P(DMAA) on GB80-AE after polymerization with different

initiator concentrations.

Page 97: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 83

1E-4 1E-3 0.01 0.10.0

0.1

0.2

0.3

0.4

0.5

graf

ted

PDM

AA (g

/g S

iOx)

c(I) (mol/L)

Figure 3-49. The amount of grafted P(DMAA) on LC700-ME after polymerization with different

initiator concentrations.

As shown by Figure 3-50 and Figure 3-51 the molecular averages within an initiator

concentration range from 0.1 mol·L-1 down to 0.1 mmol·L-1 are almost constant in both cases,

GB80-AE glass beads and LC700-ME silica. For the given initiator concentration range one

would expect a change in molecular weight by a factor of approximately 30 but the

molecular weight is not influenced by initiator concentration in a way it is the case for PS.

This result shows that transfer reactions are dominating. If the transfer reactions are strongly

dominating the polymerization the terms representing the transfer reactions in the Mayo

equation (cf. equation 3-1) get much larger than the term 0X (degree of polymerization in

absence of a solvent) that is directly dependent from the initiator concentration because that

value is proportional to the squareroot of the initiator concentration. In this case the degree

of polymerization nX and the molecular weight respectively is not affected by the initiator

concentration.

Page 98: through Surface-attached Monomers

84 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

1E-4 1E-3 0.01 0.10

102030405060708090

Mw Mn

mol

ecul

ar w

eigh

t (10

3 g/m

ol)

c(I) (mol/L)

Figure 3-50. The molecular weight averages of P(DMAA) from polymerizations on glass beads

(GB80-AE) with different initiator concentrations as obtained from GPC measurements.

1E-4 1E-3 0.01 0.10

10

20

30

40

50

60

70 Mw Mn

mol

ecul

ar w

eigh

t (10

3 g/m

ol)

c(I) (mol/L)

Figure 3-51. The molecular weight averages of P(DMAA) from polymerizations on silica gel (LC700-

ME) with different initiator concentrations as obtained from GPC measurements.

Having a constant amount of grafted P(DMAA) on the substrates with varying initiator

concentrations is within the range of reasonable expectations. Now polymerizing in a

monomer-polymer-solvent system that is not, averaged over the long polymerization time,

Page 99: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 85

in a steady state puts the initiator concentration influence aside. As above-mentioned

transfer reactions are dominant and initiation reaction and termination reactions that have to

be balanced to get a steady state do not play a role. Therefore the resulting graft densities as

shown in Figure 3-52 and Figure 3-53 are constant as well.

1E-4 1E-3 0.01 0.10.00

0.02

0.04

0.06

0.08gr

aft d

ensi

ty (µ

mol

/m2 )

c(I) (mol/L)

Figure 3-52. Graft density of P(DMAA) on glass beads (GB80-AE) after polymerization with

different initiator concentrations.

1E-4 1E-3 0.01 0.10.000

0.005

0.010

0.015

0.020

graf

t den

sity

(µm

ol/m

2 )

c(I) (mol/L)

Figure 3-53. Graft density of P(DMAA) on silica gel (LC700-ME) after polymerization with different

initiator concentrations.

Page 100: through Surface-attached Monomers

86 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

Like for the polymerization time experiments a number of single experiments revealed the

initiator concentration independence for other substrates as it was observed here on the

GB80 glass beads.

3.7.3 Influence of Monomer Concentration

In the experiments with variation of the styrene monomer concentration the LC700 silica gel

was most sensitive to the changes. Therefore the highly porous substrate comes into play

again with the DMAA monomer. A set of three polymerizations was chosen to evaluate the

effect of DMAA monomer concentration. In Figure 3-54 it is shown that variation of

monomer concentration has an influence on the amount of covalently to the substrate bound

P(DMAA).

0 1 2 3 4 50.0

0.1

0.2

0.3

0.4

0.5

graf

ted

PDM

AA

(g/g

SiO

x)

c(M) (mol/L)

P(DMAA) PS

Figure 3-54. The amount of grafted P(DMAA) for the polymerization with different monomer

concentrations on silica gel (LC700-ME). The results from the experiments with PS (grey marks) are

added for comparison.

Monomer to solvent ratios of 1:9, 1:4 and 1:2 (v/v) give monomer concentrations of 1.0, 1.9

and 3.2 mol⋅L-1 DMAA. Within this range there is a gain in covalently bound P(DMAA) of

0.3 g polymer on 1 g LC700-ME which means an increase of 170 % in polymer load. This is a

significant increase of surface-bound P(DMAA) with increasing DMAA concentration in

DMF as solvent. Because of the analogy of these results with those of polystyrene the PS data

Page 101: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 87

points were added in Figure 3-54 to Figure 3-56 for easier comparison. Figure 3-54 shows

that the increase of polystyrene amount on silica levels off with higher monomer

concentrations whereas the increase of P(DMAA) amount has a tendency for an amplified

increase at high monomer concentration.

Analysis of the “free” P(DMAA) molecular weight distribution exhibits a proportional

increase of the molecular weight averages with rising DMAA concentration (Figure 3-55).

The number average of the molecular weight distribution grows from 18300 g·mol-1 for the

lowest to 59400 g·mol-1 for the highest DMAA concentration. The weight average increases

from 51600 g·mol-1 to 190700 g·mol-1. So the initial DMAA concentration has a strong impact

on the molecular weight of the formed P(DMAA).

0 1 2 3 40

50

100

150

200 Mn Mw

mol

ecul

ar w

eigh

t (10

3 g/m

ol)

c(M) (mol/L)

Figure 3-55. The molecular weight averages of the formed “free” P(DMAA) for the polymerization on

LC700-ME using different DMAA concentrations.

Again the early gelation of the polymerization system plays a major role. Where as the

movements of growing chain ends in a highly viscous solution are strongly hindered the

small monomer molecules keep their mobility until the system reaches the glassy state. This

difference arranges for the fact that the monomer concentration influences the molecular

weight as well as the amount of grafted P(DMAA) for our polymerization system and the

initiator concentration does not. A glassy state is not reached because the solvent portion

keeps the mixture at least gelatinous. The calculated graft densities for P(DMAA) on silica

Page 102: through Surface-attached Monomers

88 3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers

gel are almost constant because the amount of grafted P(DMAA) increases nearly

proportional with the molecular weights. Only a slight decrease in graft density with longer

polymer chains for higher DMAA concentration is observed (Figure 3-56). This effect might

be due to the fact that the calculated theoretical P(DMAA) layer thicknesses on LC700-ME

are 0.5 nm and below and these thin layers do not repel larger P(DMAA) coils that

effectively. Then the monomer concentration influences the molecular weight but the impact

on graft density is small.

0 1 2 3 40.000

0.002

0.004

0.006

0.008

0.010

0.012

0.014

graf

t den

sity

(µm

ol/m

2 )

c(M) (mol/L)

Figure 3-56. The graft densities of P(DMAA) on LC700-ME are constant for different DMAA

concentrations.

3.7.4 Discussion & Conclusion for P(DMAA) on Silica

If the polymerization of N,N’-dimethyl acrylamide is compared using the same system

with styrene there are some differences to mention. The first difference is that P(DMAA)

reaches its final amount of covalently bound polymer on porous substrates very fast whereas

PS needs more than 24 hours to reach a somewhat constant polymer loading. This is due to

much higher propagation constants (~50x) for DMAA than for styrene under same

polymerization conditions with comparable termination and transfer constants[182-184]. The

result is consistent with the concept of fast growing polymer chains in the vicinity of the

surface catching quickly surface-bound monomer units. Once this nearly complete P(DMAA)

monolayer is formed on the silica surface the approach of further growing P(DMAA) coils is

strongly hindered. This lack of mobility is even intensified by increasing viscosity and finally

Page 103: through Surface-attached Monomers

3 Grafting of Polymers onto Silica Surfaces through Surface-attached Monomers 89

gelation of the P(DMAA) solution in the case of higher initial monomer concentrations

(> 1 mol·L-1). Only mobile monomer molecules can diffuse through and add to growing chain

ends whereas the radicals at the growing chain ends are encaged. This is the reason why the

main influence factor on polymer loading and graft density is the monomer concentration.

For P(DMAA) in our polymerization system the molecular weight variation is small for

varying initiator concentrations. That is because of the high amount (~ 90%) of N,N’-

dimethyl formamide controlling the polymerization. This high dilution of the polymerization

mixture with solvent was necessary to keep it fluid but this measure did not prevent the

Trommsdorf effect. This gel effect was even more pronounced for the higher DMAA

concentrations (1.9 mol·L-1 and 3.2 mol·L-1) but because of the above mentioned monomer

mobility a higher monomer concentration contributes to higher molecular weights. Sufficient

monomer is present close to the substrate surface[150] and so polymer chains with low

mobility can be connected to the surface by monomers with higher mobility. Overall the

P(DMAA) load is almost proportional to the molecular weight of the generated “free”

P(DMAA) and the graft density dependency from the monomer concentration is much less

pronounced compared to the PS system.

The possibility of growing polymer monolayers of different polarities onto glass substrates

leads us to tailor-made copolymer systems where different “basic polymers” and functional

copolymers are covalently attached to silica substrates to perform special jobs under

individual conditions.

Page 104: through Surface-attached Monomers

90 4 Grafting of Functional Copolymers onto Silica Surfaces

4 Grafting of Functional Copolymers onto Silica Surfaces

In the previous chapters it is shown that surface attached monomers enable one to create

covalently bound polymer monolayers of varying polarity to the surface of silica and glass

beads. By controlling the polymerization parameters the polymer loading and graft density

can be adjusted as the previous investigations in this work reveal. With polystyrene and

poly(N,N-dimethyl acrylamide) as protagonists the foundations are laid to expand our

system to radical copolymerization during copolymerization with surface attached

monomers. Combining in solution the “basic” monomers with highly sophisticated

monomers that are equipped with special functional groups copolymer monolayers can be

built. The desired monomer combination is used in simple radical polymerization in

presence of the monomer modified, solid substrates in order to create functionalized

copolymer monolayers. The functionalities within the copolymer layer can then be used for

further polymeranalog reactions towards the needed system for specific, custom-built

applications.

For the copolymerization styrene as hydrophobic and N,N’-dimethyl acrylamide as

hydrophilic monomer are used (Figure 4-1 1 & 2). As functional comonomers we use tailor-

made molecules developed and polymerized exemplarily by Murata et al.[185, 186]. Through

a two step synthesis the N-methacryloyl-β-alanine succinimide monomer (MAC2ae, Figure

4-1 3) is accessible. First β-alanine is reacted with methacryloyl chloride and then N-hydroxy

succinimide is coupled to the carboxylic moiety[186]. MAC2ae is a monomer bearing an

activated ester which is reactive towards amino groups under appropriate conditions. The

N-acryloyl-N-methyl-propyl phthalimide (AC3pht, Figure 4-1 4) unit is synthesized in a two

step process starting from N-methyl-1,3-diaminopropane reacting with phthalic anhydride

in the first and with acryloyl chloride in the second step[185]. The AC3pht monomer contains

a protected amino group which is deprotected by treatment with methylamine or hydrazine.

This pathway is necessary because (meth)acrylic compounds with amino moieties are

difficult to polymerize because they undergo Michael addition during polymerization[187].

This unwanted reaction type leads to loss of the amino functionalities. But creating primary

amino groups within a polymer this way gives us a copolymer reactive towards a broad

range of reagents diversifying the possibilities of our surface-modified materials. Forming

copolymer monolayers via radical copolymerization with this set of monomers we have a

Page 105: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 91

starting point for a number of subsequent reactions which then might be used as a toolbox

for generating functional copolymer monolayers on solid substrates.

Figure 4-1. Monomers used in polymerization of functional copolymer monolayers on silica surfaces:

Styrene 1 as main monomer for hydrophobic systems, N,N-dimethyl acrylamide 2 as main monomer

for hydrophilic systems. Two bifunctional compounds are incorporated into the polymer chains via

their (meth)acrylic site: N-methacryloyl-β-alanine succinimide ester 3 reacts as “activated ester” with

primary amines at room temperature and N-acryloyl-N-methyl-propyl phthalimide 4 is a protected

amino group.

Page 106: through Surface-attached Monomers

92 4 Grafting of Functional Copolymers onto Silica Surfaces

4.1 Polystyrene Copolymers on Silica Particles

4.1.1 Poly(styrene-co-N-acryloyl-N-methyl-propyl phthalimide) &

Poly(styrene-co-N-acryloyl-N-methyl-propyl amine)

Poly(styrene-co-N-acryloyl-N-methyl-propyl amine) (P(S-co-AC3amine)) and its precursor

poly(styrene-co-N-acryloyl-N-methyl-propyl phthalimide) (P(S-co-AC3pht)) is the copolymer

system that is polymerized the most and attached to the most different substrates including

silica gel and different glass beads in the course of our studies. It was prepared with different

compositions varying the molar part of the functional monomer unit in the copolymerization

solution from 5 % up to 20 %. Because of the relatively high number of different material

carrying surface attached P(S-co-AC3pht) the results will be listed in the following in tables.

The elemental analysis of all substrates with P(S-co-AC3pht) shows that polymer load and

graft density for different specific surfaces on the varying substrates changes like the

homopolymer PS would do under same polymerization conditions (Table 4-1). GPC

measurements in THF against PS standard samples show that the apparent molecular weight

of the copolymer coils gets larger with increasing phthalimide content. This apparent

molecular weight increase might be an artifact of the GPC measurement caused by the fact

that the calibration was carried out with a different polymer (PS) than the measured (P(S-co-

AC3pht)). 1H-NMR analysis of the free phthalimide copolymers reveals two peaks in the

region between 7.9 and 7.7 ppm caused by the four hydrogen atoms of the aromatic ring that

is connected to the imide ring to form the phthalimide moiety. The signals of the five

hydrogen atoms of the ring of the styrene moiety are localized between 7.3 and 6.3 ppm.

Because phthalimide and styrene signals are clearly separated at the NMR scale the integrals

of these peaks can be rationed to calculate the content of N-acryloyl-N-methyl-propyl

phthalimide in the copolymer. For three P(S-co-AC3pht) copolymers on glass beads the 1H-

NMR spectra are depicted in Figure 4-2 and the differences in height and area of the two

phthalimide peaks change significantly following the trend in a proportional way.

Page 107: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 93

8 7 6 5 4 3 2 1 chemical shift (ppm)

CDCl3

20'

20 19'

19

17,18

15,1613,14 5'

3

6

5

4

7,8

10 - 124'

91,2

N

NO

O

O

co

co

(4)

(3)

(2)

(1)

c)

b)

a)

Figure 4-2. 1H-NMR spectroscopy of free polystyrene copolymers with different contents of

phthalimide groups.

In the 1H-NMR spectra of the P(S-co-AC3pht) copolymers the peaks in the aromatic region

Figure 4-2(1) belong to the four phthalic protons 19, 19’, 20 and 20’ and they diminish as the

content of the functional monomer within the copolymer is reduced from 20 mol% (Figure

4-2a), over 10 mol% (Figure 4-2b), down to 5 mol% (Figure 4-2c). The peaks in the aromatic

region Figure 4-2 (2) represent the five styrene protons 4-6, 4’ and 5’. The peaks in the

aliphatic region Figure 4-2 (3) are caused by methylene and methyl protons next to the

nitrogen atoms and also exclusive for the functional unit. The backbone protons and the

middle methylene protons of the propyl spacer give the peaks of region Figure 4-2 (4).

Page 108: through Surface-attached Monomers

94 4 Grafting of Functional Copolymers onto Silica Surfaces

The results listed in Table 4-1 show that after a conversion of 30 to 40% the P(S-co-AC3pht)

copolymers have nearly the same composition of subunits as the monomers in the initial

polymerization solution. In most cases the percentage value of phthalimide units lies

minimal below the expected amount derived from the monomer mixture before

polymerization start.

Table 4-1. Elemental analysis results, polymer load and graft density of the styrene-co-N-acryloyl-N-

methyl-propyl phthalimide monolayer modified substrates. Molecular weight and content of

phthalimide units of the free copolymer.

product C*

(%)

H*

(%)

N*

(%)

mʹpoly#

(g)

nM †

(g·mol-1)

Γ0

(µmol·m-2)

ϕ‡

(%)

LC700-ME-PS-AC3pht(5%) 29.27 2.83 1.32 0.409 36,000 0.016 4.8

LC700-ME-PS-AC3pht(10%) 27.01 2.94 0.64 0.391 33,000 0.017 11.8

GB80-ME-PS-AC3pht(5%) 11.63 0.92 0.15 0.126 69,000 0.023 4.7

GB80-ME-PS-AC3pht(10%) 12.99 1.05 0.30 0.152 78,500 0.024 9.4

GB80-ME-PS-AC3pht(20%) 12.76 0.92 0.52 0.154 90,500 0.021 15.0

GB80-AA-PS-AC3pht(10%) 8.39 0.96 0.50 0.088 21,500 0.051 9.5

GB250-AA-PS-AC3pht(20%) 22.29 2.25 1.33 0.323 34,500 0.038 19.4

* obtained from elemental analysis, # per determined from elemental analysis, 2SiO

g † determined

from GPC analysis of the free polymer, ‡ calculated from 1H-NMR integrals of the free polymer

After having successfully generated covalently bound P(S-co-AC3pht) copolymer

monolayers the next step is the cleavage of the phthalimide to set the amino group free for

further reactions. 1H-NMR measurements of the free P(S-co-AC3amine) reveal that cleaving

the phthalimide with methylamine off from the amino group does not lead to complete

conversion. The two typical phthalimide peaks between 7.9 and 7.7 ppm are still present

with reduced intensity. A reaction of P(S-co-AC3pht) with hydrazine instead gives

quantitative cleavage confirmed by missing phthalimide hydrogen peaks in the 1H-NMR

spectrum (Figure 4-3 (1)) in combination with the presence of signals of methylene hydrogen

next to nitrogen atoms (3.7-2.5 ppm; H 10-14, 17, 18; Figure 4-3 (3)).

Page 109: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 95

8 7 6 5 4 3 2 1

17,18

15,1613,14

10 - 12

97,8

65'5

4 4'

31,2

chemical shift (ppm)

NH2

NO

co

co

(3)(1)CDCl3

(4)

(2)

Figure 4-3. 1H-NMR spectrum of the phthalimide functionalized copolymer P(S-co-AC3pht(10%))

after reaction with hydrazine.

Elemental analysis after deprotection of the amino group reveals the loss of the phthalic

acid unit, because the carbon content drops, the nitrogen content raises and the overall

copolymer load is reduced during this polymeranalog step.

Table 4-2. Elemental analysis results and polymer load for styrene-co-N-acryloyl-N-methyl-propyl

amine monolayer modified substrates after polymer analogue reaction of the phthalimide containing

substrates with hydrazine.

product C*

(%)

H*

(%)

N*

(%)

mʹpoly#

(g)

δ(NH2)†

(µmol/g) EA UV

δ(NH2)‡

theor. (µmol/g)

LC700-ME-PS-AC3amine(10%) 25.60 3.43 0.53 0.350 - 20 370

GB80-ME-PS-AC3amine(5%) 11.00 0.86 0.22 0.115 - - 50

GB80-ME-PS-AC3amine(10%) 11.17 0.86 0.33 0.120 - 90 120

GB80-ME-PS-AC3amine(20%) 11.84 1.14 0.66 0.139 - 90 180

GB80-AA-PS-AC3amine(10%) 7.83 1.19 0.60 0.081 80 50-80 70

GB250-AA-PS-AC3amine(20%) 19.46 2.51 1.62 0.242 150 120-160 470

* obtained from elemental analysis, # per determined from elemental analysis, 2SiO

g † determined

from elemental analysis of coupled DMT-disulfide units[188] and UV spectroscopy of cleaved DMT[189], ‡ calculated from actual load and actual functional group content of P(S-co-AC3pht)

Page 110: through Surface-attached Monomers

96 4 Grafting of Functional Copolymers onto Silica Surfaces

Coupling of a 13-bis(4-methoxyphenyl)-13-phenyl-4-oxo-8,9-dithia-5,12-dioxa-tridecanoic

acid unit (see Figure 4-4 for chemical structure) to the amino groups[190] within the

copolymer monolayer enables quantification of the thereby obtained amine functions via

elemental analysis of the sulfur of the disulfide linkage and additionally photometric

analysis after cleavage of DMT ([1,1-bis(4-methoxyphenyl)-1-phenyl]methyl group). To

ensure that there are no residual amino groups the Kaiser test, a colorimetric test for the

presence of amino groups, is done after the coupling step[191]. This test is based on the

reaction of ninhydrin with amino groups to form a blue adduct and therefore an incomplete

coupling leads to a positive Kaiser test.

NH2

OH

O

O

OS S

O

MeO

OMe

OMe

NH

O

O

OS S

O

MeO

OMe

OMe

silica particle

polymer monolayer

+

cleavage

UV analysis

connecting step:DCC, DMAP, TEA

cleaving step:HClO4 / EtOH

Figure 4-4. Detection and quantification of the obtained amino groups within the surface-attached

copolymer monolayer (PS-AC3pht).

It turns out that the photometric approach gives varying results (± 10 to 20%) on the

amounts of amino groups whereas the results of the elemental analysis are constant.

Regarding the amino loading values in Table 4-2 this has to be considered. GB80-ME-PS-

AC3pht(10%) & (20%) have been transformed into the amine using methylamine. As we

recognize for the free polymer this reaction is not quantitative and this explains the lower

Page 111: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 97

measured values for δ(NH2) on GB80-ME-PS-AC3amine(10%) & (20%) compared to the

theoretical values. In addition it has to be mentioned that with smaller pores and thicker

copolymer layers and higher loading respectively (LC700-ME-PS-AC3amine(10%) &

GB250-AA-PS-AC3amine(20%)) the gap between the theoretically possible value for δ(NH2)

and the measured actual value becomes bigger.

FT-IR measurements of the functionalized solids as well as the free copolymers allow a

characterization of the obtained surface modified materials. The copolymers with PS as main

component show the specific aromatic C-H stretching vibrations between 3100 and 3000 cm-1

(Figure 4-5). Surface-attached copolymers containing the phthalimide unit give sharp bands

at 1772 and 1715 cm-1 that are specific for C=O stretching of a five ring imide system in

conjugation to an aromatic ring. Out of plane vibration of the four aromatic hydrogen atoms

leads to a band at 722 cm-1. After phthalimide moieties being converted into amine functions

the phthalimide vibrations bands are not present anymore. The N-H stretching vibration

band of the amino groups is too weak to be recognized in the spectra, so that full conversion

of the phthalimide can only be inferred from the absence of the phthalimide bands.

4000 3000 2000 1000

(pht)

ν

b)

a)(Aryl-H)ν

ν

ν

(SiOSi)

(C=O)

(CH)

ν

trans

mitt

ance

(a.u

.)

wavenumber (cm-1)

Figure 4-5. Fourier infrared measurements of PS-copolymer covered silica particles carried out in

diffuse reflection mode (DRIFT): PS-AC3pht (a) and PS-AC3amine (b).

Page 112: through Surface-attached Monomers

98 4 Grafting of Functional Copolymers onto Silica Surfaces

1000 800 600 400 200 00

10

20

30

40

50

60

70

b)

c)

a)

O KLL

Si 2pSi 2s

C 1s

N 1s

O 1s

coun

ts (1

000/

s) (a

.u.)

binding energy (eV)

Figure 4-6. X-ray photoelectron spectra of a bare glass bead (a), a glass bead modified with a

N-acryloyl-N-methyl-propyl trimethoxysilane layer (b) and a poly(styrene-co-N-acryloyl-N-methyl-

propyl amine) covered glass bead (c).

Table 4-3. Ratios of different elements on the surface of silica particles modified by PS and PS

copolymer containing amino groups polymers derived from XPS measurements.

product C*

(%)

N*

(%)

F*

(%)

O*

(%)

Si*

(%)

GB80-AA-P(S-co-AC3amine(10%)) 38.5 4.1 - 38.9 18.5

GB80-AA-PS 45.9 - - 35.6 18.5

* obtained from XPS and the integral of the XPS signals (normalized with the efficiency factor)

The XPS measurements confirm the composition of the P(S-co-AC3amine) layer that was

attached to glass beads (Figure 4-6, Table 4-3). To test the amino functions within the surface-

attached copolymer monolayer FITC was applied to the glass beads and bound irreversibly

to functional sites as detected after intense extraction with various solvents by a CCD camera

and fluorescence microscopy (Figure 4-7). The confocal fluorescence microscopy shows a

strong fluorescence on the surface facets of the glass bead. Both visualizations suggest a

homogeneous coverage of the surface with amino groups respective copolymer monolayer.

Page 113: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 99

Figure 4-7. CCD image shows that glass beads with a P(S-co-AC3amine) monolayers get colored after

treatment with FITC (a). A slice of the examination with a confocal fluorescence microscope is shown

and exhibits that glass beads with a surface-attached P(S-co-AC3amine) monolayer show fluorescence

after treatment with FITC (b).

400 450 500 5500.0

0.1

0.2

a)

511 nm

482 nm

458 nm

433 nm

PS

P(S-co-AC3amine(10%))

P(S-co-AC3amine(5%))

P(S-co-MAC2ae)

FITC

abs

(a.u

.)

wavelength (nm)

Figure 4-8. Different free copolymers in DMF were investigated with an UV/vis spectrometer after

treatment with fluorescein isothiocyanate (FITC, see reference extinction curve). FITC can only be

recognized when covalently bound to the polymer, here: when amino groups were present. PS and

P(S-co-MAC2ae) as negative references indicate no dye bands.

Page 114: through Surface-attached Monomers

100 4 Grafting of Functional Copolymers onto Silica Surfaces

The results of the labeling experiments on copolymer modified substrates are confirmed by

reactions of different free copolymers with FITC as labeling dye. Whereas PS copolymers

equipped with amino groups is colored even after reprecipitation and washing, pure PS or

PS copolymers containing activated ester groups reveal no extinction bands of the

fluorescein after the same procedure (cf. Figure 4-8). There is a shift of the intensity of

extinction maxima of the polymer bound fluorescein to longer wavelengths compared to the

unreacted fluorescein dissolved in methanol. This is due to the missing solvent interaction of

the dye within the polymer. A color change from green to yellow to orange and red,

dependent from the fluorescein content in the copolymer, is the result.

4.1.2 Poly(styrene-co-N-methacryloyl-β-alanine succinimide ester)

Poly(styrene-co-N-methacryloyl-β-alanine succinimide ester) (P(S-co-MAC2ae)) is a

copolymer with active ester units. The conversion of N-succinimide activated ester groups

with reactive functional groups, especially primary amines, is well known and widely used

in preparations and analytics, e.g. cell labeling with dyes[192], biotinylation[193, 194],

biomolecule immobilization on polymers[195, 196], or patterned surface

functionalization[197]. IR measurements of surface-attached copolymer monolayers reveal

the specific C=O stretching vibration at 1826, 1788 and 1732 cm-1 of the succinimide (Figure

4-9).

4000 3000 2000 1000

(ae)ν (Aryl-H) νν

ν

(SiOSi)

(C=O)(CH)

ν

tran

smitt

ance

(a.u

.)

wavenumber (cm-1)

Figure 4-9. DRIFT survey of silica particles with P(S-co-MAC2ae) layer display indicative C-H- and

C=O-stretching bands.

Page 115: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 101

8 7 6 5 4 3 2 1

6

NHO

co

co

O O

NO

O16',17'

16,17

14,1512,13

9-117,8

5'5

4 4'

31,2

(1)

chemical shift (ppm)

(3)

CDCl3

(2)

Figure 4-10. 1H-NMR spectrum of P(S-co-MAC2ae) the active ester units.

In the 1H-NMR spectrum of P(S-co-MAC2ae) the methylene hydrogen located next to

carbonyl groups (14-17’) show up with peaks within region (2) in Figure 4-10. Best for the

determination of the MAC2ae content from the peak integral ratios in the “free” P(S-MAC2ae)

copolymer are the four succinimidyl hydrogen protons. In a 1H-NMR spectra they appear

between 2.9 and 2.6 ppm. MAC2ae copolymerized with styrene gives a spectrum where no

other peaks interfere with the succinimidyl proton peaks (Figure 4-10, region 2). 1H-NMR

analysis of the free copolymers generated during the polymerizations shows that a content of

2.1% MAC2ae was obtained for LC700-ME-P(S-co-MAC2ae) and 1.6% MAC2ae for GB80-ME-

P(S-co-MAC2ae). 95 % styrene and 5 % MAC2ae were used as monomers in solution. The

succinimidyl content of the analyzed copolymer is lower than expected from the initial

comonomer ratio in the polymerization solution due to insufficient solubility of the MAC2ae

constituent. LC700-ME-P(S-co-MAC2ae) has a loading of 474 mg copolymer per gram silica

and with 2.1 % active ester groups the functionality load is 93 µmol·g-1 and we have

22 µmol·g-1 of activated esters on GB80-ME-P(S-co-MAC2ae) glass beads. The activated ester

groups were tested after the polymerization procedures by binding dyes with primary

amines at room temperature. The primary amino groups of the dye react with the activated

ester groups by forming an amide group and N-hydroxysuccinimide is leaving the reaction.

Figure 4-11 shows a successfully dyed copolymer monolayer on glass beads dyed by

Page 116: through Surface-attached Monomers

102 4 Grafting of Functional Copolymers onto Silica Surfaces

DY-635-NH2 (for chemical structure see Figure 4-13). For the free P(S-co-MAC2ae) labeling

with DY-635-NH2 was successful as shown by UV spectroscopy (Figure 4-12).

Figure 4-11. Glass beads with a grafted P(S-co-MAC2ae) monolayer were labeled with blue

DY-635-amine.

500 550 600 650 7000.0

0.1

0.2

DY-635-NH2

P(S-co-MAC2ae(5%))

648 nm

663 nm

wavelength (nm)

abs

(a.u

.)

Figure 4-12. For the case of active ester functions the dye DY-635 equipped with an aliphatic primary

amine function (DY-635-NH2) was bound and found.

Page 117: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 103

N O+

NNH2

NH

O

SO3-

Figure 4-13. Chemical structure of DY-635-NH2. The primary amino group readily reacts with

activated esters at room temperature.

4.1.3 Two-step Grafting Processes

To get an impression of the impact of functionalities inside the ultrathin copolymer layer on

following polymeranalog reactions, a second generation grafting reaction were performed

with glass beads with an polystyrene-amine layer (GB80-AA-P(S-co-AC3amine(10%)), Figure

4-14 a). Two different systems were used to investigate the grafting onto the first polymer.

The first system for a second grafting step was based on acrylic moieties. The acryloyl

chloride reacted with the amino groups to introduce new double bonds on the copolymer

chain (Figure 4-14 b). The C=C vibration bands of these attached monomers were clearly

visible in DRIFT experiments. The thus obtained acrylic moieties took then part in a radical

polymerization of styrene to form PS branches on the copolymer layer (Figure 4-14 c).

Page 118: through Surface-attached Monomers

104 4 Grafting of Functional Copolymers onto Silica Surfaces

Figure 4-14. Two step polymerization on amino-polymer modified glass beads leading to surface

attached comb polymers.

For the second system 2-brom-propionic acid bromine was added to the first layer amino

groups. This allows to obtain covalently attached units that can start a “grafting-from”

reaction via atom transfer radical polymerization (ATRP) mechanism (Figure 4-14 d). In a

subsequent ATR polymerization of styrene, polystyrene chains start to grow from the

copolymer layer (Figure 4-14 e). The results obtained from elemental analysis and GPC

measurements show that both second generation polymerizations add a significant amount

of polymer-grafted polymer.

The ATRP case, as a grafting-from technique results in much higher amounts of grafted

polymer because the monomer diffusion is less restricted than the diffusion of growing

polymer coils and also the probability of a monomer hitting the grafting-from radical is

much higher than a polymer coil radical hitting a grafted double bond.

Page 119: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 105

Table 4-4. Resulting polymer amounts and graft densities before and after a second generation

polymerization with styrene on glass beads with different functionalized PS-monolayers.

modified substrate PS*, overall

(g·(g SiO2)-1)

PS*, 2nd

(g·(g SiO2)-1)

Γ0, surface

(µmol·m-2)

Γ0, polymer

(µmol·m-2)

nM 2nd

(g·mol-1)

GB80-AA-P(S-co-AC3amine(10%)) 0.082 - 0.051 - -

GB80-AA-P(S-co-AC3amine(10%)) + PS

0.086 0.004 0.052 - 32,500

GB80-AA-P(S-co-AC3AA-PS) 0.100 0.018 §0.052 0.007 33,000

GB80-AA-P(S-co-AC3PA-PS) 1.230 1.148 0.051 0.871 †16,500

* calculated from elemental analysis, † free polymer from an ATR polymerization that was started with starter molecules in solution under same conditions (cf. 10.1.4), § estimation from the blind test

Compared with the initially used functional beads (GB80-AA-P(S-co-AC3amine)) a slight

increase of surface-bound polymer can be found by just dispersing the amino-functionalized

glass beads during a radical polymerization of styrene (GB80-AA-P(S-co-AC3amine(10%)) +

PS, blind test) although the amino functionalities are not polymerizable. Due to remaining

double-bonds from the MPS-layer below the polymer layer there is a small amount of

polymer that increases the overall amount of grafted PS little compared to the

GB80-AA-P(S-co-AC3amine(10%)) before (cf. Table 4-4). This behavior is consistent with the

results shown in 3.3.8. The acrylic double-bonds along the primarily grafted PS-chain give an

additional increase in covalently bound polymer (GB80-AA-P(S-co-AC3AA-PS)) as the

additional polymer amount lies 18 mg per gram·SiO2 PS above that of the starting material

and 14 mg per gram·SiO2 PS above that of blind test (cf. Table 4-4). The best performance

gives the “grafting-from” approach with PS-chains branching out from the first polymer

layer (GB80-AA-P(S-co-AC3PA-PS)). The PS loading is with 1230 mg per gram·SiO2 PS

roughly twelve times higher than with the acrylic function within the polymer layer (cf.

Table 4-4).

Comparing the overall graft densities of the blind test and the two different approaches

one becomes aware that the blind test with 52 nmol·m-2 has an insignificantly higher graft

density than the glass beads with just the polystyrene-amine layer (51 nmol·m-2). The

polymerization that incorporates the monomer functions within the first polymer layer

Page 120: through Surface-attached Monomers

106 4 Grafting of Functional Copolymers onto Silica Surfaces

shows an additional graft density of 7 nmol·m-2 on the polymer, whereas the surface bound

portion should increase like it is the case for the blind test, because below the previously

formed monolayer are also surface-attached monomer present. Due to the smaller molecular

weight of the generated PS the ATRP approach delivers a graft density of 871 nmol·m-2 onto

the first polymer layer where the initiating sites are. The surface graft density stays constant

because in this experiment no polymer chains are started in solution that could diffuse

through the polymer monolayer and connect to the surface-attached monomers beneath.

When the density of amino functions, as determined before, is around 90 µmol·g-1 there are

1125 nmol amino groups per square meter surface. Compared with the final ATRP graft

density there is an efficiency of 77% totalized for all synthesis steps. This means

approximately three of four amino functions grew into a PS-chain whereas the rest did either

not connect to an ATRP-starter molecule, did not start during ATR polymerization or just

terminated already at lower molecular weights due to sterically hindered functionalities or

lack of space for propagation of the polymer chain.

Page 121: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 107

4.2 Poly(N,N-dimethyl acrylamide) copolymers on silica

particles

4.2.1 Poly(N,N-dimethyl acrylamide-co-N-acryloyl-N-methyl-

propyl phthalimide) & Poly(N,N-dimethyl acrylamide-co-N-

acryloyl-N-methyl-propyl amine)

The copolymers with P(DMAA) as main component on the silica solids exhibit aliphatic C-

H stretching vibrations in the range between 3100 and 2800 cm-1 and a characteristic carbonyl

band at 1642 cm-1 (Figure 4-15).

4000 3000 2000 1000

(pht)

b)

a)

ν

ν

ν

(SiOSi)(C=O)

(CH)

ν

tran

smitt

ance

(a.u

.)

wavenumber (cm-1)

Figure 4-15. DRIFT spectra of (a) P(DMAA-co-AC3pht) and (b) P(DMAA-co-AC3amine), the

elimination of phthalimid carbonyl vibration bands after conversion into an amine shows the extent of

the reaction.

In the 1H-NMR spectra of the “free” poly(N,N-dimethyl acrylamide-co-N-acryloyl-N-

methyl-propyl phthalimide) (P(DMAA-co-AC3pht)) copolymer the four phthalic hydrogen

show a broad peak between 7.8 and 7.5 ppm (Figure 4-16 (1)). The integral of this peak is

used to quantify the amount of the functional monomer in the copolymer monolayer when

compared to the integral of the methylene hydrogen in the backbone between 2.1 to 1.0 ppm

(Figure 4-16 (4)). The signals of the N-methyl hydrogen show up between 3.4 to 2.6 ppm

Page 122: through Surface-attached Monomers

108 4 Grafting of Functional Copolymers onto Silica Surfaces

(Figure 4-16 (2)) together with the peaks of the AC3pht propyl spacer between 3.7 to 2.1 ppm.

After cleavage of the phthalic group with hydrazine the 1H-NMR spectra of the resulting

poly(N,N-dimethyl acrylamide-co-N-acryloyl-N-methyl-propyl amine) (P(DMAA-co-

AC3amine)) show no phthalimide hydrogen signals. This indicates a successful

polymeranalog conversion of the phthalimide into amino groups as it was shown for the

styrene copolymer. For amino group quantification of GB80-AA-P(DMAA-co-AC3amine) the

DMT-disulfide linker[190] was used as described for the experiments in 4.1.1. As result of the

quantification from elemental analysis of sulfur we get 90 to 100 µmol amino groups per

gram glass beads.

8 7 6 5 4 3 2 1

13 - 15

23'

2322'

22

20,21

18,1916,17

1210,11

7 - 9

4 - 6

31,2

N

O NNO

O

O

co

co

(2)(1)(4)

(3)

chemical shift (ppm)

D2O

Figure 4-16. In case of 1H-NMR spectra of the phthalimide co-units the aromatic phthalic hydrogen

are located in region (1) whereas the peak ranges of the hydrogen next to the amide groups (2) and the

N-methyl hydrogen of the dimethyl acrylamide (3) are closely connected.

4.2.2 Poly(N,N-dimethyl acrylamide-co-N-methacryloyl-β-alanine

succinimide ester)

Substrates with poly(N,N-dimethyl acrylamide-co-N-methacryloyl-β-alanine succinimide

ester) (P(DMAA-co-MAC2ae)) have a hydrophilic, water swellable polymer layer combined

with a functional group that reacts well with amino groups. As amino groups are present in

many biological molecules, this opens interesting perspectives. In a copolymer with

P(DMAA) the succinimide C=O stretching gives slightly shifted signals at 1817, 1786 and

1740 cm-1 (Figure 4-17).

Page 123: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 109

4000 3000 2000 1000

(ae)

ν

ν

ν

(SiOSi)(C=O)

(CH)

ν

tran

smitt

ance

(a.u

.)

wavenumber (cm-1)

Figure 4-17. DRIFT spectrum of grafted P(DMAA-co-MAC2ae) on LC700-ME.

The content of functional groups cannot be unambiguously obtained by 1H-NMR analysis

of the “free” P(DMAA-co-MAC2ae) copolymer because the N-methyl hydrogen of P(DMAA)

give a broad and intense signal between 3.0 to 2.6 ppm that superimposes the succinimidyl

hydrogen peaks between 3.1 to2.8 ppm (Figure 4-18).

8 7 6 5 4 3 2 1

NNHO

co

co

O

O ON

O

O19',20'

19,20

17,18 15,16

12-1410,11

4 - 6

7 - 9

31,2

(1) (3)

(2)

chemical shift (ppm)

DMSO

Figure 4-18. 1H-NMR spectrum of free P(DMAA-co-MAC2ae); the results show that the functional

succinimide ester units cannot be distinguished from the dimethyl acrylamide units because the peaks

of the active ester (1) have nearly the same shift as the N-methyl proton peaks of the main subunits (2).

Page 124: through Surface-attached Monomers

110 4 Grafting of Functional Copolymers onto Silica Surfaces

Figure 4-19. Fluorescence detection (detection wavelength from 671 nm to 693 nm) on P(DMAA-co-

MAC2ae) monolayers and P(DMAA) monolayers on glass beads after treatment with DY-635-NH2.

If the fluorescence of the different samples is recorded, the difference between a reactive

polymer monolayer functionalized with activated ester groups and an inert homopolymer

monolayer becomes evident. Both monolayers on the same substrate are exposed to a

solution of DY-635-NH2. After filtering off and washing the solids the glass beads with the

P(DMAA) homopolymer layer give no fluorescence response, only some reflection from the

glass surface (Figure 4-19b). The P(DMAA-co-MAC2ae) copolymer layer, however, reacts

with the amino dye, binds it covalently (Figure 4-19a), which leads to a strong fluorescence

of the resulting material if illuminated with light of an appropriate wavelength.

Page 125: through Surface-attached Monomers

4 Grafting of Functional Copolymers onto Silica Surfaces 111

4.3 Discussion & Conclusion Copolymers As a first example, four different copolymer systems have been used for monolayer

formation on silica substrates to demonstrate the potential of this grafting system. We have

the ability to incorporate functionalities into the surface-attached polymer monolayer by

using appropriate comonomers in the polymerization process. To characterize these

copolymer systems the substrates with the surface-attached copolymer monolayer as well as

the free copolymers have been examined by various techniques. Basic monomers and

functional monomers in the copolymer layers generate specific FT-IR bands what gives an

easy handle to distinguish between all systems presented here. Elemental analysis and XPS

of the polymer loaded substrates and 1H-NMR spectroscopy of the free monomer give an

idea of the composition of the copolymers and the loading of functionalities can be

calculated. Statistical distribution of the different repeat units inside the generated

copolymers is generally recognized for styrene acrylic copolymers[198-201] and for systems

built from acrylic or methacrylic monomers[60, 202] always assuming monomers without

extraordinary steric or electronic hindrances. General accessibility and activity of the desired

functionalities within the copolymer layer have been successfully proven by different means,

qualitively as well as quantitively. The layer composition concerning the subunits can

controlled by the monomer quantities in the initial polymerization solution. When realizing

copolymer monolayers insufficient monomer solubility can be a limiting factor for the

incorporation of certain monomer species.

Somewhat delicate is the polymeranalog step converting the phthalimide into a primary

amine. With the right choice of reagent and conditions as identified with experiments with

free polymer and 1H-NMR analysis this conversion can be carried out quantitatively using

hydrazine (9.5.2). We identified that there are differences between theoretical and actual

value of the amine loading δ(NH2) for substrates with a high polymer load and small pores.

There are two possible explanations that were not explored any further: First the

deprotection of the amine group may be incomplete because the agent is not able to diffuse

into pores that got clogged with polymer. The second reason may be that because of clogged

pores the quantification of amino groups with the DMT-disulfide label is not completely

possible.

That a relatively high number of functionalities, amino groups as well as activated ester

groups, within the surface attached copolymer monolayer are accessible and reactive is

Page 126: through Surface-attached Monomers

112 4 Grafting of Functional Copolymers onto Silica Surfaces

shown by the various experiments with different colored dyes and fluorescence labels

analyzed by (fluorescence) microscopy.

The amino functions in the copolymer monolayer are successfully used for binding

monomers or ATRP starters in order to perform a second polymer grafting step

subsequently as radical solution polymerization or surface polymer started ATRP. These

experiments have shown that the ATRP as a “grafting-from” technique boosts the graft

density by a factor of 20 compared with our standard copolymerization with immobilized

monomers. Besides this advantage the drawback of the ATRP is the problem of getting rid of

unwanted green color of the silica substrates due to copper complexes. Nevertheless both

polymerization paths give us the ability to graft a second polymer layer onto the first. This

second layer may be of different chemical nature or bear different functional groups. The

ATRP approach is of interest when the first polymer layer just acts as compatibilizer and has

to be masked by the second polymer layer for the final usage.

In a next step application possibilities have to be realized by using the final functionalized

spherical substrates in different chemical experiment sets and environments. Of great

interest are solutions for carrier materials for catalyzed chemical reactions and

biotechnological processes where the reactive sites are located on solid substrates.

Page 127: through Surface-attached Monomers

5 Applications for Functional Copolymers on Silica Substrates 113

5 Applications for Functional Copolymers on Silica

Substrates

5.1 Enzyme Immobilization Target of the set of experiments with glass beads having a monolayer of copolymer with

active ester units (GB80-AA-P(DMAA-co-MAC2ae(10%))) and glucose oxidase (GOD) is to

explore the immobilization behavior of a common enzyme onto these tailor-made substrates.

The active ester units on the supports surface are capable of linking covalently to the amino

group of lysine units within the amino acid sequence of the protein. Different aspects of the

enzyme immobilization like washing stability, immobilization time and enzyme

concentration during immobilization were subject of the following investigations.

Figure 5-1. Three steps for the enzyme experiments: a) preparation of the support material, b)

immobilization of the enzyme (yellow) and c) measurement of the substrate (blue) conversion with an

assay procedure.

Generally the experiments on enzymes can be divided into three steps: The first step is to

prepare the glass beads according to 4.2.2, to get the functionalized support material (Figure

5-1a). Afterwards the important immobilization step of the enzymes takes place in aqueous

media (Figure 5-1b).

During this reaction step procedure variations are made to investigate the immobilization

behavior of GB80-AA-P(DMAA-co-MAC2ae(10%)) and GOD. Right after the immobilization

the GOD assay is carried out to determine the enzyme activity (Figure 5-1c). To make the

important enzyme immobilization step visible, at least qualitatively, for one immobilization

Page 128: through Surface-attached Monomers

114 5 Applications for Functional Copolymers on Silica Substrates

experiment dyed GOD was used and analyzed by a fluorescence reading device (Figure 5-2).

DY-635-NH2 reacted with the carboxyl groups in the enzyme structure of GOD forming the

dyed species.

Figure 5-2. Fluorescence labeled glucose oxidase immobilized on glass beads with P(DMAA-co-

MAC2ae) copolymer monolayer visualized by a biochip fluorescence reader.

5.1.1 Glucose Oxidase Assay

The first task while observing enzymes is to track their activity using a quantitative

visualization method. These so-called assays allow to quantify the enzymes activity. It is

important that activity of the enzymes is the limiting factor in the assay and that there is

enough substrate for the enzyme to convert. In our case we use glucoseoxidase. It is a

dimeric protein with a molecular weight of around 160000 g·mol-1 containing the cofactor

flavin adenine dinucleotide (FAD). Each enzyme unit has two FAD-sites which are not

covalently bound to the enzyme. GOD transforms β-D-Glucose into D-Glucono-1,5-lactone

and forms hydrogen peroxide. The FAD cofactor produces from oxygen present in the

aqueous (buffer) solution via an intermediate the hydrogen peroxide. 1 mole hydrogen

peroxide is used by the horseradish peroxidase, which is also present in excess, to oxidize 2

moles of 2,2’-Azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) dye (ABTS) from the colorless

ABTS2- species into the colored ABTS- species (Figure 5-3) which has an absorption

maximum at 405 nm[203-206].

Page 129: through Surface-attached Monomers

5 Applications for Functional Copolymers on Silica Substrates 115

N

SS

NN

SS O

OO

O

O

O

SSO

O

O

NN

N

S SO

O

O

O

OHOH

OH

OHO

OHOH

OH

OH

O2H2O2

-2 OH

ABTS 2- (reduced, colorless)

ABTS .- (oxidized, dark green)

β−D-Glucose D-Glucono-1,5-Lactone

peroxidase(POD)2

2

Glucoseoxidase(GOD)

N+

O

N

OH

Figure 5-3. Chemical pathway for the visualization of the GOD conversion rate with peroxidase and

ABTS.

As long as all participants except for GOD of the assay are used in excess the velocity

determining step is the conversion carried out by GOD. Thus the activity is displayed

indirectly by the absorption of the dye. The absorption maximum of ABTS- is at a

wavelength of 405 nm with an extinction coefficient of ε405 = 36.8 L·mmol-1·cm-1 [207-209].

According to Lambert-Beer’s law (equation 5-1) the chronological change in concentration of

the colored dye is directly proportional to the change in extinction (equation 5-2) recorded by

a UV/Vis spectrometer at the wavelength of the absorption maximum. The resulting graph of

the extinction as a function of time is linear. The slope represents the activity of the enzyme

and can be calculated with equation 5-3.

Page 130: through Surface-attached Monomers

116 5 Applications for Functional Copolymers on Silica Substrates

dεcII

logE λ0

λ ⋅⋅=

−= equation 5-1

1dt

dEdtdc

λ

λ

⋅⋅= equation 5-2

dε2

1dt

dEvλ

λ

⋅⋅⋅= equation 5-3

with

Eλ = extinction at a certain wavelength λ I = intensity of the transmitted light through the sample cuvette in W·m-2 I0 = intensity of the transmitted light through a blank cuvette in W·m-2 c = dye concentration in mmol·L-1

ελ = extinction coefficient at a certain wavelength λ in L·mmol-1·cm-1 d = distance through sample in centimeters t = time in seconds v = conversion rate in mmol·L-1·s-1

Based the chosen assay procedure and on these calculations we will have a look on the

enzyme behavior when immobilized to our polymer coated, functionalized glass beads.

5.1.2 Stability against Leaching

To get the full benefit from an enzyme immobilization, the enzyme has to stay on the

support and must be stable against leaching into the surrounding medium. On the other

hand one has to know how many washing steps have to be performed to wash away non-

bound enzymes. Therefore after the immobilization procedure (10.3.2) several portions of

enzyme-modified glass beads with the identical amounts of glass support have been washed

with buffer media once, twice, three or four times with equal amounts of buffer and

afterwards the activity has been monitored via measuring the extinction as a function of time

as it is shown in Figure 5-4. The washing procedure was kept simple by flushing, shaking

with buffer and decanting the liquid after settlement of the beads.

Page 131: through Surface-attached Monomers

5 Applications for Functional Copolymers on Silica Substrates 117

0 20.0

0.5

1.0

1.5

2.0

2.5

3.0

1st wash 2nd wash 3rd wash 4th wash

extin

ctio

n (a

bs)

Figure 5-4. Extinction as a function of tim

405 nm (extinction maximum). The leachin

rate between the succeeding washing steps

There is a strong decrease of the con

decrease between steps 2 and 3 and alm

3 and 4. This first strong decrease is du

glass beads and in the pores that give

more effective ways of washing suppo

this was the most practicable one. It sh

caused by leaching of GOD from the

washing step. For the following enzy

comparable results.

v = 18.6 nmol·L-1·s-1

v = 3.8 nmol·L-1·s-1

v = 2.1 nmol·L-1·s-1

v = 2.0 nmol·L-1·s-1

4 6 8 10 12 14time (min)

e monitored by a UV/Vis spectrometer at a wavelength of

g experiments show a declining decrease of the conversion

version rate between washing step 1 and 2, a slight

ost no decrease in enzyme activity between the steps

e to residual enzyme containing solution between the

s an additional conversion of glucose. There may be

rt material but for the small samples of 10 mg each

ows that there is no reduction of glucose conversion

glass beads. The activity stays constant after the 3rd

me experiments 3 washing cycles were used to get

Page 132: through Surface-attached Monomers

118 5 Applications for Functional Copolymers on Silica Substrates

5.1.3 Influence of the Immobilization Time on the Conversion

Rate of the Enzyme Modified Glass Beads

It is important to know what the appropriate span of time for the enzyme immobilization

step is to get optimal loading. Several bead samples were exposed to the enzyme solution for

varying periods of time. After washing, the activity of the samples were measured. For an

immobilization time span variation from 10 minutes to 4 hours no significant change in

activity nor any trend in conversion rate change was observed (Figure 5-5). Under the given

conditions the conversion rate of the immobilized enzyme just varies between 2 and

3 nmol·L-1·s-1.

0 2 4 6 8 10 12 140.0

0.2

0.4

0.6

0.8

1.0

10 min 30 min 60 min 120 min 240 min

extin

ctio

n (a

bs)

time (min)

v = 2.3 nmol·L-1·s-1 v = 1.9 nmol·L-1·s-1 v = 3.0 nmol·L-1·s-1 v = 2.5 nmol·L-1·s-1 v = 2.0 nmol·L-1·s-1

Figure 5-5. Extinction as a function of time monitored by a UV/Vis spectrometer at a wavelength of

405 nm. The variation of the immobilization time at equal enzyme concentrations in the

immobilization solution shows no significant change or trend regarding the conversion rate.

Page 133: through Surface-attached Monomers

5 Applications for Functional Copolymers on Silica Substrates 119

5.1.4 Influence of the Immobilization Concentration on the

Conversion Rate of the Enzyme Modified Glass Beads

Here we deal with the variation of the enzyme concentration of the solution and its

influence on the enzyme immobilization. In the initial step the activated ester sites on the

glass beads are exposed to different concentrations of enzymes in solution. With a higher

enzyme concentration the probability of binding an enzyme successfully to the surface is

higher than with lower concentration. It is observed that the activity increases with increase

of the GOD concentration. With our standard GOD concentration of 10-6 mol·L-1 we get an

activity of around 2 nmol·L-1·s-1 as it was observed in the previous experiments.

0 2 4 6 8 10 12 140.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

extin

ctio

n (a

bs)

time (min)

2 *10-6 M 10 *10-6 M 100*10-6 M

v = 1.8 nmol·L-1·s-1 v = 15.1 nmol·L-1·s-1 v =100.9 nmol·L-1·s-1

Figure 5-6. Different GOD concentrations in the immobilization solution have an influence on the

resulting conversion rate of the enzyme-support-system

With two tenfold increases in GOD concentration of 10-5 mol·L-1 and 10-4 mol·L-1 we get

according to Figure 5-6 conversion rates of 15 nmol·L-1·s-1 respectively 101 nmol·L-1·s-1. This

means an increase in glucose conversion rate by a factor of 50 while increasing the enzyme

concentration in solution by a factor of 100.

Combined with the results from the immobilization time variation where no influence was

observed, we get the picture that the enzyme immobilization in the aqueous buffer system is

not diffusion controlled. Under given conditions many of the active ester moieties hydrolyze

Page 134: through Surface-attached Monomers

120 5 Applications for Functional Copolymers on Silica Substrates

before they can react with a possible enzyme site. The probability of getting hydrolyzed is in

an aqueous solution much higher than catching an enzyme because they are highly diluted.

Therefore a higher enzyme concentration in aqueous solution increases this reaction

probability and leads to higher enzyme loadings on the glass support and finally to

increased conversion rates using the material in the assay procedure. To get an idea how

much GOD actually has been immobilized on the glass beads in the case of using a 10-

4 mol⋅L-1 GOD solution the GOD treated glass beads (51 mg and 150 mg) were heated to

reflux in hydrochloric acid. The enzyme decomposed to its single amino acids which, after

filtering off the remaining solids, can be quantified by liquid chromatography[210] (cf.

10.3.6). The measured values give the concentration of each amino acid in the analyte

solution which was for both samples 1 mL of buffer solution so that the amount of the amino

acids for each sample was calculated by dividing by 1000. As we know the amount of each

amino acid in one GOD molecule[211] and the mass of the modified glass beads the amino

acid were cleaved off from, we are able to calculate the specific GOD loading according to

equation 5-4:

1000mN

cLGOD ⋅⋅= equation 5-4

with

LGOD = specific loading of GOD on the glass beads in nmol⋅g-1

c = amount of the amino acid in the analyte solution in nmol, determined by LC N = number of the amino acid units in 1 mol of GOD m = mass of the analyzed, GOD modified glass beads in g

Page 135: through Surface-attached Monomers

5 Applications for Functional Copolymers on Silica Substrates 121

Because of the low concentrations of the single amino acids in the analyzed solution only

the three most frequent amino acids are listed and used for calculation of the immobilized

GOD amount on the glass beads (Table 5-1).

Table 5-1. Listed results from amino acid analysis of GOD hydrolysates.

amino acid concentration amino acid sample 1

(51 mg GB) (µmol⋅L-1)

concentration amino acid sample 2

(150 mg GB) (µmol⋅L-1)

# of amino acid units

in one GOD

molecule[211]

calc. GOD amount sample 1 (nmol⋅g-1)

calc. GOD amount sample 2 (nmol⋅g-1)

glycine 21.4 57.9 114 3.7 3.4

serine 17.3 42.8 74 4.6 3.9

glutamic acid 11.4 39.3 60 3.7 4.4

Both samples give the same picture so that an overall average amount of 4.0±0.5 nmol

GOD per gram glass beads is determined as specific loading of GOD on GB80-AA-P(DMAA-

co-MAC2ae(10%)) using an immobilization solution with 100 µmol⋅L-1 GOD. Regarding the

absolute amounts of GOD we used for our immobilization procedure (cf. 10.3.6) in solution

(10-6 mol GOD in the whole solution volume) and the amounts we finally immobilized on

sample 1 (0.2·10-9 mol GOD on 50 mg solid substrate) and on sample 2 (0.6·10-9 mol GOD on

150 mg solid substrate) show that the immobilization quota is below 1%. With 6.022·1023

GOD molecules per mole, an amount of 4.0 nmol GOD per gram glass beads and their

specific surface of 80 m2·g-1 a density of 5·1012 GOD molecules per m2 or 5 GOD molecules per

µm2 is calculated. It is also shown, that an increase of GB80-AA-P(DMAA-co-MAC2ae(10%))

in equal volumes of GOD solution does not influence the final specific GOD loading. The

action to take to improve the immobilization quota is to put as many glass beads into the

immobilization solution as possible to offer the largest number of activated ester moieties per

volume that is possible. That makes immobilization of GOD on the functionalized glass

beads more likely as we already mentioned the competition with the hydrolysis reaction of

the MAC2ae moieties in aqueous solutions.

Page 136: through Surface-attached Monomers

122 5 Applications for Functional Copolymers on Silica Substrates

5.2 Biomolecule Immobilization The use of immobilized biomolecules is important especially in the design of analytical

methods. Therefore we have tested if solids that were equipped with a functional copolymer

monolayer are capable to bind molecules like biotin or deoxyribonucleic acid fragments

(oligonucleotides) for further use in analytical assays.

5.2.1 Biotin Labeling

Biotin together with avidin is a well known and widely used reagent for many different

analytical purposes. Main field of use is in enzymatic immunoassays[212-214] for detection

of a wide variety of substances, e.g. nucleic acids[215], proteins[216] or food additives[217].

There are also methods for immobilization of nucleotides using a biotin system[218] which

generates the possibility of designing DNA/RNA-chips. Because of the high potential of

biotin systems a short straight forward experimental study shall give an answer to the

question if the active ester sites in a copolymer monolayer as we use it are capable of binding

the biotin and if a more complex reaction system with subsequent reaction steps can be

applied to our copolymer monolayer equipped substrates.

In a first step the functionalized glass beads (Figure 5-7a) are biotinylated by reacting

N-hydroxysuccinimido-biotin in solution with the amino units in the copolymer layer

(Figure 5-7b) on the support surface. Afterwards the solids are treated with an avidin

solution to get glass beads where bound avidin units offer unoccupied binding sites on the

copolymer surface (Figure 5-7c). In a final step from solution biotin-fluorescein binds to the

avidin on the glass beads (Figure 5-7d). If all binding steps are successfully completed

visualization by excitation with a fluorescence reader will show fluorescence of the bound

dye molecules (Figure 5-8).

Page 137: through Surface-attached Monomers

5 Applications for Functional Copolymers on Silica Substrates 123

Figure 5-7. Steps of a fluorescence labeled biotin assay on glass beads: a) Preparation of the support

material, b) immobilization of N-hydroxysuccinimido-biotin onto the active ester functions, c) linking

the biotin to streptavidin and d) labeling the streptavidin with fluorescence (fluorescein) labeled biotin.

The biotin assay procedure was done on glass beads with varying specific surface area and

with a polystyrene monolayer containing different amounts of amino groups (~ 80 and

~ 150 µmol⋅g-1). The two basic polymer systems (PS & P(DMAA)) containing amino groups

were also compared. In the first comparison it can been seen that with higher content of

functional groups and a higher specific surface area the glass beads are brighter when the

dye bound to the copolymer is fluorescing (Figure 5-8b). A direct quantification via

fluorescence intensity is not possible because of the non-uniform geometry of the beads. In a

second experiment beads with hydrophilic and hydrophobic surface layer are compared.

Interestingly the beads equipped with the more hydrophobic P(S-co-AC3amine) layer show

higher fluorescence intensity after treatment with the biotin-fluorescein assay (Figure 5-9a)

than the beads with a more hydrophilic P(DMAA-co-AC3amine) layer (Figure 5-9b) though

the determined amino contents are nearly the same.

Page 138: through Surface-attached Monomers

124 5 Applications for Functional Copolymers on Silica Substrates

Figure 5-8. GB80-AA-P(S-co-AC3amine(10%)) (a), GB250-AA-P(S-co-AC3amine(20%)) (b) as glass

bead substrates treated with the aforesaid assay and visualized by fluorescence reading.

Figure 5-9. Fluorescence reading of a substrate with different amino-functionalized polymer layers

treated with the biotin assay: GB80-AA-P(DMAA-co-AC3amine(10%)) (a), GB80-AA-P(S-co-

AC3amine(10%)) (b)

The expected behavior would have been that using polar solvents for the assay procedure

would result in better swelling of the more polar P(DMAA co-AC3amine) monolayer while

carrying out the reactions. Hence the functional groups of the copolymer layer should be

better accessible and the fluorescence would be more intense than for the fluorescein treated

glass beads with a P(S-co-AC3amine) monolayer. This unusual behavior we found may be

due dye-polymer interaction causing fluorescence quenching with a reduced fluorescence

intensity as result like it is suggested by Haugland et al. [219]. Interactions of dyes with

polymers are described by Neumann et al. [220] and it is shown that these interactions

significantly influence the fluorescence behavior of dyes absorbed on or bound to polymers

in terms of a wavelength shift of the adsorped and emitted light.

Page 139: through Surface-attached Monomers

5 Applications for Functional Copolymers on Silica Substrates 125

5.2.2 DNA Immobilization & Hybridization

An important molecule for biological and medical purposes is deoxyribonucleic acid

(DNA), particular single-stranded oligo nucleotides. DNA molecules attached to planar or

particular SiO2 surfaces are important for DNA analysis. We realize in a setup with glass

beads with an active ester copolymer layer (Figure 5-10a) an oligo nucleotide immobilization

(Figure 5-10b) and subsequent hybridization with an antisense oligonucleotide which

contains a fluorescence label (Figure 5-10c).

For the immobilization step GB80-AA-P(DMAA-co-MAC2ae(10%)) or LC700-ME-

P(DMAA-co-MAC2ae(10%)) as modified supports are used. An amino functionalized oligo

nucleotide with 23 specific bases and 18 thymine units (NH2-DNA) is bound to the active

ester units in the surface-attached copolymer monolayer. To ensure that the immobilization

step is successful, parallel experiments under same conditions with coupling control DNA

(CC-DNA) were carried out and the fluorescence was checked. CC-DNA is a bifunctional,

single-stranded oligo nucleotide with an amino group at the 3’-end and a Cy5 label at the

5’-end.

Figure 5-10. Steps from a solid substrate with a surface-attached copolymer layer with activated ester

groups (a) via immobilized oligo nucleotides (o.n.) (b) to supports with hybridized oligo nucleotides

(h.o.n.) indicated by Cy5 fluorescence (c)

Page 140: through Surface-attached Monomers

126 5 Applications for Functional Copolymers on Silica Substrates

Upon completing the coupling control experiments solids with a homopolymer P(DMAA)

layer exhibit no fluorescence whereas those equipped with a functional P(DMAA-co-

MAC2ae) layer show fluorescence (Figure 5-11a, b). The beads where the CC-DNA has the

chance to bind to the copolymer layer show bright fluorescence whereas the beads with the

inert homopolymer layer just show minimal reflectance. For the solids with surface

copolymer bound sense DNA treatment with antisense-DNA with 23 bases (AS-DNA) is the

next step. The detailed sequences of the used oligonucleotides are listed in Table 10-1 in the

experimental section. Within this step the AS-DNA hybridizes with the covalently bound

NH2-DNA and binds the Cy5 label to the copolymer monolayer on the support. This leads to

blue colored solids, as Cy5 is a blue dye,. indicating a successful hybridization. Also

examination with the fluorescence reading device reveals bright fluorescence on the beads

(Figure 5-11c, Figure 5-12).

Figure 5-11. Fluorescence detection on glass beads: GB80-AA-P(DMAA) treated with CC-DNA (a),

GB80-AA-P(DMAA-co-MAC2ae(10%)) treated with CC-DNA (b) GB80-AA-P(DMAA-co-

MAC2ae(10%)) treated with NH2-DNA and AS-DNA.

Figure 5-12. False color picture of glass beads with successfully hybridized oligo nucleotides attached

to the copolymer monolayer show fluorescence from Cy5 labeled anti-sense oligo nucleotides.

Page 141: through Surface-attached Monomers

5 Applications for Functional Copolymers on Silica Substrates 127

Again quantification of the surface fluorescence is difficult because of interfering geometry

of the beads and slight intensity differences between single beads. For glass beads and silica

similar results were observed. Because the silica particles are very small they can hardly be

detected even with a fluorescing surface (Figure 5-13, right bottom).

Figure 5-13. Fluorescence image of glass beads GB80-AA-P(DMAA-co-MAC2ae(10%)) and silica

LC700-ME-P(DMAA-co-MAC2ae(10%)) after successful DNA immobilization & hybridization.

5.3 Catalyst Immobilization Another huge application field for functionalized solids besides biomolecule

immobilization is the immobilization of organometallic species as catalysts for synthesis. All

experiments described here have been carried out in close collaboration with the group of

Prof. Dr. W. Bannwarth. After preparation of the amino modified glass beads further

reaction steps have been carried out by H. Glatz and F. Michalek.

Figure 5-14. Steps for the immobilization of a Ruthenium-metathesis catalyst onto a silica support:

Immobilization of a Hoveyda-type ligand (1) onto the support (b). Ligand exchange of Grubbs catalyst

(2 & 3) binds the species onto the support (c & d).

Page 142: through Surface-attached Monomers

128 5 Applications for Functional Copolymers on Silica Substrates

Starting point for the catalyst immobilization are glass beads (GB80 & GB250) with a

surface-attached layer of P(S-co-AC3amine) (Figure 5-14a). The amines form an amide link

with the carboxylic acid groups of a Hoveyda-type ligand (Figure 5-14, 1) using N,Nʹ-

dicyclohexyl-carbodiimide (DCC), tert-butanol and Huenig’s base. The ligand is covalently

bound to the support copolymer surface (Figure 5-14b)[221]. Then free Grubbs catalysts of 1st

(Figure 5-14, 2) and 2nd generation (Figure 5-14, 3) in solution is immobilized onto the

support using copper(I)chloride for activation and dichloromethane as solvent giving bound

Hoveyda-type catalyst of 1st (Figure 5-14c, Figure 5-15a) and 2nd generation (Figure 5-14d,

Figure 5-15b)[188].

Figure 5-15. Glass beads with immobilized Ruthenium catalyst species within the functional

copolymer monolayer give a specific color. GB80-AA-P(S-co-AC3amine) as carrier material for Grubbs

catalyst 1st generation (a) and Grubbs catalyst 2nd generation (b) is shown.

To check the performance of this new catalyst hybrid material an exemplary ring closing

metathesis (RCM) reactions were run. This benchmark reaction is displayed in Figure 5-16

where the RCM of N,N-diallyl tosylamide is done by a Ruthenium-metathesis catalyst.

Investigations on olefin metathesis catalysts immobilized on silica have shown that in

organic solvents the immobilized catalyst can be released to perform the reaction as catalytic

carbene species and afterwards may be recaptured by the immobilized ligand[222].

Figure 5-16. RCM reaction of N,N-diallyl tosylamide (a) to N-tosyl-3-pyrrolin (b) catalyzed by an

active Ru-metathesis species.

Page 143: through Surface-attached Monomers

5 Applications for Functional Copolymers on Silica Substrates 129

In the first RCM experiments the solid phase-bound catalysts are used in dichloromethane

as organic solvent under reflux conditions for 3 h with an catalyst amount of 5 mol% related

to the amount of N,N-diallyl tosylamide. The first two RCM cycles show quantitative

conversion for both catalyst generations. The third cycle reveals for both, immobilized

Grubbs catalyst of 1st and 2nd generation, less activity due leaching of small amounts of the

Ru-metathesis species. But the conversion is for both cases still above 95% conversion. After

each cycle the immobilized catalyst can be easily filtered off the reaction solution and used

for another reaction cycle. To reload the leached solid phase it has to be placed into a

solution of free Grubbs catalyst. There it regains the original loading with surface copolymer-

bound Ru-metathesis species.

After these first promising experiments in organic solvents further experiments where

supercritical carbon dioxide is used as solvent for the RCM reaction were carried out. In

these experiments our functionalized silica polymer hybrids with immobilized Grubbs

catalyst species (Hoveyda type) show even better results concerning conversion and

leaching. The results of these experiments are published and discussed in detail in [188].

Page 144: through Surface-attached Monomers

130 5 Applications for Functional Copolymers on Silica Substrates

5.4 Discussion & Conclusions Applications The first part of the applications section is concerned with enzyme immobilization on solid

supports. Our investigations are based on the well-known enzyme glucose oxidase (GOD)

and we like to investigate its immobilization behavior on glass beads modified with a

reactive, hydrophilic copolymer layer. In a first set of experiments where the substrates with

the immobilized GOD are washed to test the stability against leaching, it is recognized that

after several washing cycles there is no loss in rate of the conversion of glucose. This

confirms the picture of covalently bound enzymes that are not easily leached out by flushing

the enzyme substrate. The behavior encountered upon variation of the immobilization time

suggests that the final amount of enzyme on the surface is reached fast within a few minutes.

Allowing for additional immobilization time does not add to the activity of the enzyme

modified support as it would be expected for a clearly diffusion based process where the

reactant molecules have enough time to approach the reacting sites. The binding of the

enzyme towards the N-succinimide ester competes with the hydrolysis reaction of the active

ester in the aqueous immobilization solution. Because the hydrolysis step is irreversible

enzymes can only attach as long as the active ester moieties are intact. For the enzyme the

probability of finding a still activated ester site decreases with time. The surrounding

aqueous buffered solution (pH = 7.4) quenches active ester groups quickly before they can

bind any enzyme. The observed conversion rate increase while increasing the initial enzyme

concentration in the immobilization solution is due to the fact that in the initial state more

enzyme entities encounter more intact active ester groups. The probability of binding GOD

covalently to an active ester site increases with GOD concentration while the competitive

hydrolysis reaction of the N-succinimide esters is still quick. This corresponds with the

finding that the increase the glucose conversion rate is not proportional to the increase of

GOD in the immobilization solutions. Caused by the fact that the amount of enzyme

molecules is low (≤ 10-4 mol·L-1) and the active ester moieties are quite prone to hydrolysis

under the given conditions only few enzyme molecules get immobilized. This result is

underlined by the experiments where the immobilized GOD is hydrolyzed to quantify its

amount via liquid chromatography. The immobilization quota is below 1% and because the

aqueous system and the pH-value are necessary to retain the structure and activity of the

enzyme this immobilization quota could only be increased by increasing the amount of

functionalized glass beads in the GOD solution. It is clearly shown that immobilization of

enzymes with our covalently bound, reactive copolymer system on glass beads is feasible.

Page 145: through Surface-attached Monomers

5 Applications for Functional Copolymers on Silica Substrates 131

We get stable systems with firmly fixed enzymatic entities showing significant activity and

we are able to control the amount of loaded enzyme. To get detail information about the

interaction between enzymes and our copolymer monolayer system it will be necessary to

carry out further experiments where the specific activity of free enzymes and enzymes

bound to the polymer-substrate system will be compared. It is probable that retention of the

enzyme structure is enhanced by a polymer monolayer acting like a cushion on the

immobilization substrate. Also an interesting point for further investigations in this field is to

look if there is a dependence between the active ester loading/density and possible enzyme

loading or specific enzyme activity.

The experiments with the amino functionalized polymer layers on glass beads and the

biotin/avidin assay have shown that more complex labeling reactions with several steps can

be successfully carried out with both basic systems, the PS and the P(DMAA) system. In

direct comparison with the P(DMAA)-amine system, the PS-amine system shows the

brighter fluorescence. This is a surprising behavior because the used solvents in the assay

pathway and the affinity of biotin towards hydrophilic environments suggest a better

swellability of the P(DMAA) layer, a better access to the amino reaction sites and a better

penetration of the layer by the biotin and streptavidin. These factors should finally result in a

higher dye density within the copolymer layer. A possible factor that influences the

fluorescence is the interaction between polymer layer and dye. As we have seen in dye

labeling experiments before (Figure 4-8 & Figure 4-12) this interaction provokes shift of

extinction maxima and hence the fluorescence intensity in the observed range. Also polymer-

dye interaction quenching the fluorescence and reducing the observed intensity might be

possible. For a detailed understanding of functional group distribution and accessibility

within polymer monolayers of different polarity further efforts are necessary. Overall, the

feasibility of our functional modified beads for biotin based systems is proven and the

system may be adapted to the particular application.

The complementary pairing of an oligo nucleotide via hydrogen bonds with the

complement forms a reversible link between two DNA strands. Using elevated temperatures

the double-strand can be molten and is split into two complementary species again. In our

experiments oligonucleotides are efficiently immobilized on glass beads and silica gel

equipped with a functional copolymer layer with activated ester moieties. It is shown that

oligonucleotides can be bound specifically to the support and undergo hybridization

reactions in the polymeric environment built by the copolymer monolayer. This possibility

Page 146: through Surface-attached Monomers

132 5 Applications for Functional Copolymers on Silica Substrates

offers the use of our hybrid system for the whole range of applications in chromatography

and analytics based on the specificity and reversibility as outstanding features of nucleotide

sequences. In addition the surface bound polymer monolayer system is not restricted to

spherical and porous systems but can also be applied to flat substrates as we will show in the

next chapter. This allows one to create DNA-chips based on covalently bound copolymer

monolayers.

For the last application we leave the field of biomolecules and life science and get into

inorganic catalysts for organic synthesis. By covalent attachment of ligands onto the

functional copolymer monolayer it is possible to immobilize transition metal complexes with

catalytic behavior. In our case the immobilized Grubbs catalysts showed quantitative

conversion of the substrate and little leaching during ring closing metathesis reactions in

organic solvents. Also simple reloading of the support with catalyst species can be observed.

This example shows that we have here a potent system offering a large number of

possibilities. It has to be explored which ligands, catalytic species, catalyzed reactions and

reaction conditions can be used with our system. In most cases the polymerization step in the

hybrid material buildup allows tailoring the properties of the interacting layer. This is also

true for the other application fields. For all applications in common are the following

advantages: First to have an inert, non-swellable, easy to handle SiO2 core, second a tailor-

made copolymer monolayer which covers the core and allows to control the interaction with

the environment, third high loading of functional groups within the copolymer and forth

improved mobility of functional groups and immobilized species on a swellable “polymer

cushion” attached to a rigid substrate surface. The advantage last mentioned is important

because many chromatographic and catalytic reactions take place in packed bed reactors

where flowing fluids cause significant shear forces on the surface of the solids. These shear

forces may rip off any immobilized functionality or species when just being surface-attached

by a short, more or less rigid spacer molecule. Having a semi-mobile copolymer layer could

absorb the shear forces by polymer segment movement[223] and could prevent losing active

functionalities due to shear induced rupture.

Page 147: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 133

6 Polymer Systems on Silicon Wafers

6.1 Grafting of Poly(N,N-dimethyl acrylamide) Layers onto

Silicon Wafers through Surface-attached Monomers In previous chapters the surface modification of silica substrates, silica gel and glass beads

by using a “grafting-through” approach was used. Bulk polymerization of styrene on silicon

wafers modified with methacrylate groups was investigated by M. Bialk and O.

Prucker[153]. To close the gap we present results dealing with the radical polymerization of

N,N-dimethyl acrylamide in the presence of monomers attached to silicon wafer surfaces.

The monomers were immobilized using trialkoxy silane anchors. The experiments were

carried out in DMF as solvent because polymerization in bulk DMAA leads into problems

caused by gelation. This effect is for DMAA much more pronounced and occurs already at

much lower conversion than it is the case for styrene. The parameters that were kept

constant are polymerization temperature (60 °C), polymerization time (16 h), AIBN

concentration (3 mmol·L-1) and monomer concentration (3.2 mol·L-1 DMAA). The

experiments are supplemented with similar ones but using styrene as monomer. There we

have as polymerization conditions a temperature of 60 °C, a polymerization time of 24 h and

an initiator concentration of 3 mmol·L-1 AIBN. Variation of single parameters is described in

the following paragraphs. For each set parameter two different wafers are measured with at

least three surface points to get representative values and to calculate the standard deviation.

6.1.1 Influence of Monomer Concentration

The first parameter to investigate is the monomer concentration because it has the

strongest influence on grafting P(DMAA) on silica (3.7.3). As it is known from the

experiments with silica the polymerization solution tends to get viscous and gel at early

stages when conversion has not advanced very far. Using wafers we do not have the

necessity to stir or shake the polymerization vessel in order to prevent concentration or

temperature gradients. After 16 h polymerization time the solutions that started with a

DMAA concentration below 2 mol·L-1 are viscous, but still o.k. Above that concentration the

solution gels. A careful and extensive Soxhlet hot extraction frees the wafer from any non-

covalently bound P(DMAA). This set of experiments gives a clear picture: the higher the

monomer concentration the thicker the P(DMAA) monolayer (Figure 6-1).

Page 148: through Surface-attached Monomers

134 6 Polymer Systems on Silicon Wafers

0 2 4 6 8 100

10

20

30

40

laye

r thi

ckne

ss (n

m)

c(M) (mol/L)0 2 4 6 8 10

0

10

20

30

40

Figure 6-1. Increase of the P(DMAA) layer thickness with increasing DMAA concentration compared

to similar experiments with styrene (grey values).

At low monomer concentration where the polymerization stays fluid throughout the

polymerization only very thin layers below 3 nm are generated whereas polymerization in

bulk DMAA leads to 30 nm layers. However, at high DMAA concentrations the samples

show a much larger variance in layer thickness. This is an indication for inhomogeneities due

to local gelation at any stage of the polymerization. P(DMAA) builds thicker layers in

comparison with PS, because at comparable monomer concentrations up to bulk

polymerization the PS layer is always thinner.

Page 149: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 135

0 2 4 6 8 100

50100150200250300350400450

mol

ecul

ar w

eigh

t (10

3 g/m

ol)

c(M) (mol/L)

Mn P(DMAA) Mw P(DMAA)

0 2 4 6 8 100

50100150200250300350400450

Mn PS Mw PS

Figure 6-2. Molecular weight averages for the polymerization of DMAA at different monomer

concentrations in comparison with values for PS (grey values).

The measured molecular weights match those of the experiments with DMAA

concentration variation using silica as substrate. The molecular weights increase with higher

initial monomer concentrations for P(DMAA) as well as for PS (Figure 6-2). The number

averages of the molecular weight for both polymers rise linearly with increasing monomer

concentration. The molecular weight distribution for P(DMAA) is much broader than for PS

and the P(DMAA) from bulk polymerization could not be measured properly by our GPC

system because the molecular weight of the polymer was beyond the exclusion volume.

Page 150: through Surface-attached Monomers

136 6 Polymer Systems on Silicon Wafers

0 2 4 6 8 100.00

0.05

0.10

0.15

graf

t den

sity

(µm

ol/m

2 )

c(M) (mol/L)0 2 4 6 8 10

0.00

0.05

0.10

0.15

Figure 6-3. Graft densities for P(DMAA) on silicon wafers after polymerization with varying

monomer concentrations compared to those of PS on wafers (grey values).

In contrast to graft polymerization on porous substrates P(DMAA) and PS show on MPS-

modified silicon wafers that the increased loading/layer thickness is not altered by higher

molecular weights. For both polymer systems graft densities increase on polymerization

with higher monomer concentrations. For P(DMAA) this effect is more pronounced than for

PS (Figure 6-3). This behavior is strong evidence for a difference between our polymerization

system with surface attached monomers and pure “grafting-to” systems and will be

discussed in paragraph 6.1.4.

6.1.2 Polymerization Time Influence

For the development of the P(DMAA) layer thickness on silicon wafers which is equivalent

with grafted polymer amount on porous substrates we get a bit different result compared

with the latter ones. On glass beads and silica gel the plateau of the final polymer load was

reached quite fast in 30 minutes. After 2 hours of polymerization a level of around 5 nm layer

thickness is reached (Figure 6-4). Apparently for polymerization times above 15 hours the

layer thickness increases and reaches a maximum of 7 nm. With the chosen monomer

concentration of 1/2 (v/v) DMAA/DMF in order to get significant layer thicknesses the

polymerization already suffers from gelation which renders the values obtained not reliable

because of inhomogeneous covering of the wafer.

Page 151: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 137

0 4 8 12 16 20 240123456789

10

laye

r thi

ckne

ss (n

m)

polymerization time (h)

Figure 6-4. Layer thicknesses of P(DMAA) in the course of polymerization.

The development of the molecular weight distribution with polymerization time shows a

similar behavior as observed during the experiments on the porous substrates: In the

beginning we have a higher molecular weights which reach their final level after 4 hours

with a number average of around 60000 g·mol-1 and a weight average of the molecular

weight distribution of around 210000 g·mol-1 (Figure 6-5).

0 4 8 12 16 20 240

50

100

150

200

250

300

350 Mn Mw

mol

ecul

ar w

eigh

t (10

3 g/m

ol)

time (h)

Figure 6-5. Molecular weight averages of P(DMAA) for different periods of polymerization time.

Page 152: through Surface-attached Monomers

138 6 Polymer Systems on Silicon Wafers

From the layer thickness and the molecular weight the graft density can be calculated. In

the case of P(DMAA) on silicon wafers the graft density increases steadily with

polymerization time (Figure 6-6). Although the differences in film thickness between the

individual samples is small (± 1 nm) the trend appears to be quite stable.

0 4 8 12 16 20 240.00

0.05

0.10

0.15

0.20

graf

t den

sity

(µm

ol/m

2 )

polymerization time (h)

Figure 6-6. Graft densities of P(DMAA) on MPS-modified silicon wafers in the course of

polymerization.

6.1.3 Variation of Initiator Concentration

A change in P(DMAA) layer thickness is observed upon initiator concentration variation.

Using AIBN concentrations in the range from 0.1 to 100 mmol·L-1 the layer thickness slightly

decreases with higher concentrations from 13 nm to 7 nm (Figure 6-7). For P(DMAA) on

porous substrates and for PS on wafers we did not see any influence of the initiator

concentration on the polymer layer thickness.

Page 153: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 139

0.1 1 10 1000

4

8

12

16

20

laye

r thi

ckne

ss (n

m)

c(I) (mmol/L)

Figure 6-7. With increasing AIBN concentration in the polymerization solution the layer thickness of

P(DMAA) on wafers slightly decreases.

As observed during the initiator variation experiments on silica the number average

molecular weight stays constant around 60000 g·mol-1 whereas the weight average has a

trend to higher values with higher initiator concentration thus leading to a broader

molecular weight distribution (Figure 6-8). Again this behavior can be explained by fast

conversion and viscosity effects leading to auto-acceleration of the polymerization reaction.

1 10 1000

50

100

150

200

250

300

mol

ecul

ar w

eigh

t (10

3 g/m

ol)

c(I) (mmol/L)

Mn Mw

Figure 6-8. Molecular weight averages for P(DMAA) after polymerization at different AIBN

concentrations.

Page 154: through Surface-attached Monomers

140 6 Polymer Systems on Silicon Wafers

The factors described above lead to a decrease in apparent graft density while increasing

the initiator concentration for P(DMAA) on silicon wafers (Figure 6-9). For P(DMAA) on

porous SiO2 there was no influence of initiator concentration on graft density and for PS on

silicon wafers an opposite effect is observed. The different polymerization behavior of

DMAA with its tendency to form highly viscous polymerization solutions may lead to a high

number of unattached but blocking polymer coils for high initiator concentrations.

0.1 1 10 1000.0

0.1

0.2

0.3

0.4

graf

t den

sity

(µm

ol/m

2 )

c(I) (mmol/L)

Figure 6-9. The graft density of P(DMAA) decreases for higher AIBN concentrations.

6.1.4 Discussion & Conclusion P(DMAA) on Wafers

The impact of the varied polymerization parameters such as monomer concentration,

polymerization time and initiator concentration on the polymer loading was differently

pronounced for the single parameters. The monomer concentration variation has the

strongest influence on the polymer loading respective layer thickness like it was the case for

the graft polymerization on porous substrates. Going into more detail there are differences

between the two basic surface geometries: Porous substrates show decreasing graft density

with increasing monomer concentration whereas flat substrates have an increasing graft

density. These trends are the same as investigated with experiments on polystyrene. So this

seems to be a general behavior of systems using surface attached monomers. For a pure

“grafting-to” system one expects decreasing graft densities with higher monomer

concentration. Generated larger polymer coils attached to the surface are more effective in

Page 155: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 141

blocking the way to the surface for further polymer molecules that try approach the

surface[224]. This unusual behavior leads to the conclusion that we deal here with a system

where the “grafting-to” step is most important but the “grafting-from” component is not

negligible. For the latter step it is necessary to provide monomer to the site where the

growing chain is. The diffusion of whole polymer coils through the surface attached polymer

layer is not needed. Higher monomer concentrations supply on the one hand a better

probability of growing the already attached polymer chain further and on the other hand

represses polymer to solvent transfer reactions of the radical that terminate polymer growth.

Another factor is the surface geometry: Monomer concentration in the vicinity of flat

substrates is easier balanced with the surrounding solution than porous substrates because

they get easily clogged pores if a system like P(DMAA) in DMF shows up with high

viscosities. In polymer blocked pores fluidity is more effectively hindered and concentration

may be less balanced with the surrounding media. That is why we observe for porous

substrates for our system indeed decreasing graft densities with increasing monomer

concentrations but this influence is not as strong as for “grafting-to” systems[225]. On flat

substrates the classic “grafting-to” effect is altered in a way that the trend is reversed and we

get increasing graft densities with increasing monomer concentrations. Therefore sometimes

we call systems using surface attached monomers jokingly as “grafting-in-between” because

in fact these systems got components of both “grafting-to” and “grafting-from”. The reason

why for P(DMAA) the final layer thickness on wafers is reached later than the final polymer

load on silica is a combination of monomer balance and diffusion through the polymer layer

and stuck pores due to high viscosity. The latter cause an early final stage for P(DMAA) on

porous substrates. For the polystyrene system we get a stronger diffusion influence as the

viscosity and the propagation rate is not as high as for styrene/polystyrene. This is proven by

the fact that for PS the final layer thickness on wafers is reached faster than the final polymer

loading on porous substrates[153]. A similar picture we can draw from the results of the

initiator concentration variation although the influence is not as strong as for the monomer

variation. With higher initiator concentrations the layer thickness is reduced because the

more initiator molecules start polymer chains the earlier a certain viscosity is reached and

slows down monomer diffusion, polymer mobility and the probability of binding polymer to

the surface (Figure 6-10). For PS the viscosity under similar polymerizing conditions is less

and enables the system to level the layer thickness by diffusion resulting in equal layer

thicknesses over a range of different initiator concentrations[153].

Page 156: through Surface-attached Monomers

142 6 Polymer Systems on Silicon Wafers

Figure 6-10. Variation of the initiator concentration for the polymerization of DMAA in presence of

surface attached monomers leads to different layer thicknesses/graft densities. With a lower initiator

concentration the growing chains stay mobile for a longer time and their radical and can diffuse

through the monolayer to attach to the surface (a). With a higher initiator concentration the polymer

chains lose their mobility to diffuse through the polymer monolayer earlier because of faster conversion

(b). After extraction all non-attached polymer chains that got stuck in the polymer layer but did not

attach to the surface are removed (lower pictures) and in the case of higher initiator concentration a

lower film thickness/graft density is observed.

Page 157: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 143

6.2 Microstructured Surfaces Target of the experiments described in this chapter is to explore the possibilities of forming

microstructures and micropatterns based on polymer systems using surface attached

monomers. Two different approaches with MPS-modified silicon wafers as base material are

introduced. Both are photolithographic structuring methods that use UV light in the

patterning step. The first one creates a pattern by destroying the MPS units at the surface.

The pattern image is created directly on the flat substrate as first step. Method two creates a

pattern by building a polymer network after a polymer film has been cast onto the wafer

substrate.

6.2.1 Photoablation of Surface-attached Monomers Followed by

Polymerization

For direct patterning of the surface-attached MPS high energetic UV light with

wavelengths λ < 200 nm is used. It is known that deep UV light destroys carbon-carbon

bonds[167] such as those present in the used surface-attached monomers as well as surface-

attached polymer monolayers. In simple contact irradiation experiments the mask is in direct

contact to the substrate while photodecomposition takes place (Figure 6-11a). The MPS in the

covered areas stays intact whereas the irradiated areas lose their monomer functionality

(Figure 6-11b). This method can be performed quickly and requires only short illumination

times below 30 min because the MPS layer is very thin (~1 nm). On the silicon wafer the

residual intact MPS units give a positive image of the mask. The so patterned MPS layer is a

latent image that develops when exposed to a radical polymerization solution creating a

polymer monolayer in the non-irradiated areas (Figure 6-11c). Thicknesses of the built

polymer monolayers are according to our results in this thesis and may be varied up to a

maximum of 10-20 nm for PS and 20-30 nm for P(DMAA) choosing appropriate

polymerization conditions.

Page 158: through Surface-attached Monomers

144 6 Polymer Systems on Silicon Wafers

Figure 6-11. Direct patterning of MPS-functionalized silicon wafers (a & b) and subsequent

copolymerization of the remaining surface attached double bonds in solution polymerization.

The chosen example for the method depicted in Figure 6-11 is a polymerization of styrene

on silicon wafers with a structured MPS layer. For structure formation different TEM metal

masks are used. The PS monolayer is analyzed by imaging ellipsometry. As result we get

complementary images where on the one hand the intensity of the reflecting light from the

polymer areas (Figure 6-12a, c) and on the other hand from the substrate areas (Figure 6-12b,

d) is minimized. As we use grids as masks the grown polymer monolayer forms also thin

lines to a grid. These polymeric grid lines have a width of a few micrometers as obtained by

ellipsometry. The observed dry PS layer thickness is ~10 nm. These ultrathin layers can be

measured by imaging ellipsometry but the contrast between polymer layer and substrate is

lower than for example polymer brushes with a much higher layer thickness[168].

Page 159: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 145

Figure 6-12. Complementary imaging ellipsometry micrographs of structured polystyrene monolayers

with a dry thickness of ~10 nm on silicon wafers. The grid size of the upper example (a, b) is 128 µm

and of the lower example (c, d) it is 64 µm.

6.2.2 Patterning through Polymer Network Formation

The first step is here the already described immobilization procedure of MPS on silicon to

get the surface-attached monolayers (Figure 6-14a). Then a copolymer film containing

methacryloyl-4-oxy-benzophenone (MABP) units is cast onto the modified substrate. This is

transferred into a polymer network by UV-cross-linking (Figure 6-14b). Different

concentrations of the polymer solution and different coating techniques such as dip coating

and spin casting have been applied in order to get a thick layer on the modified wafers (cf.

Table 6-1). It is also important to get homogeneous polymer layers in order to develop the

intact polymer microstructures with light-guiding masks. By UV-illumination of the

unmasked copolymer areas a network is formed by carbon-carbon linkage of the

benzophenone unit within the polymer. In addition the formed polymer network is fixed to

Page 160: through Surface-attached Monomers

146 6 Polymer Systems on Silicon Wafers

the wafer by carbon-carbon linkage between the benzophenone in the copolymer and the

MPS molecules on the surface. Network formation through mask guided UV illumination

takes place by using wavelengths of 350 nm to activate the benzophenone units within the

polymer. Copolymerized MABP monomers have been used in the past to form polymer

networks as well as benzophenone with silane anchor group has been immobilized to

surfaces in order to attach polymers covalently to the surface[60, 226]. Here a MPS layer on

the surface is serving as attachment point for the benzophenone units in the copolymer

(Figure 6-13).

Diligent extraction with a good solvent for the copolymer in a Soxhlet extractor leaves the

cross-linked fraction of copolymer in the UV illuminated areas as a network structure. In the

non-illuminated areas the copolymer is washed away down to the wafer surface where still

the attached MPS units are (Figure 6-14c). These surface monomers are then used to bind an

ultrathin copolymer monolayer in another radical copolymerization to the surface (Figure

6-14d). This copolymer monolayer contains functional units for further use.

Figure 6-13. Detailed view on the attachment of a benzophenone copolymer to a surface-attached MPS

group by UV exposure. In the coated copolymer some benzophenone units rest in close proximity to

the MPS-covered surface and form a C-C-bond upon activation with UV light (λ = 350 nm).

Page 161: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 147

Figure 6-14. Four steps leading to a micropatterned wafer with a functional copolymer within: a)

Immobilization of MPS onto the silicon surface, b) casting a benzophenone copolymer on the MPS-

modified wafer and immediately UV illumination through a mask, c) the latent pattern evolves by

extraction of free copolymer chains from the crosslinked & surface bound copolymer network, partially

uncovering the MPS-modified surface, d) another copolymerization with these surface monomers is

possible.

Page 162: through Surface-attached Monomers

148 6 Polymer Systems on Silicon Wafers

To apply cross-linkable MABP-copolymers to MPS-modified silicon wafers common

coating techniques such as spin casting and dip coating are used. Different coating

conditions are checked and the resulting dry layer thicknesses are listed in Table 6-1 and

Table 6-2:

Table 6-1. Average layer thickness and standard deviation attained by different coating techniques

with P(S-co-MABP) with different amounts of cross linker.

P(S-co-MABP) concentration

(mg⋅mL-1)

benzophenone content (mol%)

coating technique layer thickness average

(nm)

layer thickness deviation

(nm)

20 1 dip coating 64 5

50 1 dip coating 292 8

70 1 dip coating 341 18

100 1 dip coating 315 20

50 1 spin casting 314 8

20 5 spin casting 80 10

20 10 spin casting 94 10

20 5 dip coating 132 126

20 10 dip coating 172 105

Table 6-2. Average layer thickness and standard deviation attained by different coating techniques

with P(DMAA-co-MABP) with different amounts of cross linker.

P(DMAA-co-MABP) concentration

(mg⋅mL-1)

benzophenone content (mol%)

coating technique layer thickness average

(nm)

layer thickness deviation

(nm)

50 1 dip coating 675 34

70 1 dip coating 742 16

25 5 dip coating 347 1

50 5 dip coating 770 3

25 10 dip coating 325 1

Page 163: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 149

Goal of the diverse coating experiments is to determine conditions to get homogeneous

and relatively thick (> 200 nm) copolymer layers. Dip coated P(S-co-MABP) layers show in

most cases dewetting on silicon wafers whereas spin cast P(S-co-MABP) layers are stable.

MABP-copolymer solutions are influenced by even traces of daylight and so crosslinking and

viscosity changes may occur. The higher the MABP content the more difficult the handling of

the solution and the application of the coating. Based on these insights 50 mg⋅mL-1 of

P(S-co-MABP(1%)) are spin cast and 70 mg⋅mL-1 of P(DMAA-co-MABP(1%)) are dip coated

onto MPS-modified silicon wafers for use in the following photolithographic experiments. It

is important to mention that by 1H-NMR measurements the benzophenone content of

P(S-co-MABP(1%)) was determined as 9 mol% and for P(DMAA-co-MABP(1%)) the actual

value was 1 mol%.

Different masks, self-made printouts and commercially available transmission electron

microscopy (TEM) grids, have been applied to P(S-co-MABP)- and P(DMAA-co-MABP)-

coated silicon wafers prior to UV exposure to create polymer network microstructures.

Details on used masks and patterns are given in paragraph 11.3 of the experimental section.

From this great variety some examples are chosen to illustrate the possibilities and limits of

the described photolithographic technique.

Figure 6-15 shows images of polymer network structures evolved from UV illumination

through self-made masks. These masks are printouts of a laser printer using common

transparencies for overhead projectors. The line width and space between two lines are the

same. The top view with a CCD camera shows that structures down to a feature size of

200 µm are clearly separated and only the printer resolution limits the structure dimensions.

The color gradient of the polymer structures reveals different thicknesses within one section.

Page 164: through Surface-attached Monomers

150 6 Polymer Systems on Silicon Wafers

Figure 6-15. CCD image of a cross-linked P(S-co-MABP) network. The colored areas show the

polymer. A self-made mask with a line width of 400 µm (a) and 200 µm was used. The black line

marks the taken profile of Figure 6-16.

Profilometric measurements allow to have a more detailed look on a cross section of these

structures. Figure 6-16 shows polymer network walls with a maximum height of nearly

400 nm. The structures have no sharp edge between areas where the copolymer has been

irradiated (wall) and where not (ground level). This is due to the fact that the transparencies

as carrier of the printed masks do not totally overlie on the coated copolymer and the prints

neither have a sharp contour nor they are completely black and impermeable for light. These

factors have the effect that the polymer network structure got a rounded silhouette on the

top edge. Upon inspection of the Figure 6-16 should be noted that the y-axis is given in

nanometer, whereas the x-axis is in µm, so that all structures look much more rounded then

they actually are.

Page 165: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 151

0 200 400 600 800 1000 12000

100

200

300

400

500

heig

ht (n

m)

width (µm)

Figure 6-16. Profilometric cross section of a PS network created by using a mask with 400 µm line

width (cf. Figure 6-15a).

The effects of the mask design on the resulting polymer network with its less pronounced

contours is also tangible with imaging ellipsometry. When alternately minimizing the

intensity of the reflecting light from the top of the polymer structure(Figure 6-17a) and from

the substrate (Figure 6-17b) there is a small area in between that is not dark on both images.

This is the transition area where the polymer network structure has a slope.

Figure 6-17. Complementary imaging ellipsometry micrographs of a PS network: reflectance

minimized on polymer (a) and on substrate(b). The line width of the mask was 200 µm.

Page 166: through Surface-attached Monomers

152 6 Polymer Systems on Silicon Wafers

Mask patterns printed on flexible plastic have the drawback that they become distorted

upon exposure to a UV lamp as they get significantly hot. If the mask support or material

does not fit to the copolymer layer closely there are penumbra regions where the copolymer

layer is only partially irradiated by the UV light. This results in rounded profiles instead of

sharp edged ones. A comparison between a microstructure built with a closely fit mask and a

distorted mask is given in Figure 6-18. For image a) the metal mask fitted nicely to the PS

copolymer layer surface and the resulting structures have sharp edges. On image b) the

metal was slightly twisted and did not lie tight on the PS copolymer surface which leads to

structures that mingle with each other.

Figure 6-18. TEM grids with different mesh sizes (mesh 200 = 128 µm (a), mesh 400 = 64 µm (b)) are

used to pattern PS networks. The profiles of Figure 6-19 and Figure 6-20 are marked as black lines.

The impressions from the top view of Figure 6-18 are confirmed by profilometry. A cross

section taken from the mesh 200 polymer microstructure reveals a relatively flat top with a

height of around 414 nm and a width of about 80 µm and a groove down to the base

substrate (Figure 6-19). Another picture gives the cross section of the mesh 400 polymer

microstructures (Figure 6-20). The top of the structures is more rounded and the grooves

between the polymer microstructures do not reach down to the substrate.

Page 167: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 153

0 25 50 75 100 1250

100

200

300

400

500

heig

ht (n

m)

width (µm)

Figure 6-19. Profilometric cross section of a PS network created by using a metal grid with 128 µm

grid size (cf. Figure 6-18a).

0 25 50 75 100 1250

100

200

300

400

500

heig

ht (n

m)

width (µm)

Figure 6-20. Profilometric cross section of a PS network created by using a TEM grid with a mesh size

of 64 µm (mesh 400) (cf. Figure 6-18b).

Page 168: through Surface-attached Monomers

154 6 Polymer Systems on Silicon Wafers

Figure 6-21. Square ‘pillars’ are constructed from PS network using mesh 200-type TEM grids as

shown in the complementary ellipsometry images.

Figure 6-22. From ellipsometric data computed 3-dimensional view of the structured PS network. The

z-axis (height) is exaggerated to pronounce the polymer network structures and reveals that single

structure elements look like pillars.

The ellipsometric data from mesh 200 PS microstructures shown in Figure 6-21 is used to

calculate a three-dimensional computer image where the axis perpendicular to the substrate

base is stretched for a better visualization of the polymer structures (Figure 6-22). This three-

dimensional view shows, as already seen in the graphs and images before, that with even an

Page 169: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 155

apparently sharp edged microstructure there is a slope from top level down to the base. In

the 3D view the polymeric pattern looks like single pillars.

Polymer micropatterns built by cross-linked P(DMAA-co-MABP) copolymers are

investigated as well. P(DMAA) networks are very attractive for applications where good

swellability in polar solvents is needed. With the chosen dip coating conditions relatively

thick copolymer layers are deposited on the wafer surface. These layers can be structured in

an exact way with a clear cut between the single pattern elements as it is depicted in the

following images (Figure 6-23 to Figure 6-25).

Figure 6-23. CCD image of an evolved P(DMAA) network after UV irradiation with a mesh 200

TEM grid.

Figure 6-24. Complementary imaging ellipsometry micrographs of the P(DMAA) network shown in

Figure 6-23.

Page 170: through Surface-attached Monomers

156 6 Polymer Systems on Silicon Wafers

The evaluation of the ellipsometric data (Figure 6-24) by calculating a 3D image (Figure

6-25) corroborates the assumption that the P(DMAA) networks are structured more clearly

with definite contours than PS networks with their rounded profiles.

Figure 6-25. This image shows a 3D relief computed from the ellipsometric data shown in Figure 6-24.

The pattern elements have rectangular shape and well defined edges.

Figure 6-26. CCD image of a P(DMAA) network built by photolithography using a TEM grid with

line bars having a spacing of 64 µm.

That cross-linked P(DMAA-co-MABP) copolymers have a much sharper confined

structure after the photolithography process is confirmed in the next example where a TEM

grid with line bars was used. For the sample shown in Figure 6-26 profilometric

Page 171: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 157

measurements give the following dimensions: The height of the polymer structures is

~670 nm and the lateral dimensions are 38 µm for the polymer network and 26 µm for the

base substrate crossing perpendicular to the orientation of the polymer walls.

In this case the profilometric findings are supported by data from imaging ellipsometry

(Figure 6-27). Steeper slopes at the edges of the polymeric structures provide better contours

(Figure 6-28) than it is the case for the PS network structures.

Figure 6-27. Complementary ellipsometry images focused on P(DMAA) network structures built by

photo cross-linking with a TEM grid with line bars.

Page 172: through Surface-attached Monomers

158 6 Polymer Systems on Silicon Wafers

Figure 6-28. A three-dimensional image has been calculated from the ellipsometric data shown in

Figure 6-27. The polymer walls are well confined.

With the P(DMAA-co-MABP) copolymers we are in the position to create small, separated

and surface-attached microstructures with a simple photolithographic process as the last

example in Figure 6-29 and Figure 6-30 shows.

Figure 6-29. Microstructures formed by UV cross-linking through a mesh 400 TEM grid are well

applied to P(DMAA) copolymer. Each square has a side length of nearly 40 µm.

Page 173: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 159

Figure 6-30. Complementary imaging ellipsometry pictures mapping the P(DMAA) network

structures shown in Figure 6-29.

The last step of our proposed method is the application of a ultrathin functional copolymer

layer onto the substrate copolymerizing with the surface-attached MPS. This step has been

chosen because this functionalization is not accessible by simple printing of a copolymer

onto a substrate and subsequent fixation with any technique. Several problems were

encountered during realization of this reaction step. Preparation of the polymerization

solution for the final polymerization step and degassing using freeze & thaw cycles leads to

fissured and detached polymer networks. The microstructures could still be identified but

they were not fully intact anymore. This problem is successfully solved by putting the wafer

samples into the polymerization solution after the degassing step. The functional groups of

the second ultrathin copolymer layer between the first copolymer networks are tested by

reaction with appropriate fluorescence dyes. Upon analysis with the fluorescence reader it

turned out that fluorescence has the highest intensity in the areas of the network structures

and not between the patterned network where only the functional copolymer is attached to

the surface. This leads to the assumption that the copolymer network structures do not

prevent attachment of the functional copolymer in the final polymerization step in their

areas. That is plausible because photo attachment of the first copolymer via benzophenone

groups onto the MPS-modified silicon wafer is strong enough to fix the polymer network

layer to the surface but the most MPS units are still active underneath the polymer network

after the photolithographic step. The other way round the attached polymer network even

Page 174: through Surface-attached Monomers

160 6 Polymer Systems on Silicon Wafers

enhances the probability of entanglements of functional free copolymer coils within the

network besides the surface attached polymer chains. This leads to higher fluorescence

intensities in areas where the polymer network is attached.

6.2.3 Discussion & Conclusion Micropatterning

In the first set of experiments it is shown that surface attached monomers can be actively

used for building polymer microstructures. Upon deep UV irradiation through a mask we

get fast ablation of the MPS layer. This is not surprising because this layer is according to the

ellipsometric measurements just around 1 nm thin. The subsequent radical polymerization

creates a polymer monolayer in the non-illuminated areas. The thickness of the resulting

polymer monolayer is according to the polymerization conditions as it has been determined

in [153] and paragraph 6.1. The size of single pattern elements can be varied down to a few

micrometers as we have shown with imaging ellipsometry. This patterning technique may

be interesting when using a combination of monomers introducing functionalities which

lead to applications we already have shown as examples. Especially bioanalytical purposes

using fluorescence are applicable to this micropatterned system. The advantage of this

system is the simplicity of generating micropatterns with a broad range of different (co-

)polymers and the stability of the polymer monolayer as it is covalently bound to the

substrate surface. Drawback of this method is on the one hand that immobilization of MPS is

limited to oxidic surfaces and in lab analytics glass is more and more substituted by plastic

substrates. On the other hand there are significant amounts of polymer generated which are

not bound to the substrate surface even when the substrates are densely packed inside the

polymerization solution through appropriate holders.

With the combination of a benzophenone containing copolymer and a methacryloyl propyl

trimethoxysilane one is able to form micropatterned polymer networks and attach them

covalently to the surface. The advantage of MPS layers is that they are very stable compared

to the so far often for surface attachment used benzophenone silane monolayers. MPS-

modified silicon wafers stay active for copolymerization with other monomers or attach by

benzophenone after weeks of storage in air under normal conditions in the presence of

oxygen and humidity and exposure to daylight. In the past it has been shown that a

benzophenone content of 1 mol% (and below) in the copolymer is sufficient to build stable

polymer networks[60]. But here we discover that such a benzophenone content is also

sufficient to fix a polymer network via covalent bonds effectively to surface that offers

Page 175: through Surface-attached Monomers

6 Polymer Systems on Silicon Wafers 161

aliphatic carbon-carbon bonds. In principle also alkyl silanes would do the trick but if as the

surface attached silane layer a monomer is used, it can be employed for another

polymerization step.

The evaluation of the micrographs reveals that the quality of the mask is important for the

quality of the polymer network’s shape. That means a mask for the patterning of

benzophenone copolymer layers has to be strictly planar and must lie closely on the layer in

order to get detailed and sharp edged pattern elements. We have observed a persistent

difference between PS and P(DMAA) networks regarding the slope between polymer

network surface and substrate surface. Other factors that effect the section profile are the

molecular weight in combination with the benzophenone content of the copolymer. The used

P(S-co-MABP) got higher molecular weight and higher benzophenone content by nearly a

factor of 10 in comparison with the used P(DMAA-co-MABP). These two copolymer inherent

factors provide for an extension of the polymer network into the threshold between

illuminated and non-illuminated areas. This accounts for the previously mentioned

differences documented and proven by profilometry and imaging ellipsometry. In general

well defined structures in the micrometer range are possible to build by polymer cross-

linking.

The second polymerization step was not successful the way it was intended. Blocking an

area with surface attached polymer networks for functionalization with a second copolymer

is not possible with the chosen network density and polymer mesh size respectively. But

these findings show the way to go: Application of patterned polymer networks onto a

surface where the apparent density of functional/reactive groups can be increased by using

three dimensional structures. With irradiation of functional terpolymers (base unit +

benzophenone unit + reactive unit) the combined advantages of patterns and polymer

networks on solid substrates can be used as it is similarly filed by Johnson et al.

(Motorola)[227] for immobilizing biomolecules and cross-linking copolymers in a single step.

However for simple deposition of functional copolymers patterns printing is faster and more

effective concerning resources, e.g. monomers, polymers and cross-linking agent and

photoresist techniques for patterning[228] are widely acknowledged as well. For future

research this has to be taken in account and it should come into focus if substrates modified

with immobilized monomers bring improvements for applicable systems based on

imprinting.

Page 176: through Surface-attached Monomers

162 7 Outlook

7 Outlook

This work has described the mechanism of the buildup of polymer monolayers using

surface-attached monomers. Having explored these details for two basic polymer (PS and

P(DMAA)) on various porous and flat substrates it has turned out that the influence of the

polymerization parameters on the resulting polymer monolayers is strongly dependent from

the used monomer-solvent system. Therefore we suggest also to have more detailed looks on

further monomers building polymer monolayers and if possible allocate them to certain

polymerization groups dependent of their polymerization behavior (e.g. “styrene type”, or

“acrylic type”). Although efforts to reduce the amount of generated free polymer during the

grafting polymerization process have been successful, there is still some room left for

improvements. In this regard one may speculate/dream about performing a polymerization

with monomer modified substrates where the monomer is nearly completely consumed in

the generation of the surface-attached chains.

If one wants to study copolymer monolayer systems in more detail it has to be checked

which other monomer units bearing desired functional groups can be incorporated into the

polymer layer and to overcome solubility or copolymerization problems or other restrictions,

in order to drive the specific loading of the surface-attached functional groups to even higher

values.

Especially in the field of applications for functional hybrid system many possible

applications have been tested in experiments and completed successfully. A promising case,

immobilization of Grubbs catalyst and subsequent ring closing metathesis, is investigated in

great detail by our cooperation partners at the Institute for Organic Chemistry, University of

Freiburg. As life science and bio analysis are future markets similar endeavors should be

attempted for the suggested biomolecule immobilizations. When combined with spatial

patterning through photolithographic techniques the “grafting-through” system might

evolve as a good toolbox for various chip applications, e.g. “lab on a chip”.

Page 177: through Surface-attached Monomers

8 Experimental I – Materials & Methods 163

8 Experimental I – Materials & Methods

8.1 Materials

8.1.1 Material List Acetone .................................................................................................................................................................. (p.a., Fluka)

Acryloyl chloride ................................................................................................................................................. (96%, Fluka)

Aerosil 300 ......................................................................................................................................................... (Degussa AG)

Aluminum oxide, basic ............................................................................................................................................... (Fluka)

2,2’-Azinobis(3-ethylbenzothiazoline-6-sulfonic acid) diammonium salt................................. (99.9%, Sigma-Aldrich)

2,2’-Azobisisobutyronitrile ................................................................................................................................(p.a., Merck)

Benzene ............................................................................................................................................................... (>96%, Fluka)

Fluorescein biotin ............................................................................................................................................... (90%, Sigma)

Calciumhydride .................................................................................................................................................... (p.a., Fluka)

Chloroform ........................................................................................................................................... (p.a., Riedel de Haën)

Chloroform-d3 ................................................................................................................................................ (>99.8%, Fluka)

Copper-I-Chloride ................................................................................................................................................ (p.a., Fluka)

Dichloromethane .................................................................................................................................................. (p.a., Fluka)

N,N’-Dicyclohexylcarbodiimide .................................................................................................................. (puriss., Fluka)

Diethylether ......................................................................................................................................(technical grade, Fluka)

N,N’-Dimethylacrylamide .................................................................................................................................. (p.a., Fluka)

Dimethylsulfoxide ............................................................................................................................................ (puriss., Roth)

Dimethylsulfoxide-d6 .................................................................................................................................... (>99.8%, Fluka)

DY-635-NH2 .................................................................................................................................................(Dyomics GmbH)

Ethanol ...............................................................................................................................................(technical grade, Fluka)

Ethanol ..................................................................................................................................................................(p.a., Merck)

Fluorescein ........................................................................................................................................................ (p.a., Aldrich)

Fluorescein-5-isothiocyanate .......................................................................................................................(Sigma-Aldrich)

Glass beads (details see below) ..................................................................................... (research samples, Grace GmbH)

3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10-Heptadecafluorodecylacrylat ............................................................(purum, Aldrich)

Hydrazine hydrate .........................................................................................................................................(50-60%, Fluka)

4-Hydroxybenzophenone ............................................................................................................................. (purum, Fluka)

N-Hydroxysuccinimido biotin .......................................................................................................................... (98%, Sigma)

Isopropanol .......................................................................................................................................... (p.a., Riedel de Haën)

N-Isopropyl acrylamide ...................................................................................................................................... (97%, Fluka)

LiChrospher Si60.......................................................................................................................................................... (Merck)

Methacrylic acid ................................................................................................................................................... (98%, Fluka)

Methanol ......................................................................................................................... (technical grade, Riedel-de-Haën)

Methanol ..............................................................................................................................................................(p.a., Merck)

Methyl methacrylate............................................................................................................................................ (99%, Fluka)

Page 178: through Surface-attached Monomers

164 8 Experimental I – Materials & Methods

Methylamine ................................................................................................................................................... (puriss., Fluka)

Oligonucleotides (details see below)....................................................................... (purified by HPLC, Thermo Hybaid)

Potassium peroxodisulfate ................................................................................................................................(>98%, Fluka)

Sodium ..........................................................................................................................................................................(Merck)

Sodium carbonate ............................................................................................................................................... (p.a., Merck)

Sodium hydroxide ........................................................................................................................................ (purum, Merck)

Sodium sulfate .................................................................................................................................................... (p.a., Merck)

Streptavidin ................................................................................................................. (from Streptomyces avidinii, Sigma)

Styrene ..................................................................................................................................................................(99%, Fluka)

Toluene ................................................................................................................................................(p.a., Riedel-de-Haën)

1,1,2-Trichloro trifluoro ethane .................................................................................................(Freon© 113, 99.5 %, Fluka)

Triethylamine .................................................................................................................................................... (>99.5., Fluka)

8.1.2 Material Preparation

The silica gels LiChrospher Si60 (LC700, Merck) and Aerosil300 (AE300, Degussa AG) were

dried at elevated temperature and reduced pressure for 3 hours (120 °C, 10-2 mbar). The

mesoporous glass beads of different sizes and specific surfaces (GB40, GB80, GB250, Grace

GmbH & Co. KG) were dried under the same conditions for 1 hour. The specifications of the

different solids are listed in Table 8-1:

Table 8-1. Specifications of the different silica gels and glass beads used for grafting polymer

monolayers

solid LiChrospher Si60

LC700

Aerosil300

AR300

glass bead

GB250

glass bead

GB80

glass bead

GB40

specific surface

(m2·g-1)

700 300 250 80 40

particle size

(µm)

~25 10-20 70-100 100-300 90-130

pore size

(nm)

< 2 - 20 50 100

3-methacryloylpropyl trimethoxysilane (MPS, Fluka), 3-acryloylpropyl trimethoxysilane

(APS, Fluka), octyl trimethoxysilane (OS, Fluka), and azobis(isobutyronitrile) (AIBN, Fluka)

were used as received. Styrene (Fluka) was chromatographically purified over basic

Page 179: through Surface-attached Monomers

8 Experimental I – Materials & Methods 165

aluminum oxide, distilled under reduced pressure from copper(I)chloride (60 °C, 50 mbar)

and stored under dry nitrogen at -20 °C until used. N,N’-Dimethylacrylamide was also

chromatographically purified and destabilized over basic aluminum oxide, distilled under

reduced pressure from copper(I)chloride (80 °C, 20 mbar) and stored under dry nitrogen at

-20 °C until used. Toluene (Fluka) was dried and distilled using molten sodium and

benzophenone as an indicator. After distillation it was stored over molecular sieve under dry

nitrogen. Triethyl amine (Fluka) was dried and distilled over calcium hydride prior to use.

N-methacryloyl-N-methyl-propyl trimethoxysilane (MNPS) and N-acryloyl-N-methyl-

propyl trimethoxysilane (ANPS) were synthesized from 3-amino-N-methyl-propyl

trimethoxysilane (ABCR) and methacryloyl chloride (Fluka), or acryloyl chloride (Fluka).

Coupling of the methoxysilane and the acrylic unit was performed in dichloromethane at

0 °C under nitrogen. Triethylamine was used to capture hydrochloric acid. After completion

of the reaction the solvent is removed and residuals resolved in diethylether. Solids were

filtered off. After removal of ether the product was obtained as a brown oil. Further

purification was done with a small silica gel column and leaves a yellow oil. The

characterization was done in the analytical labs of the Department for Organic Chemistry,

Albert-Ludwigs-University, Freiburg.

1H-NMR (500 MHz, CDCl3, TMS):

δ = 5.51 (bs, 1H, vinyl), 5.00-5.03 (m, 1H, vinyl) 3.55-3.60 (m, 9H, -OCH3), 3.27-3.45 (m, 2H, -CH2N), 2.90-3.03 (m, 3H, -NCH3), 1.96 (s, 3H, -CH3), 1.53-1.75 (m, 2H, -CH2-), 0.50-0.68 (m, 2H, -CH2Si).

13C-NMR (100 MHz, CDCl3, TMS):

δ = 168.8; 141.3; 115.0; 53.2; 50.6; 36.5; 21.8; 20.7; 6.3.

MS (EI, 200 °C, 70 eV):

m/z (%): 261 (22) [M+], 229 (37) [M+-MeOH], 160 (37), 121 (77) [MeO)3Si+], 112 (100).

Page 180: through Surface-attached Monomers

166 8 Experimental I – Materials & Methods

8.2 Methods

8.2.1 Elemental Analysis

All elemental analysis measurements were carried out on a Vario EL

(Elementaranalysensysteme GmbH, Germany) at the Institute for Inorganic and Analytical

Chemistry, University of Freiburg.

8.2.2 Infrared Spectroscopy

The Fourier transform infrared (FT-IR) spectra were recorded on a BioRad Excalibur FTS

3000 spectrometer collecting 128 scans within a wavenumber range of 4000 to 400 cm-1 and a

resolution of 4 cm-1. For measurements on modified silica the device was equipped with a

DRIFT compartment and purged with nitrogen to reduce carbon dioxide and water bands

from the gas phase. The incidence angle was 45° and dry potassium bromide was used for

sample preparation. The spectra evaluation was done with the ʺWin-IR Proʺ (v2.95) software.

8.2.3 Gel Permeation Chromatography (GPC)

For GPC measurements an Agilent 1100 system with ʺWinGPC scientific (v6.20)ʺ software

from PSS (Polymer Standards Service, Mainz) was used. In the case of polystyrene samples the

SDV oligo column system was run with tetrahydrofurane as eluent and calibrated with

polystyrene standard samples with polystyrene with a narrow molecular weight

distribution. For the poly(N,N-dimethyl acrylamide) samples the GRAM column system was

used with DMF as eluent and calibrated with PMMA standard samples. The separations

were carried out at a flow rate of 1 mL·min-1. All calibration polymers were bought from PSS.

8.2.4 Ellipsometry & Imaging Ellipsometry

An ELX-2 ellipsometer (Riss, Germany) was used for null ellipsometry. With an incident

angle of 70° the planar substrates were exposed to a Helium-Neon-laser with a wavelength

of 632.8 nm. The refractive indices used for the layer models for the software to perform

correct calculations were as follows: a) air nD = 1.000, b) silicon nD = 3.882, c) SiO2 nD = 1.457,

d) MPS nD = 1.431, e) PS nD = 1.450.

Page 181: through Surface-attached Monomers

8 Experimental I – Materials & Methods 167

For imaging ellipsometry a Nanofilm I-elli 2000 device was used. With an incidence angle of

50° a He-Ne-laser with a wavelength of 632.8 nm was led onto the patterned, planar

substrates.

8.2.5 X-ray Photoelectron Spectroscopy

X-ray photoelectron spectra (XPS) were recorded by using Mg Kα radiation using a Perkin

Elmer PHI 5600 spectrometer. The analyzer was positioned at an angle of 45° relative to the

substrate holder. For survey scans the step width while collecting the spectra was 0.8 eV with

a pass energy of 187.85 eV and a step time of 100 ms. For detail scans a step width of 0.06 eV

and a pass energy of 23.50 eV were used.

8.2.6 Nuclear Magnetic Resonance Spectroscopy 1H and 13C nuclear magnetic resonance (NMR) spectra were obtained from a Bruker Avance

250 MHz spectrometer. All substances and polymers were dissolved in deuterated solvents

(CDCl3, D2O or (CD3)2SO) and the solutions were measured at room temperature.

8.2.7 UV/vis Spectroscopy

For the UV/visible absorption measurements a Varian Cary 50 Bio spectrometer was used at

a wavelength range from 800 down to 200 nm with variable scan rates. In most cases the

“medium” speed program with 1 nm interval at 10 nm per second was used. In some cases

the “slow” speed program with 0.5 nm interval and 5 nm per second was used for better

resolution.

8.2.8 Fluorescence Microscopy & Reading

An Olympus Fluoview confocal scanning laser microscope including the Fluoview software

(v1.26) was used for microscopical fluorescence investigations. It is equipped with an

Krypton-Argon laser unit that emits light at 488 nm and 568 nm. For the FITC-stained

sampled the 488 nm excitation wavelength is used. FITC has its emission peak at 518 nm and

is exclusively detected by the microscope optics using colored filters. The laser intensity was

set to 20% output to minimize bleaching of the dye and not to overload the detector. To have

a high vertical resolution an aperture size of 60 µm (Aperture 1) was used. Result of the

measurements are digital pictures.

Page 182: through Surface-attached Monomers

168 8 Experimental I – Materials & Methods

For determination of fluorescence intensities on glass beads (and silica) the biochip

analyzer BioDetect 645/4 (BioChip Technologies GmbH, Freiburg) was used. This multicolor

(4 colors) reading device can be used to excite different dyes. For our measurements FITC,

Cy5 and DY-635 were used as chromophores. For FITC the biochip analyzer uses an

excitation range from 485 to 495 nm and a detection range from 515 to 545 nm. Using Cy5 or

DY-635 the ranges are 635 to 645 nm for excitation and 671 to 693 nm for detection. The

emitted light intensities are detected by a CCD array which gives a gray scale picture. For

better distinction of the different intensities the gray scale pictured are calculated into false

color pictures (Figure 8-1).

Figure 8-1. Chart for transfer from a CCD detector gray scale image to a false color image.

8.2.9 Profilometry

Profilometric line scans on patterned surfaces were done on a Tencor Alpha-Step IQ

profilometer. The survey scan width was 1000 µm with a vertical resolution of 0.8 nm and

the detail scan width was 130 µm with a vertical resolution of 0.08 nm.

8.2.10 Titration

The amount of amino groups on copolymer modified silica was determined by titration

with a TitroLine 96 (Merck Eurolab, Germany) pH titration device using a Blueline pH11

(Schott, Germany) pH-electrode which was calibrated with buffer solutions at pH = 4 and

pH = 7.

Page 183: through Surface-attached Monomers

9 Experimental II – Polymerizations 169

9 Experimental II – Polymerizations

In order to obtain a more detailed understanding of the reaction during the polymerization

reactions always one parameter was varied while the other were kept constant. Therefore the

preparation of the individual samples differs and is described for each varied parameter

separately. Degassing, isolation and drying procedures were carried out as described in

paragraph 9.3. Ingredients and reaction vessels were all strictly kept under dry nitrogen.

Larger scale polymerizations (> 50 mL reaction volume) were carried out in appropriate

Schlenk flasks with reflux condenser. 50 mL-Schlenk tubes were used as reaction vessel for

smaller scale polymerization reactions. The silica gel with the attached polymer monolayer

was separated through centrifugation and repeated washing with a good solvent for the

polymer. The free polymer was precipitated from the reaction mixture through a slow

addition of the solution to 10x excess methanol, filtered off and dried. The molecular weights

were determined by GPC. After several washings the modified silica gel was dried from

benzene and analyzed qualitively by FT-infrared spectroscopy and quantitively by elemental

analysis.

9.1 Immobilization of Silanes

9.1.1 Immobilization of Trimethoxy Silanes onto Silica Gel

In an atmosphere of dry nitrogen 6 g silica gel were suspended in 150 mL toluene. A

solution of 2.08 g MPS (8.4 mmol) in 50 mL toluene and 10 mL triethyl amine (~74 mmol)

were added to the mixture. The reaction vessel was heated until the solution started to reflux

(120°C) and stirred for 3 h. To avoid grinding of the silica beads with magnetic stir bars the

reaction mixture was agitated with a circular shaking device (IKA KS 260 control) at 160

rounds per minute (rpm).

The modified silica gels were centrifuged (Sorvall Super T21, 12500 rpm, 16.750 x g) for 15

minutes in the case of LiChrospher and 30 minutes for Aerosil300. Then washed with toluene,

ethanol, ethanol/water (1/1 v/v, acidified with HCl), ethanol/water (1/1 v/v), ethanol and

diethyl ether. The remaining colorless solid was dried for 18 h at 10-2 mbar. The reaction

conditions for immobilization of all monomers are summarized in Table 9-1.

Page 184: through Surface-attached Monomers

170 9 Experimental II – Polymerizations

Table 9-1 Reaction conditions for the immobilization of different silanes on varying solid supports.

substrate n (silane)

(mmol)

m (silica)

(g)

V (toluene)

(mL)

n (Et3N)

(mmol)

LC700 + MPS 8.4 6.0 200 54

LC700 + MNPS 2.4 2.0 66 18

LC700 + APS 9.0 6.0 200 50

LC700 + ANPS 12.5 9.0 300 50

LC700 + OS† 1.1 0.5 20 6

AR300 + MPS 8.4 6.0 200 54

GB250 + MPS 2.5 5.0 60 21

GB80 + MPS 4.2 10.0 125 36

GB80 + MNPS 1.0 2.0 25 7

GB80 + APS 2.1 4.0 50 14

GB80 + ANPS 4.2 10.0 80 28

GB40 + MPS 1.0 4.0 25 7

GB40 + MNPS 0.5 2.0 13 4

† reference sample with surface-attached octyl silane monolayer

SiO2-MPS. IR (DRIFT) ν [cm-1]: 3112, 2986, 2961, 2899, 2857, 1724, 1704, 1637, 1250-

1000 cm-1. elemental analysis: LC700-ME: C 7.30 %, H 1.15 %, AR300-ME: C 6.55%, H 0.92%,

GB250-ME: C 4.76 %, H 0.84%, GB80-ME: C 1.95 %, H 0.30 %, GB40-ME: C 1.33 %, H 0.26 %

SiO2-APS. IR (DRIFT) ν [cm-1]: 3113, 2959, 2901, 2851, 1733, 1714, 1627, 1250-1000 cm-1.

elemental analysis: LC700-AE: C 6.02 %, H 1.05 %, GB250-AE: C 4.76 %, H 0.84%, GB80-AE:

C 1.71 %, H 0.43 %

SiO2-MNPS. IR (DRIFT) ν [cm-1]: 3091, 2981, 2944, 2892, 1647, 1603, 1250-1000 cm-1.

elemental analysis: LC700-MA: C 7.94 %, H 1.38 %, N 1,31 %, GB80-MA: C 2.22 %, H 0.24%,

N 0.24%, GB40-MA: C 1.89 %, H 0.28 %, N 0.31 %

Page 185: through Surface-attached Monomers

9 Experimental II – Polymerizations 171

SiO2-ANPS. IR (DRIFT) ν [cm-1]: 3108, 2947, 2896, 2851, 1645, 1590, 1250-1000 cm-1.

elemental analysis: LC700-AA: C 6.45 %, H 1.00 %, N 0.84 %, GB250-AA: C 5.18 %, H 0.96%,

N 0.70%, GB80-AA: C 1.92 %, H 0.47 %, N 0.24 %

SiO2-OS. IR (DRIFT) ν [cm-1]: 2986, 2961, 2932, 2899, 2857 cm-1. elemental analysis:

LC700-OS C 6.50 %, H 1.42 %

9.1.2 Immobilization of Trimethoxy Silanes onto Glass Beads

Under dry nitrogen glass beads were suspended in toluene and a toluene solution of the

silane monomer unit and triethyl amine were added to the mixture. The flask with the

reaction mixture was heated until the solution started to reflux (120°C). Agitation for 3 h was

done with a circular shaking device at 160 rpm. The immobilization recipes are summarized

in Table 9-2. In case of the glass beads sedimentation was fast enough to skip centrifugation.

Washing with toluene, ethanol, ethanol/water (1/1 v/v, acidified with HCl), ethanol/water

(1/1 v/v), ethanol and diethyl ether removes excess triethyl amine. The remaining colorless

solids were dried for 18 h at 10-2 mbar.

Table 9-2. Mixtures for the immobilization of different silanes onto glass beads.

substrate & surface

modification*

n (silane)

(mmol)

m (silica)

(g)

V (toluene)

(mL)

n (Et3N)

(mmol)

GB250-ME 2.5 5 60 21

GB250-AA 2.5 7 60 21

GB80-ME 4.2 10 125 36

GB80-MA 3.9 10 100 35

GB80-AE 2.1 4 50 14

GB80-AA 4.2 10 80 28

GB40-ME 1.0 4 25 7

GB40-MA 0.5 2 13 4

Page 186: through Surface-attached Monomers

172 9 Experimental II – Polymerizations

9.1.3 Co-Immobilization of MPS and OS onto Silica Gel

MPS and OS were added to the reaction solution in certain molar ratios (4/1, 1/1, 1/4, 1/19,

pure OS). The composition of each batch is given in Table 9-3. The mixtures, heated to reflux

by an oil bath (120 °C), were agitated by circular shaking for 3 h at 160 rpm. After the

reaction was completed and between each washing step during work up, the modified silica

gels were centrifuged (15 minutes at 12500 rpm). They were washed with toluene, ethanol,

an mixture of ethanol/water (1/1 v/v, acidified with HCl), ethanol/water (1/1 v/v), ethanol

and diethyl ether. Remaining colorless solids were dried for 18 h (overnight) at <10-2 mbar.

Table 9-3. Reaction mixtures for the co-immobilization of MPS and OS.

batch (MPS/OS)

(n/n)

V (MPS)

(µL)

V (OS)

(µL)

m (LiChrospher)

(mg)

V (toluene)

(mL)

n (Et3N)

(mL)

1/0 2000 - 6000 200 10.0

4/1 190 50 715 28 1.0

1/1 120 130 715 28 1.0

1/4 50 200 715 28 1.0

1/19 15 240 715 28 1.0

0/1 - 300 509 20 0.7

The actual ratio between immobilized “active” silane (with double bond) and immobilized

“inactive” silane (without double bond) was examined by FT-infrared spectroscopy To get

an idea of the real silane composition on the surface, as it that may differ from the

composition of the initial immobilization mixture, the DRIFT data were compared to FT-IR

data of liquid mixtures with know ratios of MPS and OS. The quotient of the integral of the

C=O-valence vibration band divided by the integral of the C-H-valence vibration band of

methylene gave the extinction ratio. From the transmission-IR measurements of the liquid

silanes a calibration curve was generated. The initial DRIFT data points were aligned to the

calibration curve in order to get the surface composition of the silanes. For details

cf. paragraph 3.3.7.1.

Page 187: through Surface-attached Monomers

9 Experimental II – Polymerizations 173

9.2 Homopolymerizations

9.2.1 Synthesis of Poly(3,3,4,4,5,5,6,6,7,7,8,8,9,9,10,10,10-

heptadecafluorodecyl acrylate)

1.0 g of the solid phase with surface attached monomers is suspended in 7.5 mL Freon®-113

(1,1,2-trichloro-trifluoro ethane) and 2.5 mL distilled HDFDA are given to the suspension.

Then 24.6 mg AIBN as initiator are added and dissolved in the polymerization solution. The

whole mixture then is degassed in vacuum by six freeze & thaw cycles and finally heated to

50.0±0.1 °C with a water thermostat. After 24 h the polymerization was stopped by cooling

down. The suspensions with LiChrospher were centrifuged at 12500 rpm for 20 min for

separation of the silica gel and the glass beads were isolated by sedimentation. The solids

were put into a filter inside of an extraction case and extraction at 48 °C with Freon®-113 was

carried out over night (~ 18 h). The substrates with the perfluoro polymer monolayer then

were qualitively analyzed by x-ray photoelectron spectroscopy (cf. Table 3-3) and DRIFT and

quantitively by elemental analysis.

SiO2-P(HDFDA). IR (DRIFT) ν [cm-1]: 2955, 2851, 1738, 1193 cm-1, elemental analysis:

LC700-AE-P(HDFDA) C 15.48 %, H 1.19 %, GB80-AA-P(HDFDA) C 11.13 %, H 0.76 %,

N 0.17 %.

9.2.2 Synthesis of Polystyrene & Poly(methyl methacrylate)

First the solid phase is suspended in toluene and the monomer is added. Afterwards AIBN

is dissolved in the polymerization solution. The whole mixture is then degassed in vacuum

through five freeze & thaw cycles and finally heated with a water thermostat (60.0±0.1 °C).

After 25 h of polymerization for styrene and 16 h for MMA the polymerizations were

stopped and cooled down. Then the polymerization solutions with LiChrospher were

centrifuged at 12500 rpm for 20 minutes in order to separate the silica gel. In the mixtures

with glass beads sedimentation was fast. After decanting, the remaining solids were put into

a filter inside of an extraction case and hot extraction with toluene was carried out over night

(~ 18 h). The substrates modified with polystyrene and poly(methyl methacrylate)

monolayers were freeze-dried from benzene. All samples were analyzed by DRIFT and

elemental analysis. The different free polymers were precipitated from the reaction solution

through a slow addition to 10x excess of methanol. The precipitates were filtered off and

dried. The molecular weights were determined by gel permeation chromatography (GPC).

Page 188: through Surface-attached Monomers

174 9 Experimental II – Polymerizations

A glass bead sample with surface-attached polystyrene monolayer was also characterized by

XPS (cf. Table 3-3).

SiO2-PS. IR (DRIFT) ν [cm-1]: 3080, 3061, 3028, 2920, 2852, 1724, 1602, 1493, 1450, 1200-1000,

762, 697 cm-1, elemental analysis: LC700-ME-PS C 32.85 %, H 3.45 %, GB80-AA-PS C 11.13 %,

H 0.76 %, N 0.17 %.

SiO2-P(MMA). IR (DRIFT) ν [cm-1]: 2993, 2951, 2846, 1730, 1481, 1450, 1200-1000 cm-1,

elemental analysis: LC700-ME-P(MMA) C 27.51 %, H 4.04 %, GB80-ME-P(MMA) C 16.46 %,

H 2.38 %.

9.2.3 Synthesis of Poly(N-isopropyl acrylamide)

In 12 mL of MilliQ purified water 3.0 g NIPAAm and as water-soluble initiator 29 mg

(108 µmol) potassium peroxodisulfate were dissolved within 30 minutes under rapid

stirring. Then 300 mg of either LC700-ME, or GB80-ME are suspended in the solutions for the

polymerization reaction. By ultrasonification under reduced pressure the mixtures were

degassed. For polymerization the reaction vessels were heated at 60.0±0.1 °C with a water

thermostat. After a time period of 16 hours gelation of the solutions was observed and the

polymerization reactions were terminated by cooling down. The gelatinous mixture

including the solids was dissolved in methanol to reveal the substrates with the surface-

attached P(NIPAAm) monolayers. The solids were filtered off, extracted with hot methanol

for 2 days to remove free adsorbed P(NIPAAm) and freeze-dried from water. The resulting

solids were characterized by elemental analysis and DRIFT. The free polymer was

precipitated from the filtrates through slow addition to a 10fold excess of a mixture of

diethyl ether and acetone (2/1 v/v), filtered off and dried. GPC measurements of the free

P(NIPAAm) with the SUPREMA column system were not possible, because even very dilute

solutions in water could not be filtered with a 0.2 µm filter.

SiO2-P(NIPAAm). IR (DRIFT) ν [cm-1]: 2972, 2938, 2881, 1650, 1530, 1460, 1200-1000 cm-1,

elemental analysis: LC700-ME-P(NIPAAm) C % 20.60, H % 3.57, N % 3.75,

GB80-ME-P(NIPAAm) C % 16.51, H % 3.15, N % 2.34.

Page 189: through Surface-attached Monomers

9 Experimental II – Polymerizations 175

9.2.4 Synthesis Poly(N,N-dimethyl acrylamide)

300 mg LiChrospher or 400 mg glass beads, both modified with surface-attached

monomers, were given into a solution of DMAA in DMF. The ratio of the two liquid

components was in most cases 1 part DMAA in 9 parts DMF. The initiator AIBN was added

to the polymerization solution (10 mmol·L-1 for most experiments). The solution then was

degassed in vacuum by five freeze & thaw cycles. A water thermostat kept the

polymerization at 60.0±0.1 °C. After 24 to 25 h polymerization time the reactions were

stopped by cooling down and air exposure. For separating the silica gel the polymerization

solutions with LiChrospher inside were centrifuged at 12500 rpm for 20 minutes. The

sedimentation of the glass beads was fast although the reaction solutions were quite viscous.

After decanting, the remaining solids were washed once with DMF and the supernatant

solution was discarded. The solids were collected into a filter paper inside of an extraction

case and extracted with DMF inside of a Soxhlet apparatus overnight (~ 18 h). Last step was

freeze-drying from water. Elemental analysis and DRIFT were used to characterize the

substrates with the P(DMAA) monolayer. The decanted polymerization solution was slowly

poured under fast stirring into 10fold excess of diethyl ether to precipitate the free

P(DMAA). The free polymers were filtered off, dried under vacuum and characterized by

molecular weight determination via GPC. If any amount of substrate, monomer, solvent,

initiator or the polymerization conditions themselves is different from this description it will

be described separately in the paragraph 9.4.

SiO2-P(DMAA). IR (DRIFT) ν [cm-1]: 2928, 1641, 1496, 1399, 1353, 1250-1000 cm-1. elemental

analysis: LC700-ME-P(DMAA) C % 11.34, H % 2.24, N % 1.33; GB80-AE-P(DMAA) C % 6.43,

H % 1.17, N % 1.11.

9.2.5 Synthesis of Poly(methacrylic acid)

300 mg of a solid compound, with monomers attached to the surface, were suspended in a

solution of 3 mL methacrylic acid and 6 mL pure water (MilliQ). Before degassing via

ultrasonification under reduced pressure for 30 minutes, 22 mg (81 µmol) potassium

peroxodisulfate were added. The temperature of 60.0±0.1 °C during the polymerization

period of 16 hours is controlled by a water thermostat. In the end gelation occurred and the

substrates with P(MAA) layer were recovered by washing with methanol and filtration.

Then the modified substrates were extracted with hot methanol in a Soxhlet apparatus for

Page 190: through Surface-attached Monomers

176 9 Experimental II – Polymerizations

2 days to remove all free P(MAA) and finally freeze-dried from water. DRIFT and elemental

analysis characterized the surface-attached polymer monolayers. The filtrate of the reaction

mixture was slowly poured into a well agitated mixture of acetone and diethyl ether (1/1 v/v)

to precipitate the free P(MAA). The polymer was filtered off, dried but GPC measurements

were not possible because the aqueous solution of the free P(MAA) went not through a

0.2 µm filter.

SiO2-P(MAA). IR (DRIFT) ν [cm-1]: 2992, 2955, 2630, 1710, 1480, 1451, 1391, 1200-1000 cm-1,

elemental analysis: LC700-ME-P(MAA) C 16.58 %, H 2.81 %; GB80-ME-P(MAA) C 19.20 %,

H 2.87 %.

9.3 Polystyrene – Polymerization Parameters

9.3.1 Influence of Reaction Time

Under dry nitrogen a monomer-modified silica substrate was suspended in a mixture of

styrene and toluene and AIBN as initiator was added (Table 9-4). The solution was degassed

in vacuum through four freeze & thaw cycles. The mixture was heated with an oil bath at

60±1 °C. After chosen time periods samples were withdrawn from the reaction under

nitrogen (for silica gels 20 mL, for glass beads 10 mL). The silica gel samples were

centrifuged at 12500 rpm (15 min LiChrospher Si60, 30 min Aerosil300) to separate the silica

gel. For all types of glass beads centrifugation was not necessary because they settle fast at

the bottom. After decanting, the remaining solids were washed with toluene (~ 40 mL)

resuspended and again centrifuged. These steps were repeated (usually five times) until no

precipitate formed when the decanted washing solution was added dropwise to excess

methanol. Finally the silica gel with the polystyrene monolayer was filtered off and freeze-

dried from benzene (7 wt.% silica gel ≈ 100 mg silica gel in 1.63 mL benzene) The

supernatant reaction mixture containing the nonattached polymer chains was slowly poured

in excess methanol (>10x volume) under rapid stirring. The white precipitate was filtered off

and dried in vacuum (<10-2 mbar) overnight.

Page 191: through Surface-attached Monomers

9 Experimental II – Polymerizations 177

Table 9-4. Mixtures for the investigations on the influence of polymerization time for the

polymerization of styrene onto different modified silica substrates.

substrate & surface modification*

m (solid X)

(g)

V (styrene)

(mL)

V (toluene)

(mL)

c (AIBN)

(mmol·L-1)

LC700-ME 2.0 66 133 9

AR300-ME low Mn 2.0 66 133 9

AR300-ME high Mn 1.2 60 60 3

GB80-ME 2.1 23 46 9

GB80-AE 1.8 30 60 3

GB80-AA 1.0 16 32 3

GB40-ME 0.9 10 20 3

9.3.2 Variation of the Overall Concentration of the Solid

The investigations were divided into two sets of experiments where one was made with

high specific surface LiChrospher silica gel and a second with porous glass beads with a lower

specific surface.

In 50 mL Schlenk tubes five different amounts of MPS-modified LiChrospher (LC700-ME =

0.15 g; 0.6 g; 1.2 g; 2.4 g and 4.8 g) were suspended under dry nitrogen in 15 mL aliquots of a

mixture consisting of 30 mL styrene, 60 mL toluene and 132 mg AIBN (0.8 mmol). The

Schlenk tubes with the degassed suspensions were mounted on a shaking device and heated

in a thermostat to 60°C for 24 hours. Afterwards polymer modified solids were separated

from the polymer solution.

Under dry nitrogen, in 50 mL Schlenk tubes, four different amounts of MPS-modified glass

beads (GB80-ME = 0.6 g, 1.2 g, 2.4 g and 4.8 g) were suspended each in a reaction mixture of

12 mL. These aliquots were taken from a mixture of 16 mL styrene, 32 mL toluene and 24 mg

AIBN (3 mM). The vessels were degassed, kept in motion by a shaking device (160 RPM) and

heated in a thermostat to 60°C for 24 hours. Then the solids were separated from the free

polymer solution.

Page 192: through Surface-attached Monomers

178 9 Experimental II – Polymerizations

9.3.3 Monomer Concentration Variation

In each of the five experiments with silica gel the total volume of solution was 40 mL with

400 mg MPS-modified LiChrospher and 20 mg (3 mM) AIBN. The input volumes of monomer

and solvent (styrene/toluene, v/v) were 1/0, 1/1, 1/2, 1/4 and 1/8. All mixtures were heated at

60±1 °C for 38.5 hours. Three experiments with styrene concentration variation and glass

beads were carried out. The total volume of solution was 20 mL with 200 mg MPS-modified

GB80 and 10 mg (3 mM) AIBN. The input volumes of styrene and solvent (v/v) were 1/0, 1/1

and 1/2. All mixtures were heated at 60±1 °C for 24 hours. Separation, isolation and analysis

follow the usual procedure.

9.3.4 Initiator Concentration Variation

The polymerizations were carried out in 40 mL styrene/toluene mixtures (1/2 v/v) with

400 mg MPS-modified silica gel each. Five different initiator (AIBN) concentrations were

delivered to the solutions: 10-4 mol⋅L-1 (0.7 mg), 10-3 mol⋅L-1 (6.6 mg), 3⋅10-3 mol⋅L-1 (19.7 mg),

10-2 mol⋅L-1 (65.7 mg), 10-1 mol⋅L-1 (656.9 mg). Then all batches were heated at 60 °C for a

duration of 26.2 hours.

The reaction mixture compositions for the investigations on the initiator influence are

summarized in Table 9-5. To the solutions five (LC700-ME, GB80-ME), four (GB80-AE) and

two (GB250-ME) different initiator (AIBN) concentrations were added and after degassing all

batches were heated to 60 °C and agitated via shaking for a duration of at least 24 hours.

Separating and isolating PS-modified substrates as well as the free monomer lead to the

results of elemental analysis and GPC.

Page 193: through Surface-attached Monomers

9 Experimental II – Polymerizations 179

Table 9-5. Ingredients list for the experiments on initiator concentration influence on polymerization

of styrene onto different solids.

m (solid X)

(mg)

V (styrene)

(mL)

V (toluene)

(mL)

substrate &

surface

modification* per concentration

c (AIBN)

(mmol)

LC700-ME 400 13.3 26.4 0.1 1 3 10 100

AR300-ME low Mn 2000 66 133 - - - 9 -

GB250-ME 600 8 4 - - 3 10 -

GB80-ME 600 8 4 0.1 1 3 10 100

GB80-AE 600 8 4 0.1 1 - 10 100

9.3.5 Variation of the Polymerization Temperature

In each reaction vessel 300 mg of MPS-modified silica gel were suspended in 20 mL

toluene, 10 mL styrene and 43 mg (0.26 mmol) AIBN under dry nitrogen. The mixtures were

heated to chosen temperatures in a range from 60 to 90 °C. The heating bath temperature

was in all experiments controlled by a contact sensor inside the Schlenk tubes. For one set of

polymerization reactions at different temperatures the polymerization time was varied so

that it amounted to exactly one half life time of the initiator. For another set of experiments

the polymerizations were performed for a constant period of 21.5 hours. For details see

Table 9-6.

Page 194: through Surface-attached Monomers

180 9 Experimental II – Polymerizations

Table 9-6 Parameters used for the investigation of the temperature dependence of the grafting process.

temperature

(°C)

decomposition rate

(10-6⋅s-1)

polymerization time

(h)

constant initiator conversion (50%)

60 9 21.50

62 12 16.05

65 19 10.85

70 40 4.00

80 150 1.05

90 490 0.40

constant polymerization time

60 9 21.50

62 12 21.50

65 19 21.50

70 40 21.50

80 150 21.50

90 490 21.50

9.3.6 Influence of MPS Surface Concentration on the Polystyrene

Graft Density

The essential step is the co-immobilization of MPS and OS as described above (9.1.3.). The

resulting different modified silica gels were suspended in a polymerization solution under

equal reaction conditions. Each 300 mg MPS/OS-modified LiChrospher were placed in 30 mL

styrene/toluene (1/2 v/v) with 43 mg (9 mM) AIBN and heated at a temperature of 60 °C for

25 hours.

Page 195: through Surface-attached Monomers

9 Experimental II – Polymerizations 181

9.3.7 Variation of the Overall Concentration of Glass Beads

Under dry nitrogen, in 50 mL Schlenk tubes, four different amounts of MPS-modified glass

beads (GB80-ME: 0.6 g, 1.2 g, 2.4 g and 4.8 g) were suspended in 12 mL aliquots of a mixture

consisting of 16 mL styrene, 32 mL toluene and 24 mg AIBN (3 mM). The vessels were

degassed, kept in motion by a shaking device (160 RPM) and heated in a thermostat to 60°C

for 24 hours. Then the solids were separated from the free polymer solution.

9.3.8 Two-step Grafting

LiChrospher samples with different polystyrene monolayers that were previously attached

to the surface with this immobilized monomer approach, were used in this set of

experiments (Table 9-7). The polystyrene monolayers, covalently bound to the surface,

consist of polymer coils of different molecular weight.

Table 9-7. Molecular weights of the PS monolayers on LiChrospher used for the two-step experiments

sample PS1 PS2 PS3 PS4 PS5

nM PS (1st) (g/mol) 230000 100000 60000 32000 6500

Five different polystyrene/silica samples (300 mg/batch) were placed into a polymerization

solution consisting of 20 mL toluene, 10 mL styrene and 246 mg (50 mM) AIBN as initiator.

Polymerization took place at 60 °C for 24 hours. After separation and freeze-drying, the

solids were again characterized by elemental analysis to evaluate the changes in polymer

load in comparison to the silica after the first polymerization step.

9.3.9 Polymerization Reactions without Surface-attached

Monomers (blind tests)

300 mg LiChrospher (used as received) and 300 mg LiChrospher with immobilized octyl

trimethoxysilane were each placed into a polymerization mixture consisting of 10 mL styrene

20 mL toluene and 43 mg (0.26 mmol) AIBN and were separately heated at 60°C for 24 hours.

The workup was carried out in the same way as described before. The results of the

experiments are listed in Table 3-4.

Page 196: through Surface-attached Monomers

182 9 Experimental II – Polymerizations

9.4 Poly(N,N-dimethyl acrylamide) – Polymerization

parameters

9.4.1 Influence of Reaction Time

Under dry nitrogen 1.2 g GB250-ME was suspended in a solution of 20 mL DMAA, 40 mL

DMF and 98.4 mg AIBN as initiator were added in a first series. For a second series 984 mg

AIBN were used in the polymerization mixture. The mixture was degassed in vacuum by

five freeze & thaw cycles. Then the Schlenk flask was heated with an oil bath at 60±1 °C and

agitated with a shaking device at 180 rpm. After chosen time periods (0.5, 1, 2, 4, 8 and 24 h)

samples of 10 mL were withdrawn from the reaction under nitrogen. The glass beads settled

at the bottom and after decanting, the solids were washed with THF and put into an

extraction thimble. After Soxhlet extraction overnight with dichloromethane the P(DMAA)-

modified glass beads were freeze-dried from water (10 wt.% glass beads) and analyzed by

DRIFT and elemental analysis. The decanted polymerization solution was added dropwise

to excess methanol under rapid stirring to precipitate the nonattached polymer chains. The

white precipitate was filtered off, dried in vacuum (<10-2 mbar) overnight and analyzed by

GPC.

GB250-ME-PDMAA. IR (DRIFT) ν [cm-1]: 2930, 1643, 1499, 1400, 1355, 1250-1000 cm-1.

elemental analysis: cf. Table 9-8.

Table 9-8. Elemental analysis and GPC measurements after polymerization for time dependence of

P(DMAA) formation on glass beads.

carbon

(%)

Γ0

(µmol·m-2)

substrate & surface

modification

c (AIBN)

(mmol·L-1)

monomer P(DMAA) monomer P(DMAA)

nM

(g·mol-1)

GB250-ME 10 4.8 20.3 2.0 0.07 9600

GB250-ME 100 4.8 20.7 2.0 0.19 25700

Page 197: through Surface-attached Monomers

9 Experimental II – Polymerizations 183

9.4.2 Monomer Concentration Variation

For each experiment the total volume of solution was 10 mL with partial input volumes of

monomer and solvent (DMAA/DMF, v/v) of 1/2, 1/4 and 1/9. 300 mg MPS-modified

LiChrospher and 16.4 mg (10 mM) AIBN were added under an atmosphere of dry nitrogen

into the Schlenk tube. All mixtures were degassed by five freeze & thaw cycles and then

heated in a thermostat at 60±1 °C and agitated by a shaker at 160 rpm for 24 hours. After

stopping the polymerization by adding air and cooling down the P(DMAA) layered silica

and the free polymer were separated, isolated and analyzed as described in 9.4.1.

LC700-ME-PDMAA. IR (DRIFT) ν [cm-1]: 2928, 1635, 1497, 1403, 1357, 1250-1000 cm-1.

elemental analysis: see Table 9-9.

Table 9-9. Elemental analysis and GPC measurements after the monomer concentration dependent

polymerization experiments for P(DMAA) on glass beads.

carbon

(%)

Γ0

(µmol·m-2)

substrate & surface

modification

DMAA/DMF

(v/v)

monomer P(DMAA) monomer P(DMAA)

nM

(g·mol-1)

LC700-ME 1/9 5.7 11.3 0.85 0.011 18300

LC700-ME 1/4 5.7 14.1 0.85 0.010 30400

LC700-ME 1/2 5.7 19.3 0.85 0.009 59400

9.4.3 Initiator Concentration Variation

The polymerizations were carried out in 40 mL-Schlenk tubes under an atmosphere of dry

nitrogen. Five different initiator concentrations were used in the experiments: 10-4 mol⋅L-1,

10-3 mol⋅L-1, 10-2 mol⋅L-1 and 10-1 mol⋅L-1 For experimental details see Table 9-10. All batches

were heated at 60 °C for a duration of 24 hours and again the P(DMAA) layered silica and

the free polymer were separated, isolated and analyzed as described in 9.4.1.

Page 198: through Surface-attached Monomers

184 9 Experimental II – Polymerizations

Table 9-10. Recipes and investigated initiator concentrations for the initiator concentration variation

experiments.

m (solid)

(mg)

V (DMAA)

(mL)

V (DMF)

(mL)

substrate &

surface

modification per AIBN concentration

c (AIBN)

(mmol)

LC700-ME 300 1 9 0.1 1 10 100

GB250-ME* 1200 20 40 - - 10 100

GB80-AE 600 1 9 0.1 1 10 100

* larger amounts because time dependence was examined in parallel

DRIFT results were shown exemplarily for glass beads and silica in 9.4.1 and 9.4.2. The

results of the elemental analysis and GPC measurements were summarized in Table 9-11.

Table 9-11. Measured carbon contents from elemental analysis, calculated graft densities and

molecular weights from GPC for the initiator concentration variation experiments.

carbon

(%)

Γ0

(µmol·m-2)

substrate & surface

modification

AIBN

(mmol·L-1)

monomer P(DMAA) monomer P(DMAA)

nM

(g·mol-1)

LC700-ME 0.1 5.7 11.1 0.9 0.011 16800

LC700-ME 1.0 5.7 11.3 0.9 0.011 18300

LC700-ME 10.0 5.7 12.0 0.9 0.011 20700

LC700-ME 100.0 5.7 11.6 0.9 0.011 18800

GB250-ME 10.0 4.8 20.3 2.0 0.069 9600

GB250-ME 100.0 4.8 20.7 2.0 0.186 25700

GB80-AE 0.1 1.7 6.4 2.3 0.068 17250

GB80-AE 1.0 1.7 6.4 2.3 0.065 18320

GB80-AE 10.0 1.7 6.4 2.3 0.055 21281

GB80-AE 100.0 1.7 6.4 2.3 0.059 20125

Page 199: through Surface-attached Monomers

9 Experimental II – Polymerizations 185

9.5 Copolymerizations

9.5.1 Synthesis of Poly(styrene-co-N-acryloyl-N-methyl-propyl

phthalimide), P(S-co-AC3pht)

The substrate with surface-attached monomers was put into a Schlenk tube and toluene

and styrene in given volumes were added. Then the desired amount of N-acryloyl-N-

methyl-propyl phthalimide (AC3pht) and AIBN as initiator were dissolved in the liquid

phase by circular shaking. The solution was degassed in vacuum by five freeze and thaw

cycles and the mixture was heated by an thermostat to 60.0±0.1 °C. After a certain period of

time the polymerization reaction was stopped by air exposure and cooling down.

Sedimentation of glass beads and centrifugation of silica gel to separate the polymer

monolayer modified solid product from the polymer solution. The supernatant solution

containing the non-attached copolymer was added dropwise to 10x excess methanol to

precipitate the free copolymer, filtered off and dried in vacuum. The solids were extracted

overnight in a Soxhlet apparatus using toluene and freeze-dried from benzene. All

polymerization conditions are listed in Table 9-12. Solids were characterized by elemental

analysis (Table 9-13) and DRIFT, free copolymers by 1H-NMR, FT-IR spectroscopy and GPC

(SDV oligo system).

Table 9-12 Ingredients and reaction times for the copolymerization of styrene as main monomer (M)

and N-acryloyl-N-methyl-propyl phthalimide as functional monomer (fM) in toluene as solvent (S)

onto different silica substrates.

substrate & surface

modification

m (solid)

(mg)

m(fM)

(mg)

c(fM)

(mol%)

V (M)

(mL)

V (S)

(mL)

m (I)

(mg)

c (I)

(mmol·L-1)

t

(h)

LC700-ME 240 946 5 8.0 16 36 9 24

LC700-ME 5000 1320 10 5.0 10# 22 9 30

GB80-ME 2000 1426 5 12.0 24 18 3 24

GB80-ME 2000 2836 10 10.8 24 17 3 25

GB80-ME 2000 5673 20 9.6 24 17 3 25

GB80-AA 13500 3160 10 12.0 100 80 5 30

GB250-AA 5000 2370 20 4.0 8# 18 9 26

Page 200: through Surface-attached Monomers

186 9 Experimental II – Polymerizations

Table 9-13. Elemental analysis results of the styrene-co-N-acryloyl-N-methyl-propyl phthalimide

monolayer modified substrates.

product C*

(%)

H*

(%)

N*

(%)

LC700-ME-PS-AC3pht(5%) 29.27 2.83 1.32

LC700-ME-PS-AC3pht(10%) 27.01 2.94 0.64

GB80-ME-PS-AC3pht(5%) 11.63 0.92 0.15

GB80-ME-PS-AC3pht(10%) 12.99 1.05 0.30

GB80-ME-PS-AC3pht(20%) 12.76 0.92 0.52

GB80-AA-PS-AC3pht(10%) 8.39 0.96 0.50

GB250-AA-PS-AC3pht(20%) 22.29 2.25 1.33

* obtained from elemental analysis

1H-NMR (250 MHz, CDCl3, TMS) δ [ppm]:

l), m) (4H, d) 7.6 – 8.0

c), d), e) (5H, m) 6.3 – 7.4

i), k) (4H, b) 3.3 – 3.7

h) (3H, b) 3.0 – 3.3

g) (1H, b) 3.0 – 2.8

j) (2H, b) 2.5 – 2.8

b) (1H, m) 1.8 – 2.5

N

NO

O

O

co

co

a)b)

c) c)

d) d)e)

f)g)

h)

j)

k)

m)m)

i)

l)

l)

a), f) (2H + 2H, m) 0.8 – 1.8

Page 201: through Surface-attached Monomers

9 Experimental II – Polymerizations 187

LC700-ME-P(S-co-AC3pht). IR (DRIFT) ν [cm-1]: 3083, 3062, 3027, 2926, 2850, 1771, 1715, 1602, 1493, 1450, 1395, 1250-1000,

759, 723, 698 cm-1.

elemental analysis: cf. Table 9-13.

GB80-AA-P(S-co-AC3pht).

IR (DRIFT) ν [cm-1]: 3084, 3065, 3029, 2929, 2854, 1773, 1716, 1642, 1609, 1491, 1450, 1400, 1364,

1250-1000, 721, 699 cm-1.

elemental analysis: cf. Table 9-13.

GB250-AA-P(S-co-AC3pht).

IR (DRIFT) ν [cm-1]: 3085, 3063, 3029, 2927, 2853, 1772, 1715, 1643, 1604, 1491, 1450, 1399, 1370,

1250-1000, 721, 699 cm-1.

elemental analysis: cf. Table 9-13.

9.5.2 Synthesis of Poly(styrene-co-N-acryloyl-N-methyl-propyl

amine), P(S-co-AC3amine)

Silica substrates with a poly(styrene-co-N-acryloyl-N-methyl-propyl phthalimide)

monolayer were covered with THF and a sufficient amount (up to 50fold excess) of

hydrazine monohydrate (N2H4·1.5H2O) was added. The mixture was then shaken with

180 rpm in a one-neck flask with reflux condenser for 18 h at 60 °C. Afterwards the reaction

solution is filtered with a Büchner funnel and the remaining solids were washed with

toluene, dichloro methane and again with toluene. Last step was freeze-drying from

benzene. Then the substrates were analyzed by DRIFT and elemental analysis (cf.

Table 9-14). To quantify the number of amino groups on the solids a [1,1-bis(4-

methoxyphenyl)-1-phenyl]methyl (DMT) containing dye[229] (λmax = 495 nm, ε = 71700[230])

with disulfide linker was coupled to the amino function[190] for further characterization of

the solids. The glass beads with the amino functions in the copolymer monolayer were also

treated with 1 mM HClaq. and the filtrate then was titrated against 1 mM NaOHaq. to calculate

the concentration of converted amino groups. 500 mg of the phthalimide copolymer were

dissolved in a emulsion of 1 mL (17.6 mmol) hydrazine hydrate in 5 mL THF and well stirred

for 18 h at 60 °C to deprotect the amine functions. The whole emulsion is evaporated until

Page 202: through Surface-attached Monomers

188 9 Experimental II – Polymerizations

there is a dry residue. This residue was dissolved in toluene and precipitation of the free

amino copolymer took place by dropwise addition of the copolymer solution in excess (10x)

methanol. The colorless polymer is filtered off and dried carefully in vacuum for

examination with 1H-NMR.

The first phthalimide conversions in the case of GB80-ME-P(S-co-AC3pht(10%)) and GB80-

ME-P(S-co-AC3pht(20%)) were carried out with methyl amine. A suspension of 1.9 g solid

substrate and 20 mL of a 2 M MeNH2 solution in THF was heated to 60 °C and shaken

(160 rpm) for 18 hours. During cooling down the glass beads settled and the supernatant

solution was decanted. Six washing steps with 40 mL THF each and two washing steps with

40 mL diethyl ether purified the glass beads before they were dried in vacuum for 8 hours.

To deprotect the amino groups in the free phthalimide copolymer 125 mg of the copolymer

were dissolved in 5 mL of a 2 M MeN2 solution in THF and stirred for 18 h at 60 °C. Then the

solution is evaporated until dry. The residues were dissolved again in toluene and the free

amino copolymer was precipitated by dropwise addition of the polymer solution in excess

(10x) methanol. The copolymer is filtered off, dried in vacuum and finally characterized with 1H-NMR.

1H-NMR (250 MHz, CDCl3, TMS) δ [ppm]:

c), d), e) (5H, m) 6.8 – 7.3

i) (2H, b) 3.1 – 3.3

h) (3H, b) 2.8 – 3.1

k) (2H, b) 2.7 – 2.8

g) (1H, b) 2.6 – 2.7

j) (2H, b) 2.4 – 2.6

b) (1H, m) 1.7 – 2.4

NH2

NO

co

co

a)b)

c) c)

d) d)e)

f)g)

h)

j)

k)

i)

a), f) (2H + 2H, m) 1.0 – 1.7

Page 203: through Surface-attached Monomers

9 Experimental II – Polymerizations 189

LC700-ME-P(S-co-AC3NH2).

IR (DRIFT) ν [cm-1]: 3083, 3063, 3027, 2925, 2850, 1724, 1602, 1493, 1450, 1250-1000, 758,

699 cm-1.

elemental analysis: cf. Table 9-14.

GB80-AA-P(S-co- AC3NH2).

IR (DRIFT) ν [cm-1]: 3086, 3066, 3030, 2929, 1632, 1494, 1450, 1250-1000, 699 cm-1. elemental

analysis: cf. Table 9-14.

GB250-AA-P(S-co- AC3NH2).

IR (DRIFT) ν [cm-1]: 3085, 3065, 3029, 2926, (2855) 1638, 1491, 1450, 1250-1000, (758), 699 cm-1.

elemental analysis: cf. Table 9-14.

Table 9-14. Elemental analysis results for styrene-co-N-acryloyl-N-methyl-propyl amine monolayer

modified substrates after polymer analogue reaction of the phthalimide containing substrates with

hydrazine.

product C*

(%)

H*

(%)

N*

(%)

LC700-ME-PS-AC3amine(10%) 25.60 3.43 0.53

GB80-ME-PS-AC3amine(5%) 11.00 0.86 0.22

GB80-ME-PS-AC3amine(10%) 11.17 0.86 0.33

GB80-ME-PS-AC3amine(20%) 11.84 1.14 0.66

GB80-AA-PS-AC3amine(10%) 7.83 1.19 0.60

GB250-AA-PS-AC3amine(20%) 19.46 2.51 1.62

* obtained from elemental analysis

9.5.3 Synthesis of Poly(styrene-co-N-methacryloyl-β-alanine

succinimide ester), P(S-co-MAC2ae)

884 mg (3.5 mmol, 5 mol%) N-methacryloyl-β-alanine succinimide ester (MAC2ae) were

dissolved in a solution of 8 mL styrene (66 mmol, 95 mol%) in 16 mL DMF. Then 240 mg

Page 204: through Surface-attached Monomers

190 9 Experimental II – Polymerizations

LC700-ME substrate and 35 mg (0.2 mmol, 9 mM) AIBN were added. Degassing in vacuum

by five freeze & thaw cycles was followed by a heating period of 24 h at 60 °C and shaking

with 160 rpm. The polymerization reaction was terminated by air exposure and cooling. The

mixture was centrifuged to separate the polymer modified silica gel The solid product was

extracted in a Soxhlet apparatus with DMF and toluene, 8 hours each and in the end freeze-

dried from benzene. The free copolymer precipitated on dropwise addition of the

polymerization solution to excess (10fold) methanol.

Into a solution of 1,397 mg MAC2ae (5.5 mmol, 5 mol%), 12 mL styrene (10.4 mmol,

95 mol%), 24 mL DMF 2 g GB80-ME were added. 18 mg (0.1 mmol, 3 mM) AIBN were added

as initiator. The mixture was degassed in vacuum by five freeze & thaw cycles. The

polymerization reaction was started by heating to 60 °C and the temperature was kept

constant for 24 h by an oil bath. Afterwards the polymerization reaction was stopped by air

exposure and cooling down. After the glass beads had settled, the supernatant solution was

added dropwise to an excess (10fold) of cold methanol to precipitate the free copolymer. The

beads were extracted with DMF an toluene (8 h each) in a Soxhlet apparatus and finally

freeze-dried with benzene.

1H-NMR (250 MHz, CDCl3, TMS) δ [ppm]:

c), d), e) (5H, m) 6.3 – 7.3

h) (2H, b) 3.1 – 3.4

j), i) (6H, b) 2.6 – 2.9

b) (1H, m) 1.7 – 2.4

a), f), g) (2H + 5H, m) 0.9 – 1.7

NHO

co

co

O O

NO

O

h)i)

a)b)

c) c)

d) d)e)

f)g)

j)j)

SiO2-ME-P(S-co-MAC2ae).

IR (DRIFT) ν [cm-1]: 3083, 3063, 3028, 2925, 2851, 1826, 1788, 1732, 1603, 1493, 1450, 1250-1000,

699 cm-1.

elemental analysis: LC700-ME-P(S-co-MAC2ae) C 27.37 %, H 2.68 %, N 1.61 %

GB80-ME-P(S-co-MAC2ae) C 11.21 %, H 0.87 %, N 0.50 %.

Page 205: through Surface-attached Monomers

9 Experimental II – Polymerizations 191

9.5.4 Synthesis of Poly(N,N-dimethyl acrylamide-co-N-acryloyl-N-

methyl-propyl phthalimide), P(DMAA-co-AC3pht)

First 1.47 g AC3pht (4.8 mmol, 10 mol%) were dissolved in a solution of 5 mL DMAA

(48.5 mmol, 90 mol%) in 10 mL DMF. Then 5 g of GB80-AA and 22 mg (0.13 mmol, 9 mM)

were added and the mixture was degassed in vacuum by five freeze & thaw cycles. The

polymerization reaction was run for 6 h at 60 °C (oil bath). After that time period the

polymerization solution was highly viscous and dichloro methane was added to make

filtration of the glass particles possible. The organic solvents of the filtrate were removed by

rotational evaporation. The residual substance was dissolved again in THF and dropwise

addition to cold diethyl ether precipitates the free copolymer. The filtered glass beads were

extracted with dichloro methane in a Soxhlet apparatus over night and finally dried in high

vacuum (< 10-3 mbar) to give the product.

1H-NMR (250 MHz, CDCl3, TMS) δ [ppm]:

j), k) (4H, b) 7.5 – 7.8

g), i) (4H, b) 3.4 – 3.7

c), f) (3H + 3H, m) 2.6 – 3.4

b), e), h) (1H + 3H, b) 2.1 – 2.7

a), d) (2H + 2H, b) 1.0 – 2.1

N

O NNO

O

O

co

co

a)b)

c)

c)

d)e)

g)h)

j)

j)k)k)

i)

f)

GB80-AA-P(DMAA-co-AC3pht).

IR (DRIFT) ν [cm-1]: 2928, 1643, 1772, 1715, 1495, 1398, 1356, 1250-1000, 724 cm-1.

elemental analysis: C 16.90 %, H 2.87 %, N 3.51 %.

Page 206: through Surface-attached Monomers

192 9 Experimental II – Polymerizations

9.5.5 Synthesis of Poly(N,N-dimethyl acrylamide-co-N-acryloyl-N-

methyl-propyl amine), P(DMAA-co-AC3amine)

6.4 g GB80-AA-P(DMAA-co-AC3pht(10%)) were covered with 15 mL dry ethanol. After

addition of 2 mL of hydrazine monohydrate (N2H4·1.5H2O) the reaction suspension was kept

for 18 h at 60 °C. At the end of the reaction the glass beads were separated from the reaction

solution by sedimentation and decanting. Several washing steps were done analogous. The

washing steps were in detail: 4x ethanol/water (v/v 1/1), methanol and 2x dichloro methane.

The particles were filtered off and dried in high vacuum (< 10-3 mbar) overnight. For a

quantitative determination of the amino groups a DMT-dye with disulfide linker was

coupled to them, followed by elemental analysis and cleavage of the dye for quantitative

UV/vis spectroscopy (λ = 499 nm, ε = 71900). The glass beads with the amino functions in the

copolymer monolayer were also treated with 1 mM HClaq. and the filtrate then was titrated

against 1 mM NaOHaq. to obtain the amount of amino groups.

1H-NMR (250 MHz, CDCl3, TMS) δ [ppm]:

g) (2H, b) 3.1 – 3.3

c), f), i) (6H + 5H, m) 2.6 – 3.1

b), e), h) (1H + 3H, b) 2.1 – 2.7

a), d) (2H + 2H, b) 1.0 – 2.1

GB80-AA-P(DMAA-co-AC3NH2).

IR (DRIFT) ν [cm-1]: 2928, 1641, 1496, 1399, 1353, 1250-1000 cm-1.

elemental analysis: C 15.42 %, H 3.01 %, N 3.68 %.

Page 207: through Surface-attached Monomers

9 Experimental II – Polymerizations 193

9.5.6 Synthesis of Poly(DMAA-co-N-methacryloyl-β-

alanine succinimide ester), P(DMAA-co-MAC2ae)

A solution of 1.36 g MAC2ae (5.4 mmol, 5 mol%) and 5 mL DMAA (48.5 mmol, 95 mol%)

as monomers in 10 mL DMF was filled under dry nitrogen in a Schlenk flask. Then

5 g GB80-AA and 25 mg (10 mM) AIBN were added before degassing in vacuum by five

freeze & thaw cycles. The mixture was heated for 25 h at 60 °C and rapid circular agitation at

200 rpm. All solids were filtered off. The filtrate solution was freed from DMF by rotational

evaporation and the residues were dissolved in THF. The free copolymer was obtained by

dropwise addition of the solution to cold diethyl ether where a white substance precipitates.

After an extraction with dichloro methane in a Soxhlet apparatus for 24 hours the copolymer

modified glass beads were dried in high vacuum (< 10-3 mbar).

1.36 g MAC2ae (5.4 mmol, 5 mol%) were dissolved in 5 mL DMAA (48.5 mmol,

95 mol%)and 10 mL DMF. 5 g LC700-ME and 22 mg (0.13 mmol, 9 mM) AIBN were added to

this solution. After that the suspension was degassed in vacuum by five freeze & thaw cycles

and the polymerization reaction kept running for 30 h at 60 °C. The reaction leaves a highly

viscous solution and in order to make a filtration of the silica gel possible dichloro methane

was added. The CH2Cl2 and DMF of the filtrate solution were removed by rotational

evaporation. The polymeric residues were dissolved in THF and dropwise addition to cold

diethyl ether precipitates the free copolymer. The filtered silica particles were extracted with

dichloro methane in a Soxhlet apparatus for 25 hours and as last step dried in high vacuum

(< 10-3 mbar).

Page 208: through Surface-attached Monomers

194 9 Experimental II – Polymerizations

1H-NMR (250 MHz, CDCl3, TMS) δ [ppm]:

f) (2H, b) 3.3 – 3.5

g), h) (6H, b) 2.8 – 3.1

c) (6H, m) 2.6 – 3.0

b) (2H, b) 2.2 – 2.6

a), d) (2H + 2H, b) 1.3 – 1.9

NNHO

co

co

O

O ON

O

O

a)b)

c)

c)

d)e)

g)

h)h)

f)

e) (3H, b) 0.9 – 1.3

SiO2-P(DMAA-co-MAC2ae).

IR (DRIFT) ν [cm-1]: 2938, 1817, 1786, 1740, 1644, 1496, 1399, 1356, 1250-1000 cm-1.

elemental analysis: GB80-AA-P(DMAA-co-MAC2ae) C 14.07 %, H 2.36 %, N 3.07 %

LC700-ME-P(DMAA-co-MAC2ae) C 18.51 %, H 3.49 %, N 3.51 %.

Page 209: through Surface-attached Monomers

10 Experimental III – Reactions with Functional Groups 195

10 Experimental III – Reactions with Functional Groups

10.1 Two Step Grafting on Functionalities In this section two methods of grafting polystyrene onto functional copolymer layers on

glass beads are described. In both cases the starting substrate is GB80-AA-P(S-co-AC3amine)

with 10 mol% of amino groups within the copolymer layer.

10.1.1 Immobilization of Acryloyl Chloride as Monomer

1 g of GB80-AA-PS-AC3amine were suspended in 3 mL dichloromethane (distilled and

dried) and brought to a temperature of 0 °C with an ice-bath. According to the estimated

amount of amino groups 14 µL (139 µmol) acryloyl chloride and 24 µL (172 µmol) Et3N were

added drop wise to the suspension that was agitated by a shaking device. After 20 minutes

the ice-bath was removed and the suspension let shaken for 4 hours at room temperature.

Afterwards the reaction solution is removed by a porous glass filter. Glass beads a washed

with acidified ethanol/water (pH = 3, v/v 1/1) to remove excess base, then washed neutral

with ethanol/water. Basic ethanol/water treatment (pH = 11) to recovered amino groups and

again washed neutral with ethanol/water. Two washing cycles with pure ethanol and diethyl

ether followed for easy removable volatiles. Drying overnight under reduced pressure yields

colorless glass beads (GB80-AA-P(S-co-AC3acrylamide)) that were characterized by DRIFT

and elemental analysis:

GB80-AA-P(S-co-AC3acrylamide).

IR (DRIFT) ν [cm-1]: 3086, 3066, 3030, 2928, 1624, 1609, 1491, 1450, 1408, 1250-1000, 799,

700 cm-1.

elemental analysis: C 7.27 %, H 1.03 %, N 0.57 %.

10.1.2 Immobilization of 2-Brom-Propionic Acid Bromine as Starter

for ATRP

In a 25 mL-Schlenk flask with a rubber septum a suspension of 500 mg GB80-AA-P(S-co-

AC3amine) and 3 mL acetone (p.a.) was agitated at 160 rpm in a dry nitrogen atmosphere.

Via syringe a few drops triethyl amine and 2-brom-propionic acid bromine were added to

Page 210: through Surface-attached Monomers

196 10 Experimental III – Reactions with Functional Groups

the suspension. While shaking at room temperature the mixture turns yellow. After a

reaction time of six hours the glass beads were filtered off and washed with acetone. Then

the modified glass beads are subsequently washed with acidified (pH = 3) ethanol/water (v/v

1/1), ethanol/water, ethanol (3x) and diethyl ether (3x). Drying under high vacuum for 6

hours gives GB80-AA-P(S-co-AC3(2-brom-propionic amide)) what was immediately used for

the polymerization (10.1.4).

10.1.3 Polymerization of Styrene with Polymer-attached Monomers

In a 40 mL-Schlenk tube 200 mg GB80-AA-P(S-co-AC3acrylamide) were suspended in 8 mL

dry toluene. 4 mL styrene and 18.6 mg (9 mM) AIBN were added. The mixture was degassed

by five freeze & thaw cycles and finally heated to 60 °C and shaken at a speed of 160 rpm for

26 hours. After stopping the polymerization by cooling down the supernatant liquid was

decanted and added dropwise to 120 mL methanol to precipitate the free polystyrene. The

glass bead were flushed into a filter and extracted with toluene overnight in a Soxhlet

apparatus. Then the glass beads were freeze dried in 3 mL benzene to get GB80-AA-P(S-co-

AC3AA-PS) as colorless product that was characterized per elemental analysis and DRIFT.

GB80-AA-P(S-co-AC3AA-PS).

IR (DRIFT) ν [cm-1]: 3086, 3066, 3030, 2929, 1630, 1494, 1450, 1250-1000, 699 cm-1.

elemental analysis: C 9.53 %, H 1.08 %, N 0.56 %.

As blind test without polymer attached monomers the same polymerization procedure as

described above was carried out with the precursor support GB80-AA-P(S-co-AC3amine) to

get GB80-AA-P(S-co-AC3amine)-PS as product and afterwards characterized in the same

manner:

GB80-AA-P(S-co-AC3amine)-PS.

IR (DRIFT) ν [cm-1]: 3085, 3065, 3029, 2928, 1629, 1495, 1450, 1250-1000, 699 cm-1.

elemental analysis: C 8.33 %, H 1.06 %, N 0.65 %.

Page 211: through Surface-attached Monomers

10 Experimental III – Reactions with Functional Groups 197

10.1.4 Atom Transfer Radical Polymerization of Styrene with

Polymer-attached Initiators

In a 40 mL-Schlenk tube 48 mg (0.217 mmol) copper(II)bromide were suspended in 25 mL

(0.217 mol) styrene and 376 µL (2.17 mmol) PMDETA were added and agitated until a slight

green color of the copper complex appears. Then 230 mg GB80-AA-P(S-co-AC3(2-brom-

propion amide)) were mixed into the solution followed by two freeze & thaw cycles for

degassing. Under an atmosphere of dry nitrogen 311 mg (2.17 mmol) copper(I)bromide was

added and again two degassing cycles took place. The polymerization reaction was started

by raising the temperature to 90 °C. For 21 hours the mixture was shaken at 190 rpm at

constant temperature. The viscosity of the suspension increased slightly during the

polymerization. The reaction was stopped by cooling down and flooding the vessel with air.

Parallel a degassed mixture of 25 mL (0.217 mol) styrene with 376 µL (2.17 mmol) PMDETA,

230 mg GB80-AA-P(S-co-AC3(2-brom-propion amide)), 296 µL (2.17 mmol) PEBr (as free

initiator) and 311 mg (2.17 mmol) copper(I)bromide was polymerized for 2 hours in order to

get free polystyrene under same polymerization conditions for GPC analysis. For this case

viscosity of the reaction mixture increased much faster during polymerization. The glass

beads were filtered off and washed with toluene. Then the solids were extracted with THF

for 22 hours. The glass beads remained light green in color even after treatment with solvents

of different polarity. Finally the polystyrene grafted glass beads were freeze-dried in benzene

and characterized by elemental analysis. Free polystyrene was isolated by first reducing the

volume of filtrate with a rotary evaporator and adding drop wise to excess (10fold) methanol

to precipitate the polymer. Because of slight blue color of the polystyrene it was dissolved in

THF and again precipitated in methanol. After drying under high vacuum overnight the free

PS is characterized by GPC.

GB80-AA-P(S-co-AC3(propionic amide)PS).

elemental analysis: C 50.95 %, H 4.46 %, N 0.30 %.

Page 212: through Surface-attached Monomers

198 10 Experimental III – Reactions with Functional Groups

10.2 Coupling of Dyes to the Functionalities of the Copolymer

10.2.1 Coupling of Fluorescein Isothiocyanate

A solution of 10 mg FITC (fluorescein isothiocyanate) in 10 mL distilled and dried DMF

was prepared and glass beads with different surface-attached polymer monolayers as well as

the corresponding free polymers were mixed or dissolved at room temperature. The amount

of dye molecules was always calculated so that an excess compared to all functionalities

within the polymer (monolayer) was added. 20 mg of GB80-ME-P(S-co-AC3amine(10%))

were put into 1 mL of FITC solution and shaken for 2 hours. Afterwards the supernatant

FITC solution was discarded and the glass beads were washed repeatedly with methanol

until no further coloring of the solution takes place. Then the now yellow glass beads were

washed with toluene and diethyl ether and dried in a high vacuum overnight. The stained

glass beads remained yellow and the fluorescence of the dye was examined a confocal

scanning laser microscope to have many slices in z-direction. As blind test experiments

20 mg GB80-ME-P(S-co-MAC2ae) and 20 mg GB80-ME-PS were also put into 1 mL of FITC

solution and received the same treatment as the amino-functionalized glass beads. During

the washing steps the color faded away and after drying both substrates are again colorless.

100 mg P(S-co-AC3amine(5%)), P(S-co-AC3amine(10%)) and P(S-co-AC3amine(20%)) were

each dissolved in a solution of 60 mg FITC in 20 mL DMF and stirred for one hour at room

temperature. Then the DMF of each solution was nearly removed and the remaining solution

was added dropwise to excess (10x) cold methanol to precipitate the colored copolymers.

Filtration and washing with methanol removed unbound dye. The colored copolymers were

dried in vacuum overnight. It was not possible to dissolve the colored P(S-co-AC3thiourea-

fluorescein(10%)) and P(S-co-AC3thiourea-fluorescein(20%)) again for UV/vis spectroscopy

characterization in any solvent or mixture of solvents. Due to this insolubility only the

P(S-co-AC3thiourea-fluorescein(5%)) was examined with the photospectrometer at a

concentration of 1 mg·mL-1in DMF. For the other two polymer species only a description of

the optical properties was done.

Page 213: through Surface-attached Monomers

10 Experimental III – Reactions with Functional Groups 199

10.2.2 Coupling of DY-635-NH2

A solution of 1 mg (1.356 mmol) DY-635-NH2 (8-(3-(1-(5-(2-amino-ethylcarbamoyl)-

pentyl)-3,3-demethyl-5-sulfonato-1,3-dihydro-indol-2-ylidene)-10-tbutyl-2,3,5,6,11b,11c-hexa-

hydro-1H,4H-11-oxonia-3a-aza-benzo(de)anthracene) in 0.5 mL EtOHabs. was prepared and

100 µL of this solution were diluted with 170 µL ethanol to a 1 mmol·L-1 solution. Then some

milligram of silica solids with a monolayer of P(DMAA-co-MAC2ae) or P(S-co-MAC2ae) were

suspended in a few milliliters of the dye solution. After 20 minutes shaking at room

temperature the solids were filtered off, washed three times with methanol and hot extracted

in THF for 18 h to remove unbound dye molecules. Finally all particles were dried in

vacuum.

A solution of 50 mg P(S-co-MAC2ae(5%)) in 5 mL dry toluene was mixed well with 50 µL

of a 1 mM DY-635-NH2 ethanol solution. After a reaction time of 10 minutes at room

temperature the mixture was added dropwise to cold methanol to precipitate the labeled

copolymer. It was filtered off, washed with methanol and dried in vacuum. The product

again was dissolved in toluene and analyzed with UV/vis spectroscopy at a concentration of

1 mg·mL-1.

10.3 Enzyme Immobilization

10.3.1 Activity Assay

The activity assay for glucose oxidase (GOD) is performed using four different solutions:

The glucose solution contains the substrate, the GOD solution delivers the enzyme to the

immobilization site on the glass beads, the horseradish peroxidase solution transforms the

GOD product hydrogen peroxide into electrons that reduce the dye 2,2‘-Azino-

bis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) for a color change from colorless to deep

green in the fourth solution.

1 liter of phosphate buffered saline stock solution (10x PBS) at pH = 7.4 consists of 0.19 mol

NaH2PO4, 0.81 mol Na2HPO4, 1.38 mol NaCl and 0.027 mol KCl in distilled water and was

used in all described handling steps with 10 mmol·L-1 phosphate concentration (1x PBS).

To prepare the glucose solution 301.4 mg D-(+)-glucose were dissolved in 3 mL phosphate

buffered saline (PBS, pH=7.4). To get the 1 µmol·L-1 GOD solution 16.1 mg (0.1 µmol) GOD

Page 214: through Surface-attached Monomers

200 10 Experimental III – Reactions with Functional Groups

were dissolved in 10 mL PBS and 1 mL of this solution was diluted with 9 mL PBS and kept

cool (~4 °C). 15 mg horseradish peroxidase were dissolved in 5 mL PBS and kept cool. The

dye solution consisted of 274.5 mg ABTS in 10 mL PBS and was kept cool and stored in the

dark.

To measure the GOD activity the freshly prepared solutions were mixed in a cuvette with

the following dosages: 1 mL glucose solution, 0.2 mL peroxidase solution, 0.2 mL dye

solution and 1.6 mL PBS (pH=7.4). Upon addition of GOD loaded active ester polymer layers

on glass beads the assay turned green with time and a UV spectrometer measured the

extinction change at a wavelength of λ=405 nm. The cuvette was equipped with a magnetic

stirring bar and stirred for equal color distribution. For comparison also experiments with

30 µL of 1 µmol·L-1 GOD solution and experiments with inert solids with a polystyrene layer

were carried out.

For each experiment 10 mg of dry solid support (GB80-AA-P(DMAA-co-MAC2ae(10%)))

were put into 1 mL GOD solution for a certain time period. Then the glass beads were

decanted and washed several cycles with PBS. After filtering off the beads were immediately

used in the GOD activity assay. Specific details of the enzyme experiments are explained in

the following paragraphs.

10.3.2 Washing Cycles

After 30 minutes of enzyme immobilization in 1 µmol·L-1 GOD solution four samples of

GB80-AA-P(DMAA-co-MAC2ae(10%)), 10 mg each, were washed with PBS for a different

number of cycles. For one washing cycle the filtered solids from the prior solution were

taken, put into 1 mL PBS, agitated for approximately 5 minutes and allow to settle. The

supernatant liquid was discarded. The washed, enzyme containing supports were then used

in the GOD activity assay.

10.3.3 Varied Immobilization Time

Five samples of 10 mg GB80-AA-P(DMAA-co-MAC2ae(10%)) were put into 1 µmol·L-1

GOD solution for immobilization and kept in motion for different periods of time. The

examined time periods were 10, 30, 60, 120 and 240 minutes. After the immobilization

procedure all samples were washed with three PBS washing cycles to get rid of free GOD

and finally measured with the GOD activity assay.

Page 215: through Surface-attached Monomers

10 Experimental III – Reactions with Functional Groups 201

10.3.4 Varied Initial Immobilization Concentration

Immobilization of the enzyme on GB80-AA-P(DMAA-co-MAC2ae(10%))took place for 30

minutes in GOD solution with three different GOD concentrations. In each of the solutions,

1·10-6, 10·10-6, 100·10-6 mol·L-1, 10 mg of the solid support were put and after immobilization

washed with PBS three times. The activity of the immobilized GOD enzyme was measured

by UV spectroscopy of the activity assay.

10.3.5 Immobilization of Dyed Enzymes

In order to get colored GOD enzymes, a DY-635-NH2 solution was added to a GOD

solution where reaction with free accessible carboxyl groups at room temperature on the

enzyme took place. This deep blue colored enzyme solution (6.2 10-5 mol·L-1) was used as

stock solution for the following enzyme immobilization steps. For 10-6 mol·L-1 GOD solution

161 µL of the stock solution were diluted with 839 µL PBS and for 10-6 mol·L-1 GOD solution

16 µL of the stock solution were diluted with 9846 µL PBS. Each immobilization

concentration was used with either GB80-AA-P(DMAA-co-MAC2ae(10%)) or LC700-ME-

P(DMAA-co-MAC2ae(10%)). To immobilize the colored enzyme 50 mg of solid support were

suspended in 500 µL GOD solution and agitated for 30 minutes at room temperature. Four

subsequent washing cycles with 1 mL deionized water and vivid agitation for 15 minutes

were used to remove non-covalently bound GOD. The enzyme modified solids were

investigated with optical methods (CCD camera and fluorescence reading device

(Biodetect®).

10.3.6 Hydrolysis of Immobilized Enzymes

16.2 mg GOD were dissolved twice in 10 mL PBS (pH = 7.4) to get two 10-4 mol·L-1

solutions. 51 mg and 150 mg GB80-AA-P(DMAA-co-MAC2ae(10%)) as separate samples

were suspended in these solutions and agitated for 30 minutes in order to immobilize the

enzyme on the support surface. After the immobilization procedure the solids were

separated by decanting the supernatant solution and washed three times with distilled water

for 10 minutes.

In a 25 mL flask equipped with a reflux condenser the GOD-covered glass beads were

suspended into 5 mL of 6 mol·L-1 hydrochloric acid, stirred and heated to reflux (115 °C oil

bath) for 24 hours. The solids were filtered off and the filtrate was concentrated in a rotary

Page 216: through Surface-attached Monomers

202 10 Experimental III – Reactions with Functional Groups

evaporator until dry. Then the residue was incorporated in 1 mL of a buffer (pH = 2.0)

provided by the Pediatric hospital Freiburg where the analysis via liquid chromatography of

the single amino acid units took place[210].

10.4 Oligo Nucleotide Immobilization For the DNA experiments different oligo nucleotides with functional groups and

fluorescent dyes were used as supplied by Thermo Electron GmbH, Ulm (Table 10-1). The

oligo-nucleotide reacted with the active ester surface groups on the solid supports got at the

5’-end an amino group with a C6-spacer (“aminolink”) and the following nucleotide

sequence: 5’-TTT-TTT-TTT-TTT-TTT-TTT-TTA-TGG-TGT-AAA-CTT-GTA-CCA-GT-3’. This

oligo nucleotide is called NH2-DNA. To check if the immobilization procedures were

successful an oligo-nucleotide that bears an amino group as linker and a fluorescent dye for

detection was used. This connection control DNA has a Cy5 dye as 5’-modification, an amino

group with C7-spacer as 3’-modification and a nucleotide sequence as follows: 5’-CAA-

GAA-CTA-TGC-TTC-AAG-CAG-ACA-TCA-ACC-GTT-TTT-TTT-TTT-3’. This oligo-

nucleotide is called CC-DNA (coupling control). The complementary oligo-nucleotide for

hybridization on the NH2-DNA got a Cy5 fluorescent dye as 5’-modification and 5’-ACT-

GGT-ACA-AGT-TTA-CAC-CAT-AA-3’ as nucleotide sequence. This oligo-nucleotide is

called AS-DNA (anti-sense). All Cy5-labelled oligo-nucleotides have a deep blue color.

Table 10-1. List of functionalized oligonucleotides used in the DNA-immobilization experiments.

abbreviation structure

NH2-DNA aminolink-C6-(5’-TTT-TTT-TTT-TTT-TTT-TTT-TTA-TGG-TGT-AAA-CTT-

GTA-CCA-GT-3’)

CC-DNA Cy5-(5’-CAA-GAA-CTA-TGC-TTC-AAG-CAG-ACA-TCA-ACC-GTT-

TTT-TTT-TTT-3’)-C7-aminolink

AS-DNA Cy5-(5’-ACT-GGT-ACA-AGT-TTA-CAC-CAT-AA-3’)

For the handling of DNA oligo-nucleotides special salt solutions were used. NaPi is an

sodium phosphate buffer. For pH = 8.0 with an overall salt concentration of 300 mmol·L-1 it

Page 217: through Surface-attached Monomers

10 Experimental III – Reactions with Functional Groups 203

consists of 285 mmol Na2HPO4 and 15 mmol NaH2PO4. A 10-fold sodium salt citrate (SSC)

stock solution consists of 1.5 mol NaCl and 0.15 mol sodium citrate in 1 liter distilled water,

has a pH-value of 7.0 and. The aqueous sodium dodecyl sulfate stock solution had 10 wt.%

of the surfactant.

10.4.1 Immobilization Reaction

To handle the oligo-nucleotides 100 µmol·L-1 solutions were made from 0.05 µmol DNA

and 500 µL NaPi buffer at a pH-value of 8. The solids used for the oligo-nucleotide

immobilization were GB80-AA-P(DMAA-co-MAC2ae(10%)) and LC700-ME-P(DMAA-co-

MAC2ae(10%)). For every immobilization reaction 40 mg solid support were suspended in

400 µL SSC (10x)inside of an Eppendorf vial and 100 µL of DNA-solution (= 10 nmol DNA)

were added. For each support, glass beads and silica gel, one immobilization with NH2-DNA

and one immobilization with CC-DNA were carried out. The immobilization lasted two

hours at a temperature of 50 °C in a water bath and the vials were agitated with a shaking

device at 230 rpm. After reaction the solids were let settle and the supernatant solution was

decanted. The supports reacted with the coupling control oligo-nucleotide were colored blue.

All remaining solids were washed three times with SSC buffer (2x) including 0.1 wt.% SDS

using a vortex apparatus for 10 minutes each cycle. Then three washing cycles with

deionized water (MilliQ, 18.2 MΩ) followed. After the washing procedure the glass beads

and the silica gel with the immobilized CC-DNA stayed blue. They were dried for detection

of the fluorescent DNA-label with a DNA-fluorescence-reading device. Solids with the

immobilized NH2-DNA (GB80-AA-P(DMAA-co-MAC2amido-DNA) and LC700-ME-

P(DMAA-co-MAC2amido-DNA)) were prepared for the hybridization step.

10.4.2 Hybridization

Each sample with immobilized NH2-DNA, 50 mg of GB80-AA-P(DMAA-co-MAC2amido-

DNA) and 40 mg of LC700-ME-P(DMAA-co-MAC2amido-DNA), was suspended in 900 µL

SSC (6x, 0.5 wt.% SDS). Then 100 µL of AS-DNA solution were added to each solution. Both

reaction mixtures were agitated at room temperature for 30 minutes on a vortex apparatus.

Subsequent washing cycles (same agitation, 15 minutes each) were carried out three times

with 1 mL SSC buffer (2x, 0.1 wt.% SDS) and finally with 1 mL SSC buffer (0.5x). Filtration

and drying led to blue colored solids.

Page 218: through Surface-attached Monomers

204 10 Experimental III – Reactions with Functional Groups

10.5 Biotin assay GB80-AA-P(S-co-AC3amine(10%)), GB250-AA-P(S-co-AC3amine(20%)) and GB80-AA-

P(DMAA-co-AC3amine(10%)) as solid supports equipped with a covalently bound

N-acryloyl-N-methyl-propyl amine containing copolymer were used for the following biotin

assay procedure:

15.4 mg (45 mmol) N-hydroxysuccinimido-biotin were dissolved in 3 mL dry DMF. 1 mL

of this solution was given to 100 mg of each solid. Reaction took place at room temperature

overnight. Then the supernatant was removed from the settled solids with filter paper. The

residuals were suspended in methanol and washed two times (30 minutes vortex agitation).

After drying in air overnight the remaining solids were put each into 1 mL of 1 µmol·L-1

solution of streptavidin in PBS (1x) for 30 minutes at room temperature and 6 hours at 4 °C.

The supernatant was discarded. 0.2 mg of fluorescein labeled biotin (M = 644.7 g·mol-1) was

dissolved in 10 mL PBS (1x) giving 30 µmol·L-1 solution. Every solid was suspended in 1 mL

of the biotin-fluorescein solution and kept at 4 °C overnight. Again the supernatant was

discarded and the solid were washed two times with distilled water and in a final step with

acetone. After drying the solids were analyzed by a fluorescence reading device (Biodetect®).

10.6 Catalyst Immobilization Basic support material for the catalyst immobilization was GB80-AA-P(S-co-

AC3amine(10%)). The covalent attachment of the Ruthenium catalyst species onto the

support was realized by immobilization of a styrene based Hoveyda type ligand[221] via

reaction of its carboxyl group with the amino groups of the support material[188] using

DCC, tBuOH and Huenig’s base. Subsequently free Grubbs catalyst complexes were fixed to

the immobilized ligands in CH2Cl2 under reflux conditions with Cu(I)Cl as activator. The

ring closing metathesis of N,N-bis(allyl-tosyl amide) was used as benchmark test for the

reactivity respectively the effectivity of the whole catalyst system. The reaction was carried

out in dried dichloromethane under reflux conditions with an amount of 5 mol% of

immobilized Grubbs catalyst species related to the amount of N,N-bis(allyl-tosyl amide). The

substrate with the catalyst was filtered off after each cycle and used for at least three reaction

cycles. The product of the ring closing metathesis was isolated by distillation to determine

the conversion.

Page 219: through Surface-attached Monomers

11 Experimental IV – Silicon Wafers 205

11 Experimental IV – Silicon Wafers

11.1 Silane Immobilization Standard silicon wafers with a 2.5 nm silicon oxide layer (polished face) were used as

substrates. For silane immobilization and polymer deposition the silicon wafers were cut into

pieces between 2 x 2 and 2 x 4 cm. The pieces were cleansed of organic traces by subsequent

ultrasound exposure in methanol, toluene and acetone. Silicon dust particles were then

removed by SnowJet treatment using a carbon dioxide snow plume.

For the silane immobilization the wafer pieces were placed in a Schlenk tube equipped

with a reflux condenser in an inert atmosphere of dry nitrogen. Under nitrogen counter flow

dried toluene was added until the silicon was covered. Some drops of freshly distilled

triethylamine were added as catalyst. 1 mL of methacryloyl propyl trimethoxysilane MPS

was added in excess amount to react most of the surface groups. The immobilization reaction

was carried out at 120 °C for 3.5 hours heated by an oil bath and the solution was stirred

with a magnetic stir bar. After cooling down the silicon wafer pieces were removed and

rinsed with toluene and methanol. Finally they were dried with dry nitrogen and

characterized by ellipsometric measurements to determine the silane layer thickness. Until

the use for polymerization experiments the surface modified silicon was stored under dry

nitrogen.

11.2 P(DMAA) Layers Generally the MPS-modified silicon wafers were put into a Schlenk tube under an

atmosphere of dry nitrogen, containing the polymerization mixture consisting of the

monomer N,N-dimethyl acrylamide, the solvent DMF and the initiator AIBN. Five freeze-

thaw cycles degassed the polymerization system carefully. The polymerization was started

by placing the reaction vessel into a thermostat at 60±0.1 °C. After the chosen polymerization

time the silicon sample was withdrawn from the polymerization solution, rinsed with DMF

and extracted in a Soxhlet apparatus for at least 18 h with dichloromethane. Finally the

samples were dried under vacuum and measured with ellipsometry to determine the

polymer film thicknesses. At least two silicon wafer samples were prepared for every

polymerization condition variation.

Page 220: through Surface-attached Monomers

206 11 Experimental IV – Silicon Wafers

The polymerization in the solution was stopped by cooling down to ambient temperature

and air exposure. The produced free P(DMAA) was precipitated by dropwise adding the

polymerization solution to cold excess (10x) diethyl ether and filtrated. If the precipitate was

still viscous due to embedded DMF, the polymer was dissolved in THF and again

precipitated in cold diethyl ether. After filtration the remaining colorless polymer is dried

under high vacuum and characterized by GPC measurements.

11.2.1 Influence of the Polymerization Time

For the time dependent polymerization the polymerization mixture was prepared of 3 mL

N,N-dimethyl acrylamide in 6 mL DMF and 4.5 mg AIBN (3 mmol·L-1). The chosen periods

of time were 10 min, 30 min, 1 h, 2 h, 4 h, 14 h, 16 h and 21.5 h. Silicon wafer preparation,

polymerization and isolation of the wafers and free polymers were carried out as described

above.

11.2.2 Influence of the Monomer Concentration

The overall volume of the polymerization solution was in every case 10 mL. Chosen ratios

between monomer and solvent were 1:0, 1:1, 1:2, 1:4 and 1:9 to vary the monomer

concentration. 5 mg (3 mmol·L-1) AIBN was used as initiator. For the polymerization the

reaction vessels were kept at 60 °C for 16 hours.

11.2.3 Influence of the Initiator Concentration

For the variation of the initiator concentration the polymerization mixture was prepared of

3 mL N,N-dimethyl acrylamide in 6 mL DMF. 0.15 mg (0.1 mmol·L-1), 1.5 mg (1 mmol·L-1),

4.5 mg (3 mmol·L-1), 15 mg (10 mmol·L-1) and 150 mg (100 mmol·L-1) were the chosen

amounts of initiator. For 16 hours the polymerization was kept at 60 °C.

11.3 Microstructured Layers For all microstructure experiments MPS-modified silicon wafers were used. There were

different methods to microstructure the silicon surfaces with polymer layers and two

different kind of masks have been used. The first type of masks were double printed laser

images on common overhead projector transparencies. Here the patterns were designed with

a vector oriented graphics program (CorelDRAW 9.0) and they include squares, lines and

grids with feature widths of 100, 200 and 400 µm (Figure 11-1).

Page 221: through Surface-attached Monomers

11 Experimental IV – Silicon Wafers 207

Figure 11-1. Template used for laser printed masks on plastic with different widths and structures.

Squares give a polymer grid with dimples, lines give polymer walls and grids give polymer pillars.

For sizes of 100 µm and below the printing procedure by printing twice on the same plastic

sheet to get a deep black pattern is too inaccurate and the resolution of the printer (300 dpi)

was insufficient. The second type of masks were small metal nets with different patterns

normally used for transmission electron microscopy (TEM) These patterns were lines and

grids with mesh sizes of 128 µm (mesh 200 = 200 grid bars per inch) and 64 µm (mesh 400 =

400 grid bars per inch).

11.3.1 UV Photoablation of Surface-attached Monomers

Metal electron microscopy nets were put on MPS-modified silicon wafers. The UV light

source was a pen ray lamp (high pressure Hg, 500 W, L.O.T. Oriel) that emits in the far

(deep) UV range (<200nm). Exposure to the emitted light for 20 minutes decomposed the

MPS in the illuminated areas. Afterwards the silicon samples were rinsed with 1 M KOHaq.,

distilled water and methanol to remove the molecular debris of the ablation.

11.3.2 Grafting of Polystyrene

Silicon samples with an ablation patterned MPS-layer were put into a Schlenk tube under

an atmosphere of dry nitrogen. The polymerization solution, consisting of 20 mL styrene,

20 mL toluene and 19.6 mg (3 mmol·L-1) AIBN was added. Four freeze-thaw cycles removed

trace oxygen from the polymerization system. By placing the reaction vessel into a

thermostat at 60±0.1 °C the polymerization was started. After 24 hours the silicon samples

were withdrawn from the polymerization solution, rinsed with toluene and extracted in a

Soxhlet apparatus for 24 h with toluene. Finally the samples were dried under vacuum.

Ellipsometric measurements determined film thickness and the lateral extension of the

Page 222: through Surface-attached Monomers

208 11 Experimental IV – Silicon Wafers

structures (imaging ellipsometry). The free PS was characterized by GPC measurements with

the SDV poly columns in THF:

Mn = 154000 g·mol-1 , Mw = 301000 g·mol-1.

11.4 Combined Functional Layers & Networks To prepare a combined system of a polymer network and a covalently bound, thin

polymer layer with chemical functionalities based on a planar MPS-modified silicon wafer a

sequence of several steps was necessary. First the modified wafers were coated with a UV-

cross-linkable copolymer with benzophenone units with layer thicknesses ranging from

60 nm to several hundred nanometers. Then the coated wafers were equipped with a mask.

Exposition to UV light cross-linked the polymer layer and bound it to the MPS-layered

surface in the illuminated areas. The patterned polymer layer evolved during extraction with

a good solvent for the copolymer. Finally the prepared wafers were used in a

copolymerization reaction. A covalently bound, functional copolymer layer of P(S-co-

AC3pht) was formed on the wafer and by treatment with hydrazine solution transferred to

P(S-co-AC3amine). Thickness could not be measured because the scale towards the polymer

network was too small for proper detection but a detection of fluorescence after a subsequent

reaction of the amino groups with FITC was carried out.

11.4.1 Preparation of Copolymers for Network Formation

All polymerization solutions were prepared in Schlenk tubes under an atmosphere of dry

nitrogen and degassed by 3 freeze & thaw cycles to remove residual oxygen traces. For

polymerization the reaction vessels were heated in a thermostat to 60 °C for the given

amount of time. The recipes are listed in Table 11-1:

Page 223: through Surface-attached Monomers

11 Experimental IV – Silicon Wafers 209

Table -1. Ingredients and reaction times for the copolymerization of styrene and N,N’-dimethyl

acrylamide with benzophenone-methacrylic ester.

11

MABP styrene toluene AIBN t

(mol%)

m (mg)

n (mmol)

V (mL)

n (mmol)

V (mL)

m (mg)

c (mmol·L-1)

(h)

1 540 2.0 22.76 198.0 0.0 11.2 3.00 19

5 666 2.5 4.48 47.5 15.5 0.5 0.15 16

10 1332 5.0 4.25 45.0 15.7 0.5 0.15 16

MABP DMAA DMF AIBN t

(mol%)

m (mg)

n (mmol)

V (mL)

n (mmol)

V (mL)

m (mg)

c (mmol·L-1)

(h)

1 107 0.4 4.11 39.6 25.0 13.1 3 16

5 666 2.5 4.53 47.5 31.5 5.9 1 21

10 1332 5.0 4.29 45.0 30.0 5.6 1 21

All polymerization solutions were prepared in Schlenk tubes under an atmosphere of dry

nitrogen and degassed by 3 freeze & thaw cycles to remove residual oxygen traces. For

polymerization the reaction vessels were heated in a thermostat to 60 °C for the given

amount of time. To precipitate the polymers the reaction mixture of P(S-co-MABP) was

slowly poured into cold methanol while stirring and for the P(DMAA-co-MABP) polymers

the precipitation solvent was cold diethyl ether. In both cases the colorless polymer was

filtered off and vacuum dried overnight. All copolymers were analyzed by 1H-NMR to verify

the incorporation of benzophenone units (Table 11-2). Molecular weights were determined

by GPC measurements:

P(S-co-MABP(1%)) GPC: Mn = 220000 g·mol-1 , Mw = 415000 g·mol-1.

P(DMAA-co-MABP(1%)) GPC: Mn = 28400 g·mol-1 , Mw = 95300 g·mol-1.

Page 224: through Surface-attached Monomers

210 11 Experimental IV – Silicon Wafers

1Table 11-2. H-NMR measurements of the copolymers reveal the amount of incorporated

benzophenone after polymerization.

Copolymer 1H-NMR integrals Benzophenone content*

5H arom. styrene 9H arom. MABP (mol%)

P(S-co-MABP(1%)) 82.7 13.9 9.3

P(S-co-MABP(5%)) 102.1 31.1 16.9

P(S-co-MABP(10%)) 44.2 24.1 30.3

6H+1H methyl+methin

DMAA

9H arom. MABP

P(DMAA-co-MABP(1%)) 145.0 2.0 1.0

P(DMAA-co-MABP(5%)) 43.4 3.0 5.3

P(DMAA-co-MABP(10%)) 40.8 5.1 9.8

* the P(S-co-MABP) copolymers show a preferred MABP incorporation

11.4.2 Coating & UV-cross-linking

The use of the following copolymers and following coating methods worked out best (cf.

Table 6-1 & Table 6-2): 0.5 g P(S-co-MABP(1%)) were dissolved in 10 mL toluene and applied

to the MPS-modified silicon surface via spin casting with a speed of 2000 rpm (B.L.E.

Delta10). The spin casting procedure was chosen because the PS-co-BP layers tend to dewet

upon dip casting. 0.7 g P(DMAA-co-MABP(1%)) were dissolved in 10 mL isopropanol and

MPS-modified wafer pieces were dipped into the solution and slowly retracted with a speed

of 1 mm·s-1 (Zwick Z2.5). The copolymers with 5 mol% and 10 mol% benzophenone were not

used for further experiments because they had a strong tendency to crosslink early and give

inhomogeneous polymer layers. The air dried copolymer layer thickness was

elipsometrically measured for each sample (cf. Table 6-1 & Table 6-2). Then masks as

described in 11.3 were put onto the coated wafer pieces and illuminated for 100 minutes with

light of a wavelength of 365 nm (Stratalinker, Stratagene). The masks were removed and the

wafers were extracted in either toluene or isopropanol for at least 8 hours. The originated

copolymer network patterns were measured by imaging ellipsometry, profilometry and

visually magnified with a CCD camera.

Page 225: through Surface-attached Monomers

11 Experimental IV – Silicon Wafers 211

11.4.3 Functional layers

In a Schlenk tube under dry nitrogen a polymerization solution consisting of 72 mg AC3pht

(1% based on the styrene amount) and 4.5 mg AIBN dissolved in 6 mL DMF and 3 mL

styrene was prepared. The solution was degassed by 5 freeze & thaw cycles. Then the silicon

wafers with an attached and patterned P(DMAA) polymer network were given under dry

nitrogen into this polymerization mixture. The polymerization reaction was run for 16 h at

60 °C in a water thermostat. After the chosen time period the silicon wafers were taken out of

the polymer solution and extracted with toluene in a Soxhlet apparatus for 16 h. The patterns

were observed with a CCD camera. The generated free copolymer was analyzed with 1H-

NMR and GPC. The patterned surfaces were treated with hydrazine solution for 3 hours in

order to convert the phthalimide function into an amine. The samples were well rinsed with

acetone and methanol. Finally one prepared wafer sample was brought to reaction overnight

in a solution of 5 mg FITC in 5 mL CH2Cl2 and another wafer sample was used for biotin

assay treatment (same reactant solutions used as in 5.2.1). These samples were analyzed with

the biochip fluorescence reader.

The silicon wafers with an attached and patterned PS polymer network were placed in a

degassed polymerization solution of 75 mg MAC2ae (1% based on DMAA amount) and

4.5 mg AIBN dissolved in 3 mL DMAA and 6 mL DMF. The polymerization mixture was

kept at 60 °C for 6 h. GPC measurements on PSS GRAM columns calibrated with PMMA

standards of the free P(DMAA-co-MAC2ae(1%)) GPC gave the following molecular weight

averages: Mn = 64000 g·mol-1 , Mw = 279000 g·mol-1. After copolymerization the wafers were

retracted from the polymer solution and extracted with dichloromethane in a Soxhlet

apparatus for 16 h. Afterwards prepared wafer samples were put overnight into a 1 mmol·L-1

solution of DY-635-NH2 in EtOHabs. to bind the dye to the functional layer. These samples

were again analyzed with the biochip fluorescence reader.

Page 226: through Surface-attached Monomers

212 12 Summary

12 Summary

In this thesis the surface modification of various solid substrates with monomer-bearing

silanes and subsequent use in radical polymerization is described. The surface-attached

monomers provide during the polymerization process the buildup of a covalently bound

polymer monolayer. A variety of homopolymers with different polarities from a strongly

hydrophobic perfluorinated polymer to a very hydrophilic polymer like poly(methacrylic

acid) have been successfully grafted onto solids. Styrene and N,N-dimethyl acrylamide have

been investigated in detail describing the influence of different polymerization parameters

on polymer loading and graft density. In order to determine these values the hybrid

materials were characterized by elemental analysis and GPC. The achieved polymer load,

layer thickness and graft densities are above of those for “grafting-to” systems and below of

those for “grafting-from” systems. As a result of our investigations we find in most cases the

“grafting-to” step dominating. For characterization of the chemical identity of the silanes and

polymer layers DRIFT and XPS measurements were carried out as well as 1H-NMR

spectroscopy of the free polymer. Generally the polymer formation and the graft process

follow the expected behavior for simple radical polymerization as there is much more free

monomer in solution compared to the number of surface-attached monomers on the

substrates. The investigated polymerization parameters are polymerization time, initiator

concentration, monomer concentration in solution and on the substrate and polymerization

temperature. It turns out that this polymerization system is very robust in respect to run-to-

run variations, e.g. monomer density on the surface, residual oxygen in solution, and in

respect to certain polymerization parameter variations depending on the polymerization

behavior of the used free monomer. An influence which cannot be neglected is that of the

surface topology of the substrate onto the amount of polymer attached in the immobilization

reaction. The substrate topology was varied from microporous to mesoporous spherical

substrates and also silicon wafers as typical flat substrates were used. It is shown that the

size of the pores influences the surface accessibility during the graft polymerization process.

Combining the results with substrates of different pore size and with silicon wafers we find

evidence that though the “grafting-to” step has superior influence on the whole grafting

process there is a significant portion of grafted polymer created by the “grafting-from” step.

Also the fact that at each stage of the polymerization new polymer chains are generated that

are small enough to diffuse through the existing polymer monolayer contributes significantly

to the deposition of surface-attached polymer. Therefore we like to call this process using

Page 227: through Surface-attached Monomers

12 Summary 213

surface-attached monomers in radical solution polymerization as a “grafting-through”

technique. Starting from the theoretical considerations of Cohen et al.[149] and Chakrabarti

et al.[152] and the experiments of Hamann et al.[147] we got an detailed insight in the

mechanism of such systems where radical polymerization is carried out in the presence of

surface-attached monomers.

Furthermore with this technique we are able to create copolymer monolayers including

units with functional groups. The buildup of these copolymer layers follows the rules

determined by the polymerization conditions and the used base polymer. The amount of

integrated functional units can be controlled just by using an adequate amount of functional

monomer in solution. We exemplified the high potential of the “grafting-through” process

by combining a base monomer with a functional monomer. As base monomer served mostly

two monomers in our study for which we looked into polymerization behavior in some

detail: styrene and N,N-dimethyl acrylamide. This delivers a functionalized hydrophobic or

hydrophilic polymer monolayer attached to the substrate surface and gives the possibility to

use the hybrid material in different solvent environments later on. The monomers bearing a

functional group are N-acryloyl-N-methylpropyl phthalimide as an amine group precursor

and N-methacryloyl-β-alanine succinimide ester as an activated ester. These two functional

groups are reactive towards a large number of reagents and thus give us a whole palette of

possibilities for further polymeranalogous reactions. This approach can be used as a toolbox

system for a broad range of applications based on the “grafting-through” approach. The

copolymer coated materials were analyzed by elemental analysis and GPC in order to

determine the copolymer loading on the substrates and the corresponding graft densities.

The chemical nature of the surface layers is investigated using DRIFT and XPS

measurements on the solids and 1H-NMR spectroscopy on the free copolymers. Accessibility

and reactivity of the functional groups is tested by labeling and fluorescence experiments. In

the case of N-acryloyl-N-methylpropyl phthalimide where the amino groups are created

after film formation by a deprotection step, additional tests and reaction have been

performed to determine the actual number of amino groups within the copolymer layer on

the hybrid material.

As we are able to generate significant amounts of hybrid material consisting of a SiO2 core

and a functional copolymer monolayer as shell, we use this material in different applications.

In the field of biooriented applications we use glucose oxidase as an example for an enzyme,

biotin for assay procedures and oligonucleotides for DNA analysis. All these compounds are

Page 228: through Surface-attached Monomers

214 12 Summary

immobilized on the functional copolymer layer of the hybrid material with good success. For

the enzyme immobilization the results of the study suggest a quick immobilization reaction.

A stable covalent linkage prevents leaching of the enzyme and the amount of immobilized

enzyme can be controlled. A very promising use of this new functional hybrid material is the

use for catalyst immobilization. Via ligand linkage to a system with polystyrene-amino layer

one is able to immobilize catalytic species. In our case we used different Grubbs catalyst for

immobilization and found very good conversions for exemplary ring closing metathesis

reactions. Despite that leaching was not a problem, the solid material can be easily reloaded

with free Grubbs species in solution.

Based on the studies with silicon wafers where the “grafting-through” approach has been

applied to using styrene or N-N-dimethyl acrylamide the system is expanded to polymer

network formation and fixation. To form polymer networks copolymers including

benzophenone moieties are irradiated. Here the monomers attached to the surface of silicon

wafers are used as attachment points for benzophenone during irradiation. The layer of

surface-attached monomers is much more robust than a layer of benzophenone silane so far

used for such purposes. Not just an alternate attachment for polymer networks onto silicon

wafers has been studied but also photolithographic methods to form patterned polymer

monolayers and networks. The microstructures are characterized by profilometric,

ellipsometric and microscopic measurements. The spatial dimensions of the generated

patterns are in the micrometer range and the layer thicknesses in the nanometer range. For

such systems an application in analytic chip design is possible especially when combined

with functional monomers to terpolymer network layers.

The bottom line is: The “grafting-through” process is a very simple way to attach

functional molecules to a surface. It allows to attach a wide spectrum of polymers to

surfaces, irrespective of the details of the topography of the sample. The recognition of

details of the mechanism is worthwhile as it enables one to built polymer monolayers in a

desired manner within the physically and chemically given boundaries. The docking step of

growing polymer chains (“grafting-to”) in combination with the buildup of a kinetic barrier

by polymer monolayer formation mostly dominates, but partly polymer growth from the

surface (“grafting-from”) comes into play when shorter, growing polymer chains penetrate

the polymer monolayer and successfully bind to the surface, before growing further. Precise

tailoring of chemical and physical properties of the surface of the substrates is possible by

careful choice of the monomers used in solution (co-)polymerization. It is also possible to

Page 229: through Surface-attached Monomers

12 Summary 215

build surface microstructures using surface-attached monomers combined with different

irradiation techniques.

The investigated graft technique is simple and robust and gives nicely reproducible

results. It can be hoped that the “grafting-through” technique becomes an equal partner for

surface decoration with polymers next to the attachment of preformed polymers (“grafting-

to”) and growth of polymers through surface-initiated polymerization reactions (“grafting-

from”).

Page 230: through Surface-attached Monomers

216 13 Zusammenfassung

13 Zusammenfassung

In dieser Arbeit wird die Oberflächenmodifizierung verschiedener fester Substrate mit

Silanen, die eine Monomergruppe tragen, beschrieben. Die so modifizierten Substrate

werden anschließend in einer radikalischen Polymerisation verwendet, wodurch die

oberflächengebundenen Monomere dafür sorgen, dass sich während der Polymerisation eine

kovalent an das Substrat gebundene Polymermonolage aufbaut. Eine Vielfalt von

Homopolymeren mit unterschiedlichen Polaritäten wird erfolgreich auf die Feststoffe

gepfropft. Dabei reicht die Bandbreite vom stark hydrophoben Perfluoropolymer bis hin zu

sehr hydrophilen Polymeren wie Poly(Methylacrylsäure). Styrol und N,N-

Dimethylacrylamid werden im Detail untersucht und der Einfluss verschiedener

Polymerisationsparameter auf Polymerbeladung und Propfdichte der Substrate beschrieben.

Um diese Größen zu bestimmen werden diese Hybridmaterialien mittels Elementaranalyse

und Gelpermeationschromatographie charakterisiert. Die erzielten Polymerbeladungen,

Schichtdicken und Pfropfdichten liegen über denen, die durch Aufpfropfen reaktiver

Polymere („grafting-to“) auf vergleichbare Substrate erreicht werden und unter denen, die

durch oberflächeninitiiertes Polymerwachstum („grafting-from“) erzielt werden. Ein

Ergebnis unserer Untersuchungen ist es, dass in den meisten Fällen der „grafting-to“-Schritt

dominierend ist. Um die Silan- und Polymermonolagen chemisch zu charakterisieren

werden sowohl Diffuse-Reflexion-IR-Spektroskopie (DRIFT) und Röntgen-Photoelektronen-

Spektroskopie (XPS) an den modifizierten Substraten, als auch Protonen-Kernresonanz-

Spektroskopie (1H-NMR) am freien Polymer durchgeführt. Generell folgen die

Polymerbildung und der Pfropfprozess dem erwarteten Verhalten einer einfachen

radikalischen Polymerisation, da im System wesentlich mehr freies Monomer in Lösung als

oberflächengebundenes Monomer auf den Substraten existiert. Die untersuchten

Polymerisationsparameter sind Polymerisationszeit, Initiatorkonzentration,

Monomerkonzentration sowohl in Lösung, als auch auf der Substratoberfläche sowie

Polymerisationstemperatur. Es zeigt sich, dass das untersuchte Polymerisationssystem sehr

unempfindlich gegenüber Schwankungen einzelner Versuchsdurchläufe ist, wie z.B.

Monomerdichte auf der Substratoberfläche, oder Restsauerstoff in der

Polymerisationslösung. Der Einfluss der Oberflächentopologie auf die Menge des während

der Polymerisation gepfropften Polymers kann nicht vernachlässigt werden. Die

Oberflächentopologie der Substrate wird von mikroporösen zu makroporösen kugeligen

Substraten variiert, aber auch planare Substrate, wie z.B. Siliziumscheiben, kommen zum

Page 231: through Surface-attached Monomers

13 Zusammenfassung 217

Einsatz. Es wird gezeigt, dass die Porengröße die Verfügbarkeit der Oberfläche für den

Pfropfprozess beeinflusst. Kombiniert man die Resultate der Substrate mit verschiedenen

Porengrößen und die der Siliziumscheiben miteinander, so finden sich Belege dafür, dass

trotz des überlegenen „grafting-to“-Einflusses, eine signifikante Menge des gepfropften

Polymers durch den „grafting-from“-Schritt erzielt wird. Da in jeder Phase der

Polymerisation neue Polymerketten generiert werden, die klein genug sind, um durch die

bereits existierende Polymermonolage hindurchzudiffundieren, trägt dieser Anteil

maßgeblich mit zur Pfropfmenge bei. Daher schlagen wir vor, den Prozess, der

oberflächengebundene Monomere in einer radikalischen Polymerisation verwendet, als

„grafting-through“-Technik zu bezeichnen. Ausgehend von den theoretischen

Betrachtungen von Cohen et al.[149] und Chakrabarti et al.[152] sowie den Experimenten

von Hamann et al.[147] haben wir einen detaillierten Einblick in die Mechanismen eines

System erlangt, bei dem eine radikalische Polymerisation in der Gegenwart von

oberflächengebundene Monomeren durchgeführt wird.

Weiterhin ist man in der Lage mit dieser Technik oberflächengebundene

Copolymermonolagen herzustellen, die funktionelle Gruppen beinhalten. Der Aufbau dieser

Copolymerschichten erfolgt nach den Regeln für die Polymerisation des verwendeten

Basismonomers. Der Anteil an eingebauten funktionellen Gruppen kann über eine

entsprechende Menge an funktionellem Monomer in Lösung gesteuert werden. Das hohe

Potential des „grafting-through“-Prozesses durch Kombination funktioneller Monomere mit

Basismonomeren wird beispielhaft erläutert. Als Grundmonomere dienen meist zwei

Monomere in unserer Studie, für die das Polymerisationsverhalten im Detail untersucht

wurde: Styrol und N,N-Dimethylacrylamid. Somit wird eine hydrophobe oder hydrophile

funktionelle Copolymermonolage auf die Substratoberfläche gebracht. Dies ermöglicht es,

das Hybridmaterial später in unterschiedlichen Lösungsmitteln zu verwenden. Bei den

verwendeten funktionellen Monomeren handelt es sich um N-Acryloyl-N-Methylpropyl-

Phthalimid als eine Vorstufe für Aminogruppen und N-Methacryloyl-β-Alanin-

Succinimidester als einen aktivierten Ester. Die genannten funktionellen Gruppen sind

gegenüber einer Vielzahl von Reagenzien reaktiv und eröffnen uns eine breite Palette an

polymeranalogen Umsetzungen. Dieser Ansatz, basierend auf der „grafting-through“-

Technik kann somit als Baukastensystem für einen breiten Bereich an Anwendungen genutzt

werden. Die copolymerbeschichteten Materialien werden per Elementaranalyse untersucht,

um die Polymerbeladung zu bestimmen und GPC des freien Copolymers zur

Molekulargewichtsbestimmung ermöglicht eine Berechnung der entsprechenden

Page 232: through Surface-attached Monomers

218 13 Zusammenfassung

Pfropfdichten. Chemisch charakterisiert werden die copolymeren Oberflächenschichten

mittels DRIFT und XPS Messungen. Die freien Copolymere werden mit 1H-NMR-

Spektroskopie untersucht. Zugänglichkeit und Reaktivität der copolymerisierten

funktionellen Gruppen werden durch Markierung mit Farbstoffen und

Fluoreszenzuntersuchungen getestet. Im Falle des N-Acryloyl-N-Methylpropyl-Phthalimids,

bei dem nach der Copolymerisation die Aminogruppen entschützt werden, werden

zusätzliche Tests durchgeführt, um die tatsächliche Anzahl der Aminogruppen in der

oberflächengebundenen Copolymerschicht zu bestimmen.

Da wir in der Lage sind einiges an Hybridmaterialien herzustellen, die aus einem SiO2-

Kern und einer Schale, gebildet durch eine Copolymermonoschicht, bestehen, verwenden

wir diese Materialen für unterschiedliche Anwendungen. Im Bereich der biologisch

orientierten Anwendungen verwenden wir folgende Beispiele: Glucoseoxidase für ein

Enzym, Biotin für einen Assay und ein Oligonukleotid für die DNA-Analyse. All diese

Moleküle werden erfolgreich durch die funktionellen Copolymerschichten auf den

Hybridmaterialien gebunden bzw. immobilisiert. Für die Enzymimmobilisierung lässt sich

aus den Ergebnissen eine schnelle und stabile Bindung an die Oberfläche folgern. Die Menge

des Enzyms auf der Oberfläche kann gesteuert werden und es verrichtet weiterhin seine

enzymatisch Arbeit. Ein sehr vielversprechendes Einsatzgebiet für die neuen

funktionalisierten Hybridmaterialien ist die Katalysatorimmobilisierung. Die katalytische

Spezies bindet auf die Oberfläche per Ligandenaustausch an einen Liganden, der an die

vorhandenen Aminogruppen geknüpft wurde. In unserem Fall verwenden wir für die

Immobilisierung verschiedene Grubbs-Katalysatoren. Die entstandenen

katalysatorbeladenen Materialien zeigen bei exemplarisch durchgeführten Ringschluss-

Metathese-Reaktionen sehr gute Umsätze. Obwohl Auswaschen des Katalysators bei diesem

System kein Problem ist, kann das Feststoffmaterial durch Zugabe in eine Lösung des

entsprechenden Katalysators leicht wieder beladen werden.

Basierend auf den Studien mit Siliziumscheiben als Substrat, werden

oberflächengebundene Monomere durch Belichtung (Ablation) mikrostrukturiert, um in der

Folge mikrostrukturierte Polymerschichten durch den „grafting-through“-Ansatz zu

generieren. Weiterhin können oberflächengebundene dazu eingesetzt werden, um polymere

Netzwerke auf der Oberfläche anzubinden, die durch Bestrahlung von einpolymerisierten

Benzophenon-Einheiten gebildet werden. Es stellt sich heraus, das die Schicht

oberflächengebundener Monomere sich robuster handhaben lässt als die für gleiche Zwecke

Page 233: through Surface-attached Monomers

13 Zusammenfassung 219

eingesetzten Benzophenon-Silan-Schichten. Es werden nicht nur polymere Netzwerke

gebildet, sondern auch photolithographische Methoden angewendet, um mikrostrukturierte

Netzwerkarchitekturen aufzubauen. Die räumlichen Abmessungen der hergestellten

Polymerstrukturen werden durch Profilometrie, Ellipsometrie und Mikroskopie

charakterisiert. Die Ausdehnung der Strukturen in der Ebene liegt im Mikrometerbereich

und die Schichtdicken im Nanometerbereich. Ausgestattet mit funktionellen Co- oder

Terpolymeren ist für solche Systeme eine Anwendung im Bereich analytischer Chips

denkbar.

Die Quintessenz ist, dass der „grafting-through“-Prozess ein sehr einfacher Weg ist, um

funktionale Moleküle an eine Oberfläche zu binden. Er erlaubt das Anknüpfen eines weiten

Spektrums an Polymeren auf Oberflächen, unabhängig von Topologiedetails der

verwendeten Substrate. Die Erkenntnis der Details des Mechanismus ist lohnenswert und

hilfreich insofern dies es ermöglicht polymere Monoschichten in gewünschter Form,

innerhalb der physikalisch und chemisch gegebenen Grenzen, aufzubauen. Der

Andockschritt wachsender Polymerketten („grafting-to“), der den Aufbau einer kinetischen

Barriere in Form einer polymeren Monoschicht mit sich bringt, ist meist dominierend. Aber

teilweise wird Polymerwachstum von der Oberfläche („grafting-from“) mit ins Spiel

gebracht, dadurch dass kürzere, wachsende Polymerketten die Polymermonoschicht

durchdringen und erfolgreich an die oberflächengebundenen Monomere anbinden, um dann

von der Oberfläche wegzuwachsen. Präzises Maßschneidern von chemischen und

physikalischen Eigenschaften der Oberflächen der Substrate ist durch eine durchdachte

Wahl der eingesetzten Monomere für die Copolymerisation in Lösung möglich. Es ist auch

möglich Oberflächenmikrostrukturen mittels oberflächengebundener Monomere durch

verschiedene Belichtungsmethoden zu erzeugen.

Die untersuchte Pfropftechnik ist einfach, robust und ergibt gut reproduzierbare Resultate.

Es bleibt zu hoffen, dass die „grafting-through“-Technik es schafft, ebenbürtig neben der

Anbindung von reaktiven, fertigen Polymeren („grafting-to“) und dem Polymerwachstum

ausgehend von oberflächengebundenen Initiatoren („grafting-from“), für die Ausstattung

von Oberflächen mit Polymeren eingesetzt zu werden.

Page 234: through Surface-attached Monomers

220 14 References

14 References

1. Rasmussen, K.E., Quantitative determination of heptacosane and nonacosane in Norwegian-grown Cannabis plants. Medd. Nor. Farm. Selsk., 1975. 37(2): p. 128.

2. Grant, W.F. and Whetter, J.M., The cytogenetics of Lotus. XI. The use of thin-layer chromatography in the separation of secondary phenolic compounds in Lotus. J. Chromatogr., 1966. 21(2): p. 247.

3. Millot, M.C. and Sebille, B., Rapid preparation of bovine mercaptalbumin by means of covalent chromatography on silica-based materials. J. Chromatogr., 1987. 408: p. 263.

4. Lobarzewski, J., Fiedurek, J., Ginalska, G., and Wolski, T., New matrices for the purification of pectinases by affinity chromatography. Biochem. Biophys. Res. Commun., 1985. 131(2): p. 666.

5. Pugin, B. and Mueller, M., Enantioselective metal complex catalysts immobilized on inorganic supports via carbamate links. Stud. Surf. Sci. Catal., 1993. 78(Heterogeneous Catalysis and Fine Chemicals III): p. 107.

6. Ohnaka, T., Komiya, K., and Moriyama, H., Packing material for reversed phase chromatography and process for its preparation, in Eur. Pat. Appl. 1991, (Tosoh Corp., Japan). Ep. p. 9 pp.

7. Kamiusuki, T., Morioka, K., Kadode, T., and Okawa, N., Hydrophilic silane coupling agent for liquid chromatograph packing material, in Jpn. Kokai Tokkyo Koho. 2000, (Neos Co., Ltd., Japan). Jp. p. 6 pp.

8. Williams, D.E., Preparation of liquid chromatography packing materials having ketal-blocked diol groups on porous support surfaces, in U.S. 1988, (Dow Corning Corp., USA). Us. p. 5 pp.

9. Hauck, H.E. and Jost, W., Stationary-phase material for thin-layer chromatography, in Ger. Offen. 1986, (Merck Patent G.m.b.H., Fed. Rep. Ger.). De. p. 29 pp.

10. Bente Iii, P.F., Zerenner, E.H., and Dandeneau, R.D., Chromatographic column, in Ger. Offen. 1980, (Hewlett-Packard Co., USA). De. p. 16 pp.

11. Imura, S. and Fukano, K., Support for chromatography, in Jpn. Kokai Tokkyo Koho. 1977, (Toyo Soda Mfg. Co., Ltd., Japan). Jp. p. 7 pp.

12. Kanda, T., Sakamoto, A., Ohta, T., and Ohtsu, Y., Manufacture of column packing materials for chromatography using porous supports and silicone polymers, in Eur. Pat. Appl. 1991, (Shiseido Co., Ltd., Japan). Ep. p. 11 pp.

13. Komiya, Y., Sasaki, H., and Fukano, K., Hydrophilic packing materials, in Jpn. Kokai Tokkyo Koho. 1980, (Toyo Soda Mfg. Co., Ltd., Japan). Jp. p. 10 pp.

14. Yang, X., Dai, J., and Carr, P.W., Analysis and critical comparison of the reversed-phase and ion-exchange contributions to retention on polybutadiene coated zirconia and octadecyl silane bonded silica phases. J. Chromatogr., A, 2003. 996(1-2): p. 13.

15. Ascah, T.L., Kallury, K.M.R., Szafranski, C.A., Corman, S.D., and Liu, F., Characterization and high performance liquid chromatographic evaluation of a new amide-functionalized reversed phase column. J. Liq. Chrom. Relat. Tech., 1996. 19(17-18): p. 3049.

Page 235: through Surface-attached Monomers

14 References 221

16. Jackson, T., Clark, J.H., Macquarrie, D.J., and Brophy, J.H., Base catalyst immobilised on silica coated reactor walls for use in continuous flow systems. Green Chem., 2004. 6(4): p. 193.

17. Bruckner, H. and Wachsmann, M., Liquid chromatographic separation of amino acid enantiomers on a silica-bonded chiral s-triazine column. J. Chromatogr. A, 1996. 728(1-2): p. 447.

18. Silva, C.R., Jardim, I., and Airoldi, C., Development of new urea-functionalized silica stationary phases - Characterization and chromatographic performance. Journal of Chromatography A, 2001. 913(1-2): p. 65.

19. Chen, L., Zhang, L.F., Ching, C.B., and Ng, S.C., Synthesis and chromatographic properties of a novel chiral stationary phase derived from heptakis(6-azido-6-deoxy-2,3-di-O-phenylcarbamoylated)-beta-cyclodextrin immobilized onto amino-functionalized silica gel via multiple urea linkages. J. Chromatogr. A, 2002. 950(1-2): p. 65.

20. Ruffing, F.J., Lux, J.A., Roeder, W., and Schomburg, G., Chiral stationary phases for LC and SFC obtained by ʺpolymer coatingʺ. Chromatographia, 1988. 26: p. 19.

21. Wang, L., Lu, S., Gao, P., Li, S., and Huang, Q., Preparation and characterization of chiral packing materials for high performance liquid chromatography: cellulose tris(4-methylbenzoate) coated on silica gel. Sepu, 1998. 16(4): p. 285.

22. Wang, L., Lu, S., Shen, R., and Li, S., Preparation and characterization of chiral packing materials for high performance liquid chromatography: cellulose triphenylcarbamate coated on silica gel. Fenzi Cuihua, 1997. 11(6): p. 439.

23. Felix, G. and Zhang, T., Chiral packing materials for high-performance liquid chromatographic resolution of enantiomers based on substituted branched polysaccharides coated on silica gel. J. Chromatogr., 1993. 639(2): p. 141.

24. Okamoto, Y., Kawashima, M., and Hatada, K., Chromatographic resolution. 7. Useful chiral packing materials for high-performance liquid chromatographic resolution of enantiomers: phenylcarbamates of polysaccharides coated on silica gel. J. Am. Chem. Soc., 1984. 106(18): p. 5357.

25. Okamoto, Y., Kawashima, M., Yamamoto, K., and Hatada, K., Chromatographic resolution. 6. Useful chiral packing materials for high-performance liquid chromatographic resolution. Cellulose triacetate and tribenzoate coated on macroporous silica gel. Chem. Lett., 1984(5): p. 739.

26. Cao, S.G., Liu, Z.B., Feng, Y., Ma, L., Ding, Z.T., and Cheng, Y.H., Esterification and transesterification with immobilized lipase in organic solvent. Appl. Biochem. Biotechnol., 1992. 32: p. 1.

27. Yee, L.N., Akoh, C.C., and Phillips, R.S., Pseudomonas sp. lipase-catalyzed synthesis of geranyl esters by transesterification. J. Am. Oil Chem. Soc., 1995. 72(11): p. 1407.

28. Fratzke, A.R., Jr., Factors affecting the activity and stability of glucose isomerase immobilized on porous glass. 1974, Dep. Chem. Eng., Iowa State Univ., Ames, IA, USA. p. 127 pp.

29. Ramachandran, K.B. and Perlmutter, D.D., Effects of immobilization on the kinetics of enzyme-catalyzed reactions. II. Urease in a packed-column differential reactor system. Biotechnol. Bioeng., 1976. 18(5): p. 685.

Page 236: through Surface-attached Monomers

222 14 References

30. Ramachandran, K.B. and Perlmutter, D.D., Effects of immobilization on the kinetics of enzyme-catalyzed reactions. I. Glucose oxidase in a recirculation reactor system. Biotechnol. Bioeng., 1976. 18(5): p. 669.

31. Pollak, A., Blumenfeld, H., Wax, M., Baughn, R.L., and Whitesides, G.M., Enzyme immobilization by condensation copolymerization into crosslinked polyacrylamide gels. J. Am. Chem. Soc., 1980. 102(20): p. 6324.

32. Parthasarathy, R.V. and Martin, C.R., Synthesis of polymeric microcapsule arrays and their use for enzyme immobilization. Nature, 1994. 369(6478): p. 298.

33. Taqieddin, E. and Amiji, M., Enzyme immobilization in novel alginate-chitosan core-shell microcapsules. Biomaterials, 2004. 25(10): p. 1937.

34. Atallah, M.T., Wasserman, B.P., and Hultin, H.O., Increasing catalyst activity by immobilization of enzyme conjugates. Biotechnol. Bioeng., 1976. 18(12): p. 1833.

35. Dixon, J.E., Stolzenbach, F.E., Berenson, J.A., and Kaplan, N.O., Immobilized enzymes. Catalytic properties of lactate dehydrogenase covalently attached to glass beads. Biochem. Biophys. Res. Commun., 1973. 52(3): p. 905.

36. Mody, H.M. and Mody, K.H., Immobilization of alpha-amylase on porous silica. Indian J. Chem. A, 1999. 38(11): p. 1200.

37. Tiller, J.C., Rieseler, R., Berlin, P., and Klemm, D., Stabilization of activity of oxidoreductases by their immobilization onto special functionalized glass and novel aminocellulose film using different coupling reagents. Biomacromolecules, 2002. 3(5): p. 1021.

38. Kim, J.-K., Shin, D.-S., Chung, W.-J., Jang, K.-H., Lee, K.-N., Kim, Y.-K., and Lee, Y.-S., Effects of polymer grafting on a glass surface for protein chip applications. Colloids Surf., B, 2004. 33(2): p. 67.

39. Nakamura, M., Kumazawa, T., and Kogure, E., Insoluble chemical composite material suitable as a biocatalytic or affinity chromatographic carrier and method for its preparation, in Eur. Pat. Appl. 1987, (Olympus Optical Co., Ltd., Japan). Ep. p. 13 pp.

40. Abraham, M., Boross, L., and Szajani, B., Immobilization of glucose oxidase on silica-based Prikl. Biokhim. Mikrobiol., 1988. 24(4): p. 499. supports.

41. Bulmus, V., Ayhan, H., and Piskin, E., Modified PMMA monosize microbeads for glucose oxidase immobilization. Chem. Eng. J., 1997. 65(1): p. 71.

42. Yang, X.H., Hua, L., Gong, H.Q., and Tan, S.N., Covalent immobilization of an enzyme (glucose oxidase) onto a carbon sol-gel silicate composite surface as a biosensing platform. Anal. Chim. Acta, 2003. 478(1): p. 67.

43. Kumar, P., Agrawal, S.K., Misra, A., and Gupta, K.C., A new heterobifunctional reagent for immobilization of biomolecules on glass surface. Bioorg. Med. Chem. Lett., 2004. 14(5): p. 1097.

n adsorption and cell growth on glass surfaces bearing immobilized glycerol monosaccharides.

45. Sackmann, E. and Tanaka, M., Supported membranes on soft polymer cushions: fabrication, characterization and applications. Trends Biotechnol, 2000. 18(2): p. 58.

46. Loidl-Stahlhofen, A., Schmitt, J., Noller, J., Hartmann, T., Brodowsky, H., Schmitt, W., and Keldenich, J., Solid-supported biomolecules on modified silica surfaces - a tool for fast

44. Bain, J.R. and Hoffman, A.S., Glycophase glass revisited: protei Biomaterials, 2002. 23(16):

p. 3347.

Page 237: through Surface-attached Monomers

14 References 223

physicochemical characterization and high-throughput screening. Adv. Mater. (Weinheim, Ger.), 2001. 13(23): p. 1829.

47. Mateo, C., Abian, O., Fernandez-Lafuente, R., and Guisan, J.M., Reversible enzyme immobilization via a very strong and nondistorting ionic adsorption on support-polyethylenimine composites. Biotechnol. Bioeng., 2000. 68(1): p. 98.

48. Watsuji, T. and Cass, A., Biomolecular switch for biosensors or data acquisition/processing devices containing immobilized biological macromolecules, in Brit. UK Pat. Appl. 1993, (Sharp Kabushiki Kaisha, Japan). Gb. p. 36 pp.

49. Barie, N., Development of a universal process for biofunctionalization of surfaces to be used in sensor technology by surface acoustic wave (SAW) devices. 2000, Inst. Instrumentelle Analytik, FZ Karlsruhe, Germany. p. 1.

50. Walsh, M.K., Wang, X.W., and Weimer, B.C., Optimizing the immobilization of single-stranded DNA onto glass beads. J. Biochem. Biophys. Meth., 2001. 47(3): p. 221.

51. Ghosh, S.S. and Musso, G.F., Covalent Attachment of Oligonucleotides to Solid Supports. Nucleic Acids Res., 1987. 15(13): p. 5353.

52. Jin, L. and Horgan, A., Preparation of end-tethered DNA monolayers on siliceous surfaces using heterobifunctional cross-linkers. Langmuir, 2003. 19(17): p. 6968.

53. Thompson, M. and McGovern, M.E., High-surface-density covalent immobilization of oligonucleotide monolayers, in PCT Int. Appl. 1999, (Sensorchem International Corporation, Can.). Wo. p. 68 pp.

54. Kirschning, A., Altwicker, C., Drager, G., Harders, J., Hoffmann, N., Hoffmann, U., Schonfeld, H., Solodenko, W., and Kunz, U., PASSflow syntheses using functionalized monolithic polymer/glass composites in flow-through microreactors. Angew. Chem. Int. Ed., 2001. 40(21): p. 3995.

55. Gewehr, M., Nakamura, K., Ise, N., and Kitano, H., Gel-Permeation Chromatography Using Porous-Glass Beads Modified with Temperature-Responsive Polymers. Macromol. Chem. Phys., 1992. 193(1): p. 249.

56. Kapustin, D.V., Yagudaeva, E.Y., Zavada, L.L., Zhigis, L.S., Zubov, V.P., Yaroshevskaya, E.M., Plobner, L., Leiser, R.M., and Brem, G., A composite polyaniline-containing silica sorbent for DNA isolation. Russ. J. Bioorg. Chem., 2003. 29(3): p. 281.

57. Kim, J., Park, H., Jung, D., and Kim, S., Protein immobilization on plasma-polymerized ethylenediamine-coated glass slides. Anal. Biochem., 2003. 313(1): p. 41.

58. Pathak, S., Singh, A.K., McElhanon, J.R., and Dentinger, P.M., Dendrimer-Activated Surfaces for High Density and High Activity Protein Chip Applications. Langmuir, 2004. 20(15): p. 6075.

59. Pohl, C.A. and Papanu, S.C., Latex based adsorbent chip and its preparation and use in (bio)assays, in U.S. Pat. Appl. Publ. 2003, (Ciphergen Biosystems, Inc., USA). Us. p. 39 pp.

60. Freidank, D., Oberflächengebundene funktionelle Polymernetzwerke als Matrix für Nukleinsäure-Microarrays, in Institute for Microsystemtechnology. 2005, Albert-Ludwigs-University: Freiburg im Breisgau.

61. Oh, S.J. and Cho, S.J., Characteristics of DNA microarrays fabricated on various aminosilane layers. Langmuir, 2002. 18(5): p. 1764.

Page 238: through Surface-attached Monomers

224 14 References

62. Snyder, M., Reed, M., Zhu, H., and Klemic, J.F., High density protein arrays for screening of protein activity, in PCT Int. Appl. 2001, (Yale University, USA). Wo. p. 69 pp.

63. Melnyk, O., Gras-Masse, H., Olivier, C., Durand, J.-O., Auriault, C., Duburcq, X., Bouzidi, A., Garcia, J.-M., and El-Mahdi, O., Preparation and use of a peptide biochip with semicarbazone linkage, in PCT Int. Appl. 2002, (Societe DʹEtude et de Developpement des Antigenes Combinatoires - Sedac Therapeutics, Fr.; Centre National de la Recherche Scientifique CNRS; Institut Pasteur de Lille; Universite de Lille.): Wo. p. 49 pp.

64. Weber, A., Schiestel, T., Tovar, G., Brunner, H., Rupp, S., and Hauser, N., Microarrays with improved structured-functional bonding matrices for biomolecules, in PCT Int. Appl. 2003, (Fraunhofer-Gesellschaft Zur Foerderung Der Angewandten Forschung E.V., Germany). Wo. p. 92 pp.

65. Dickinson, T., Coblentz, K.D., and Carlson, E., Alternative substrates and formats for bead-based array of arrays for detection of biomolecules, in U.S. Pat. Appl. Publ. 2002, (Illumina, Inc., USA). Us. p. 56 pp.

66. Arai, H., Kaneko, T., and Kunugi, T., Catalytic vapor phase hydroformylation of olefins over polymer-immobilized rhodium complexes. Chem. Lett., 1975(3): p. 265.

67. Hong, L. and Ruckenstein, E., Immobilization of alkoxylated phosphine ligands and their Rh complexes to a silica surface coated with an organic mono- or multilayer. J. Mol. Catal., 1994. 90(3): p. 303.

68. Berry, M., Champaneria, R.K., and Howell, J.A.S., The Preparation and Catalytic J. Mol. Catal., 1986.

(2-3): p. 243. Activity of Rhodium(I) Complexes of Isonitrile-Functionalized Silica.37

69. Kurusu, Y. and Neckers, D.C., Functionalization of Silica-Gel - Application for the Catalytic-Oxidation of Alkanes. J. Org. Chem., 1991. 56(6): p. 1981.

70. Kurusu, Y., Functionalization of Silica .2. Oxidation of Cyclohexanes in the Presence of Functionalized Silica Complexed by Iron and Copper. Makromol. Chem. Macromol. Symp., 1992. 59: p. 313.

71. Sarkany, A., Beck, A., Horvath, A., Revay, Z., and Guczi, L., Acetylene hydrogenation on sol-derived Pd/SiO2. Appl. Catal. Gen., 2003. 253(1): p. 283.

72. Pan, Y., Zhang, Z.Y., and Hu, H.W., Amino-Functionalized Silica-Supported Palladium Catalysts and Their Applications in Heck Olefination. J. Mol. Catal., 1990. 62(3): p. 297.

73. Hultman, H.M., de Lang, M., Nowotny, M., Arends, I., Hanefeld, U., Sheldon, R.A., and Maschmeyer, T., Chiral catalysts confined in porous hosts. 1. Synthesis. J. Catal., 2003. 217(2): p. 264.

74. Hultman, H.M., de Lang, M., Arends, I., Hanefeld, U., Sheldon, R.A., and Maschmeyer, T., Chiral catalysts confined in porous hosts. 2. Catalysis. J. Catal., 2003. 217(2): p. 275.

75. Kingsbury, J.S., Garber, S.B., Giftos, J.M., Gray, B.L., Okamoto, M.M., Farrer, R.A., Fourkas, J.T., and Hoveyda, A.H., Immobilization of olefin metathesis catalysts on monolithic sol-gel: practical, efficient, and easily recyclable catalysts for organic and combinatorial synthesis. Angew. Chem. Int. Ed., 2001. 40(22): p. 4251.

76. Buchmeiser, M.R., Nuyken, O., and Krause, J.O., Preparation of heterogeneous metathesis catalysts by fixation to supports, in Austrian. 2004, (Austria). At. p. 18 pp.

Page 239: through Surface-attached Monomers

14 References 225

77. Mingotaud, A.-F., Reculusa, S., Mingotaud, C., Keller, P., Sykes, C., Duguet, E., and Ravaine, S., Ring-opening metathesis polymerization on well defined silica nanoparticles leading to hybrid core-shell particles. J. Mater. Chem., 2003. 13(8): p. 1920.

78. Krause, J.O., Lubbad, S., Nuyken, O., and Buchmeiser, M.R., Monolith- and silica-supported carboxylate-based Grubbs - Herrmann-type metathesis catalysts. Adv. Synth. Catal., 2003. 345(8): p. 996.

79. Hollemann, A.F. and Wiberg, E., Lehrbuch der Anorganischen Chemie. 101. ed. 1995, Berlin: de Gruyter.

80. Dislich, H. and Hinz, P., History and Principles of the Sol-Gel Process and some new Multicomponent Oxide Coatings. J. Non-Cryst. Sol., 1982. 48(1): p. 11.

81. Roy, R., Sol-Gel Process - Origins, Products, Problems. Am. Ceram. Soc. Bull., 1981. 60(3): p. 363.

82. Puyane, R. and James, P.F., Preparation of Silica and Soda-Silica Glasses by the Sol-Gel Process. J. Non-Cryst. Sol., 1980. 41(1): p. 105.

83. Kistler, S.S., Coherent expanded aerogels. J. Phys. Chem., 1932. 36(1): p. 52. 84. Teichner, S.J. and Nicolaon, G.A., Method of preparing inorganic aerogels, in U.S. Pat.

Appl. Publ. 1972, France Etat: US. 85. Hüsing, N. and Schubert, U., Aerogels - Airy Materials: Chemistry, Structure and

Properties. Angew. Chem. Int. Ed., 1998. 37(1/2): p. 22. 86. Hench, L.L. and West, J.K., The Sol-Gel Process. Chem. Rev., 1990. 90: p. 33. 87. Tewari, P.H., Hunt, A.J., and Lofftus, K.D., Ambient-temperature supercritical drying of

transparent silica aerogels. Mat. Lett., 1985. 3(9-10): p. 363. 88. Stöber, W., Fink, A., and Bohn, E., Controlled Growth of Monodisperse Silica Spheres in

the Micron Size Range. J. Colloid Interface Sci., 1968. 26: p. 62. 89. HDK pyrogene Kieselsäure. 2005, Wacker Chemie AG. 90. Perry, R. and Adams, M., Tospearl: Silicone Resin for Industrial Applications, in Silicones

and Silicone-modified Materials (ACS Symp. Ser., S. Clarson, Editor. 2000, American Chemical Society: Washington, DC. p. 533.

91. Cheng, Y., Walter, T., Lu, Z., Iranota, P., Alden, B., Gendreau, L., Neue, U., Grassi, J., Carnody, J., OʹGara, J., and Fisk, R., LC-GC, 2000. 18: p. 1162.

92. Rich, J., Cella, J., Lewis, L., Stein, J., Singh, N., Rubinstajn, S., and Wengrovius, J., in Kirk-Othmer Encyclopedia of Chemical Technology. 1997, John Wiley and Sons: New York. p. 114.

93. Lentz, C., Inorg. Chem., 1964. 4: p. 574. 94. Sobieski, L., in Handbook of Pressure Sensitive Adhesive Technology, D. Satas, Editor.

1989, van Nostrand Reinhold: New York. 95. Lin, S., J. Appl. Polym. Sci., 1994. 54: p. 2135. 96. Menz, W., Mohr, J., and Paul, O., Mikrosystemtechnik für Ingenieure. 3rd ed. 2005,

Weinheim: Wiley-VCH. 574. 97. Mescheder, U., Mikrosystemtechnik. Konzepte und Anwendungen. 2nd ed. 2004,

Wiesbaden: Teubner Verlag. 292. 98. Samuel, J.D.J.S., Ruther, P., Frerichs, H.-P., Lehmann, M., Paul, O., and Rühe, J., A

simple route towards the reduction of surface conductivity in gas sensor devices. Sens. Act. B, 2005. 110(2): p. 218.

Page 240: through Surface-attached Monomers

226 14 References

99. Deelder, R.S., Claassen, P.J.W.M., and Hendricks, P.J.H., Chemically bonded stationary phases on diatomite supports for use in high-performance liquid chromatography. J. Chromatogr., 1974. 91: p. 201.

100. Asmus, P.A., Low, C.-E., and Novotny, M., Preparation and chromatographic evaluation

J. Chromatogr., 1976. 123(1): p. 109. of chemically bonded ion-exchange stationary phases. II. Weak and strong cation exchangers.

101. Gawdzik, J., Suprynowicz, Z., and Jaroniec, M., Studies of chromatographic packings comprising chemically bonded phases obtained from porous glass beads. I. Sorption properties and surface heterogeneity. J. Chromatogr., 1977. 131: p. 7.

organic material from seawater with reversed-phase techniques.

103. Velazquez, A., Leonard, J., and Cote, J.E., Surface free energy of silane-covered materials measured by gas chromatography. Chromatographia, 1989. 27(9-10): p. 455.

104. Tripp, C.P. and Hair, M.L., Chemical Attachment of Chlorosilanes to Silica - A Two-Step Amine-Promoted Reaction. J. Phys. Chem., 1993. 97(21): p. 5693.

105. Hair, M.L. and Tripp, C.P., Alkylchlorosilane Reactions at the Silica Surface. Colloid. Surface. A, 1995. 105(1): p. 95.

106. Kurth, D.G. and Bein, T., Surface-Reactions on Thin-Layers of Silane Coupling Agents. Langmuir, 1993. 9(11): p. 2965.

107. Eklund, T. and Britcher, L., Thermogravimetric analysis of gamma-aminopropyl-trimethoxysilane adsorbed on silica support. J. Therm. Anal., 1999. 58(1): p. 67.

108. Trens, P. and Denoyel, R., Adsorption of (Gamma-Aminopropyl)Triethoxysilane on Silica from Aqueous Solution - a Microcalorimetric Study. Langmuir, 1995. 11(2): p. 551.

109. Moon, J.H. and Shin, J.W., Formation of uniform aminosilane thin layers: An Imine formation to measure relative surface density of the amine group. Langmuir, 1996. 12(20): p. 4621.

110. Kallury, K.M.R., Macdonald, P.M., and Thompson, M., Effect of Surface-Water and Base Catalysis on the Silanization of Silica by (Aminopropyl)Alkoxysilanes Studied by X-Ray Photoelectron-Spectroscopy and C-13 Cross-Polarization Magic-Angle-Spinning Nuclear-Magnetic-Resonance. Langmuir, 1994. 10(2): p. 492.

111. Yamaguchi, M. and Nakamura, Y., Reactivity of silane coupling agent with surface of silica particles. Polymer. Polymer Compos., 1998. 6(2): p. 85.

112. McGovern, M.E., Kallury, K.M.R., and Thompson, M., Role of Solvent on the Silanization of Glass with Octadecyltrichlorosilane. Langmuir, 1994. 10(10): p. 3607.

113. Berendsen, G.E. and Galan, L.d., Preparation and Chromatographic Properties of Some Chemically Bonded Phases for Reversed-Phase Liquid Chromatography. J. Liq. Chromatogr., 1978. 1(5): p. 561.

114. Ulmann, A., An introduction to ultrathin Organic Films. 1991, New York: Academic Press.

115. Halperin, A., Tirrell, M., and Lodge, T.P., Tethered Chains in Polymer Microstructures. Adv. Polym. Sci., 1992. 100: p. 31.

102. Derenbach, J.B., Ehrhardt, M., Osterroht, C., and Petrick, G., Sampling of dissolved Mar. Chem., 1978. 6(4): p.

351.

116. Halperin, A., Tirrell, M., and Lodge, T.P., Macromol. Rep. Suppl., 1992. 29: p. 1.

Page 241: through Surface-attached Monomers

14 References 227

117. Fleer, G.J., Cohen-Stuart, M.A., Scheutjens, J.M.H.M., Cosgrove, T., and Vincent, B., Polymers at Interfaces. 1993, London: Chapman & Hall.

118. Tsubokawa, N., Hosoya, M., Yanadori, K., and Sone, Y., Grafting onto Carbon-Black - Reaction of Functional-Groups on Carbon-Black with Acyl Chloride-Capped Polymers. J. Macromol. Sci. Chem., 1990. A27(4): p. 445.

119. Krenkler, K.P., Laible, R., and Hamann, K., Polyreactions on Pigment Surfaces .7. Reactions of Polymers with Chlorosilane End Groups on Silicon Dioxide Surfaces. Angew. Makromol. Chem., 1976. 53(JUL): p. 101.

120. Benouada, H., Hommel, H., Legrand, A.P., Balard, H., and Papirer, E., Organization of the Layers of Polyethylene Oxide Grafted with Different Densities on Silica. J. Colloid Interface Sci., 1988. 122(2): p. 441.

121. Bridger, K. and Vincent, B., The Terminal Grafting of Poly(Ethylene Oxide) Chains to Silica Surfaces. Eur. Polymer J., 1980. 16(10): p. 1017.

122. Boven, G., Oosterling, M., Challa, G., and Schouten, A.J., Grafting Kinetics of Poly(Methyl Methacrylate) on Microparticulate Silica. Polymer, 1990. 31(12): p. 2377.

123. Carlier, E., Guyot, A., and Revillon, A., Functional Polymers Supported on Porous Silica .2. Radical Polymerization of Vinylbenzyl Chloride from Grafted Precursors. Reactive Polymers, 1992. 16(2): p. 115.

124. Fery, N., Hamann, K., and Hoene, R., Graft Reactions on SiO2 Surfaces. Angew. Chem. Int. Ed., 1972. 11(4): p. 337.

125. Prucker, O. and Rühe, J., Synthesis of poly(styrene) monolayers attached to high surface area silica gels through self-assembled monolayers of azo initiators. Macromolecules, 1998. 31(3): p. 592.

126. Prucker, O. and Rühe, J., Mater. Res. Soc. Symp. Proc., 1993. 304: p. 1675. 127. Tsubokawa, N. and Ishida, H., Graft-Polymerization of Methyl-Methacrylate from Silica

Initiated by Peroxide Groups Introduced onto the Surface. J. Polym. Sci. A, 1992. 30(10): p. 2241.

128. Haddleton, D.M., Duncalf, D.J., Kukulj, D., and Radigue, A.P., 3-aminopropyl silica supported living radical polymerization of methyl methacrylate: Dichlorotris(triphenylphosphine)ruthenium(II) mediated atom transfer polymerization. Macromolecules, 1999. 32(15): p. 4769.

129. Kong, X.X. and Kawai, T., Amphiphilic polymer brushes grown from the silicon surface by atom transfer radical polymerization. Macromolecules, 2001. 34(6): p. 1837.

Brushes on Silicon Surfaces by Atom Transfer Radical Polymerization.

131. Tu, H. and Heitzmann, C.E., Patterned poly(N-isopropylacrylamide) brushes on silica surfaces by microcontact printing followed by surface-initiated polymerization. Langmuir, 2004. 20(19): p. 8313.

132. Advincula, R.C., Brittain, W., Caster, K., and Rühe, J., Polymer Brushes: Synthesis, Characterization and Applications. 2004, New York: Wiley. 496.

133. Hubbard, A.T., Encyclopedia of Surface & Colloid Science. Adsorption of Polymers at Colloid Surfaces. 2002, New York: Marcel Dekker Inc. 5640.

130. Ramakrishnan, A., Dhamodharan, R., and Rühe, J., Growth of Poly(methyl methacrylate) J. Polym. Sci. A,

2006. 44: p. 1758.

Page 242: through Surface-attached Monomers

228 14 References

134. Negievich, L.A. and Kachan, A.A., Photochemical Vapor-Phase Graft-Polymerization on Aerosil with a Phosphorus-Containing Coating. Vysokomolekulyarnye Soedineniya Seriya B, 1976. 18(11): p. 809.

135. Negievich, L.A., Photochemical Graft-Polymerization of Vinyl Monomers on Aerosil Induced by Surface Peroxide Radicals. Ukrainskii Khimicheskii Zhurnal, 1977. 43(2): p. 176.

136. Ivanchev, S.S., Dmitrenko, A.V., Polozov, B.V., Demidova, V.A., Dabagova, A.K., and Sergienko, N.V., Characteristics of Graft-Polymerization of Methyl Methacylate on Aerosil Modified with (Methacryloyloxymethyl)Methyldiethoxysilane. Journal of Applied Chemistry of the USSR, 1983. 56(11): p. 2368.

137. Ivanchev, S.S., Dmitrenko, A.V., and Krupnik, A.M., Graft-Polymerization of Styrene Initiated by Redox Decay of Hydroperoxide Fixed on Aerosil. Vysokomolekulyarnye Soedineniya Seriya A, 1988. 30(9): p. 1951.

138. Yushkova, S.M., Tager, A.A., Adamova, L.V., Zubov, V.P., and Olenin, A.V., Thermodynamics of Interaction of Components in Polymerization-Filled Systems on the Base of (Meth)Acrylates and Aerosil. Vysokomolekulyarnye Soedineniya Seriya A, 1992. 34(7): p. 146.

139. Porcar, I., Garcia, R., Soria, V., Campos, A., and Figueruelo, J.E., Porous Fractal Gels - Secondary Effects in Sec. J. Non-Cryst. Sol., 1992. 147: p. 170.

140. Sworn, G., Marrs, W.M., and Hart, R.J., Characterization of Carrageenans by High-Performance Size-Exclusion Chromatography Using a Lichrospher 1000 Diol Column. J. Chromatogr., 1987. 403: p. 307.

141. Herbreteau, B., Lafosse, M., Morinallory, L., and Dreux, M., High-Performance Liquid-Chromatography of Raw Sugars and Polyols Using Bonded Silica-Gels. Chromatographia, 1992. 33(7-8): p. 325.

142. Forlay-Frick, P., Nagy, Z.B., and Fekete, J., Comparative study of reversed-phase columns for evaluation of monolithic stationary phase. J. Liq. Chrom. Relat. Tech., 2002. 25(9): p. 1431.

143. Baeyens, W.R.G., Van der Weken, G., Smet, E., Garcia-Campana, A.M., and Remon, J.P., Comparison of morphine and hydromorphone analysis on reversed phase columns with different diameters. J. Pharmaceut. Biomed. Anal., 2003. 32(4-5): p. 913.

144. Poletti, M., Lichrospher 60 Rp Select-B - Universal Support for Reversed-Phase Chromatography with Guaranteed Reproducibility. Analysis, 1991. 19(8): p. M33.

145. Guillaume, M. and Sebille, B., Study of a polymer immobilized on silica gel as stationary phase for chiral liquid chromatography. European Polymer Journal, 1996. 32(1): p. 19.

146. Kurganov, A., Puchkova, Y., Davankov, V., and Eisenbeiss, F., Polyvinylpyrrolidone-Coated Silica Packings for Chromatography of Proteins and Peptides. J. Chromatogr. A, 1994. 663(2): p. 163.

147. Laible, R. and Hamann, K., Formation of Chemically Bound Polymer Layers on Oxide Surfaces and Their Role in Colloidal Stability. Adv. Colloid Interface Sci., 1980. 13(1-2): p. 65.

148. Trachenko, V.I., Zilberman, Y.N., Shatskaya, T.F., and Pomerantseva, E.G., Polymerization of Methyl-Methacrylate in the Presence of Active Fillers. Vysokomolekulyarnye Soedineniya Seriya A, 1986. 28(3): p. 580.

Page 243: through Surface-attached Monomers

14 References 229

149. Chaimberg, M. and Cohen, Y., Kinetic Modeling of Free-Radical Graft-Polymerization. AIChE Journal, 1994. 40(2): p. 294.

150. Chakrabarti, A. and Toral, R., Density profile of terminally anchored polymer chains: A Monte Carlo study. Macromolecules, 1990. 23(7): p. 2016.

151. Kopf, A., Baschnagel, J., Wittmer, J., and Binder, K., On the Adsorption Process in Polymer Brushes: A Monte Carlo Study. Macromolecules, 1996. 29: p. 1433.

152. Zajac, R. and Chakrabarti, A., Irreversible polymer adsorption from semidilute and moderately dense solutions. Phys. Rev. E, 1995. 52(6): p. 6536.

153. Bialk, M., Prucker, O., and Ruhe, J., Grafting of polymers to solid surfaces by using immobilized methacrylates. Colloid. Surface. A, 2002. 198: p. 543.

154. Ligoure, C. and Leibler, L., Thermodynamics and kinetics of grafting end-functionalized polymers to an interface. J. Phys. France, 1990. 51: p. 1313.

155. Goustouridis, D., Misiakos, K., Petrou, P.S., and Kakabakos, S.E., Protein patterning by micromachined silicon embossing on polymer surfaces. Appl. Phys. Lett., 2004. 85(26): p. 6418.

156. Chen, C., Imanishi, Y., and Ito, Y., Macromolecules, 1998. 31: p. 4379. 157. Chan-Park, M.B., Yan, Y., Neo, W.K., Zhou, W., Zhang, J., and Zue, C.Y., Langmuir,

2003. 56: p. 4371. 158. Katzmüller, T., Appelhans, D., and Braun, H.-G., Adv. Polym. Sci., 1999. 11: p. 555. 159. Xia, Y. and Whitesides, G.M., Angew. Chem. Int. Ed., 1998. 37: p. 550. 160. deGans, B.J., Duinefeld, P.C., and Schubert, U.S., Adv. Mater., 2004. 16: p. 203. 161. Tekin, E., deGans, B.J., and Schubert, U.S., J. Mater. Chem., 2004. 17: p. 2627. 162. Morra, M. and Cassinelli, C., Cell Adhesion Micropatterning by Plasma Treatment of

Alginate coated Surfaces. Plasma. Polymer., 2002. 7(2): p. 3971. 163. Stamou, D., Musil, C., Ulrich, W.-P., Leufgen, K., Padeste, C., David, C., Gobrecht, J.,

Duschl, C., and Vogel, H., Langmuir, 2004. 20: p. 3495. 164. Gupta, B. and Anjum, N., Adv. Polym. Sci., 2003. 162: p. 35. 165. Kim, S.N., Shim, S.C., Kim, D.Y., Rhee, Y.H., and Kim, Y.B., Photochemical crosslinking

and enzymatic degradation of poly(3- hydroxyalkanoate)s for micropatterning in photolithography. Macromol. Rapid Commun., 2001. 22(13): p. 1066.

166. Prucker, O., Schimmel, M., Tovar, G., Knoll, W., and Rühe, J., Microstructuring of molecularly thin polymer layers by photolithography. Adv. Mater., 1998. 10: p. 1073.

167. Prucker, O., Habicht, J., Park, I.-J., and Rühe, J., Photolithographic structuring of surface-attached polymer monolayers. Mater. Sci. Eng. C, 1999. 8-9: p. 291.

168. Konradi, R., Weak Polyacid Brushes: Synthesis, Swelling Behavior, Complex Formation and Micropatterning, in Institute for Microsystemtechnology. 2005, Albert-Ludwigs-University: Freiburg.

169. Revzin, A. and Russell, R.J., Fabrication of Poly(Ethylene Glycol) Hydrogel Microstructures using Photolithography. Langmuir, 2001. 17(18): p. 5440.

170. Elias, H.-G., An Introduction to Polymer Science. 1st ed. 1997, Weinheim: Wiley-VCH. 492.

171. Odian, G., Principles of Polymerization. 4th ed. 2004, Hoboken, New Jersey: John Wiley & Sons. 832.

172. Tieke, B., Makromolekulare Chemie. 2nd ed. 2005, Weinheim: Wiley-VCH. 368.

Page 244: through Surface-attached Monomers

230 14 References

173. Browne, T., Chaimberg, M., and Cohen, Y., Graft Polymerization of Vinyl Acetate onto Silica. J. Appl. Polym. Sci., 1992. 44: p. 671.

174. Brunauer, S., Emmet, P.H., and Teller, E., Adsorption of Gases in Multimolecular Layers. J. Am. Chem. Soc., 1938. 60: p. 309.

175. Legrange, J.D. and Markham, J.L., Effects of Surface Hydration on the Deposition of Silane Monolayers on Silica. Langmuir, 1993. 9(7): p. 1749.

176. Sindorf, D.W. and Maciel, G.E., Solid-State NMR-Studies of the Reactions of Silica Surfaces with Polyfunctional Chloromethylsilanes and Ethoxymethylsilanes. J. Am. Chem. Soc., 1983. 105(12): p. 3767.

177. Angst, D.L. and Simmons, G.W., Moisture Absorption Characteristics of Organosiloxane Self-Assembled Monolayers. Langmuir, 1991. 7(10): p. 2236.

178. Trommsdorf, E., Über Polymerisation. Angew. Chem., 1954. 66(17): p. 567. 179. Mayo, F.R., Chain Transfer in the Polymerization of Styrene: The Reaction of Solvents with

Free Radicals. J. Am. Chem. Soc., 1943. 65: p. 2324. 180. Barson, C.A., in Comprehensive Polymer Science, G. Allen and J.C. Bevington, Editors.

1989, Pergamon Press: Oxford. p. Chapter 13. 181. Brandrup, J., Immergut, E.H., and Grulke, E.A., Polymer Handbook. 4th Edition ed.

2003: Wiley & Sons. 182. Matheson, M.S., Auer, E.E., Bevilacqua, E.B., and Hart, E.J., J. Am. Chem. Soc., 1951.

73: p. 5395. 183. Misra, G.S. and Chadha, R.N., Makromol. Chem., 1957. 50: p. 72. 184. North, A.M. and Scallan, A.M., Polymer, 1964. 5: p. 447. 185. Murata, H. and Rühe, J., Polymer Preprints, 2003. 44: p. 558. 186. Murata, H. and Rühe, J., Polymer Preprints, 2003. 44: p. 560. 187. Bergmann, E.D., Ginsburg, D., and Pappo, R., The Michael Reaction. Org. React., 1959.

: p. 179. 10188. Michalek, F., Mädge, D., Rühe, J., and Bannwarth, W., Ring-closure metathesis in

supercritical carbon dioxide as sole solvent with use of covalently immobilized ruthenium catalysts. Eur. J. Org. Chem., 2006: p. 577.

189. Glatz, H., Darstellung neuartiger Festphasenmaterialien für die kombinatorische Synthese, in Faculty for Chemistry & Pharmacy. 2003, Albert-Ludwigs-University: Freiburg.

190. Lack, O., Zbinden, H., and Woggon, W.-D., A Useful Disulfide Linker for Single-Bead Analysis of Peptide Libraries. Helv. Chim. Acta, 2002. 85(2): p. 495.

191. Kaiser, E., Colescott, R.L., Bassinger, C.D., and Cook, P.J., Color test for detection of free terminal amino groups in the solid-phase synthesis of peptides. Anal. Biochem., 1970. 34(2): p. 595.

192. Fazekas de St Groth, B., Smith, A.L., Koh, W.P., Girgis, L., Cook, M.C., and Bertolino, P., Carboxyfluorescein diacetate succinimidyl ester and the virgin lymphocyte: a marriage made in heaven. Immunol Cell Biol, 1999. 77(6): p. 530.

193. Miller, B.T., Collins, T.J., Rogers, M.E., and Kurosky, A., Peptide biotinylation with amine-reactive esters: differential side chain reactivity. Peptides, 1997. 18(10): p. 1585.

194. Graf, R. and Friedl, P., Detection of immobilized proteins on nitrocellulose membranes using a biotinylation-dependent system. Anal. Biochem., 1999. 273(2): p. 291.

Page 245: through Surface-attached Monomers

14 References 231

195. Lecolley, F., Tao, L., Mantovani, G., Durkin, I., Lautru, S., and Haddleton, D.M., A new approach to bioconjugates for proteins and peptides (ʺpegylationʺ) utilising living radical polymerisation. Chem. Commun. (Camb), 2004(18): p. 2026.

196. Yan, M., Cai, S.X., Wybourne, M.N., and Keana, J.F., N-hydroxysuccinimide ester functionalized perfluorophenyl azides as novel photoactive heterobifunctional cross-linking reagents. The covalent immobilization of biomolecules to polymer surfaces. Bioconjug. Chem., 1994. 5(2): p. 151.

197. Yin, H.B., Brown, T., Wilkinson, J.S., Eason, R.W., and Melvin, T., Submicron patterning of DNA oligonucleotides on silicon. Nucleic Acids Res., 2004. 32(14): p. e118.

198. Fernandez-Garcia, M., Fernandez-Sanchez, M., Madruga, E.L., and Fernandez-Monreal, C., Macromol. Chem. Phys., 1999. 200: p. 199.

199. Dube, M.A., Penlidis, A., and OʹDriscoll, K.F., Can. J. Chem. Eng., 1990. 68: p. 974. 200. Maxwell, I.A., Aerdts, A.M., and German, A.L., Macromolecules, 1993. 26: p. 1956. 201. Coote, M.L., Johnston, L.P.M., and Davis, T.P., Macromolecules, 1997. 30: p. 8191. 202. Berchtold, B., Oberflächengebundene Polymernetzwerke für die Re-Endothelialisierung von

porcinen Herzklappenbioprothesen, in Institute for Microsystemtechnology. 2005, Albert-Ludwigs-University: Freiburg im Breisgau.

203. Szutowicz, A., Anal. Biochem., 1984. 138: p. 86. 204. Al-Kaissi, E. and Mostratos, A., J. Immunol. Meth., 1983. 58: p. 127. 205. Engvall, E., Meth. Enzymol., 1980. 70: p. 419. 206. Faizyme, Assay procedure for Glucose Oxidase. 2003, Faizyme Laboratories Ltd. 207. Keesey, J., in Biochemica Information. 1987, Boehringer Mannheim Biochemicals:

Indianapolis (IN, USA). p. 58. 208. Pütter, J. and Becker, R., in Methods of Enzymatic Analysis, H.U. Bergmeyer, Editor.

1983, Verlag Chemie: Deerfield Beach (FL, USA). p. 286. 209. Sigma-Aldrich, Enzymatic Assay of PEROXIDASE (EC 1.11.1.7) - 2,2ʹ-Azino-bis(3-

Ethylbenzthiazoline-6-Sulfonic Acid) as a Substrate. 1996. 210. Moore, S. and Stein, W.H., Chromatography of amino acids on sulfonated polystyrene

resins. J. Biol. Chem., 1951. 192: p. 663. 211. Hecht, H.J., Kalisz, K., Hendle, J., Schmid, R.D., and Schomburg, D., Crystal structure

of glucose oxidase from aspergillus niger: Refined at 2.3 Angstroms resolution. J. Mol. Biol., 1993. 229: p. 153.

212. Kurstak, E., Progress in enzyme immunoassays: production of reagents, experimental design and interpretation. Bull. World Health Organ., 1985. 63(4): p. 793.

213. Wisdom, G.B., Enzyme and biotin labeling of antibody. Methods Mol. Biol., 1994. 32: p. 433.440.

214. Shiuan, D., Wu, C.H., Chang, Y.S., and Chang, R.J., Competitive enzyme-linked immunosorbent assay for biotin. Methods Enzymol., 1997. 279: p. 321.

215. Guesdon, J.L., Immunoenzymatic techniques applied to the specific detection of nucleic acids. J. Immunol. Meth., 1992. 150: p. 33.

216. Dunn, M.J., Detection of proteins on blots using the avidin-biotin system. Methods Mol. Biol., 1994. 32: p. 227.

217. Yoshida, A. and Takagaki, Y., Detection of food additives by enzyme immunoassay. Nippon Rinsho, 1995. 53(9): p. 2316.

Page 246: through Surface-attached Monomers

232 14 References

218. Szakacs, J.G. and Livingston, S.K., mRNA in-situ hybridization using biotinylated oligonucleotide probes. Ann. Clin. Lab. Sci., 1994. 24(4): p. 324.

219. Haugland, R.P. and Zhou, M., Dye labeled polymers as reagents for measuring polymer degradation, in U.S. 1998, (Molecular Probes, Inc., USA). Us. p. 20 pp.

220. Neumann, M.G., Gessner, F., and De Sena, G.L., The adsorption of basic dyes by poly(styrene sulfonate). Dyes Pigm., 1993. 20(4): p. 255.

221. Koç, F., Michalek, F., Rumi, L., Bannwarth, W., and Haag, R., Catalysts on Functionalized Polymer Chips as Recyclable Entities. Synthesis, 2005: p. 3362.

222. Kingsbury, J.S. and Hoveyda, A.H., Regarding the Mechanism of Olefin Metathesis with Sol-Gel-Supported Ru-Based Complexes Bearing a Bidentate Carbene Ligand. Spectroscopic Evidence for Return of the Propagating Ru Carbene. J. Am. Chem. Soc., 2005. 127: p. 4510.

223. Angelescu, D.E., Waller, J.H., Adamson, D.H., Deshpande, P., Chou, S.Y., Register, R.A., and Chaikin, P.M., Macroscopic Orientation of Block Copolymer Cylinders in Single-Layer Film by Shearing. Adv. Mater., 2004. 16: p. 1736.

224. Lehmann, T. 1998, Johannes Gutenberg-University: Mainz. 225. Tsubokawa, N., Shirai, Y., and Hashimoto, K., Effect of polymerization conditions on the

molecular weight of polystyrene grafted onto silica in the radical graft-polymerization initiated by azo or peroxyester groups introduced onto the surface. Coll. Polym. Sci., 1995. 273(11): p. 1049.

226. Chang, B.-J., Control of cell adhesion on heart valve implants through ultrathin surface-attached polymer layers, in IMTEK. 2002, Albert-Ludwigs-University: Freiburg. p. 160.

227. Johnson, T., McGowen, J., Beuhler, A., Brush, C.K., and Lajos, R.E., Methods and compositions for attachment of biomolecules to solid supports, hydrogels, and hydrogel arrays, in PCT Int. Appl. 2001, (Motorola Inc., USA). Wo. p. 46 pp.

228. Moser, I., Jobst, G., and Gamp, T., Method for production of a lab on a chip from photoresist material for medical diagnostic applications, in PCT Int. Appl. 2003, (Universitaet Freiburg, Germany). WO. p. 24 pp.

229. Asseline, U. and Nguyen, T.T., Tetrahedron Lett., 1989. 30: p. 2521. 230. Gait, M.J., Matthes, H.W.D., Singh, M., Sproat, B.S., and Titmas, R.C., Nucleic Acids

Res., 1982. 10: p. 6243.

Page 247: through Surface-attached Monomers

Acknowledgements

In first place I like to thank my supervisor, Prof. Dr. Jürgen Rühe, for the chance to work

on an interesting research topic that evolved many different facettes during my

investigations. The helpful discussions with him and his support were the foundation of this

thesis. Further I want to thank him for the excellent working conditions regarding

infrastructure as well as the great team spirit within the whole working group.

PD Dr.-Ing. Thomas Hanemann for kindly act as co-referee for this thesis.

Dr. Oswald “Ossi” Prucker for his introduction to the “grafting-in-between” topic and his

continuous support on all questions I had, including long or latenight discussions.

Of great importance were my cooperation partners at the department for organic

chemistry: I want to thank Prof. Dr. Willi Bannwarth for many fruitful discussions in a

pleasant atmosphere, how to drive my research results to application. This especially

concerns Heiko Glatz, who synthesized some of the special monomers and started using

“G.I.B.” material for solid phase supports. Thanks to Florian Michalek the work on catalyst

immobilization was continued and evolved to a real success story. Last but not least

cooperation member Tobias Cramer did fluorous phase stuff while grafting somewhere in

between.

Bernd “Barney the butcher” Berchtold, Daniel “das kommt auf den Hut” Freidank,

Rupert “there can be only one Rupilander” Konradi, Ulrike “sÜrface design” Mock,

Jörg “phosphonate anchor-man” Pahnke for helping me in many different issues of my work

and of course for their friendship exceeding working time.

Sidar Loschonsky and Marc Loschonsky for their great support during “the final

countdown” to the thesis submission.

Bong Jun Chang, Doungporn Sirikittikul, Dr. “JP” J. D. Jeyaprakash S. Samuel, Kamlesh

Shroff and Haining Zhang for great international colleagueship spirit and fantastic culinary

cooking events.

In the end I would like to thank all the people who spent work and time together with me,

supported, encouraged and enlightened me and therefore share an essential part of this

thesis with me.

Page 248: through Surface-attached Monomers

Natalia Schatz for her great work on “G.I.B.” in the lab and spending many hours with the

self-developped shaking-device; “Shaken not stirred!”

Dorothea Freidank for the good experiments with enzymes and their UV measurements.

Daniela Mössner and Martin Schönstein for the XPS measurements on porous and flat

substrates.

Kirstin Seidel for her help on enzyme assay development and Dr. Holger Klapproth for his

valuable hints on DNA preparations.

Dr. J. D. Jeyaprakash S. Samuel for his introduction and help on ATRP experiments.

Dr. Jörn O. Saß for amino acid analysis of immobilized enzymes on “G.I.B.” supports and

Dipl.-Ing. (FH) Michael Reichel for introducing profilometry to me using my micropatterned

samples.

Waltraud Hanser and Petra Hettich for their help “at the office desk” handling all that

bureaucratic stuff.

Gerhard Baaken, Dr. Markus Biesalski, Petra Böhringer, Julien Couet, Franziska

Degenhardt, Alexey Kopyshev, Meike Moschallski, Thorsten Neumann, Dr. Svetlana Santer,

Christian Schlemmer, Falko Stenzel, Julia Viertel, Anke Wörz and Hyun-Kwan Yang for their

good colleagueship and conversations at the coffeebreak.

Life at CPI was fun, I had a good time I’ll always have in good remembrance and so I

wanna thank again those I maybe forgot to list here and the whole funky bunch at CPI. Keep

up the fine spirit!

Finally I am deeply grateful for the continual support I got from my friend Karen, from my

parents and my grandparents, who always believed that “I’ll make it”, throughout my

studies and PhD work.