Vous êtes sur la page 1sur 74

UNDERSTANDING THE MECHANISMS BEHIND ATOM TRANSFER RADICAL POLYMERIZATION EXPLORING THE LIMIT OF CONTROL

Helena Bergenudd

AKADEMISK AVHANDLING

som med tillstnd av Kungliga Tekniska Hgskolan i Stockholm framlgges till offentlig granskning fr avlggande av teknisk doktorsexamen fredagen den 29 april 2011, kl 10.00 i sal K2, Teknikringen 28, KTH, Stockholm. Avhandlingen frsvaras p engelska. Opponent: Professor David M. Haddleton frn University of Warwick, England.

Copyright 2011 Helena Bergenudd All rights reserved

Paper I 2009 Elsevier Ltd. Paper II 2009 American Chemical Society

TRITA-CHE Report 2011:21 ISSN 1654-1081 ISBN 978-91-7415-933-2

ABSTRACT
Atom transfer radical polymerization (ATRP) is one of the most commonly employed techniques for controlled radical polymerization. ATRP has great potential for the development of new materials due to the ability to control molecular weight and polymer architecture. To fully utilize the potential of ATRP as polymerization technique, the mechanism and the dynamics of the ATRP equilibrium must be well understood. In this thesis, various aspects of the ATRP process are explored through both laboratory experiments and computer modeling. Solvent effects, the limit of control and the use of iron as the mediator have been investigated. It was shown for copper mediated ATRP that the redox properties of the mediator and the polymerization properties were significantly affected by the solvent. As expected, the apparent rate constant (kpapp) increased with increasing activity of the mediator, but an upper limit was reached, where after kpapp was practically independent of the mediator potential. The degree of control deteriorated as the limit was approached. In the simulations, which were based on the thermodynamic properties of the ATRP equilibrium, the same trend of increasing kpapp with increasing mediator activity was seen and a maximum was also reached. The simulation results could be used to describe the limit of control. The maximum equilibrium constant for controlled ATRP was correlated to the propagation rate constant, which enables the design of controlled ATRP systems. Using iron compounds instead of copper compounds as mediators in ATRP is attractive from environmental aspects. Two systems with iron were investigated. Firstly, iron/EDTA was investigated as mediator as its redox properties are within a suitable range for controlled ATRP. The polymerization of styrene was heterogeneous, where the rate limiting step is the adsorption of the dormant species to the mediator surface. The polymerizations were not controlled and it is possible that they had some cationic character. In the second iron system, the intention was to investigate how different ligands affect the properties of an ATRP system with iron. Due to competitive coordination of the solvent, DMF, the redox and polymerization properties were not significantly affected by the ligands. The differences between normal and reverse ATRP of MMA, such as the degree of control, were the result of different FeIII speciation in the two systems.

SAMMANFATTNING
Atom transfer radical polymerization (ATRP) r en av de mest anvnda teknikerna fr kontrollerad radikalpolymerisation. ATRP har stor potential fr utvecklingen av nya material tack vare mjligheten att kontrollera molekylvikt och polymerarkitektur. Fr att kunna utnyttja den fulla potentialen hos ATRP som polymerisationsteknik r det av vikt att frst mekanismen och dynamiken i ATRP-jmvikten. I denna avhandling utforskas olika aspekter av ATRP-processen genom bde laboratoriefrsk och datormodellering. Lsningsmedelseffekter, grnsen fr kontroll och jrn som mediator har underskts. Fr kopparmedierad ATRP visades ptagliga lsningsmedelseffekter p mediatorns redoxegenskaper och p polymerisationsegenskaperna. Som frvntat kade den operationella hastighetskonstanten (kpapp) med kande mediatoraktivitet, men en vre grns nddes, varefter kpapp var praktiskt taget oberoende av mediatorpotentialen. Graden av kontroll frsmrades d grnsen nrmades. I simuleringarna, som baserades p de termodynamiska egenskaperna hos ATRP-jmvikten, sgs samma trend med kande kpapp med kande mediatoraktivitet och ett maximum nddes ven hr. Simuleringsresultaten kunde ocks anvndas fr att beskriva grnsen fr kontroll. Den maximala jmviktskonstanten fr kontrollerad ATRP korrelerades till propageringshastighetskonstanten, vilket mjliggr design av kontrollerade ATRP-system. Att anvnda jrnfreningar istllet fr kopparfreningar som ATRPmediatorer r attraktivt ur miljhnsyn. Tv system med jrn utforskades. JrnEDTA som mediator undersktes, eftersom dess redoxegenskaper faller inom ett lmpligt intervall fr kontrollerad ATRP. Polymerisationen av styren var heterogen och det hastighetsbestmmande steget var adsorptionen av de passiva specierna p ytan av mediatorn. Polymerisationerna var inte kontrollerade och det r mjligt att de hade viss katjonkaraktr. I det andra jrnsystemet var tanken att underska hur olika ligander pverkar egenskaperna hos ett ATRP-system med jrn. Varken redoxegenskaperna eller polymerisationsegenskaperna pverkades nmnvrt av liganderna p grund av konkurrerande koordination frn lsningsmedlet (DMF). Skillnaderna mellan normal och omvnd ATRP, t.ex. graden av kontroll, berodde p skillnader i FeIII speciering i de tv systemen.

LIST OF PAPERS
The thesis is a summary of the following papers:

Heterogeneous iron(II)-chloride mediated radical polymerization of styrene H. Bergenudd, M. Jonsson, D. Nystrm, E. Malmstrm, Journal of Molecular Catalysis A: Chemical, 2009, 306, 69-76

II

Solvent Effects on ATRP of Oligo(ethylene glycol) Methacrylate. Exploring the Limits of Control H. Bergenudd, G. Coullerez, M. Jonsson, E. Malmstrm, Macromolecules, 2009, 42, 3302-3308

III

Investigation of Iron Complexes in ATRP. Indications of Different Iron Species in Normal and Reverse ATRP H. Bergenudd, M. Jonsson, E. Malmstrm, Journal of Molecular Catalysis A: Chemical, submitted

IV

Investigation of the ATRP Process Through Simulations Predicting the Limit of Control H. Bergenudd, M. Jonsson, E. Malmstrm, Macromolecular Theory and Simulations, submitted

The contribution of the author of this thesis to the appended papers is: I, III II IV All the experimental work and the preparation of the manuscripts. Some of the experimental work, all of the evaluation of the results and the preparation of the manuscript. All the simulations and the preparation of the manuscript.

ABBREVIATIONS
AGET ATRP AIBN ATRP bipy CRP DMF DMSO DP E1/2 ECu ERX EBiB EDTA ESR / EPR EtOAc FRP GPC HMTETA IDA IPA KATRP (lim) ka kd Kx kp kpapp kt M Mn,th Mn MBriB Me6-TREN Me4-cyclam MeCN activators generated by electron transfer ATRP azobisisobutyronitrile atom transfer radical polymerization 2,2-bipyridine controlled radical polymerization N,N-dimethylformamide dimethylsulfoxide degree of polymerization half-wave potential copper complex reduction potential alkyl halide reduction potential ethyl 2-bromoisobutyrate ethylenediaminetetraacetic acid electron spin/paramagnetic resonance ethyl acetate free radical polymerization gel permeation chromatography (analogous to SEC) 1,1,4,7,10,10-hexamethyltriethylenetetraamine iminodiacetic acid isophthalic acid the ATRP equilibrium constant (limiting ditto) activation rate constant deactivation rate constant halidophilicity propagation rate constant apparent rate constant termination rate constant monomer (unless it is as a unit) theoretical molecular weight number average molecular weight methyl 2-bromoisobutyrate tris[2-(dimethylamino)-ethyl]amine 1,4,8,11-tetraaza-1,4,8,11-tetramethylcyclotetradecan acetonitrile

MeOH MMA-Br MMA NIPAAM NMR OEGBr OEGMA PA PDA PDI PEBr PMDETA PPh3 2-PrOH RX R SCE SEC TEMPO

methanol see MBriB methyl methacrylate N-isopropylacrylamide nuclear magnetic resonance oligo(ethylene glycol) 2-bromoisobutyrate oligo(ethylene glycol) methacrylate 2-picolinic acid dipicolinic acid polydispersity index 1-phenylethylbromide N,N,N,N,N-pentamethyldiethylenetriamine triphenylphosphine isopropanol alkyl halide (or dormant polymer chain) alkyl (or polymeric) radical saturated calomel electrode size exclusion chromatography (analogous to GPC) 2,2,6,6-tetramethylpiperidine 1-oxyl

TABLE OF CONTENTS
1 2 PURPOSE OF THE STUDY .................................................................................... 1 INTRODUCTION ..................................................................................................... 2 2.1 CONTROLLED RADICAL POLYMERIZATION ............................................................ 2 2.2 ATOM TRANSFER RADICAL POLYMERIZATION (ATRP) ......................................... 3 2.2.1 The ATRP mechanism .................................................................................. 3 2.2.2 Free radicals? .............................................................................................. 6 2.2.3 The ATRP equilibrium ................................................................................. 7
2.2.3.1 2.2.3.2 2.2.3.3 2.2.3.4 2.2.3.5 Activation and deactivation rate constants and the equilibrium constant ........... 10 The reduction potential of the mediator ............................................................. 11 Correlations between ECu and KATRP or kpapp ...................................................... 12 The halidophilicity, Kx ....................................................................................... 12 The reduction potential of the initiator ............................................................... 12

2.2.4 2.2.5 3

Simulations of ATRP kinetics ..................................................................... 14 ATRP techniques and applications ............................................................ 14

EXPERIMENTAL ................................................................................................... 17 3.1 MATERIALS ........................................................................................................ 17 3.2 CHARACTERIZATION ........................................................................................... 17 3.3 CYCLIC VOLTAMMETRY...................................................................................... 18 3.4 POLYMERIZATIONS ............................................................................................. 19 3.4.1 Solvent effects ............................................................................................ 19 3.4.2 Heterogeneous polymerization with Fe/EDTA .......................................... 19 3.4.3 Iron complexes in DMF ............................................................................. 19 3.5 SIMULATIONS ..................................................................................................... 20

RESULTS AND DISCUSSION .............................................................................. 21 4.1 SOLVENT EFFECTS ON ATRP .............................................................................. 21 4.1.1 Reduction potentials of mediators.............................................................. 22 4.1.2 Polymerization kinetics .............................................................................. 23 4.1.3 Solvent effects ............................................................................................ 23 4.1.4 Limit of control .......................................................................................... 25 4.2 SIMULATIONS ..................................................................................................... 26 4.2.1 Predicting the limit of control .................................................................... 30 4.3 ATRP SYSTEMS WITH IRON MEDIATORS ............................................................. 33 4.3.1 Styrene and iron chloride........................................................................... 34

4.3.1.1

Discussion .......................................................................................................... 38

4.3.2
4.3.2.1 4.3.2.2 4.3.2.3 4.3.2.4

ATRP with iron mediators in DMF ............................................................ 39


Reduction potentials of mediators ...................................................................... 40 Polymerizations ................................................................................................. 42 UV-vis spectroscopy .......................................................................................... 45 Discussion .......................................................................................................... 46

5 6 7 8

CONCLUSIONS ...................................................................................................... 49 FUTURE WORK ..................................................................................................... 51 ACKNOWLEDGEMENTS .................................................................................... 53 REFERENCES ........................................................................................................ 54

1 Purpose of the study

1 PURPOSE OF THE STUDY


One of the most commonly used techniques for controlled radical polymerization is atom transfer radical polymerization (ATRP). An advantage of ATRP over free radical polymerization is the ability to control the molecular weight and polymer architecture. Many aspects of ATRP in general, and the mechanism in particular, have been investigated previously. The ATRP equilibrium constant, as well as the rate constant for the activation, has been reported for different mediators and initiators. It has also been shown that the reduction potential of the mediator is one parameter that affects the kinetics of ATRP and that solvent effects on the redox properties are significant. Although these results are very interesting and can be used to compare different mediators or initiators with each other, it can be difficult to connect the magnitude of e.g. the equilibrium constant to polymerization behavior. That is, within which range of equilibrium constants can well controlled ATRP be expected? This is an important question for the development of ATRP, in particular in the search for new ATRP systems, e.g. new mediators. The purpose of this study was to increase the understanding of the ATRP mechanism and to connect the thermodynamic properties of the ATRP system with the polymerization properties, but also to explore new ATRP systems. The goal was to be able to predict the optimal conditions for well controlled ATRP, i.e. for which combination of monomer, mediator and initiator well defined polymers can be attained. To do this, the effects of solvent and ligand on the redox properties of the mediators and on the polymerization properties (i.e. the kinetics and polymer properties) had to be investigated further. Kinetic simulations can also be a useful tool to explore polymerization processes such as ATRP. Thus, based on the thermodynamics of the ATRP equilibrium, i.e. the equilibrium constant, an investigation of the ATRP kinetics was also initiated, with the aim to describe the limit of control. Another goal was to search for new ATRP mediators, particularly aiming towards environmentally friendly compounds. In this study, iron mediators were considered as an attractive alternative to the more commonly used copper based mediators.

2.1 Controlled radical polymerization

2 INTRODUCTION
2.1 CONTROLLED RADICAL POLYMERIZATION

Free radical polymerization (FRP) is the most commonly employed polymerization technique, standing for approximately 50 % of the total polymer production. It can be used for a wide range of monomers, including olefins (e.g. ethylene and propylene) and vinyl monomers (e.g. vinylidene chloride, styrene and methyl methacrylate), and is less sensitive to reactant impurities than e.g. anionic polymerization. The FRP mechanism consists of three steps initiation, propagation and termination (combination and disproportionation), equations 1-5.

Initiator dissociation Initiation Propagation Termination (comb.) Termination (disprop.)

kd

2R
ki
kp

(1)

R
Mi

M
M

M1
Mi 1
k tc
ktd

(2) (3)
m

Mn
Mn

Mm
Mm

Mn
Mn

(4)
Mm

(5)

In contrast to anionic polymerization, pure block copolymers and other advanced polymer architectures cannot be produced by free radical polymerization due to the inherent reactivity of the radical, resulting in unavoidable termination and transfer reactions. Also, as the initiation is a continuous process and terminations are always present, the molecular weight distribution will be wide and it is not possible to control the molecular weight. However, during the last two decades, several new techniques have emerged that allow for controlled radical polymerization (CRP). The most studied techniques are nitroxide mediated polymerization (NMP),[1] reversible addition-fragmentation chain transfer (RAFT) polymerization[2,3] and atom transfer radical polymerization (ATRP, vide infra). The

2 Introduction

idea behind all these techniques is to suppress radical termination and transfer reactions by keeping a very low concentration of radicals throughout the polymerization. An equilibrium is established between dormant, end-capped chains, unable to propagate or terminate, and the active species, Scheme 1. The equilibrium should be shifted towards the dormant species and the exchange between active and dormant species should be fast relative to propagation. Scheme 1. General CRP mechanism.

2.2 2.2.1

ATOM TRANSFER RADICAL POLYMERIZATION (ATRP) The ATRP mechanism

Atom transfer radical polymerization was developed in 1995 independently by Wang and Matyjaszewski[4,5] and Kato et al.[6] as an expansion of transition metal catalyzed atom transfer radical addition (ATRA). The origin is the reaction generally known as the Kharasch addition reaction, where a polyhalogenated alkane (e.g. CCl4) adds to an alkene, forming a 1:1 addition product, Scheme 2. [7] Scheme 2. The Kharasch reaction, addition of a polyhalogenated alkane to an alkene.

Radicals, produced by e.g. light, add to the double bond of the alkene, where after the halogen is abstracted from the organic halide. This reaction can also be catalyzed by transition metal complexes by a redox mechanism, which is what is used in ATRA.[8,9] Here, the radical generation and halogen transfer is catalyzed by

2.2 Atom transfer radical polymerization (ATRP)

the metal complex (Scheme 3), which makes the halogen transfer much more efficient. In practice, the halogen transfer is reversible and the stability of the intermediate alkyl radical and the ease of halogen transfer from the metal complex depicts whether selective 1:1 addition occurs or if the products are telomers/polymers. Scheme 3. The ATRA mechanism.

Polymerization, i.e. the repetitive addition of alkenes to the radical species, can occur if the reactivity of the radical species is comparable before and after the initial addition to the double bond of the alkene (step 2 in Scheme 3) and the radical species is sufficiently stabilized (e.g. as for styrene). This is the basis of what has been designated atom transfer radical polymerization, ATRP. The suggested general mechanism for ATRP is shown in Scheme 4. Scheme 4. General mechanism for ATRP.

In ATRP, a dynamic equilibrium is established between an alkyl halide (or halogen end-capped polymer chain, RX) and the corresponding radical (R) by means of a transition metal complex (Mt nY/ligand X-Mtn+1Y/ligand). As mentioned above, the equilibrium should be shifted towards the dormant species in order to keep the radical concentration to a minimum, thereby suppressing radical termination reactions and enabling control over molecular weight and

2 Introduction

polymer architecture. However, terminations cannot be completely avoided in ATRP and terminations in the beginning of the polymerization will lead to a build-up of the deactivator (i.e. X-Mtn+1Y/ligand) concentration. Consequently, the equilibrium will be shifted towards the dormant species and the radical concentration will be lowered, the polymerization is thus in a sense self regulating. This is also known as the persistent radical effect (PRE). [10] By the utilization of electron paramagnetic resonance (EPR) the concentration of deactivator during ATRP has been measured.[11-13] The results show that the deactivator concentration reaches 5-10 % of the initial activator concentration during the polymerization. The control over the ATRP reaction can be improved by the addition of small amounts of the deactivator, which thus shifts the equilibrium towards the dormant species. Commonly, 10 % deactivator relative to the activator concentration is added, which thus corresponds to the amount of deactivator formed during the polymerization due to irreversible terminations as seen above.

At the end of a polymerization in truly living polymerization systems (e.g. in anionic polymerization), the chain ends should be living, i.e. they should not be irreversibly terminated, but be able to be used for as macroinitiators for further polymerizations. In ATRP, this is realized by the reversible halogen end-capping. Thus, when the polymerization has reached completion, all polymer chains should (ideally) have a halogen atom at the chain end. They can then be utilized as macroinitiators (corresponding to the initial alkyl halide initiator) for producing block copolymers (or other advanced polymer architectures).

Several different transition metals, including copper, iron, cobolt, ruthenium and nickel (usually in the form of salts with chlorine, bromine or iodine), have been used in ATRP together with various complexing ligands, including nitrogen and phosphine based structures.[14] The most commonly used transition metal is copper, largely due to its low cost and versatility. The role of the ligand is to solubilize the metal ion, which also affects the reduction potential of the transition metal ion. Alkyl bromides and chlorides are usually used as initiators, but alkyl iodides have also been used. Some examples of commonly used ligands and alkyl halides are shown in Figure 1.

2.2 Atom transfer radical polymerization (ATRP)

Figure 1. Ligands and initiators commonly used in ATRP.

2.2.2

Free radicals?

The general view is that the propagating species in ATRP are free radicals. However, the possibility that it is radical-metal complexes instead has been debated. It is known that transition metals can trap radicals and that organometallic species can mediate radical polymerization (organometallic radical polymerization, OMRP).[15] Cobolt, which has been shown to work as an ATRP mediator as well, can be used to control radical polymerization through direct capping of the radical in OMRP.[16] Asscher and Vofsi suggested a free radical mechanism for the addition of carbon tetrachloride and chloroform to olefins mediated by iron or copper chlorides (i.e. ATRA) on the basis of, among other things, the formation of identical mixtures of diastereomers from cis- and trans-2-butene.[17,18] On the other hand, pulse-radiolytic studies have shown that short lived copper-radical

2 Introduction

complexes (CuIII-R) exist in aqueous solution[19-21] and organometallic species have been suggested for iron complexes as well.[22] The radical nature of ATRP has been investigated by e.g. measuring the reactivity ratios in copolymerizations, kinetic isotope effects, trapping reactions, the addition of known radical inhibitors, ESR measurements of trapped radicals in glassy matrices, racemization and halogen exchange reactions. [23-33] Indications of metal-radical intermediates or caged radicals were found by measuring kinetic isotope effects and reactivity ratios in both FRP and ATRP, which showed differences between the two polymerization methods. [30,33] However, the same two techniques have been used to find evidence of the opposite, i.e. the same free radical intermediates in both FRP and ATRP.[28,32] Radical trapping and chain transfer agents, such as TEMPO and thiols, effectively react with the radicals in ATRP in the same way as in FRP.[25,26] Also, free radicals were detected in highly crosslinked polymerization matrices by electron spin resonance (ESR). [27,31] These observations verify the suggested ATRP mechanism with radicals that are not associated with the transition metal of the mediator. The majority of the experimental data indicate that free radicals are indeed the reactive intermediates in ATRP, as in FRP.

2.2.3

The ATRP equilibrium

The ATRP equilibrium (A) can be divided into three parts, as shown in Scheme 5 for copper mediated ATRP. Scheme 5. The ATRP equilibrium for copper mediated ATRP.

R-X + CuIX/L R + X-CuIIX/L CuIX/L CuIIX/L+ + e R-X + e R + X X + CuIIX/L+ X-CuIIX/L

(A) (B) (C) (D)

2.2 Atom transfer radical polymerization (ATRP)

Equilibrium B is the oxidation of the CuI complex and the equilibrium constant is given by equation 6, where ECu is the reduction potential of the CuII complex.

K Cu = exp -

F ECu RT

(6)

Equilibrium C is the dissociative reduction of the alkyl halide (dormant polymer chain), for which the equilibrium constant is given by equation 7, where ERX is the reduction potential of the alkyl halide.

K RX = exp

F ERX RT

(7)

Equilibrium D is the formation of the X-CuIIX/L complex, also termed halidophilicity,[34] which equilibrium constant is given by equation 8 (where CuIIX/L+ and X-CuIIX/L are simplified to CuII and CuIIX, respectively).

Kx =

[CuII X] [CuII ][X- ]

(8)

The overall equilibrium constant is then given by

K ATRP =

ka kd

K Cu K RX K x

exp

F ( ERX RT

ECu ) K x

[CuII X] [R ] [CuI ] [RX]

(9)

where ka and kd are the activation and deactivation rate constants, respectively. Equation 9 can also be rearranged to give the radical concentration, [R], as a

2 Introduction

function of KATRP. A lower (more negative) reduction potential of the mediator and/or a higher (more positive) reduction potential of the alkyl halide results in an equilibrium shifted towards the active species. It should be noticed that the ATRP equilibrium can be divided into other partial equilibria than those shown in Scheme 5, for example bond homolysis of the alkyl halide (R-X R + X) and the mediator (CuIX/L + X X-CuIIX/L).[34,35] The overall result is of course the same, but depending on which properties of the components in the ATRP equilibrium that are known, different ways to divide the equilibrium can facilitate the investigation of different aspects of the ATRP equilibrium, e.g. solvent effects.

The rate of polymerization, Rp, is expressed in eq. 10

Rp = -

d[M] k p [M][R ] dt

(10)

where [M] is the monomer concentration, [R] is the radical concentration and kp is the propagation rate constant. Integration of this expression, assuming constant [R], gives

ln

[M]0 [M]

k p [R ] t

app kp t

(11)

where kpapp = kp [R] is the apparent rate constant. In the ideal case, i.e. ATRP without termination reactions and constant [R], a plot of ln([M]0/[M]) (i.e. ln(1/(1conversion))) versus time will give a straight line with the slope kpapp. A non-linear plot indicates a significant degree of terminations, i.e. decreasing [R]. A well controlled ATRP process will thus have apparent first order kinetics, but also molecular weights that increase linearly with conversion, following the theoretical molecular weight Mn,th = Mmon conversion [M]0/[I]0.

2.2 Atom transfer radical polymerization (ATRP)

The dynamics of the ATRP equilibrium is very important for the polymerization outcome. As mentioned above, the exchange between dormant and active species should be fast relative to propagation. The initiation should also be very fast (instantaneous). This ensures that all polymer chains grow at the same rate, leading to a narrow molecular weight distribution. A fast deactivation relative to termination is also important, minimizing the amount of irreversible terminations. One way to increase the understanding of the ATRP process is to look into the different aspects of the ATRP equilibrium. Several of the parameters that govern the equilibrium have been investigated experimentally and theoretically (vide infra), i.e. the activation and deactivation rate constants (ka and kd, respectively), the equilibrium constant (KATRP), the alkyl halide and mediator reduction potentials (ERX and ECu (for copper mediated ATRP), respectively) and the halidophilicity (Kx). Properties of the mediator and alkyl halide that will affect the magnitude of the equilibrium constant are the carbon-halide bond strength of the alkyl halide (C-Cl > C-Br), the bond strength between the transition metal and the halogen atom (e.g. Cu-Br > Cu-Cl), the radical stability (3 > 2 > 1) and steric effects on alkyl halide, radical and metal complex. The carbon-halogen bond strength and radical stability affect the alkyl halide reduction potential, i.e. a more stable radical and/or a lower bond dissociation energy (BDE) (weaker carbonhalogen bond) shifts equilibrium C in Scheme 5 towards the radical species and ERX increases. The metal-halogen bond strength will affect the halidophilicity. Steric effects on the alkyl halide and the radical will affect ERX and steric effects on the metal complex will affect ECu and Kx.

2.2.3.1

Activation and deactivation rate constants and the equilibrium constant

The activation rate constant, ka, has been experimentally determined for copper mediated ATRP systems by several groups.[36-52] The majority of the determinations has been made by isolating the activation process from the deactivation process by irreversibly trapping the radical formed in the activation step with the stable nitroxyl radical TEMPO. Excess of TEMPO ensures that the trapping is quantitative and instantaneous. The deactivation rate constant, kd, is much less accessible than ka. There are only two reports of determinations of kd, one of them via fitting of kinetic data.[39,41] However, by determining the equilibrium constant, KATRP, the deactivation can be accessed via values of ka. KATRP can be determined by measuring the concentration of deactivator in model reactions, i.e. with only mediator and alkyl halide present

10

2 Introduction

in the reaction solution (no monomer). The evaluation is based on equations originating from the persistent radical effect.[34,53-57] The majority of the determinations of ka and KATRP are made in acetonitrile at ambient temperature. The general trend in activity among copper complexes is the same for KATRP as for ka. For the alkyl halides, alkyl bromides have higher activation rate constants than alkyl chlorides, which can be attributed to the higher bond strength in alkyl chlorides, and the magnitude of KATRP follows the order 3 > 2 > 1 in general, which is also the order of radical stability. For ligands, the general order of activity for copper complexes with different ligands is tetradentate (cyclic-bridged > branched (e.g Me6-TREN) > cyclic > linear) > tridentate (e.g PMDETA) > bidentate (e.g bipy) ligands.[40,49] The activation rate constant increases with solvent polarity. The activation rate constant spans more than three orders of magnitude, from 10-3 to 3 M-1 s-1 [49,50] and KATRP varies more than seven orders of magnitude, from 10-11 to 10-4 [57] depending on the type of ligand and initiator.

2.2.3.2

The reduction potential of the mediator

The reduction potentials of mediators have been measured by cyclic voltammetry for several transition metal complexes used in ATRP, e.g. copper, iron, osmium and ruthenium.[58-69] A lower (more negative) reduction potential of a complex/mediator means that the oxidized species (e.g. CuII) are more stable (e.g. better solvated) than the reduced species (e.g. Cu I), i.e. the complex is more active (more reducing) than a complex with a higher reduction potential. For the most commonly used copper complexes, the reduction potentials in acetonitrile vary between 0.035 V vs. SCE for the less active CuBr/bipy complex and -0.30 V vs. SCE for the very active CuBr/Me 6-TREN complex. The reduction potentials are in general lower for copper chlorides than copper bromides (in acetonitrile).[58] The difference between copper chlorides and bromides can be attributed to the stronger bond between copper and bromide compared to copper and chloride. The reduction potentials of copper complexes have also been measured in other organic solvents, e.g. DMF, DMSO and MeOH, with significant solvent effects on ECu.[61]

11

2.2 Atom transfer radical polymerization (ATRP)

2.2.3.3

Correlations between ECu and KATRP or kpapp

Several reports have demonstrated that the reduction potential of the mediator is correlated to the equilibrium constant and also to the polymerization kinetics. For copper complexes, linear relationships have been shown between log(Keqapp) and E1/2 (Keqapp = KATRP/[CuII] = kpapp/(kp [CuI]0 [RX]0))[58] and between log(KATRP) and E1/2.[57] Lower (i.e. more negative) reduction potentials result in higher equilibrium constants, that is higher radical concentrations, i.e. the equilibrium is shifted towards the active species (which is in accordance with eq. 9). A linear relationship between log(kpapp) and E1/2, similar to that for log(Keqapp) vs. E1/2, was shown for copper complexes in aqueous solution (water + monomer). [59] However, the relative trend within a given set of copper complexes differ between aqueous and organic solution, demonstrating significant solvent effects on the mediator properties. Qualitative comparisons of the reduction potentials of iron and ruthenium complexes and the corresponding activity in ATRP have also been made. [62,64,65,67,69] The general trend shows that lower mediator reduction potentials result in faster polymerizations (higher kpapp).

2.2.3.4

The halidophilicity, Kx

The comparison of copper, osmium and ruthenium complexes indicated that the higher activity of the Os and Ru complexes compared to Cu complexes, despite higher (more positive) reduction potentials, originates from higher halide affinities (i.e. Kx).[68] A significantly lower halidophilicity of copper complexes in water (in the order of 10 M-1) compared to other solvents (around 10 6 M-1 in MeCN) show that halide dissociation from the deactivator contributes to lower deactivation efficiency in protic solvents.[70,71]

2.2.3.5

The reduction potential of the initiator

Reports on the reduction potential of alkyl halides (ERX) are few. Recently, ab initio calculations of ERX for various alkyl halides being structurally similar to ATRP initiators and polymer chain ends were reported. [35] For a tertiary alkyl bromide in DMF, exemplified by MMA-Br (= MBriB), the reported reduction potential was -0.80 V (vs. SCE) for adiabatic reduction (i.e. RX RX-) and -0.33 V (vs. SCE) for dissociative reduction (i.e. RX R + X-). However, in earlier reports

12

2 Introduction

of alkyl halide reduction potentials, calculated from experimentally determined thermodynamic parameters, the reduction potential for the dissociative reduction of a tertiary alkyl bromide in DMF is around -1 V (vs. SCE).[72,73] This discrepancy in ERX values has a significant impact on calculated values of KATRP (eq. 9). Note that a 100 mV difference in ERX corresponds to a factor 25 difference in KATRP. The magnitude of the three literature values of ERX can be assessed by calculating the corresponding equilibrium constants and comparing them with the experimental equilibrium constant for the same system. The system used for the comparison is CuBr/PMDETA together with a tertiary alkyl bromide in DMF. KATRP has been determined for EBiB + CuBr/PMDETA in acetonitrile (MeCN),[57] but not in DMF. However, for the similar system MBriB + CuBr/HMTETA, KATRP has been determined in both DMF and MeCN, and the ratio between the experimental values of KATRP in these two solvents is equal to six.[34] Assuming that this ratio is approximately the same for other systems in these solvents, the experimental value for EBiB + CuBr/PMDETA in MeCN (KATRP = 7.6 10-8) can be recalculated for DMF (resulting in KATRP = 4.6 10-7). Using the three literature values of ERX together with ECu = -0.2 V (vs. SCE) for CuBr/PMDETA in DMF[61] and Kx = 106 M-1,[71] the corresponding equilibrium constants can be calculated from eq. 9. The calculated values of KATRP are compared in Table 1. The results suggest that the alkyl halide reduction potential should be in the range between -1 and -0.8 V (vs. SCE), rather than around -0.3, as the equilibrium constant calculated with the latter value is far from the (recalculated) experimental value. Table 1. ATRP equilibrium constants, KATRP, calculated from eq. 9 using different values of the alkyl halide reduction potential, ERX.a ERX exp. det. KATRP b dissociative reduction adiabatic reduction exp. det. ERX (diss. red.) -0.33 -0.80 -1 KATRP 4.6 10-7 6300 7.1 10-5 3.0 10-8 log(KATRP) -6.3 3.8 -4.1 -7.5

a) ERX in V vs. SCE (from refs. [35,72,73]), ECu = -0.2 V vs. SCE (CuBr/PMDETA in DMF)[61], Kx = 106 M-1 (ref. [71]), 22 C. b) The experimentally determined KATRP for CuBr/PMDETA + EBiB in MeCN (i.e. KATRP = 7.6 10-8) was recalculated for DMF, see the text.

13

2.2 Atom transfer radical polymerization (ATRP)

2.2.4

Simulations of ATRP kinetics

Modeling of the ATRP process through numerical simulations can be a useful tool to investigate different aspects of ATRP. The influence of various parameters on the polymerization can easily be overviewed, such as the effect of addition of deactivator. Kinetic simulations have been reported by several groups. These include both Monte Carlo simulations and the use of the Predici software. In some cases, experimental values of ka and kd have been used (e.g. from the report by Ohno et al.[36]),[74-76] but there are also reports where fitting of experimental data has been performed to find values of ka and kd.[77-79] The behavior of ATRP with bifunctional initiators, surface initiated ATRP and copolymerizations in ATRP have also been explored.[74,75,79-81] In addition, the influence of the conversion dependence of the termination rate constant (kt) has been investigated from two different approaches.[82,83] The behavior of different ATRP systems (e.g. normal and reverse ATRP, vide infra) and free radical polymerization have been compared both regarding the kinetics and the concentrations of various species in the polymerizations.[84]

2.2.5

ATRP techniques and applications

ATRP can be used to polymerize a wide range of monomers, such as styrenes, (meth)acrylates and (meth)acrylamides, and can be conducted in both polar and non-polar solvents (e.g. methanol, water, diphenyl ether, anisole and toluene).[4,70,85-96] It has, however, been noted that the polymerization is generally faster and less controlled in polar solvents.[91] Two important effects of polar solvents on the mediator are disproportionation and solvolytic loss of a halide ligand. Exemplified by CuIX, disproportionation yields CuIIX2 and Cu0, i.e. loss of the activator. However, this reaction has been utilized in the relatively new polymerization technique called single electron transfer living radical polymerization (SET-LRP), which is usually conducted in polar media such as water and alcohols.[97] In e.g. water, dissociation of halide ligand from the deactivator results in a lowered effective deactivator concentration. [70]

Several modifications of the original ATRP process (Scheme 4) have emerged. Two drawbacks with conventional ATRP is that the activator is sensitive to oxygen, which necessitates thorough removal of oxygen prior to polymerization, and that equimolar amounts of initiator (alkyl halide) and mediator must be

14

2 Introduction

employed. Reverse ATRP starts from the mediator in its higher oxidation state (i.e. the deactivator) and utilizes a conventional radical initiator (e.g. AIBN).[98] The radicals produced by the initiator are reversibly trapped by the deactivator to generate the corresponding alkyl halide. An advantage of reverse ATRP is that the initial polymerization mixture is less sensitive to oxygen, which facilitates the preparation. A variation of this technique is simultaneous normal and reverse initiation (SN&RI) ATRP, where both a conventional radical initiator and an alkyl halide are added, together with the mediator in its higher oxidation state. [99] The activator, generated as in reverse ATRP by reaction with the radicals from the conventional radical initiator, can activate also the added alkyl halide. Another technique, activators generated by electron transfer (AGET) ATRP, also starts with the mediator in its higher oxidation state.[100] The polymerization starts by the reduction of the mediator by a reductant, commonly ascorbic acid or tin(II) ethylhexanoate, and the subsequent activation of the alkyl halide initiator. High mediator amounts, which is the consequence of equimolar amounts to the initiator, requires removal of the mediator after polymerization, in particular when (toxic) copper complexes are used. On laboratory scale, this procedure is not a major issue, but scaling up ATRP for industrial applications, this can present a significant challenge. The mediator amount can be reduced to some extent in conventional ATRP, but too low initial concentrations of activator will result in ceased polymerizations since all activator will be consumed due to unavoidable irreversible radical terminations. Presently, there are two techniques which reduce the required amount of mediator. The first is activators regenerated by electron transfer (ARGET) ATRP which is related to AGET ATRP. [101] The mediator is added in its higher oxidation state at low concentrations. A reductant (commonly the same type that is used in AGET ATRP, but also zero valent transition metals such as Cu0) added in excess reduces the mediator to start the reaction, but is also present to continuously regenerate the deactivator that accumulates due to irreversible terminations. The second technique, ICAR (initiators for continuous activator generation) ATRP, is related to reverse ATRP by the addition of a conventional radical initiator (in excess).[102] The amount of mediator can be lowered since the accumulated deactivator is regenerated by the slow, but continuous, generation of free radicals from the conventional radical initiator.

Two interesting applications of ATRP are surface initiated ATRP (SI-ATRP) and copolymerizations with uniform monomer distribution (i.e. all polymer chains have similar monomer sequence). A very large number of applications of ATRP has appeared utilizing SI-ATRP, where polymer chains can be grown in a

15

2.2 Atom transfer radical polymerization (ATRP)

controlled manner from a solid substrate, thus creating new (tailored) properties of the surface (e.g hydrofobicity or biocompatibility). The ATRP initiator is immobilized on the surface and polymer grafts are grown directly from the surface. Examples are surface modification of cellulose, silica (nano)particles, gold and iron nanoparticles and silica wafers. Biomedical applications is one area where this field is growing. Stimuli responsive polymers are also frequently investigated for SI-ATRP. The resulting modified surface changes properties in response to e.g. changes in pH or temperature.[103-106] Building advanced polymer architectures is also one area where ATRP has great potential. Besides block copolymers, ATRP (and CRP in general) provides conventional copolymers (e.g. random or gradient copolymers) with chain compositions very different from FRP. The polymer chains grow simultaneously throughout the polymerization and are thus subjected to the same monomer composition, i.e. all chains will have the same composition. In FRP, however, chains start growing at different times during the polymerization, at the same time as the monomer composition continuously changes, i.e. the different chains will have different compositions in the end.

16

3 Experimental

3 EXPERIMENTAL
This section briefly describes the methods and experimental procedures used in the studies for this thesis. Detailed descriptions can be found in the appended papers.

3.1

MATERIALS

All monomers were passed through a column of neutral aluminum oxide to remove inhibitors prior to use. The oligo(ethylene glycol) based initiator, OEGBr, was synthesized from monomethoxy capped oligo(ethylene glycol) and 2bromoisobutyryl bromide according to literature procedure.[85] The ligand Me6TREN was synthesized from tris(aminoethyl)amine similar to procedure presented by Ciampolini and Nardi.[107] Deionized Milli-Q water was used. All other chemicals and solvents were purchased from commercial sources and used as received.

3.2
1

CHARACTERIZATION

H NMR spectra were recorded on a Bruker Avance 400 MHz NMR instrument using the solvent signal or TMS as internal reference. NMR spectra were recorded in D2O (for OEGMA) or CDCl3. Size exclusion chromatography (SEC) was performed on two different instruments. For the poly(OEGMA) samples, SEC was performed using a Waters 717 Plus autosampler and a Waters model M-6000A solvent pump equipped with a PL-EMD 960 light scattering evaporative detector, two PLgel 10-mm mixed B columns (300 x 7.5 mm) from Polymer Laboratories and one Ultrahydrogel linear column (300 x 7.8 mm) from Waters, connected to an IBM-compatible computer. The eluent was N,N-dimethylformamide (Lab-Scan, Sweden) at a flow rate of 1.0 ml/min and narrow molecular weight linear poly(ethylene oxide) standards were

17

3.3 Cyclic voltammetry

used for calibration. Millennium software version 3.20 was used to process the data. For all other polymers, SEC using THF (1.0 mL min-1) as the mobile phase was performed at 35 C using a Viscotek TDA model 301 equipped with two T5000 columns, a VE 5200 GPC autosampler, a VE 1121 GPC solvent pump and a VE 5710 GPC degasser (all from Viscotek corp.). A conventional calibration method was created using narrow linear polystyrene standards. (The error in the molecular weight due to the difference in hydrodynamic volume between the polystyrene standards and MMA polymers was estimated to 10 %.) Corrections for the flow rate fluctuations were made using toluene as an internal standard. Viscotek OmniSEC version 4.5 software was used to process data.

3.3

CYCLIC VOLTAMMETRY

A short description of cyclic voltammetry (CV) can be justified. A potential is applied between the working electrode and the reference electrode. The potential is varied linearly with time at some scan rate (V/s) and the current between the working electrode and the counter electrode is monitored. A voltammogram is created by plotting the current versus the potential. As the reduction potential of the analyte is reached, the current increases but decreases again when the analyte concentration is depleted near the surface of the electrode. The opposite occurs upon reoxidation of the analyte. The half-wave potential for the analyte is given by E1/2 = (Eox - Ered)/2. Cyclic voltammetry was performed with a PAR 263A potentiostat/galvanostat interfaced to a base PC using the EG&G Model 270 software package. The cell was a standard three-electrode setup using a 2 mm diameter glassy carbon working electrode, a platinum coil counter electrode and a saturated calomel reference electrode. Full IR compensation was employed in all measurements. CV-measurements were performed for copper complexes in OEGMA-solvent mixtures and for iron complexes in DMF. For the copper complexes, the OEGMAsolvent contained 33 weight-% solvent. The supporting electrolyte was 0.5 M KCl in OEGMA-water and 0.1 M Bu4NBF4 in the other OEGMA-solvent mixtures. The scan rate was 200 mV/s and the potentials were measured against ferrocene. For the iron complexes in DMF, 0.1 M Bu4NBF4 was used as supporting electrolyte. The potentials of the iron complexes were measured against the saturated calomel reference electrode (SCE) at 3 mM concentration and the scan rate was 1000 mV/s.

18

3 Experimental

3.4

POLYMERIZATIONS

The polymerization experiments had some experimental details in common. Polymerizations were performed under argon atmosphere in round bottomed flasks placed in a pre-heated oil bath at the desired temperature. The initiator was added separately after thoroughly degassing the polymerization solution, before placing flask with the polymerization mixture in the oil bath. The monomer conversion was followed by withdrawing aliquots at timed intervals, which were analyzed by 1H NMR spectroscopy. The mediator was removed by passing the polymer solution through a column of neutral aluminum oxide.

3.4.1

Solvent effects

ATRP of OEGMA was performed at ambient temperature in deaerated solutions with 33 weight-% solvent. Six different solvents were utilized: 2-PrOH, MeOH, MeCN, DMF, DMSO and H2O. The targeted DP was 100 and the initiator (OEGBr) to mediator ratio was 1:1. Five different copper complexes were used as mediators CuBr with bipy, HMTETA, PMDETA, Me4-cyclam and Me6-TREN. The CuBr/ligand ratio was 1/1 for all ligands except bipy for which it was 1/2. Polymerizations were allowed to proceed for between 20 and 130 min depending on the mediator and solvent, and interrupted by exposing them to air.

3.4.2

Heterogeneous polymerization with Fe/EDTA

Heterogeneous polymerizations of styrene in p-xylene (40 weight-% solvent) were performed with FeCl2 as mediator (EDTA was also present) at 50 C. The targeted DP was 78 and the iron to initiator ratio was varied between 1 and 0.03. In most of the experiments the heterogeneous polymerization mixture was subjected to ultrasound for 5 minutes prior to the start of the polymerization. The ultrasound equipment was a Sonics Vibra-Cell VCX-750 (20 kHz) with a 13 mm probe extender which was immersed 5-10 mm into the solution. The polymerization times varied up to 6.5 hours depending on polymerization conditions.

3.4.3

Iron complexes in DMF

Homogeneous ATRP of MMA was performed with iron mediators in DMF (50 weight-%). Normal ATRP with FeCl2/ligands (at 65 C or 90 C), reverse ATRP with FeCl36 H2O/ligands and AIBN (at 90 C) and AGET ATRP with FeCl 36 H2O/ligands and ascorbic acid (at 90 C) were carried out with a targeted DP of

19

3.5 Simulations

200 and a 1:1 initiator to mediator ratio. Six ligands were used for the polymerizations: bipy, PMDETA, PPh3, IPA PA and PDA. The polymerizations were interrupted after around 23 hours by exposing the solutions to air.

3.5

SIMULATIONS

The GEPASI software (version 3.30) was used for the simulations of the ATRP kinetics. GEPASI is a software package for simulation of the kinetics of chemical and biochemical reactions, building and solving differential equations that govern the behavior of the system and can simulate both steady-state and time-course behavior.[108-110] For the simulations, the reaction steps included were the ATRP equilibrium, the propagation and the termination by combination. The rate constants used (i.e. ka, kd, kp and kt) were given appropriate values; for the propagation rate constant (kp) literature values[111] for common monomers were used, for the termination rate constant (kt) three different approaches were used either a constant kt or a conversion dependent kt (for details, see the Results section). The activation rate constant (ka) was calculated from a given value of the equilibrium constant, using a linear correlation between the two based on experimental data (for details, see Paper IV).[57] The deactivation rate constant (kd) was then calculated from eq. 9. The equilibrium constant and the propagation rate constant were varied in the simulations to mimic different ATRP systems. The outcome from kinetic simulation runs were the variations in reactant and product concentrations with time.

20

4 Results and Discussion

4 RESULTS AND DISCUSSION


4.1 SOLVENT EFFECTS ON ATRP

The solvent effects on ATRP were briefly discussed in the Introduction. Coullerez et al. investigated the solvent effects on the reduction potentials of copper complexes (ECu).[61] The solvent effects on the redox properties were significant and the solvent sensitivity appeared to be larger for complexes with ligands having higher degrees of freedom, i.e. the ligand with the most rigid structure (Me4-cyclam) displayed the weakest solvent sensitivity. The solvent effects were also analyzed in terms of Kamlet-Taft relationships. The experimentally determined values of the reduction potentials and the values estimated from the Kamlet-Taft relationships showed good correlation.[61] In this study, the investigation of the solvent effects on ATRP was extended to monomer-solvent mixtures, with measurements of both copper complex reduction potentials and polymerization properties. Previously, linear relationships between log(kpapp) and ECu have been shown for copper complexes in MeCN[58] and in aqueous solution.[59] In aqueous solution, the reduction potentials in pure water did not correlate to the ATRP kinetics, whereas ECu measured in the monomerwater mixture used in the polymerizations did. This demonstrates the importance of carefully choosing the solvent for electrochemical measurements, especially for solvents such as water, which can be expected to have very different properties compared to the polymerization solution (whether in bulk or with solvent present). The monomer used in the present investigation (as in the study by Coullerez et al.[59]) was oligo(ethylene glycol) methacrylate (OEGMA), a methacrylate with an ethylene glycol side chain on average 8 units long. To ensure similar conditions in the electrochemical measurements and the polymerizations, the same OEGMA-solvent mixtures were used in both, i.e. a 2:1 ratio of OEGMA:solvent (by weight). The solvents included in the study were 2-PrOH, MeOH, MeCN, DMF, DMSO and water (for which some of the data are from ref. [59]). Five common ligands were used to form complexes with CuBr: bipy, PMDETA, HMTETA, Me6-TREN and Me4-cyclam.

21

4.1 Solvent effects on ATRP

4.1.1

Reduction potentials of mediators

The reduction potentials of the copper complexes were measured by cyclic voltammetry. The results in pure solvents and in OEGMA-solvent mixtures are compared in Figure 2. The reduction potentials of the five complexes span more than 400 mV in both pure solvent and in OEGMA-solvent mixtures, corresponding to a more than six orders of magnitude difference in KATRP (eq. 9). The difference in reduction potentials is in many cases not very large between the pure solvent and the OEGMA-solvent mixture. The potentials in aqueous solution are more scattered and the value for CuBr/Me6-TREN in aqueous solution deviates strongly from the values for that complex in the other solvents (and mixtures). The spread in ECu values (the difference between the highest and lowest value) for a particular complex is larger in pure solvent than in the OEGMA-solvent mixtures. Since OEGMA constitutes the major part of the mixture, the solvation, and thus the electrochemical properties, of the complexes are largely governed by the properties of OEGMA.

Figure 2. Half-wave potentials (E1/2 or ECu) for copper complexes in OEGMAsolvent mixtures and in pure solvents. Potentials given in V vs. ferrocene (in pure water vs. ferrocyanide). The dotted line represents the 1:1 relationship.

22

4 Results and Discussion

4.1.2

Polymerization kinetics

From the polymerizations of OEGMA in the various solvents, the apparent rate constants (kpapp) were calculated from the initial slopes in the ln([M] 0/[M]) versus time plots. Log(kpapp) plotted against the reduction potentials in the OEGMA-solvent mixtures is shown in Figure 3. A general trend can be detected kpapp increases with decreasing ECu down to a point where it becomes essentially independent of the potential. This is caused by the persistent radical effect, PRE, i.e. the equilibrium becomes shifted towards the dormant species due to the buildup of the deactivator concentration from termination reactions.

Figure 3. The logarithm of the apparent rate constant, log(kpapp) versus the reduction potentials of the copper complexes in OEGMA-solvent mixtures, ECu. The average log(kpapp) versus average ECu for each complex is also included. The circled area contains the points for CuBr/Me4-cyclam.

4.1.3

Solvent effects

The solvent effect on kpapp is a combination of the solvent effects on kp, ECu, ERX and Kx. The solvent effects on kpapp were significant, resulting in variations up to one order of magnitude for the same complex (Figure 3). From eq. 9 and 10, it can be estimated that a 100 mV difference in ECu would result in a factor 7 difference in kpapp, but for some complexes in this study, kpapp varies a factor 10 within as little as a 20 mV difference in ECu (e.g bipy in DMSO and 2-PrOH and HMTETA in DMSO and MeCN). This indicates significant solvent effects also on other factors than ECu

23

4.1 Solvent effects on ATRP

that govern the polymerization and kpapp, e.g. the propagation rate constant. The cyclic, more rigid ligand Me4-cyclam stands out from the rest of the complexes with high apparent rate constants at relatively high (positive) potentials in the nonaqueous solvents. The reason can be that the oxidation of the complex requires reorganization of the coordination sphere, which can be more difficult with a rigid ligand, but it is also possible that the activation process proceeds through a different mechanism compared to the other complexes, e.g. electron transfer versus atom transfer. Linear free energy relationships or linear solvation energy relationships (LSER) can often be used to describe properties in solution, such as solubility, rates of reaction and enthalpy of equilibria.[112] The Kamlet-Taft expression (eq. 12) has been found to be one of the most successful relationships, where XYZ is the property of interest, XYZ0, a, b, s and h are solvent independent coefficients characteristic of the process, is the hydrogen bond donor ability of the solvent, is the hydrogen bond acceptor or electron pair donor ability to form a covalent bond, * is its dipolarity/polarizability parameter and H is the Hildebrand solubility parameter, which is a measure of the solvent-solvent interactions that are interrupted in creating a cavity for the solute.

XYZ = XYZ0 + a + b + s* + hH

(12)

For some processes, any of the coefficients XYZ0, a, b, s or h may be negligibly small, so that the corresponding terms do not play a role in the characterization of the solvent effects for these processes. Kamlet-Taft relationships have been used by Coullerez et al. [61] and Brauecker et al.[34] to analyze the solvent effects on the mediator reduction potentials and the ATRP equilibrium constants, respectively. In both cases, the correlation between the predicted and experimental values were good, demonstrating that the solvent effects are well described by Kamlet-Taft relationships. In the present study, both the reduction potentials and the apparent rate constants were analyzed in terms of Kamlet-Taft relationships. Though too few solvents were included to allow for quantitative evaluation of the Kamlet-Taft coefficients, some conclusions could be drawn from the relative importance of the parameters (i.e. , , * and H). For the reduction potentials, the ligands were affected differently by the solvent properties and the relative importance of the KT parameters were different in pure solvents and OEGMA-solvent mixtures. Thus, it

24

4 Results and Discussion

is obvious that the added solvents affect the redox properties significantly although OEGMA is the dominating part of the solvent mixtures. The relative importance of the KT parameters shifted for the apparent rate constants compared to the reduction potentials, which indicates that the solvent effects on kp and ERX have higher impact on kpapp than the solvent effects on ECu.

4.1.4

Limit of control

Besides the solvent effects on the magnitude of ECu and kpapp, the degree of control over the polymerizations is also important. For the polymerizations in four of the solvents, i.e. water, DMSO, 2-PrOH and MeCN, molecular weights and PDIvalues were evaluated in addition to the kinetics. The solvents were chosen to cover a representative range of potentials and apparent rate constants. A well controlled system is defined by having first order kinetics, linearly increasing molecular weights with conversion and low PDI-values. The systems with the lowest apparent rate constants were generally better controlled. Two examples of kinetics and molecular weights are shown in Figure 4, representing a fairly well controlled system (CuBr/PMDETA) and a poorly controlled system (CuBr/Me 6TREN) in MeCN.

Figure 4. Kinetics (left) and molecular weights and PDI-values (right) from ATRP of two OEGMA systems in MeCN. In the right plot, the filled symbols are the molecular weights (Mn in g/mol) and the empty symbols are the PDI-values, the solid line is the theoretical molecular weight.

25

4.2 Simulations

The best control was achieved with CuBr/bipy, whereas CuBr/Me 6-TREN resulted in poorly controlled polymerizations. In water, the control was generally poor. Comparing the results of the assessment of the degree of control with Figure 3 it is obvious that the degree of control decreases as the potential independent part is approached. Thus, at some point the limit of control will be reached, i.e. below a certain potential, well controlled ATRP can no longer be achieved.

4.2

SIMULATIONS

As shown in the previous section, a limit of control seems to be reached and the apparent rate constant ceases to increase as more active copper complexes are used. Although many aspects of the ATRP system have been investigated, the profile shown in Figure 3 has not been reported elsewhere. As mentioned in the Introduction, kinetic simulations of ATRP systems have been reported by several groups and can be a useful tool to investigate the ATRP process. In this study, kinetic simulations of the ATRP process were performed using the GEPASI software. The investigation was based on the magnitude of the equilibrium constant, KATRP, which was varied over more than 10 orders of magnitude. The monomer conversion and the degree of polymerization (DP = ([M]0-[M])/[RX] + 1) were used to evaluate the simulation results for systems with different propagation rate constants (corresponding to different monomers). The termination rate constant, kt, is an important parameter in the simulations, which varies with conversion and chain length. To account for the conversion and chain length dependence of kt, two different approaches were used. The first correlation between kt and the conversion and chain length used is the composite model, introduced by Smith et al.[113] and Johnston-Hall et al.[114], describing the termination rate constant between two radicals of chain length i, kti,i as a function of the termination rate constant between two monomeric radicals, kt1,1, eq. 13-16.

Dilute solution regime, for i < igel if ix < iSL if ix iSL


i k t ,i

k 1,1 i x t

(13)
S)

k ti ,i

( k 1,1 iSLL t

ix

(14)

26

4 Results and Discussion

Gel regime, for i igel if ix < iSL


k ti ,i
i k t ,i

k 1,1 i gelL t

S)

ix
(

gel

(15)
L) gel

if ix iSL

k1,1 iSLL t

S)

i gelgel

ix

(16)

The parameters in eq. 13-16 are explained in Paper IV and the values used in the simulations were taken from Johnston-Hall et al.[115] and are given in Table 2. Table 2. Parameters used in the simulations for the chain length and conversion (x) dependent termination rate constant, kti,i (eq. 13-16).[115] MA log kt1,1 S L iSL igel gel 9.0 0.78 0.15 18 6.90 x
-2.20

MMA MAN 9.1 0.65 0.15 100 0.53 x


-2.5

St 8.7 0.53 0.15 30 3.30 x-2.13 1.22 x 0.11

0.81 x 0.05

1.66 x 0.06

The second correlation used to calculate the chain length and conversion dependence of kt is given in eq. 17, where DP is calculated from ([M]0-[M])/[RX] + 1, wM,0 is the initial weight fraction of monomer and p is the monomer conversion. kt,0 is the termination rate constant at 0 % conversion, i.e. for small radicals. For kt,0, the values for kt1,1 in Table 2 were used. Eq. 17 is similar to the expression used by Shipp and Matyjaszewski in ATRP simulations[83] and is based on an empirical relationship for the normalized diffusion coefficients of oligomeric styrenes and methacrylates.[116]

k t, DP

k t,0 DP

( 0.664 2.02 p w M ,0 )

(17)

27

4.2 Simulations

In addition to the conversion and chain length dependent termination rate constants, simulations were performed with a constant value of kt for comparison. kt = 5 107 M-1 s-1 was used.

Kinetic plots for an ATRP system with MMA and two different mediators, simulated with the three different termination rate constants, are shown in Figure 5. In the simulations with constant kt, curvature in the kinetic plots was found even with the lowest equilibrium constant. With the conversion dependent termination rate constants, the kinetic plots were linear overall, but curvature appeared at low conversions. Linear kinetic plots, i.e. apparent first order kinetics, which is often seen in experimental ATRP systems, is a sign of good control in ATRP since it means a constant radical concentration with time. The simulation results with the conversion dependent kt thus resemble experimental data more than when a constant kt is used.

Figure 5. Simulation results with different termination rate constants for two ATRP systems where [M]0 = 4.5 M, [RX]0 = [Cu(I)]0 = 0.045 M and kp = 1400 M-1 s-1. (A) ka = 0.015 M-1 s-1, kd = 5.7 106 M-1 s-1; (B) ka = 1.9 M-1 s-1, kd = 2.3 104 M-1 s-1

Simulations varying KATRP and kp were performed. The initial apparent rate constants for the simulated systems were calculated from the initial slopes, i.e. between 0 and 6 % conversion, in the kinetic plots. Figure 6 shows log(kpapp) from the simulations of four ATRP systems (with kt calculated with the composite model, eq. 13-16) plotted against log(KATRP). As expected, the apparent rate constant increases with KATRP, but a maximum is reached, i.e. more active mediators would eventually result in decreasing apparent rate constants. This is

28

4 Results and Discussion

very similar to what was seen for the experimental system with OEGMA in the previous section, although that trend was shown against ECu, which however is analogous to KATRP through eq. 9. The appearance of a maximum is the result of a high initial radical concentration as a consequence of the high equilibrium constant, which results in a high degree of terminations early in the polymerization. The polymerization system becomes depleted of RX and together with the high deactivator concentration, this results in an lower apparent rate constant. Curves corresponding to those in Figure 6 were very similar with kt calculated from eq. 17 and with the constant kt (see Figure 7 for an example with MMA).

Figure 6. Initial kpapp versus KATRP for four simulated ATRP systems with different values of kp. [M]0 = 4.5 M, [RX]0 = [Cu(I)]0 = 0.045 M, kt calculated with the composite model.

In a termination free ATRP system, i.e. where [R] = [Cu II], log(kpapp) is linearly dependent on log(KATRP) (eq. 9 and 11). Figure 7 shows this linear relationship together with simulated systems with different termination rate constants, under the same conditions. The data sets are linear and apparently parallel at low values of KATRP, but the simulated system deviates from linearity as KATRP increases due to an increased degree of terminations, leading to the build-up of deactivator and a shift of the equilibrium towards the dormant species. At low enough KATRP values, there would be no terminations and the data sets should have a common intercept. Since the slopes are practically the same, the common intercept appears at very low KATRP values in the order of log(KATRP) < - 4000. It is thus obvious that even for ATRP systems where the equilibrium is strongly shifted towards the

29

4.2 Simulations

dormant species, the apparent rate constant is overestimated in the termination free system. The concentration of CuII in the simulated system is in the order of ppm relative to the CuI concentration at the point with the lowest value of KATRP in Figure 7, but even that minute amount of Cu II results in a deviation from the ideal termination free system.

Figure 7. Comparison between a termination free () and simulated (symbols) ATRP systems. [M]0 = 4.5 M, [RX]0 = [Cu(I)]0 = 0.045 M, kp = 1400 M-1 s-1.

4.2.1

Predicting the limit of control

It would be very useful to be able to predict the level of control in an ATRP system from simple relationships based on thermodynamic parameters such as ECu. This section describes how the simulations of different ATRP systems can be used to predict the level of control. As mentioned in the section describing the experimental results with OEGMA, the degree of control deteriorated as the potential independent part of Figure 3 was approached. That is, at some point along the respective curves in Figure 6 the level of control over the polymerization will deteriorate significantly, the limit of control is reached. In a poorly controlled ATRP system, large amounts of irreversibly terminated polymer chains/initiators result in broader molecular weight distributions and a degree of polymerization (DP) higher than targeted. Thus, the level of control can be assessed by comparing the simulated (or experimental) DP with the theoretical DP (DPth = [M]0/[RX]0 conversion). Further, to find the limit of control, the radical concentration was used as a measure of the degree of terminations by comparing

30

4 Results and Discussion

the actual radical concentration (from simulations) with the equilibrium radical concentration ([R]eq), which is given by eq. 9 for a system without terminations. From the simulations of different ATRP systems (i.e. different values of kp and KATRP) the radical concentrations and DPs at different conversions were used to create plots of log([R]/[R]eq) versus log(DP/DPth). Figure 8 displays the plot for one system at 60 % conversion, where the slope and the intercept of the straight line vary with kp.

Figure 8. The linear relationship between log([R]/[R] eq) and log(DP/DPth) for a simulated ATRP system (kp = 1400 M-1 s-1, kt calculated with the composite model). The [R] and DP values were taken at 60 % conversion from simulations with different values of KATRP.

A ratio of 1.25 between DP and DPth results in a small but acceptable deviation of DP from DPth at high conversions, and was used to find the limit of control. From Figure 8, DP/DPth = 1.25 gives the [R]/[R]eq ratio at the limit of control, [R]/[R]eq lim, which can be used further. The radical concentration data from the simulation can also be used to create a plot of log([R]/[R] eq) versus log(KATRP), Figure 9. From this, the maximum value of the equilibrium constant that will result in controlled polymerization in this system, KATRP lim, can be found from [R]/[R]eq lim, see Figure 9.

31

4.2 Simulations

Figure 9. Log([R]/[R]eq) versus log(KATRP) for a simulated ATRP system (the same system as in Figure 8). The square symbol is the limit of control. The inset shows an enlargement of the region around the limit of control (marked with a rectangle).

Repeating this for other ATRP systems, i.e. different values of kp with constant or conversion and chain length dependent kt, the limiting KATRP can be plotted against the propagation rate constants, Figure 10. KATRP lim only depends on the type of monomer, i.e. kp, as the monomer concentration or targeted DP do not affect the limit (although KATRP lim also varies with the termination rate constant, see Figure 10). With these simulation results, it is fairly straightforward to design a controlled ATRP system. With a known value of kp for the monomer of interest, KATRP lim can be found from Figure 10 by employing the equations for the linear regression lines. Controlled ATRP will be achieved for systems with KATRP lower than this limiting value. (To be on the safe side, the lower limit should be used.) To find a suitable mediator, KATRP can be calculated from eq. 9 if ECu and ERX are known (ECu can easily be measured if unknown). Another option is to measure KATRP (see Tang et al.)[53] or to use literature values of KATRP.[57] A third possibility is to simply try a mediator in a polymerization and if the outcome is uncontrolled ATRP, a less active mediator should be chosen (i.e. with a higher (more positive) reduction potential).

32

4 Results and Discussion

Figure 10. The values of KATRP at the limit of control versus the propagation rate constants, kp, for simulated ATRP systems (at 60 % conversion) with different termination rate constants, i.e. kt = 5 107 M-1 s-1 (), kt from the composite model () and kt from eq. 17 (, ---). Equations from the linear regression: with constant kt, log(KATRP lim) = 1.01 log(kp) 10.05; and with kt from the composite model, log(KATRP lim) = 1.01 log(kp) 10.30.

The results so far have shown that for homogeneous ATRP systems, the limit of control can be described based on the thermodynamic properties of the system. The trend in kpapp with KATRP shown in the simulations were in accordance with the experimental results with OEGMA. The next part of the work deals with ATRP systems where iron is used as the mediator, but where the systems are not as easily described as shown above due to heterogeneity or competitive coordination to the mediator by the solvent.

4.3

ATRP SYSTEMS WITH IRON MEDIATORS

The results from the experiments with ATRP of OEGMA in different solvents led to the investigation of ATRP systems with iron as the mediator. The use of iron is beneficial from an environmental point of view, since copper (and many ligands) is harmful both to the environment and to humans. Exploring other ligands than the ones commonly used with copper is thus also of interest. With OEGMA in water, oxalate was successfully used as ligand for copper, [59] which makes it interesting to look at other ligands of similar structure, e.g. EDTA. The direct correlation between log(kpapp) and ECu shown for copper complexes (Figure 3

33

4.3 ATRP systems with iron mediators

and ref. [58,59]) has not been reported for iron complexes, although qualitative comparisons have been made between the polymerization kinetics and the reduction potentials of the iron complexes.[62-66] Firstly, the polymerization of styrene with iron chloride will be discussed, where after the investigation of ATRP with several iron complexes in DMF is reported.

4.3.1

Styrene and iron chloride

The intention was to investigate the use of iron complexed with EDTA in ATRP. EDTA forms 1:1 complexes with FeII and FeIII and the complex could be interesting as an ATRP mediator due to its low toxicity. In water, the reduction potential of the Fe/EDTA complex is around -0.14 V vs. SCE, which is close to the reduction potentials of many copper complexes which can be used for controlled ATRP (e.g. in Figure 3, ECu (CuBr/PMDETA in MeCN) = -0.098 V vs. SCE). Thus, using the previous results for copper complexes (vide supra) as a guide, Fe/EDTA has the potential to give reasonably well controlled ATRP. Initially, several attempts were made to perform polymerizations in aqueous solution with NIPAAM or OEGMA, but with poor results either gelation occured or no or very little polymer was generated. Previously however, the polymerization of styrene with Fe/EDTA has been reported, [117] which made us attempt to use the same system as a starting point and see if it could be developed further to other solvents and monomers. However, the mediator was not soluble in the polymerization medium (i.e. styrene + p-xylene) which resulted in a heterogeneous system. The previous report did not mention heterogeneity and the results indicated reasonably controlled ATRP,[117] wherefore we continued the investigation of the heterogeneous system, varying the mediator amounts and the temperature. Although the presence of EDTA did not affect the polymerization results due to its insolubility and consequent inability to interact with the equally insoluble iron chloride, the majority of the experiments were performed with EDTA present in a 1:1 ratio to FeCl2. The molar ratio of initiator (1-PEBr) to mediator (i.e. FeCl 2) was varied between 1:1 and 1:0.03. The variation of reaction conditions also included the use of ultrasound, initially in an attempt to increase the solubility of the mediator. Thus, prior to the start of some of the polymerizations, the reaction mixture was subjected to ultrasound by means of a probe immersed in the polymerization solution.

34

4 Results and Discussion

The kinetics of some polymerizations of styrene with FeCl 2 as mediator and with ultrasound pretreatment are shown in Figure 11. Although the experiments display first order kinetics initially, they start to deviate at around 50 % conversion, but seem to resume first order kinetics after some time.

Figure 11. Kinetics of the heterogeneous polymerizations of styrene in p-xylene with iron chloride as mediator. [St]0 = 5.1 M, [St]0:[PEBr]0 = 78/1, 50 C, ultrasound pretreatment. Mediator to initiator ratios: 0.8 (A2), 0.1 (A3), 0.05 (A5) and 0.03 (A7).

Table 3. Initial apparent rate constants from the heterogeneous experiments with FeCl2/EDTA and styrene (with ultrasound pretreatment, * = without ultrasound pretreatment). (A few of the apparent rate constants have been recalculated and differ slightly from the values in Paper I.) Exp no A2 A3 A4 A5 A6 A7 A9* A10* [Fe]0/[I]0 0.8 0.1 0.08 0.05 0.04 0.03 1 0.1 [Fe]0 (mM) 51 6.8 4.7 3.2 2.6 2.1 65 6.8 kpapp (10-5 s-1) 59.2 51.6 39.4 36.9 28.6 29.4 10.4 3.0

35

4.3 ATRP systems with iron mediators

The apparent rate constant depends only weakly on the iron concentration, Table 3. In ATRP, a direct correlation between log(kpapp) and log[mediator] can be expected (eq. 9 + 11), which has also been shown for copper mediated ATRP for mediator to initiator ratios between 0.2 and 1.[118,119] However, as seen in Figure 12A, the slope in the log(kpapp) versus log[Fe] plot is only 0.47 at the lowest iron concentrations and even lower (slope = 0.07) as the iron to initiator ratio exceeds 0.1.

Figure 12. (A) Dependence of kpapp on the mediator concentration for the heterogeneous polymerizations of styrene with FeCl 2/EDTA in p-xylene at 50 C. Slope of regression line = 0.47. (B) Simulation results dependence of kpapp on the mediator concentration for a homogeneous ATRP system. Simulation performed with GEPASI. [M]0 = 4.5 M, [RX]0 = 0.045 M, kp = 1400 M-1 s-1, ka = 0.37 M-1 s-1, kd = 1.5 105 M-1 s-1, kt = f(conversion) (eq. 17).

The profile displayed in Figure 12A, with a steeper slope at mediator to initiator ratios below 0.1, can also be seen in kinetic simulations (of homogeneous systems, Figure 12B). This can be rationalized by the effect of irreversible terminations on the relative concentration of activator. When very low initial mediator concentrations are employed, the build-up of deactivator as the result of terminations will lead to a relatively low activator to deactivator ratio compared to when high mediator concentrations are employed (as an illustrative example: assuming that 10 % of the initiator terminates initially; with a 1:1 ratio between initiator and mediator, the deactivator will constitute 10 % of the total mediator concentration (i.e. activator + deactivator); however, with a 1:0.2 ratio between

36

4 Results and Discussion

initiator and mediator, the deactivator will constitute 50 % of the total mediator concentration). The apparent rate constant will thus be lower than expected as a result of the shifted equilibrium and it also results in a different correlation between kpapp and the mediator concentration, as seen in the simulation and in Figure 12A at low mediator concentrations. The molecular weights and PDI-values develop in practically the same way in all experiments, irrespective of the mediator amount. Some typical variations in molecular weights and PDI-values with conversion are displayed in Figure 13. The molecular weight increase is nearly linear, although lower than the theoretical molecular weight (Mn,th = DP Mmon conversion). The PDI-values are relatively narrow initially, but increase at higher conversions. The deviation from first order kinetics and the increasing PDI-values indicate a significant contribution of termination and transfer reactions.

Figure 13. Molecular weights and PDI-values for the heterogeneous polymerizations of styrene in p-xylene at 50 C with iron chloride as the mediator. For experimental conditions, see Figure 11. The straight line is the theoretical molecular weight.

The variation of kpapp with temperature was also investigated. The results from polymerizations at 30, 50 and 90 C with ultrasound pretreatment were used to calculate the apparent activation energy (Eaapp) from the Arrhenius plot. The value of kpapp at 90 C given in Paper I has been recalculated and the value 77 10 -4 s-1 is used here, which leads to a revised value of the apparent activation energy, Eaapp = 86 kJ/mol (the value was 107 kJ/mol in Paper I). This is significantly higher than

37

4.3 ATRP systems with iron mediators

the apparent activation energy for ATRP in homogeneous systems, which is reported to be around 60 kJ/mol.[118-123]

4.3.1.1

Discussion

The basic steps of the reaction are outlined in Scheme 6. In this heterogeneous system, the two important rate limiting steps are the adsorption of the dormant species to the FeCl2 particles (route 1 in Scheme 6) and the activation step in the ATRP equilibrium (route 2 in Scheme 6). The weak dependence of kpapp on the mediator concentration and the higher Eaapp compared to normal ATRP systems indicate that adsorption is the rate limiting step rather than the activation, which is rate limiting in normal ATRP.

Scheme 6. Proposed mechanism for the iron mediated polymerization of styrene. RX is the alkyl halide or halogen capped polymer. * = active species/radical, TA = chain transfer agent, M = monomer. Subscripts: sol = in solution, Fe = adsorbed to the FeCl2 surface.

When the initiator (1-PEBr) was heated alone with the mediator, no dimer of the initiator could be detected, which would be expected if the activation process yielded radicals that diffuse away from the mediator surface (route 4a in Scheme 6). Instead, the addition product between the initiator and the solvent (p-xylene) was observed. Attempts to polymerize MMA with the same system were not successful, neither with EBiB nor with 1-PEBr as initiators. These observations indicate that the initiator radical is further oxidized to a carbocation by the deactivator, route 5b in Scheme 6 (cationic polymerization is not possible for

38

4 Results and Discussion

MMA). FeIII is known to have the ability to oxidize radicals to cations. However, whether the polymerization of styrene proceeds through cationic polymerization or if there is some cationic character of the propagating center has not been determined. Cationic polymerization is characterized by high propagation rates and the main chain termination process is transfer to monomer or trace amounts of water or other adventitious impurities. To favor propagation over competing reactions, the polymerizations are usually performed at sub zero temperatures. Thus, whether this heterogeneous polymerization of styrene is poorly controlled ATRP or if it has some cationic character, with a significant degree of side reactions due to high temperature and the presence of impurities, cannot be established. Attempts to make PSt-PMMA block-copolymers through ATRP (using a copper complex as mediator) showed that some chain ends were apparently halogen end-capped, since the molecular weight increased significantly for a small fraction of the polymer, resulting in bimodal GPC traces. This indicates that there is at least some character of ATRP in the polymerization mechanism. Conclusively, polymerization of styrene was achieved with iron chloride as the mediator but the heterogeneity of the system complicated the interpretation of the results. Thus, the prediction that Fe/EDTA would work as ATRP mediator could not be confirmed in this system as it is far from the ideal case with a completely homogeneous system.

4.3.2

ATRP with iron mediators in DMF

As mentioned previously, only qualitative comparisons between the reduction potentials of the mediators and the polymerization results have been made for iron mediated ATRP. The general picture is that the apparent rate constant increases as the reduction potential is lowered (i.e. becomes more negative) as observed for copper complexes. In the search for more environmentally friendly ATRP mediators, where iron is an interesting candidate, it is important to understand the effect of different ligands on the ATRP system. In this work, the reduction potentials of iron in combination with eight different ligands were measured by cyclic voltammetry and polymerizations of MMA with the complexes were conducted in DMF. The intent was to investigate the relationship between the reduction potentials and the polymerization behavior in the same way as was previously done for copper complexes, as described above.

39

4.3 ATRP systems with iron mediators

The eight ligands used are presented in Figure 14. Intending to get a good spread in reduction potentials of the complexes, ligands containing different functional groups, i.e. carboxylic acids, pyridine units, amines and phosphines, were chosen. It is important to have a homogeneous system in a mechanistic study, since heterogeneity can complicate the interpretation of the results significantly, as seen above in the study of iron/EDTA. Therefore, DMF was utilized as solvent in both the polymerizations and the measurements of the reduction potentials, since all the iron complexes were completely soluble in DMF.

Figure 14. The structures of the ligands used in combination with iron.

4.3.2.1

Reduction potentials of mediators

The redox properties were measured for the eight ligands in combination with FeCl2, but also for FeCl2 and FeCl3 without ligands for comparison. The oxidation and reduction potentials of the iron complexes and FeCl 2 did not differ significantly in most cases. The reduction potentials (Ered) were close to 0 mV vs. SCE and the oxidation potentials (Eox) approximately 360 mV vs. SCE. The peakto-peak separation values (E = Eox Ered) are much larger than the ideal 60 mV for

40

4 Results and Discussion

a one-electron transfer process, demonstrating that the redox processes were not reversible. Some examples of voltammograms are shown in Figure 15.

Figure 15. Voltammograms for FeCl2/ligand complexes in DMF (0.1 M Bu4NBF4, scan rate = 1000 mV/s, FeCl2/ligand concentration = 3 mM).

Many of the complexes display peaks and shoulders at potentials other than 0 and 360 mV. For example, FeCl2/PDA exhibits a reversible redox process at E1/2 = -280 mV vs. SCE and the appearance of the voltammograms for the complexes with bipy, PA and IPA differ depending on the FeCl2/ligand ratio, which could indicate the presence of more than one structure of the iron complex, e.g. different number of ligands bound to iron. The nature of the iron species will be further discussed below.

41

4.3 ATRP systems with iron mediators

4.3.2.2

Polymerizations

The polymerizations of MMA were performed with normal, AGET and reverse ATRP, respectively. The similar redox behavior of the different iron complexes was reflected in the polymerization results, since the initial apparent rate constants did not differ more than a factor two within each ATRP method, Table 4. The variations in the voltammograms seen for some of the complexes did not seem to affect the polymerization behavior. Table 4. Initial apparent rate constants for normal, AGET and reverse ATRP of MMA with iron complexes.a Ligand bipy IPA PA PDA PMDETA PPh3 PPh3 PPh3 PDA PPh3 5% 50 % 50 % Added FeCl3b kpapp (10-5 s-1)c 65 C 9.2 2.8 6.6 4.2 4.7 3.9 8.6 8.1 0.4 0.3 6.1d 29.3 8.8 23.1 11.8 5.2 4.9 90 C Reverse 6.0

a) Initiators = EBiB (normal and AGET ATRP, [MMA]0:[I]0:[Fe]0 = 200:1:1), AIBN (reverse ATRP, [MMA]0:[I]0:[Fe]0 = 200:0.5:1), 50 % (w/w) DMF. b) In percent of [FeCl2]0. c) kpapp for normal ATRP at 65 C and 90 C and for reverse and AGET ATRP (at 90 C). d) AGET ATRP, [Fe]0:[AscAc]0 = 1:0.5.

The degree of control over the polymerizations was significantly better in reverse ATRP compared to normal ATRP. This was manifested in both kinetics and molecular weights. First order kinetics was found in all reverse ATRP systems, whereas curvature was seen in the kinetic plots of the normal ATRP systems. The kinetics of the ATRP systems with PPh3 as the ligand are shown in Figure 16. The results from AGET ATRP were very similar to reverse ATRP.

42

4 Results and Discussion

Figure 16. The kinetics of the normal and reverse ATRP of MMA in DMF with the iron-PPh3 complex as mediator.

Figure 17. Molecular weights and PDI-values for PMMA prepared by normal and reverse ATRP using iron/PPh3 as the mediator.

The molecular weights were constant throughout the polymerization in the normal ATRP systems. With AGET and reverse ATRP, the molecular weights increased linearly with conversion and were close to the theoretical molecular weight (Mn,th). The molecular weights and PDI-values in the ATRP systems with PPh3 as the ligand are shown in Figure 17. Examples of GPC traces are shown in Figure 18. The molecular weights of the polymers prepared by normal ATRP

43

4.3 ATRP systems with iron mediators

appear to decrease with conversion (Figure 17), but as seen from the GPC traces, this is an effect of an increasing fraction of lower molecular weight chains with conversion. The bimodality of the GPC traces for the polymers from normal ATRP can be the consequence of early terminations, resulting in the shift of the bimodal curve towards higher molecular weights at lower conversions.

Figure 18. GPC traces from normal and reverse ATRP of MMA with iron/PPh 3 in DMF at different monomer conversions.

Some of the polymers prepared by normal and reverse ATRP were also chain extended with MMA, using CuBr2/PMDETA as the mediator in AGET ATRP. The macroinitiators prepared by reverse ATRP increased significantly in molecular weight, though the molecular weight distributions were broad or bimodal. The macroinitiators prepared by normal ATRP did not grow to any larger extent, however. Since the macroinitiators were isolated at the end of the polymerization, it is probable that a large fraction of the chains from normal ATRP were terminated as assessed from the strong curvature in the kinetic plots. Thus, although the results indicate that the polymers prepared by reverse ATRP are irreversibly terminated to a lesser extent than the polymers prepared by normal ATRP, it is difficult to draw any conclusions regarding the degree of control from these chain extensions. However, 1H NMR spectroscopy of the macroinitiators indicated fewer active chain ends (i.e. halogen end-capped) from normal ATRP compared to reverse ATRP.

44

4 Results and Discussion

4.3.2.3

UV-vis spectroscopy

Some iron complexes, FeCl2 and FeCl3 with PPh3 or IPA, were analyzed by UV-vis spectroscopy. The majority of the measurements were performed with the complexes in DMF, but one solution was also prepared by mixing FeCl 2/IPA with EBiB in DMF/EtOAc (50/50 by weight) where the Fe III species resulting from FeII species oxidized through the activation of EBiB were analyzed. Figure 19 shows the UV-vis spectra of solutions with Fe/IPA.

Figure 19. UV-vis spectra of iron complexes with IPA in DMF or DMF/EtOAc (FeCl2/IPA + EBiB). There is no significant difference between FeCl 3 and FeCl3/IPA, indicating that the ligand does not bind to FeIII to any appreciable extent. There appears to be a small difference between FeII and FeIII species, however. At 360 nm, the oxidized FeII species (FeCl2/IPA ox and FeCl2/IPA + EBiB) are similar to FeCl 3/IPA. At 316 nm, the shoulders displayed for the oxidized FeII samples seem to be shifted towards shorter wavelengths compared to the peak at 316 nm for FeCl 3/IPA. However, since the peaks are not well defined, it is difficult to draw definite conclusions. At wavelengths between 450 and 600 nm, however, there is a clear difference between FeCl3/IPA and the oxidized FeII species, with a shift towards shorter wavelengths for the species originating from FeCl 2/IPA and FeCl2/IPA + EBiB.

45

4.3 ATRP systems with iron mediators

4.3.2.4

Discussion

The results from the electrochemical measurements, with practically the same redox properties irrespective of ligand, suggest that the ligands do not bind to iron. This was not expected since iron complexes with PPh 3 and PMDETA have been isolated and characterized.[124-126] Also, the redox properties of copper complexes vary significantly with the ligands in DMF as seen above in the study with OEGMA and by Coullerez et al.[61] However, it is known that DMF coordinates with iron[127,128] and electrochemical studies of iron complexes in DMF have indicated that only FeIII forms complexes, since it was only the reduction peak that was affected by ligand addition. [129,130] Although it would be entropically favorable to replace DMF with bidentate ligands such as bipy, the ability of DMF to compete with the ligands for the coordination sites on iron obviously affects the properties of the complexes significantly. It thus seems that the influence over the iron complex properties from coordinated DMF molecules is larger than the influence from the other ligands in this study, if they are coordinated at all. Similar conclusions can be drawn from the UV-vis study, where the spectra with and without ligand were practically the same, at least for FeCl 3. Although the electrochemical results did not support coordination of the ligands to FeII, some interaction between the ligands and iron was indicated by observations during the polymerizations. During normal ATRP, the polymerization solution became heterogeneous after some time when no ligands were added, presumably due to limited solubility of Fe III species formed, whereas the polymerization solution stayed homogeneous when ligands were added. This difference with and without ligands was not observed during reverse ATRP, where the polymerization mixtures were homogeneous all the time both with and without ligands. Another difference between normal and reverse ATRP was the color of the reaction mixtures. In normal ATRP, the mixture turned orange as the polymerization proceeded, most likely from the formation of FeIII species. However, the initial color of the FeIII containing reaction mixture in reverse ATRP was yellow. There was also a small difference between FeII species oxidized by the activation of EBiB and FeIII species from FeCl3 in the UV-vis spectra. These observations, together with the much better control over the polymerization in reverse ATRP compared to normal ATRP, indicate that the Fe III species are different in the two systems. The degree of control in AGET ATRP was comparable to reverse ATRP and it is thus a better approach to start from FeIII rather than from FeII. Constant molecular weights with conversion, which was seen in the normal ATRP systems, can be the result of e.g. catalytic chain transfer (CCT) or other

46

4 Results and Discussion

reactions between the growing polymer chain and the iron mediator. Fe III species are known to terminate polystyrene radicals by hydrogen transfer, leaving unsaturated chain ends,[131,132] and chain transfer by iron complexes has been reported for ATRP of styrene and MMA.[62,69] Gibson et al. have shown that the metal spin state of the iron mediator affects the polymerization outcome. [69,133,134] With low spin complexes, catalytic chain transfer dominated and the molecular weights were constant with conversion, whereas high spin complexes resulted in well controlled ATRP of MMA. Thus, there is a possibility that the difference seen between reverse and normal ATRP in the present study is the result of a difference in metal spin state of the FeIII species. Good control in reverse ATRP can result from incomplete initiation. If a significant portion of the initiator molecules terminate initially, the reduction of FeIII to FeII will be incomplete and the ATRP equilibrium will be shifted towards the dormant species. However, if this was the case, the molecular weights would be significantly higher than the theoretical molecular weights. Since that is not seen in the polymerizations by reverse ATRP here (the molecular weights are very close to the theoretical molecular weights), it suggests that practically all initiator molecules have started a polymer chain. Another possibility is that additional initiating species are produced by the oxidation of MMA by Fe III species. This redox mechanism has been suggested for copper and iron mediated ATRP where the polymerization is started from the mediator in its higher oxidations state, but without the addition of a reducing agent.[135-138] However, it has not been shown that the resulting alkyl halides can actually initiate ATRP and the overall initiation efficiency would still only be 80 %, since the suggested mechanism involves two mediator molecules and one MMA molecule.[135] Although the good control in reverse ATRP might be explained by the presence of FeIII species which shift the equilibrium towards the dormant species, the addition of a substantial amount of FeCl 3 (50 % relative to the initial FeII concentration) to normal ATRP did not improve the degree of control. The molecular weights were still constant with conversion, though the apparent rate constant decreased. This observation supports the suggestion of different Fe III species, since the better control in reverse ATRP cannot solely be explained by the possibility of a shift of the ATRP equilibrium towards the dormant species due to incomplete initiation. Thus, there are indications of different FeIII species in the normal and reverse ATRP systems, but the exact nature and the role of these species in the polymerization systems, in particular regarding the molecular weights, remain unclear. However, from this study it can be concluded that the nature of the

47

4.3 ATRP systems with iron mediators

ligand is not crucial for the polymerization (or electrochemical) properties of the system when DMF is used as solvent for iron mediated ATRP and that it is a better approach to start from FeIII rather than FeII in order to achieve a well controlled ATRP system.

48

5 Conclusions

5 CONCLUSIONS
The solvent effects on ATRP in homogeneous systems were investigated and it was found that the reduction potentials of the mediators (copper complexes) and the apparent rate constants were significantly affected by the solvent. The apparent rate constant increased as the reduction potential became less positive but eventually reached an upper limit, where after it became practically independent of the potential. The degree of control over the polymerization also decreased significantly as this limit was approached. Through kinetic simulations, the same trend of increasing apparent rate constants with increasingly active ATRP systems was found and a limit was reached as in the experiments, but with a maximal apparent rate constant instead of a plateau. The simulation results could also be used to describe the limit of control. The linear correlation in the double logarithmic plot between the propagation rate constant and the limiting equilibrium constant (i.e. above which controlled ATRP will not be achieved) enables the prediction of the degree of control in an ATRP system, which can be used as a tool for the design of well controlled ATRP systems. The attempt to use Fe/EDTA as the mediator in ATRP of styrene resulted in heterogeneous systems where the apparent rate constant depended very weakly on the mediator concentration. The possibility of a cationic character of the propagating center was discussed, but has not been proved. It became obvious however, that in order to use the aforementioned design tool, the ATRP system must be homogeneous. A similar simulation based design tool for controlled ATRP can of course be developed for heterogeneous systems as well, if the required parameters are known. However, the requirement of homogeneity for the predictive tool might not be enough as seen in the investigation of ATRP of MMA using different iron complexes in DMF. Due to the competitive coordination of the solvent to the mediator, the electrochemical and polymerization properties were practically the same irrespective of the nature of the added ligand. The magnitude of the reduction potentials of the iron complexes was in the range where controlled ATRP would be expected. However, normal ATRP resulted in uncontrolled polymerizations with indications of catalytic chain transfer, whereas reverse

49

5 Conclusions

ATRP gave controlled polymerizations, which could not be explained solely by the possibility of an ATRP equilibrium shifted towards the dormant species due to incomplete initiation. The results indicated different Fe III speciation in the two ATRP systems. In conclusion, there are properties of the ATRP system which can be difficult to foresee but which will affect the polymerization outcome significantly and which are not included in the predictive tool presented in this thesis. Besides heterogeneity and differences in mediator speciation, there are many other factors in an ATRP polymerization system, such as viscosity, that can affect the possibility to predict the outcome of the polymerization. Still, the ability to design controlled ATRP based on the thermodynamic properties of the ATRP system is a valuable addition to the ATRP literature available today.

50

6 Future work

6 FUTURE WORK
The investigation and description of the limit of control performed in this work was focused on homogeneous ATRP systems. It would be interesting to expand this to heterogeneous systems, thus increasing the possibility to design ATRP systems of different types. Surface initiated ATRP (SI-ATRP) is a growing field with an increasing number of possible applications. It would be beneficial if the degree of control could be predicted for SI-ATRP similarly as shown for homogeneous ATRP, thus enabling the design of SI-ATRP systems. We started to investigate the kinetics of SI-ATRP through polymerizations of MMA and styrene from MCC (microcrystalline cellulose) and filter papers. The question was whether the kinetics were similar for polymer chains growing in solution and chains tethered to a surface. We also wanted to see if there is a similar correlation between kinetics and mediator activity as for homogeneous systems. It would also be of interest to develop a kinetic model for SI-ATRP based on the thermodynamics of the ATRP equilibrium. In order to measure the kinetics of the surface polymerization, no sacrificial initiator can be added. It is also an advantage if the polymer can be cleaved from the surface for molecular weight analysis. However, due to the low initiator concentration in SI-ATRP, it can be difficult to measure the monomer conversion. The progress of the polymerization is commonly followed by measuring the conversion in bulk (from sacrificial initiator) or by measuring the thickness of the polymer film on the surface. In the preliminary study, we could follow the monomer conversion by 1H NMR, but unfortunately this monomer consumption was the result of spontaneous polymerization in solution (possibly by thermal initiation in some cases). With the materials we had at our disposal, it was difficult to achieve an initiator concentration comparable to that in homogeneous polymerizations. Thus, with a material where a high surface loading of initiator is possible and with a polymerization system where polymerization in the bulk is suppressed, a kinetic study of SI-ATRP would be very interesting. It could also be interesting to develop a model for ARGET ATRP. Since the activator is continuously regenerated by a reducing agent, the polymerization

51

6 Future work

system is more difficult to describe than normal ATRP if the kinetics of the reaction between the deactivator and the reducing agent are unknown. In addition, the reducing agent is often (partially?) insoluble (e.g. ascorbic acid in anisole), which leads to a heterogeneous system.

52

7 Acknowledgements

7 ACKNOWLEDGEMENTS
First of all, I would like to thank my supervisors Professor Mats Jonsson and Professor Eva Malmstrm for accepting me as a PhD student and for their support and guidance throughout these years. It has been a pleasure working with you.

The Swedish Research Council is gratefully acknowledged for financial support.

I also want to thank former and present coworkers of Nuclear Chemistry and Coating Technology (ytgruppen) for creating a nice atmosphere and brightening my days at work. Thanks to all those who have taken care of the SEC instrument over the years, where would we be without your struggle with an obstinate instrument? Special thanks to roommates and fume hood neighbors over the years, its always nice to know that you are not alone in the lab or in the efforts to make computers/polymerizations do what you want them to. Susanne is thanked for the help and cooperation with the attempts to perform SI-ATRP, I hope our struggle will lead to something good in the end.

Many thanks to my family who have supported me and for always believing in me.

53

8 References

8 REFERENCES
1. 2. Georges, M. K., Veregin, R. P. N., Kazmaier, P. M., Hamer, G. K., Narrow molecular weight resins by a free-radical polymerization process. Macromolecules, 1993, 26, 2987-2988 Chiefari, J., Chong, Y. K., Ercole, F., Krstina, J., Jeffery, J., Le, T. P. T., Mayadunne, R. T. A., Meijs, G. F., Moad, C. L., Moad, G., Rizzardo, E., Thang, S. H., Living Free-Radical Polymerization by Reversible AdditionFragmentation Chain Transfer: The RAFT Process. Macromolecules, 1998, 31, 5559-5562 Mayadunne, R. T. A., Rizzardo, E., Chiefari, J., Chong, Y. K., Moad, G., Thang, S. H., Living Radical Polymerization with Reversible AdditionFragmentation Chain Transfer (RAFT Polymerization) Using Dithiocarbamates as Chain Transfer Agents. Macromolecules, 1999, 32, 6977-6980 Wang, J.-S., Matyjaszewski, K., Controlled/"Living" Radical Polymerization. Halogen Atom Transfer Radical Polymerization Promoted by a Cu(I)/Cu(II) Redox Process. Macromolecules, 1995, 28, 7901-7910 Wang, J.-S., Matyjaszewski, K., Controlled/"living" radical polymerization. Atom transfer radical polymerization in the presence of transition-metal complexes. J. Am. Chem. Soc., 1995, 117, 5614-5615 Kato, M., Kamigaito, M., Sawamoto, M., Higashimura, T., Polymerization of Methyl Methacrylate with the Carbon Tetrachloride/Dichlorotris(triphenylphosphine)ruthenium(II)/Methylaluminum Bis(2,6-di-tertbutylphenoxide) Initiating System: Possibility of Living Radical Polymerization. Macromolecules, 1995, 28, 1721-1723 van de Kuil, L. A., Grove, D. M., Gossage, R. A., Zwikker, J. W., Jenneskens, L. W., Drenth, W., van Koten, G., Mechanistic Aspects of the Kharasch Addition Reaction Catalyzed by Organonickel(II) Complexes Containing the Monoanionic Terdentate Aryldiamine Ligand System [C6H2(CH2NMe2)2-2,6-R-4]-. Organometallics, 1997, 16, 49854994 Matyjaszewski, K., From Atom Transfer Radical Addition to Atom Transfer Radical Polymerization. Curr. Org. Chem., 2002, 6, 67-82 Minisci, F., Free-Radical Additions to Olefins in the Presence of Redox Systems. Accounts Chem. Res., 1975, 8, 165-171 Fischer, H., The Persistent Radical Effect: A Principle for Selective Radical Reactions and Living Radical Polymerizations. Chem. Rev., 2001, 101, 3581-3610

3.

4.

5.

6.

7.

8. 9. 10.

54

8 References

11. 12. 13. 14.

Matyjaszewski, K., Kajiwara, A., EPR Study of Atom Transfer Radical Polymerization (ATRP) of Styrene. Macromolecules, 1998, 31, 548-550 Kajiwara, A., Matyjaszewski, K., Simultaneous EPR and Kinetic Study of Styrene Atom Transfer Radical Polymerization (ATRP). Macromolecules, 1998, 31, 5695-5701 Kajiwara, A., Matyjaszewski, K., EPR and Kinetic Studies of Atom Transfer Radical Polymerization of (Meth)acrylates. Polym. J., 1999, 31, 70-75 Ouchi, M., Terashima, T., Sawamoto, M., Transition Metal-Catalyzed Living Radical Polymerization: Toward Perfection in Catalysis and Precision Polymer Synthesis. Chem. Rev., 2009, 109, 4963-5050, and references therein Poli, R., Controlled Radical Polymerization; Relationship between One-Electron TransitionMetal Reactivity and Radical Polymerization Processes. Angew. Chem. Int. Ed., 2006, 45, 5058-5070 Wayland, B. B., Poszmik, G., Mukerjee, S. L., Living Radical Polymerization of Acrylates by Organocobalt Porphyrin Complexes. J. Am. Chem. Soc., 1994, 116, 7943-7944 Asscher, M., Vofsi, D., Chlorine activation by redox transfer. Part II. The addition of carbon tetrachloride to olefins. J. Chem. Soc., 1963, 1887-1896 Asscher, M., Vofsi, D., Chlorine activation by redox-transfer. Part III. The "abnormal" addition of chloroform to olefins. J. Chem. Soc., 1963, 3921-3927 Freiberg, M., Meyerstein, D., An intermediate with a copper-carbon bond formed by the reaction of copper ions with CH2CO2H radicals in aqueous solution. Chem. Comm., 1977, 127-128 Mulac, W. A., Meyerstein, D., Intermidates with copper-carbon bonds formed by the reaction of aliphatic free radicals with a copper-peptide complex in aqueous solutions. Chem. Comm., 1979, 893-895 Masarwa, M., Cohen, H., Meyerstein, D., Effect of Nitrilotriacetate on the Mechanism of Reduction of Copper(II) Ions by -Hydroxyalkyl Free Radicals via Complexes with CopperCarbon Bonds as Intermediates. A Pulse-Radiolytic Study. Inorg. Chem., 1986, 25, 48974900 Shaver, M. P., Allan, L. E. N., Gibson, V. C., Organometallic Intermediates in the Controlled Radical Polymerization of Styrene by -Diimine Iron Catalysts. Organometallics, 2007, 26, 4725-4730 Haddleton, D. M., Clark, A. J., Crossman, M. C., Duncalf, D. J., Heming, A. M., Morsley, S. R., Shooter, A. J., Atom transfer radical polymerisation (ATRP) of methyl methacrylate in the presence of radical inhibitors. Chem. Comm., 1997, 1173-1174

15.

16. 17. 18. 19.

20.

21.

22.

23.

55

8 References

24.

Haddleton, D. M., Duncalf, D. J., Kukulj, D., Heming, A. M., Shooter, A. J., Clark, A. J., Atom transfer polymerisation of methyl methacrylate: use of chiral aryl/alkyl pyridylmethanimine ligands; with copper(I) bromide and as structurally characterised chiral copper(I) complexes. J. Mater. Chem., 1998, 8, 1525-1532 Heuts, J. P. A., Mallesch, R., Davis, T. P., Atom transfer radical polymerization in the presence of a thiol: more evidence supporting radical intermediates. Macromol. Chem. Phys., 1999, 200, 1380-1385 Matyjaszewski, K., Paik, H.-j., Shipp, D. A., Isobe, Y., Okamoto, Y., Free-Radical Intermediates in Atom Transfer Radical Addition and Polymerization: Study of Racemization, Halogen Exchange, and Trapping Reactions. Macromolecules, 2001, 34, 3127-3129 Yu, Q., Zeng, F., Zhu, S., Atom Transfer Radical Polymerization of Poly(ethylene glycol) Dimethacrylate. Macromolecules, 2001, 34, 1612-1618 Ziegler, M. J., Matyjaszewski, K., Atom Transfer Radical Copolymerization of Methyl Methacrylate and n-Butyl Acrylate. Macromolecules, 2001, 34, 415-424 Matyjaszewski, K., Mechanistic features and radical intermediates in atom transfer radical polymerization. Macromol. Symp., 2002, 183, 71-81 Harrisson, S., Rourke, J. P., Haddleton, D. M., 13C Kinetic isotope effects in the copper(I)mediated living radical polymerization of methyl methacrylate. Chem. Comm., 2002, 14701471 Wang, A. R., Zhu, S., ESR Study on Diffusion-Controlled Atom Transfer Radical Polymerization of Methyl Methacrylate and Ethylene Glycol Dimethacrylate. Macromolecules, 2002, 35, 9926-9933 Singleton, D. A., Nowlan III, D. T., Jahed, N., Matyjaszewski, K., Isotope Effects and the Mechanism of Atom Transfer Radical Polymerization. Macromolecules, 2003, 36, 8609-8616 Lad, J., Harrisson, S., Mantovani, G., Haddleton, D. M., Copper mediated living radical polymerisation: interactions between monomer and catalyst. J. Chem. Soc. Dalton, 2003, 21, 4175-4180 Braunecker, W. A., Tsarevsky, N. V., Gennaro, A., Matyjaszewski, K., Thermodynamic Components of the Atom Transfer Radical Polymerization Equilibrium: Quantifying Solvent Effects. Macromolecules, 2009, 42, 6348-6360 Lin, C. Y., Coote, M. L., Gennaro, A., Matyjaszewski, K., Ab Initio Evaluation of the Thermodynamic and Electrochemical Properties of Alkyl Halides and Radicals and Their Mechanistic Implications for Atom Transfer Radical Polymerization. J. Am. Chem. Soc., 2008, 130, 12762-12774 Ohno, K., Goto, A., Fukuda, T., Xia, J., Matyjaszewski, K., Kinetic Study on the Activation Process in an Atom Transfer Radical Polymerization. Macromolecules, 1998, 31, 2699-2701

25.

26.

27. 28. 29. 30.

31.

32. 33.

34.

35.

36.

56

8 References

37. 38.

Goto, A., Fukuda, T., Determination of the activation rate constants of alkyl halide initiators for atom transfer radical polymerization. Macromol. Rapid Commun., 1999, 20, 633-636 Chambard, G., Klumperman, B., German, A. L., Effect of Solvent on the Activation Rate Parameters for Polystyrene and Poly(butyl acrylate) Macroinitiators in Atom Transfer Radical Polymerization. Macromolecules, 2000, 33, 4417-4421 Matyjaszewski, K., Paik, H.-j., Zhou, P., Diamanti, S. J., Determination of Activation and Deactivation Rate Constants of Model Compounds in Atom Transfer Radical Polymerization. Macromolecules, 2001, 34, 5125-5131 Matyjaszewski, K., Gobelt, B., Paik, H.-j., Horwitz, C. P., Tridentate Nitrogen-Based Ligands in Cu-Based ATRP: A Structure-Activity Study. Macromolecules, 2001, 34, 430440 Chambard, G., Klumperman, B., German, A. L., Experimental Determination of the Rate Constant of Deactivation of Poly(styrene) and Poly(butyl acrylate) Radicals in Atom Transfer Radical Polymerization. Macromolecules, 2002, 35, 3420-3425 Gromada, J., Matyjaszewski, K., Measurement of Initial Degree of Polymerization without Reactivation as a New Method To Estimate Rate Constants of Deactivation in ATRP. Macromolecules, 2002, 35, 6167-6173 Nanda, A. K., Matyjaszewski, K., Effect of [PMDETA]/[Cu(I)] Ratio, Monomer, Solvent, Counterion, Ligand, and Alkyl Bromide on the Activation Rate Constants in Atom Transfer Radical Polymerization. Macromolecules, 2003, 36, 1487-1493 Nanda, A. K., Matyjaszewski, K., Effect of Penultimate Unit on the Activation Process in ATRP. Macromolecules, 2003, 36, 8222-8224 Nanda, A. K., Matyjaszewski, K., Effect of [bpy]/[Cu(I)] Ratio, Solvent, Counterion, and Alkyl Bromides on the Activation Rate Constants in Atom Transfer Radical Polymerization. Macromolecules, 2003, 36, 599-604 Pintauer, T., Braunecker, W., Collange, E., Poli, R., Matyjaszewski, K., Determination of Rate Constants for the Activation Step in Atom Transfer Radical Polymerization Using the Stopped-Flow Technique. Macromolecules, 2004, 37, 2679-2682 Matyjaszewski, K., Nanda, A. K., Tang, W., Effect of [Cu(II)] on the Rate of Activation in ATRP. Macromolecules, 2005, 38, 2015-2018 Tang, W., Nanda, A. K., Matyjaszewski, K., Effect of [Pyridylmethanimine]/[Cu(I)] Ratio, Ligand, Solvent and Temperature on the Activation Rate Constants in Atom Transfer Radical Polymerization. Macromol. Chem. Phys., 2005, 206, 1171-1177 Tang, W., Matyjaszewski, K., Effect of Ligand Structure on Activation Rate Constants in ATRP. Macromolecules, 2006, 39, 4953-4959 Tang, W., Matyjaszewski, K., Effects of Initiator Structure on Activation Rate Constants in ATRP. Macromolecules, 2007, 40, 1858-1863

39.

40.

41.

42.

43.

44. 45.

46.

47. 48.

49. 50.

57

8 References

51.

Lin, C. Y., Coote, M. L., Petit, A., Richard, P., Poli, R., Matyjaszewski, K., Ab Initio Study of the Penultimate Effect for the ATRP Activation Step Using Propylene, Methyl Acrylate, and Methyl Methacrylate Monomers. Macromolecules, 2007, 40, 5985-5994 Seeliger, F., Matyjaszewski, K., Temperature Effect on Activation Rate Constants in ATRP: New Mechanistic Insights into the Activation Process. Macromolecules, 2009, 42, 60506055 Tang, W., Tsarevsky, N. V., Matyjaszewski, K., Determination of Equilibrium Constants for Atom Transfer Radical Polymerization. J. Am. Chem. Soc., 2006, 128, 1598-1604 Tang, W., Kwak, Y., Tsarevsky, N. V., Matyjaszewski, K., Effect of ligand and initiator structures on the equilibrium constants in atrp. Polym. Preprints, 2007, 48, 392-393 Braunecker, W. A., Tsarevsky, N. V., Gennaro, A., Matyjaszewski, K., Understanding solvent effects on thermodynamic equilibria in atom transfer radical polymerization. Polym. Preprints, 2008, 49, 376-377 Chung, H., Tang, W., Matyjaszewski, K., Effect of temperture and solvent on the ATRP equilibrium constants and activation rate constants. Polym. Preprints, 2008, 49, 97-98 Tang, W., Kwak, Y., Braunecker, W., Tsarevsky, N. V., Coote, M. L., Matyjaszewski, K., Understanding Atom Transfer Radical Polymerization: Effect of Ligand and Initiator Structures on the Equilibrium Constants. J. Am. Chem. Soc., 2008, 130, 10702-10713 Qiu, J., Matyjaszewski, K., Thouin, L., Amatore, C., Cyclic voltammetric studies of copper complexes catalyzing atom transfer radical polymerization. Macromol. Chem. Phys., 2000, 201, 1625-1631 Coullerez, G., Carlmark, A., Malmstrm, E., Jonsson, M., Understanding Copper-Based Atom-Transfer Radical Polymerization in Aqueous Media. J. Phys. Chem. A, 2004, 108, 7129-7131 Tang, H., Arulsamy, N., Radosz, M., Shen, Y., Tsarevsky, N. V., Braunecker, W. A., Tang, W., Matyjaszewski, K., Highly Active Copper-Based Catalyst for Atom Transfer Radical Polymerization. J. Am. Chem. Soc., 2006, 128, 16277-16285 Coullerez, G., Malmstrm, E., Jonsson, M., Solvent Effects on the Redox Properties of Cu Complexes Used as Mediators in Atom Transfer Radical Polymerization. J. Phys. Chem. A, 2006, 110, 10355-10360 Gibson, V. C., O'Reilly, R. K., Reed, W., Wass, D. F., White, A. J. P., Williams, D. J., Four-coordinate iron complexes bearing -diimine ligands: efficient catalysts for Atom Transfer Radical Polymerisation (ATRP). Chem. Comm., 2002, 1850-1851 Gibson, V. C., OReilly, R. K., Wass, D. F., White, A. J. P., Williams, D. J., Iron complexes bearing iminopyridine and aminopyridine ligands as catalysts for atom transfer radical polymerisation. J. Chem. Soc. Dalton, 2003, 2824-2830

52.

53. 54. 55.

56. 57.

58.

59.

60.

61.

62.

63.

58

8 References

64.

OReilly, R. K., Gibson, V. C., White, A. J. P., Williams, D. J., Design of Highly Active Iron-Based Catalysts for Atom Transfer Radical Polymerization: Tridentate Salicylaldiminato Ligands Affording near Ideal Nernstian Behavior. J. Am. Chem. Soc., 2003, 125, 8450-8451 O'Reilly, R. K., Gibson, V. C., White, A. J. P., Williams, D. J., Five-coordinate iron(II) complexes bearing tridentate nitrogen donor ligands as catalysts for atom transfer radical polymerisation. Polyhedron, 2004, 23, 2921-2928 OReilly, R. K., Shaver, M. P., Gibson, V. C., White, A. J. P., -Diimine, Diamine, and Diphosphine Iron Catalysts for the Controlled Radical Polymerization of Styrene and Acrylate Monomers. Macromolecules, 2007, 40, 7441-7452 Ando, T., Kamigaito, M., Sawamoto, M., Catalytic Activities of Ruthenium(II) Complexes in Transition-Metal-Mediated Living Radical Polymerization: Polymerization, Model Reaction, and Cyclic Voltammetry. Macromolecules, 2000, 33, 5825-5829 Braunecker, W., Brown, W. C., Morelli, B. C., Tang, W., Poli, R., Matyjaszewski, K., Origin of activity in Cu-, Ru-, and Os-mediated radical polymerization. Macromolecules, 2007, 40, 8576-8585 Gibson, V. C., OReilly, R. K., Wass, D. F., White, A. J. P., Williams, D. J., Polymerization of Methyl Methacrylate Using Four-Coordinate (-Diimine)iron Catalysts: Atom Transfer Radical Polymerization vs Catalytic Chain Transfer. Macromolecules, 2003, 36, 2591-2593 Tsarevsky, N. V., Pintauer, T., Matyjaszewski, K., Deactivation Efficiency and Degree of Control over Polymerization in ATRP in Protic Solvents. Macromolecules, 2004, 37, 97689778 Pintauer, T., McKenzie, B., Matyjaszewski, K., Determination of the equilibrium constant for bromide dissociation from ATRP active Cu(II) complexes and its effect on the overall catalyst activity. Polym. Preprints, 2002, 43, 217-218 Eberson, L., Electron transfer reactions in organic chemistry II. An analysis of alkyl halide reduction by electron transfer reagents on the basis of the Marcus theory. Acta Chem. Scand., 1982, B 36, 533-543 Saveant, J. M., A simple model for the kinetics of dissociative electron transfer in polar solvents. Application to the homogeneous and heterogeneous reduction of alkyl halides. J. Am. Chem. Soc., 1987, 109, 6788-6795 Zhang, M., Ray, W. H., Modeling of living free-radical polymerization processes. I. Batch, semibatch, and continuous tank reactors. J. Appl. Polym. Sci., 2002, 86, 1630-1662 Kim, J.-B., Huang, W., Miller, M. D., Baker, G. L., Bruening, M. L., Kinetics of SurfaceInitiated Atom Transfer Radical Polymerization. J. Polym. Sci. A Pol. Chem., 2003, 41, 386394 Al-Harthi, M., Soares, J. B. P., Simon, L. C., Dynamic Monte Carlo Simulation of AtomTransfer Radical Polymerization. Macromol. Mater. Eng., 2006, 291, 993-1003

65.

66.

67.

68.

69.

70.

71.

72.

73.

74. 75.

76.

59

8 References

77. 78.

Kwark, Y.-J., Novak, B. M., Determination of the Kinetic Parameters of Atom Transfer Radical Polymerizations. Macromolecules, 2004, 37, 9395-9401 Al-Harthi, M., Soares, J. B. P., Simon, L. C., Modeling of Atom Transfer Radical Polymerization with Bifunctional Initiators: Diffusion Effects and Case Studies. Macromol. Chem. Phys., 2006, 207, 469-483 Al-Harthi, M., Cheng, L. S., Soares, J. B. P., Simon, L. C., Atom-transfer radical polymerization of styrene with bifunctional and monofunctional initiators: Experimental and mathematical modeling results. J. Polym. Sci. A Pol. Chem., 2007, 45, 2212-2224 Al-Harthi, M., Soares, J. B. P., Simon, L. C., Mathematical Modeling of Atom-Transfer Radical Polymerization Using Bifunctional Initiators. Macromol. Theory Simul., 2006, 15, 198-214 Gao, X., Feng, W., Zhu, S., Sheardown, H., Brash, J. L., Kinetic Modeling of SurfaceInitiated Atom Transfer Radical Polymerization. Macromol. React. Eng., 2010, 4, 235-250 Johnston-Hall, G., Monteiro, M. J., Kinetic Simulations of Atom Transfer Radical Polymerization (ATRP) in Light of Chain Length Dependent Termination. Macromol. Theory Simul., 2010, 19, 387-393 Shipp, D. A., Matyjaszewski, K., Kinetic Analysis of Controlled/Living Radical Polymerizations by Simulations. 1. The Importance of Diffusion-Controlled Reactions. Macromolecules, 1999, 32, 2948-2955 Tang, W., Matyjaszewski, K., Kinetic Modeling of Normal ATRP, Normal ATRP with [CuII]0, Reverse ATRP and SR&NI ATRP. Macromol. Theory Simul., 2008, 17, 359-375, and references therein Wang, X.-S., Armes, S. P., Facile Atom Transfer Radical Polymerization of Methoxy-Capped Oligo(ethylene glycol) Methacrylate in Aqueous Media at Ambient Temperature. Macromolecules, 2000, 33, 6640-6647 Kamigaito, M., Ando, T., Sawamoto, M., Metal-Catalyzed Living Radical Polymerization. Chem. Rev., 2001, 101, 3689-3745 Matyjaszewski, K., Xia, J., Atom Transfer Radical Polymerization. Chem. Rev., 2001, 101, 2921-2990 Matyjaszewski, K., Nakagawa, Y., Jasieczek, C. B., Polymerization of n-Butyl Acrylate by Atom Transfer Radical Polymerization. Remarkable Effect of Ethylene Carbonate and Other Solvents. Macromolecules, 1998, 31, 1535-1541 Gobelt, B., Matyjaszewski, K., Diimino- and diaminopyridine complexes of CuBr and FeBr2 as catalysts in atom transfer radical polymerization (ATRP). Macromol. Chem. Phys., 2000, 201, 1619-1624

79.

80.

81. 82.

83.

84.

85.

86. 87. 88.

89.

60

8 References

90.

Matyjaszewski, K., Shipp, D. A., Wang, J.-L., Grimaud, T., Patten, T. E., Utilizing Halide Exchange To Improve Control of Atom Transfer Radical Polymerization. Macromolecules, 1998, 31, 6836-6840 Perrier, S., Haddleton, D. M., Effect of water on copper mediated living radical polymerisation. Macromol. Symp., 2002, 182, 261-272 Jewrajka, S. K., Mandal, B. M., Living Radical Polymerization. 1. The Case of Atom Transfer Radical Polymerization of Acrylamide in Aqueous-Based Medium. Macromolecules, 2003, 36, 311-317 Xia, J., Matyjaszewski, K., Controlled/Living Radical Polymerization. Atom Transfer Radical Polymerization Using Multidentate Amine Ligands. Macromolecules, 1997, 30, 7697-7700 Xia, J., Zhang, X., Matyjaszewski, K., Atom Transfer Radical Polymerization of 4Vinylpyridine. Macromolecules, 1999, 32, 3531-3533 Robinson, K. L., Khan, M. A., de Paz Banez, M. V., Wang, X. S., Armes, S. P., Controlled Polymerization of 2-Hydroxyethyl Methacrylate by ATRP at Ambient Temperature. Macromolecules, 2001, 34, 3155-3158 Masci, G., Giacomelli, L., Crescenzi, V., Atom Transfer Radical Polymerization of NIsopropylacrylamide. Macromol. Rapid Commun., 2004, 25, 559-564 Percec, V., Guliashvili, T., Ladislaw, J. S., Wistrand, A., Stjerndahl, A., Sienkowska, M. J., Monteiro, M. J., Sahoo, S., Ultrafast Synthesis of Ultrahigh Molar Mass Polymers by Metal-Catalyzed Living Radical Polymerization of Acrylates, Methacrylates, and Vinyl Chloride Mediated by SET at 25 C. J. Am. Chem. Soc., 2006, 128, 14156-14165 Wang, J.-S., Matyjaszewski, K., "Living"/Controlled Radical Polymerization. TransitionMetal-Catalyzed Atom Transfer Radical Polymerization in the Presence of a Conventional Radical Initiator. Macromolecules, 1995, 28, 7572-7573 Gromada, J., Matyjaszewski, K., Simultaneous Reverse and Normal Initiation in Atom Transfer Radical Polymerization. Macromolecules, 2001, 34, 7664-7671

91. 92.

93.

94. 95.

96. 97.

98.

99.

100. Jakubowski, W., Matyjaszewski, K., Activator Generated by Electron Transfer for Atom Transfer Radical Polymerization. Macromolecules, 2005, 38, 4139-4146 101. Jakubowski, W., Min, K., Matyjaszewski, K., Activators Regenerated by Electron Transfer for Atom Transfer Radical Polymerization of Styrene. Macromolecules, 2006, 39, 39-45 102. Matyjaszewski, K., Jakubowski, W., Min, K., Tang, W., Huang, J., Braunecker, W. A., Tsarevsky, N. V., Diminishing catalyst concentration in atom transfer radical polymerization with reducing agents. PNAS, 2006, 103, 15309-15314 103. Fristrup, C. J., Jankova, K., Hvilsted, S., Surface-initiated atom transfer radical polymerization - a technique to develop biofunctional coatings. Soft Matter, 2009, 5, 46234634

61

8 References

104. Edmondson, S., Armes, S. P., Synthesis of surface-initiated polymer brushes using macroinitiators. Polym. Int., 2009, 58, 307-316 105. Liu, P., Modification of Polymeric Materials via Surface-Initiated Controlled/Living Radical Polymerization. e-Polymers, 2007, 106. Barbey, R., Lavanant, L., Paripovic, D., Schwer, N., Sugnaux, C., Tugulu, S., Klok, H.A., Polymer Brushes via Surface-Initiated Controlled Radical Polymerization: Synthesis, Characterization, Properties, and Applications. Chem. Rev., 2009, 109, 5437-5527 107. Ciampolini, M., Nardi, N., Five-Coordinated High-Spin Complexes of Bivalent Cobalt, Nickel, and Copper with Tris(2 - dimethylaminoethyl)amine. Inorg. Chem., 1966, 5, 41-44 108. Mendes, P., GEPASI: A software package for modelling the dynamics, steady states and control of biochemical and other systems. Comput. Appl. Biosci., 1993, 9, 563-571 109. Mendes, P., Biochemistry by numbers: simulation of biochemical pathways with Gepasi 3. Trends Biochem. Sci., 1997, 22, 361-363 110. Mendes, P., Kell, D. B., Non-linear optimization of biochemical pathways: applications to metabolic engineering and parameter estimation. Bioinformatics, 1998, 14, 869-883 111. Odian, G. G. ed., Principles of polymerization. 3rd ed., 1991, Wiley-Interscience, New York 112. Marcus, Y., The properties of organic liquids that are relevant to their use as solvating solvents. Chem. Soc. Rev., 1993, 22, 409-416 and references therein 113. Smith, G. B., Russell, G. T., Heuts, J. P. A., Termination in Dilute-Solution Free-Radical Polymerization: A Composite Model. Macromol. Theor. Simul., 2003, 12, 299-314 114. Johnston-Hall, G., Stenzel, M. H., Davis, T. P., Barner-Kowollik, C., Monteiro, M. J., Chain Length Dependent Termination Rate Coefficients of Methyl Methacrylate (MMA) in the Gel Regime: Accessing kti,i Using Reversible Addition-Fragmentation Chain Transfer (RAFT) Polymerization. Macromolecules, 2007, 40, 2730-2736 115. Johnston-Hall, G., Monteiro, M. J., Bimolecular Radical Termination: New Perspectives and Insights. J. Polym. Sci. A Pol. Chem., 2008, 46, 3155-3173 116. Griffiths, M. C., Strauch, J., Monteiro, M. J., Gilbert, R. G., Measurement of Diffusion Coefficients of Oligomeric Penetrants in Rubbery Polymer Matrixes. Macromolecules, 1998, 31, 7835-7844 117. Deng, K., Zhang, Y., Liu, Y., Shi, Z., Wang, Y., Atom transfer radical polymerization of styrene catalyzed by iron(II) chloride/EDTA. Chemical Journal on Internet, 2002, 4, 59 118. Matyjaszewski, K., Patten, T. E., Xia, J., Controlled/Living Radical Polymerization. Kinetics of the Homogeneous Atom Transfer Radical Polymerization of Styrene. J. Am. Chem. Soc., 1997, 119, 674-680

62

8 References

119. Wang, J.-L., Grimaud, T., Matyjaszewski, K., Kinetic Study of the Homogeneous Atom Transfer Radical Polymerization of Methyl Methacrylate. Macromolecules, 1997, 30, 65076512 120. Haddleton, D. M., Kukulj, D., Duncalf, D. J., Heming, A. M., Shooter, A. J., LowTemperature Living Radical Polymerization (Atom Transfer Polymerization) of Methyl Methacrylate Mediated by Copper(I) N-Alkyl-2-Pyridylmethanimine Complexes. Macromolecules, 1998, 31, 5201-5205 121. Mittal, A., Sivaram, S., A Novel Tridentate Nitrogen Donor as Ligand in Copper Catalyzed ATRP of Methyl Methacrylate. J. Polym. Sci. A Pol. Chem., 2005, 43, 4996-5008 122. Zhu, S., Yan, D., Zhang, G., Reverse Atom Transfer Radical Polymerization of Methyl Methacrylate with a New Catalytic System, FeCl3/Isophthalic Acid. J. Polym. Sci. A Pol. Chem., 2001, 39, 765-774 123. Huang, J., Pintauer, T., Matyjaszewski, K., Effect of Variation of [PMDETA]0/[Cu(I)Br]0 Ratio on Atom Transfer Radical Polymerization of n-Butyl Acrylate. J. Polym. Sci. A Pol. Chem., 2004, 42, 3285-3292 124. Walker, J. D., Poli, R., FeCl3-Phosphine Adducts with Trigonal-Bipyramidal Geometry. Influence of the Phosphine on the Spin State. Inorg. Chem., 1989, 28, 1793-1801 125. Calderazzo, F., Englert, U., Pampaloni, G., Vanni, E., Tetra-, penta-, and hexacoordination in iron(II) adducts of aliphatic bidentate amines. Crystal and molecular structure of the mononuclear complexes Fe(dienMe)Cl2 and Fe(tmeda)I2. C. R. Acad. Sci. II C, 1999, 2, 311319 126. Seewald, O., Flrke, U., Henkel, G., Dichlorobis(triphenylphosphine)iron(II). Acta Cryst. E, 2005, 61, m1829-m1830 127. Kim, Y. J., Park, C. R., Analysis of Problematic Complexing Behavior of Ferric Chloride with N,N-Dimethylformamide Using Combined Techniques of FT-IR, XPS, and TGA/DTG. Inorg. Chem., 2002, 41, 6211-6216 128. Yilmaz, V. T., Topcu, Y., Preparation, characterization and thermal reactivity of N,Ndimethylformamide complexes of some transition metal chlorides. Thermochim. Acta, 1997, 307, 143-147 129. Escot, M. T., Pouillen, P., Martinet, P., Electrochemical study of the iron(III) complexation with organic diacids in N,N-dimethylformamide. C. R. Acad. Sci. II, 1981, 292, 665-668 130. Escot, M.-T., Martre, A.-M., Pouillen, P., Martinet, P., Electrochemical study of iron(III) complexation by some model ligands of biological interest. I. Acetohydroxamic acid, acetylacetone and citric acid. Bull. Soc. Chem. Fr., 1989, 3, 316-320 131. Dass, N. N., George, M. H., Effects of iron(III) chloro complexes on the polymerization of styrene. J. Polym. Sci. A Pol. Chem., 1969, 7, 269-281

63

8 References

132. Chetia, P. D., Dass, N. N., Effects of tris (phenanthroline)-iron(III) complex on the polymerization of styrene. Eur. Polym. J., 1976, 12, 165-168 133. Shaver, M. P., Allan, L. E. N., Rzepa, H. S., Gibson, V. C., Correlation of Metal Spin State with Catalytic Reactivity: Polymerizations Mediated by -DiimineIron Complexes. Angew. Chem. Int. Edit., 2006, 45, 1241-1244 134. Allan, L. E. N., Shaver, M. P., White, A. J. P., Gibson, V. C., Correlation of Metal SpinState in -Diimine Iron Catalysts with Polymerization Mechanism. Inorg. Chem., 2007, 46, 8963-8970 135. Nanda, A. K., Hong, S. C., Matyjaszewski, K., Concurrent Initiation by Air in the Atom Transfer Radical Polymerization of Methyl Methacrylate. Macromol. Chem. Phys., 2003, 207, 1151-1159 136. Becer, C. R., Hoogenboom, R., Fournier, D., Schubert, U. S., Cu(II)-Mediated ATRP of MMA by Using a Novel Tetradentate Amine Ligand with Oligo(ethylene glycol) Pendant Groups. Macromol. Rapid Commun., 2007, 28, 1161-1166 137. Xue, Z., Oh, H. S., Noh, S. K., Lyoo, W. S., Phosphorus Ligands for Iron(III)-Mediated Atom Transfer Radical Polymerization of Methyl Methacrylate. Macromol, Rapid Commun, 2008, 29, 1887-1894 138. Xue, Z., Linh, N. T. B., Noh, S. K., Lyoo, W. S., Phosphorus-Containing Ligands for Iron(III)-Catalyzed Atom Transfer Radical Polymerization. Angew. Chem. Int. Edit., 2008, 47, 6426-6429

64

Vous aimerez peut-être aussi