Accepted Manuscript: Progress in Polymer Science

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 92

Accepted Manuscript

Title: Electrochemically Mediated Atom Transfer Radical


Polymerization (eATRP)

Authors: Paweł Chmielarz, Marco Fantin, Sangwoo Park,


Abdirisak A. Isse, Armando Gennaro, Andrew J.D. Magenau,
Andrzej Sobkowiak, Krzysztof Matyjaszewski

PII: S0079-6700(16)30109-5
DOI: http://dx.doi.org/doi:10.1016/j.progpolymsci.2017.02.005
Reference: JPPS 1018

To appear in: Progress in Polymer Science

Received date: 29-11-2016


Revised date: 7-2-2017
Accepted date: 22-2-2017

Please cite this article as: Chmielarz Paweł, Fantin Marco, Park Sangwoo,
Isse Abdirisak A, Gennaro Armando, Magenau Andrew JD, Sobkowiak
Andrzej, Matyjaszewski Krzysztof.Electrochemically Mediated Atom
Transfer Radical Polymerization (eATRP).Progress in Polymer Science
http://dx.doi.org/10.1016/j.progpolymsci.2017.02.005

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Electrochemically Mediated Atom Transfer
Radical Polymerization (eATRP)
Paweł Chmielarz1,2* p_chmiel@prz.edu.pl , Marco Fantin2,3, Sangwoo Park2, Abdirisak A.Isse3*

abdirisak.ahmedisse@unipd.it , Armando Gennaro3, Andrew J.D.Magenau4, Andrzej Sobkowiak1,


Krzysztof Matyjaszewski2* km3b@andrew.cmu.edu

1
Department of Physical Chemistry, Faculty of Chemistry, Rzeszow University of Technology, Al. Powstańców Warszawy 6, 35-
959 Rzeszow, Poland
2
Department of Chemistry, Carnegie Mellon University, Pittsburgh, PA 15213, USA
3
Department of Chemical Sciences, University of Padova, via Marzolo 1, 35131 Padova, Italy
4
Department of Materials Science and Engineering, Drexel University, Philadelphia PA 19104

*
Corresponding authors

Abstract

This review covers both fundamental aspects and applications of electrochemically mediated atom transfer
radical polymerization (eATRP). eATRP setup is discussed in detail, together with the advantages and
limitations of this technique. All relevant parameters that can influence eATRP outcome are evaluated (e.g.
applied current and potential, stirring and diffusion, solvents and supporting electrolytes). Various materials
prepared by eATRP are described, including homopolymers, block copolymers, star polymers, and surface
grafted polymer brushes. In addition, other electrochemical techniques conceptually similar to eATRP are
discussed, including copper-catalyzed azide-alkyne cycloaddition, electrochemical micropatterning, reversible
addition-fragmentation chain transfer polymerization using redox-sensitive initiators, and catalyst removal by
electrochemical reduction. The increasing research activity in the last decade indicates that electrochemically
regulated methods are becoming valuable tools in the design and synthesis of advanced polymer materials.

1
Keywords: Electrochemically mediated ATRP, stimulus operated polymerization, controlled/living

polymerization, reversible deactivation radical polymerization, electropolymerization, electrochemistry,

polymer synthesis, click chemistry, on-demand polymerization

Nomenclature

AFM atomic force microscopy


AM acrylamide
ARGET activators regenerated by electron transfer
ATRP atom transfer radical polymerization
BA n-butyl acrylate
-CD -cyclodextrin
BPE bipolar electrode
BPY 2,2’-bipyridine
CC click chemistry
CFs carbon fibers
CE counter electrode
CEC click electrochemistry
CoPc cobalt phthalocyanine
CuAAC CuI-catalyzed azide–alkyne cycloaddition
CuII(OTf)2 copper(II) triflate
CV cyclic voltammetry
Ð molecular weight distribution (Mw/Mn)
DMF dimethylformamide
DMSO dimethyl sulfoxide
DP degree of polymerization
DPV differential pulse voltammetry
DVB divinylbenzene
Eo standard potential
E1/2 half-wave potential
Eapp applied potential
eATRP electrochemically mediated ATRP
Epa anodic peak potential
Epc cathodic peak potential
EBiB ethyl 2-bromoisobutyrate
EC′ electrochemical catalytic cycle
ET electron transfer
F Faraday’s constant (96485 C/mol)
Go standard Gibbs free energy change in a chemical process
2
GC glassy carbon
GCE glassy carbon electrode
GMA glycidyl methacrylate
GPC gel permeation chromatography
Hb hemoglobin
HEBiB 2-hydroxyethyl 2-bromoisobutyrate
I (direct) current
Iapp
ICAR applied current initiation for continuous activators regeneration
ka activation rate constant
kpapp apparent propagation rate constant
KATRP ATRP equilibrium constant (KATRP = ka/kda)
kda deactivation rate constant
kp propagation rate constant
kred reduction rate constant
kt termination rate constant
MeCN acetonitrile
Me6TREN tris(2-(dimethylamino)ethyl)amine
MBA N,N′-methylene-bis-acrylamide
MI macroinitiator
MIP molecularly imprinted polymer
MMA methyl methacrylate
Mn number-average molecular weight
Mw weight-average molecular weight
MW molecular weight
MWD molecular weight dispersity
n number of electrons
η overpotential (η = Eapp – E1/2)
NIPAM N-isopropylacrylamide
OEGA oligo (ethylene glycol) acrylate
OEGMA oligo(ethylene oxide) monomethyl ether methacrylate
OMC ordered mesoporous carbons
p monomer conversion
PANI polyaniline
PANI-N3 poly(4-azidoaniline)
PAM polyacrylamide
PBA poly(n-butyl acrylate)
PBMA poly(n-butyl methacrylate)
PBS phosphate-buffer saline
PDMAEMA poly(2-(dimethylamino)ethyl methacrylate)
PEDOT poly(3,4-ethylenedioxythiophene)
PEDOT-N3 poly(3,4-(1-azidomethylethylene) dioxythiophene)
3
PEO poly(ethylene oxide)
PGMA poly(glycidyl methacrylate)
PHEMA poly(2-hydroxyethyl methacrylate)
PMA poly(methyl acrylate)
PMAA poly(methacrylic acid)
PMBA poly(N,N′-methylene-bis-acrylamide)
PMDETA N,N,N’,N’,N’’-pentamethyldiethylenetriamine
PMETAC poly(2-(methacryloyloxy)ethyl trimethylammonium chloride)
PMMA poly(methyl methacrylate)
PNIPAM poly(N-isopropylacrylamide)
POEGMA poly(oligo(ethylene oxide) monomethyl ether methacrylate)
POEGA poly(oligo (ethylene glycol) acrylate)
PSBMA poly(sulfobetaine methacrylate)
PSPMA poly(potassium 3-sulfopropyl methacrylate)
PtBA poly(tert-butyl acrylate)
Q total passed charge
RAFT reversible addition-fragmentation chain transfer
RE reference electrode
Rp rate of polymerization
SARA supplemental activator and reducing agent
seATRP simplified electrochemically mediated ATRP
SCE saturated calomel electrode
SECM scanning electrochemical microscopy
TA terminal-alkynyl
tBA tert-butyl acrylate
TBABr tetrabutylammonium bromide
TBAP tetrabutylammonium perchlorate
TPMA tris(2-pyridylmethyl)amine
TPMM tris(2,4,6-methoxyphenyl)phosphine
WE working electrode

4
1. Introduction

Atom transfer radical polymerization (ATRP) is one of the most rapidly developing areas of polymer

science. It allows control over molecular weight, preparation of polymers with narrow molecular weight

distributions, incorporation of precisely placed functionalities, fabrication of various architectures, and synthesis

of well-defined hybrid composites [1-26]._ENREF_9_ENREF_2_ENREF_2_ENREF_3 ATRP can control

polymer topology, ranging from linear chains with precisely controlled dimensions and dispersity to various

branched structures [18, 21, 22]. These complex architectures are formed by copolymerizing macromonomers,

grafting to and from functionalized backbones, or polymerizing inimers and multi-functional initiators [22, 27-

29]. ATRP is also used to precisely control a polymer’s composition to prepare segmented, block, graft,

periodic, and gradient polymer microstructures [22]. The list of monomers successfully homopolymerized by

ATRP includes various styrenes, (meth)acrylates, (meth)acrylamides, N-vinylpyrrolidone, acrylonitrile, vinyl

acetate, as well as vinyl chloride [2, 10, 22, 30-44] Moreover, ATRP can be combined with other

polymerization techniques such as anionic, cationic, ring-opening metathesis, ionic ring-opening, free-radical

polymerization and step-growth polymerizations to form a vast array of compositionally diverse block

copolymers [7, 45-50].

With the advent of methodologies to (re)generate active polymerization catalysts, ATRP can now be

conducted at parts per million (ppm) levels of transition metal using activators regenerated by electron transfer

(ARGET) [51-71], initiators for continuous activator regeneration (ICAR) [51, 72-78], supplemental activator

and reducing agent (SARA) [79-97],_ENREF_21 photoinduced (photo-ATRP) [64, 98-118] and

electrochemically mediated ATRP (eATRP) [94, 119-137]. Each of these ATRP methods offers a more

environmentally benign and industrially relevant alternative for synthesizing polymers compared to normal

ATRP, while having the added benefit of tailoring the polymer’s dispersity through catalyst loading [22, 138].

The concept behind eATRP is that the ratio of activator to deactivator catalyst is precisely controlled by

an electrochemical redox process at an electrode surface, thereby rendering the eATRP process operational

5
through an external stimulus. Several parameters, such as the applied current (Iapp), applied potential (Eapp), and

total charge passed (Q) can be defined in eATRP to allow selection of the desired concentration of redox-active

catalytic species. This enhances the level of control during polymerization [119, 122, 126]. eATRP can be

stopped and (re)initiated on demand by modulating the applied potential (Eapp).

eATRP has shown versatility in synthesizing various soft materials with well-defined polymeric

architectures [120, 121, 123-125, 130, 133, 134, 136, 137, 139-143], although some limitations remain,

especially associated with eATRP’s cumbersome reaction setup. Typically, eATRP was conducted by utilizing

a three-electrode system under potentiostatic conditions, i.e. at constant applied potential. Recently, eATRP

setup has been simplified through several modifications: 1) by using a sacrificial counter electrode (CE)

inserted directly into the reaction mixture, which does not require any separation of electrode compartments, 2)

by carrying out the polymerization under galvanostatic conditions (i.e. constant current) without requirement of

a reference electrode, and 3) by the possibility of selecting less expensive conductive materials as working

electrodes compared to precious and rare metal electrodes like platinum, with little or no impact on the

polymerization outcome [126, 129, 130, 132, 133, 136, 137, 143].

The present review will survey recent advances in eATRP in relation to the synthesis of materials with

different architecture and composition. The general concept of this method will be overviewed, followed by

discussion of mechanism, apparatus, advantages and limitations.

2. Fundamentals of eATRP

2.1. Mechanism of ATRP

Copper-mediated ATRP is controlled by an equilibrium (KATRP) between active propagating species at a

low concentration and dormant polymer chains at a much higher concentration. These dormant species

periodically react with transition metal complexes in their lower oxidation state (e.g., Cu IL), through an inner

sphere electron transfer (ISET) [144], which allows to reduce the C-X bond in the dormant species, at rates

6
proportional to the rate constant of activation (ka). PnX activation generates growing radicals (Pn•) and the

transition metal complex in its higher oxidation state with a coordinated halide ligand (XCuIIL) (Scheme 1)

[22, 145]. In the case of an outer sphere electron transfer (OSET), such radicals are easier to reduce than the

dormant species [146, 147], however, the ideal ATRP catalyst does not reduce radicals to the corresponding

carbanions. Then, in the deactivation reaction, the ternary complex XCuIIL reacts with the propagating radicals

resulting in reformation of the dormant species and CuIL.

Despite Cu has been the most used metal in ATRP, this catalytic process can be mediated by complexes

originating from other transition metals including Fe, Ru, Ni, Ti, Re, Mo, Co, and Os [1, 8, 15, 16, 148-150].

The propagating radicals (Pn•), intermittently generated by the fast activation/deactivation process, add to the

monomer with rates proportional to the propagation rate constant (kp). Termination reactions also occur in

ATRP, primarily through radical coupling (PnPm) or disproportionation. In the latter case, two terminated

chains are formed through a hydrogen abstraction reaction; of the two terminated chains, one has a double bond

and the other a saturated chain-end. In a well-controlled ATRP, only a small fraction of polymer chains undergo

termination, which is diminished by the low concentration of active propagating radicals (and corresponding

higher concentration of dormant species).

The rate of ATRP (Rp) depends on kp the concentrations of monomer ([M]), and the growing radical

([Pn•]). In turn, the radical concentration depends on the ATRP equilibrium constant (KATRP = ka/kda) which is

defined by the concentrations of dormant species ([PnX]), activators ([CuIL]), and deactivators ([XCuIIL]),

according to eq. 1 [21]. To minimize radical concentrations, tsshe equilibrium constant is usually small (KATRP

<< 1).

[Pn - X][CuI L][M]


Rp  kp [M][Pn ]  kp K ATRP ((1)
[X - Cu II L]

The structure of ligands, monomers, dormant species, as well as reaction conditions (e.g. solvent,

temperature and pressure) strongly influence the values of KATRP and rate constants ka and kda [22, 25, 151-153].

7
In the typical case of a neutral amine ligand, both Cu I and CuII complexes bear a positive charge. ATRP rates

increase with catalyst activity (i.e. higher KATRP), but under some conditions may decrease due to excessive

radical termination reactions, causing a low [CuIL]/[XCuIIL] ratio to be established by accumulated XCuIIL

via the persistent radical effect [154, 155]. Polymers with narrow molecular weight distributions (Ð or Mw/Mn)

are obtained by ATRP, whereby more uniform polymer chains are provided by higher concentrations of

deactivator ([XCuIIL]) and by higher monomer conversions (p), as defined by eq. 2. [22].

Mw  kp [Pn - X]  2 
 1   II
  1 ((2)
Mn  kda [X - Cu L]  p 

2.2. From Normal ATRP to ATRP with Diminished Catalyst Concentrations

One disadvantage of “early” ATRP procedures was associated with the relatively high concentration of

catalyst often equimolar to the initiator [21, 22]. This high concentration of catalyst was required to overcome

both radical termination events and the low activity of early catalysts [80, 156]. As a consequence, extensive

purification was required necessitating polymer precipitations, columns, ion-exchange resins, or use of

heterogeneous catalysts that could be isolated after polymerization was complete [157-

167]._ENREF_91_ENREF_92_ENREF_93 In recent years, various activator (e.g. CuIL) (re)generation

methods have been developed, which allowed ATRP to be conducted with low concentrations of catalyst,

typically at or below 100 ppm [51, 64, 76, 168, 169]. In these low ppm ATRP systems, the polymerization

usually begins from the oxidized catalyst complex (deactivator) as illustrated in Scheme 2. The activator (CuIL)

is then generated by reduction of the initially present deactivator (XCuIIL), for instance, by addition of a

reducing agent. Reducing agents can continuously regenerate CuIL from XCuIIL, which would otherwise

accumulate in solution by the occurrence of radical termination reactions [42], as indicated in Scheme 2. The

reducing/reactivating cycles were also employed to eliminate oxygen from the system [22].

8
Reduction of XCuIIL complexes can occur by various methods. In ARGET ATRP, chemical reducing

agents such as organo-tin compounds, glucose, ascorbic acid, hydrazine, sulfites, or even functional monomers

are used [51, 170-172]._ENREF_94_ENREF_35_ENREF_95 In ICAR ATRP, thermal initiators, e.g.,

azobisisobutyronitrile (AIBN), are used to regenerate CuIL from XCuIIL [76]. In SARA ATRP, zerovalent

metals (Cu0, Fe0 or Ag0), act as supplemental activators and mild reducing agents [79, 173]. Instead, in

nonchemical methods, such as eATRP or photo-ATRP, a proper potential (current) or electromagnetic radiation

is applied to adjust the ratio between CuIL and XCuIIL during the reaction [119].

ARGET, ICAR, and SARA ATRP processes can generate undesirable byproducts. In ARGET ATRP,

products of the oxidized reducing agents may be toxic, e.g., dehydroascorbic acid or SnIV compounds, and may

cause unwanted color changes in the synthesized polymers [174]. In ICAR ATRP, products also contain a

fraction of polymer chains initiated by the added free radical initiator, rather than solely from the ATRP

initiator, which are considered an impurity in certain applications [51]. In SARA ATRP, copper concentrations

build up throughout the polymerization [91]. In contrast, eATRP is uniquely capable of eliminating reaction

byproducts by the direct insertion or removal of electrons through a heterogeneous electron transfer (ET)

process; it produces cleaner materials through a more environmentally friendly process, at the expense of

supporting electrolyte impurities. Moreover, eATRP is less sensitive than other low ppm techniques to changes

in the polymerization conditions because it relies on an external stimulus (Iapp, Eapp, and Q). Another newly

developed technique that relies on an external stimulus is photo-ATRP. This technique, however, differently

from eATRP requires the presence of excess amine ligand or additionally added amines.

2.3. Mechanism of eATRP

Many reactions in synthetic chemistry are based on the use of redox-active metal catalysts as part of the

initiation system; one example is cerium-based oxidation of alcohols, obtained through the controlled formation

of radicals [175]. Moreover, the use of electrochemistry to provide radicals and initiate polymerizations on the

9
electrode surface is established in the literature [175-178]. Electropolymerizations have been extensively used

to generate thin films [179] and/or hydrogel layers on working electrode (WE) surfaces [180, 181], via free

radical or reversible addition-fragmentation chain transfer (RAFT) polymerization.

An initial report of electrochemically-triggered controlled polymerization occurring in the bulk of the

solution was reported in 2009 by Amatore and coworkers [182]. The ATRP active catalyst FeII/N,N’-bis(3,5-di-

tert-butylsalicylidene)-1,2-ethyldiimine (FeIISalen) was electrochemically generated by reduction of FeIIISalen.

The FeII catalyst activated alkyl or benzyl halide initiators (e.g. 1-chloro-1-phenylethane) which led to the

production of polymers in the presence of either styrene or methyl methacrylate. Notably, polystyrene having a

narrower molecular weight distribution (Ð = 1.43) was observed when a 1/1 molar ratio of FeIIISalen/FeIISalen

was electrochemically generated.

In copper-based eATRP, a desired amount of the catalytic complex (XCuIIL) is electrochemically

reduced to its active complex to trigger controlled radical polymerization [119]. Typically, the reaction mixture

initially contains solvent, monomer, initiator, supporting electrolyte and XCuIIL. In the absence of CuIL

activator, polymerization does not occur. The polymerization begins when an appropriate potential (Eapp) or

current is applied at the WE. Reduction of XCuIIL to XCuIL occurs at the WE surface, then the reduced

activator molecules are dispersed throughout the bulk reaction mixture by vigorous stirring. XCuIL dissociates

to give CuIL, which reacts with the initiator (e.g., alkyl halide, Pn−X) to form radicals and the deactivating

complex. The radical species propagate to form polymeric chains by reacting with the monomer (M), or are

deactivated back to dormant species (Pn−X) by XCuIIL (Scheme 2). Continuous (re)generation of the activator

and modulation of the rate of polymerization (Rp) can be achieved by adjusting Eapp (see Section 2.7.1).

An electrochemical catalytic process (EC’) governs the polymerization by electrochemically generating

CuIL through electron transfer, i.e. the electrochemical step (E); then, CuIL catalytically reacts with Pn−X, i.e.

the chemical step, and regenerates the original catalyst (C’) (Scheme 3) [122, 183-185]. The driving force of the

electron transfer depends on the difference between the applied potential and the standard potential of the
10
copper complex (Eo). The standard Gibbs free energy change can be calculated as rGo  nF ( Eapp  E1/2 ) ,

where n is the number of exchanged electrons, F is the Faraday constant, and E1/2 is the half-wave potential

measured from voltammetry, which is approximately equal to the standard potential Eo (see also eq. 8).

Since ATRP is generally carried out in the presence of halide ions, e.g. by introducing copper as CuIIX2,

only XCuIIL is present at the beginning of electrolysis; as a result, the initial cathodic current decays rapidly as

XCuIIL is converted to XCuIL (Figure 1). Afterward, the current approaches a low and constant value as the

system tends toward a fixed [XCuIIL]/[XCuIL] ratio determined by the selected Eapp (see eq. 15 for more

details). After reduction, the ternary XCuIL complex dissociates into X and CuIL, the latter being the active

ATRP catalyst which, participates in the reversible atom transfer reaction with alkyl halides. These reactions

represent a perturbation to the equilibrium [XCuIIL]/[XCuIL] ratio imposed by Eapp. Therefore, whether a

constant [XCuIIL]/[XCuIL] ratio can be achieved in the bulk solution depends on the mutual rates of electro-

generation and disappearance of XCuIL.

Even if the reduction of XCuIIL occurs only at the electrode surface, during the electrolysis the solution

is usually vigorously stirred and therefore the thickness of the Nernst diffusion layer is drastically reduced; as a

result, the catalyst concentration near the electrode surface is close to that of the bulk solution. The

polymerization continues in the bulk solution resulting in some accumulation of XCuIIL, e.g., by radical-

radical termination reactions. This slow regeneration of XCuIIL is mainly responsible of the low and almost

constant current observed after the initial stage in eATRP [120].

In conclusion, the main process contributing to the current decay in Figure 1 is the reduction of the

initially introduced XCuIIL. On the other hand, radical termination (occurring throughout the polymerization)

leads to the slow formation of XCuIIL, which is continuously reduced at the working electrode. As a result, the

current does not decay to zero as in the case of electrolysis without P nX, but tends toward a constant value

when the rate of bulk regeneration of XCuIIL matches its reduction rate at the electrode.

11
According to Scheme 2, rate equations describing the changes of [XCuIIL] and [Pn•] during

polymerization can be written as follows:

d [X - Cu II L] (3)
 ka [Pn - X][Cu I L] - kda [X - Cu II L][Pn ] - kred [X - Cu II L]
dt

d [Pn ] (4)
 ka [Pn - X][Cu I L] - kda [X - Cu II L][Pn ] - k t [Pn ]2
dt

After the initial stages of electrolysis, the system reaches a stationary state, in which [XCuIIL] and [Pn]

become approximately constant. In this steady-state condition, eqs. 3 and 4 can be combined to give the

following expression defining the concentration of radicals.


kred [X - Cu II L]
[P ] 
n (5)
kt

It follows from eq. 5 that the concentration of propagating species is proportional to the square root of the

product of [XCuIIL] and kred/kt. Therefore, Rp is also proportional to the square root of [XCuIIL] (eq. 1) and to

kred, which is defined by the applied potential during electrolysis. (More details on the effect of applied potential

on eATRP are presented in Section 2.7.1).

The charge passed during electrolysis was used to evaluate the amount of radical termination events, and

therefore, the fraction of living RX chain-ends. In general, in ATRP the loss of chain end functionality is

proportional to the amount of [CuIL] regenerated from [XCuIIL] originating from termination. In eATRP, the

total consumed charge can be used to precisely calculate the amount of reduced XCuIIL, because no other

electrochemical processes occur at the relatively positive Eapp (i.e. current efficiency is high). If the amount of

Cu species in the reaction medium is small, e.g., <1 mol% versus initiator, the fraction of terminated chains can

be calculated by integration of current versus time to determine the total charge passed [42, 186].

Recently, eATRP was successfully extended to iron-based systems, with the production of well-defined

poly(methyl methacrylate) (PMMA). An alkyl halide initiator was activated by electro-generated FeIIBr2/TMPP

12
(tris(2,4,6-methoxyphenyl)phosphine) catalysts [135]. Notably, PMMA with a narrow molecular weight

distribution (Ð = 1.39) was obtained when an equimolar ratio of FeIIIBr3/TMPP was used.

2.4. Electrochemical Reaction Setup

Figure 2 represents the three-electrode setup typically used for both the electrochemical characterization

of the catalytic system and the eATRP electrolysis process [119]. In eATRP, as in any other electrochemical

system using a three-electrode configuration, the applied potential is measured between WE and a reference

electrode (RE), while the electrical current flows between WE and a counter electrode (CE).

The eATRP reaction setup consists of a pear shaped five-neck jacketed flask equipped with a magnetic

stirring element, inert gas purge, and a condenser. The flask necks are fitted with a reference electrode (e.g.

Ag/Ag+ or Ag/AgX/X- for organics and saturated-calomel electrode for aqueous conditions) [187], a working

electrode (WE), a counter electrode (CE, e.g. Pt gauze or plate), an inlet/outlet port for degassing, and an

inlet/outlet port for samples and reagents. A condenser can be also attached to the system to avoid evaporation

of volatile monomers [126, 129, 130, 137, 188, 189]. The CE should be placed in a separate fritted compartment

to minimize the diffusion of oxidized species formed at the CE to the working solution. Moreover, a

methylcellulose gel, saturated with an adequate electrolyte, can be placed over the glass frit to further isolate the

anodic and cathodic compartments. For all analytical cyclic voltammetry measurements, a Pt or glassy carbon

(GC) disc can be used as the WE, while separation of the CE is not strictly necessary [126, 190].

2.5. Cyclic Voltammetry of Copper Complexes

Cyclic voltammetry has been widely used to measure the redox properties and activity of ATRP copper

catalysts [8, 94, 125, 129, 130, 133, 137, 189, 191-193]. In cyclic voltammetry, copper-amine complexes, CuIIL

or XCuIIL, exhibit a redox couple due to the following reversible electrochemical processes (Figure 3a):

CuIIL + e– CuIL (6)

13
XCuIIL + e– XCuIL (7)

The standard reduction potential of CuII can be estimated from cyclic voltammetry as the half-sum of

cathodic and anodic peak potentials, E1/2 = (Epc + Epa)/2. Indeed, E1/2 measured from CV is defined as

1
RT  DCu I 
2
E1/2  E o'  ln (8)
nF  D II 
 Cu 

 
where E o '  E o  RT nF ln  Cu II  Cu I is the formal potential,  is activity coefficient, n is the number of

exchanged electrons (n = 1 for ATRP catalysts), R is the universal gas constant, and DCu II and DCu I are the

diffusion coefficients of CuII and CuI complexes, respectively. Typically voltammetric analysis are performed in

dilute solutions at constant ionic strength (e.g. 0.1 M strong electrolyte), so the ratio of the activity coefficients,

 Cu II  Cu I , is approximately 1 and hence Eo′  Eo. Additionally, the two forms of the catalyst may be assumed

to have similar diffusion coefficients, DCu II DCu I  1 . These approximations lead to E1/2  Eo [191].

E1/2 values are strongly dependent on the nature of the ligand (Figure 3a) and type of halogen atom [194].

One of the earliest studies established the existence of a linear correlation between E1/2 and apparent KATRP

equilibrium constant (Keqapp = KATRP/[XCuIIL] Figure 3b), providing a straightforward method to estimate

catalytic activity using the XCuIIL/XCuIL redox couple [191].

The ATRP equilibrium, and thus the activity of an ATRP catalyst, can be formally expressed as a

combination of simpler equilibria, shown below, with the relevant thermodynamic parameter indicated within

the same line [155, 193]:

RX R + X GR
o
X (9)

Eo 
X +e -
X - X / X (10)

CuIIL + e- CuIL o
ECu II 2 
L / CuI L
(11)s

K XII
CuIIL + X- XCuIIL (12)
14
where, GR
o
X is the standard Gibbs free energy of RX bond dissociation and K XII is the equilibrium constant of

the association reaction between a halide ion and the binary CuIIL complex (halidophilicity).

For the case of a stable 1:1 metal:ligand complex, the standard redox potential of the CuIIL/CuIL couple

depends on the relative stability of CuIIL and CuIL complexes, according to:

RT β I
o
ECu II
L/Cu I L
 ECu
o
2
/Cu 
 ln (13)
nF β II

where Cu2+ and Cu+ stand for solvated copper ions,  II and  I are the stability constants of CuIIL and CuIL,

respectively, and n = 1. Thus, ligands forming very stable complexes with Cu2+ lead to high βII/βI ratios, high

KATRP values, and a strongly reducing CuIIL/CuIL couple (eq. 13).

The redox potential of ternary complexes, EXo -Cu IIL/X-Cu IL , in turn, depends on the relative affinity of higher

and lower oxidation states of the metal complex for halide ion determined by their respective equilibrium

II I
constants ( K X and K X ) [150]:

RT KI (14)
EXo -CuIIL / X-CuI L  ECu
o
II
L / CuI L
 ln XII
F KX

On the basis of these thermodynamic considerations one can assess the catalytic activity of metal complexes in

ATRP for a given RX, if βII/βI and KIX/KIIX are known. Typically, in an organic solvent,

o
EX -Cu II L / X-Cu I L
 ECu
o
II
L / Cu I L
and therefore K XII > K XI , which is a desirable result as stable XCuIIL complexes are

necessary for efficient deactivation. In contrast, a high K XI results in high XCuIL concentration, which slows

polymerization because the ternary XCuIL complex is inactive and must first dissociate to provide the active

binary CuIL complex [195]. When the same ligand was used, it was found that ClCuIIL complexes have a more

negative potential than BrCuIIL complexes [191], however, the Rp was lower with ClCuIIL than when

employing BrCuIIL. This behavior can be rationalized because: 1) the activity of the catalyst is directly linked

to the standard reduction potential of the binary CuIIL complexes and not the ternary XCuIIL complex, 2) RCl

15
bonds are stronger than RBr bonds, and 3) KATRP is lower for RCl than for RBr [191, 192, 196]. In

agreement with this trend, eATRP is expected to be even slower in the presence of C-F bonds, which are

significantly less reactive than C-Cl [197].

Cyclic voltammetry can be used as a screening method to select the appropriate ATRP catalysts. The

measured redox potentials of various copper complexes were correlated to the ATRP equilibrium constant

(KATRP = ka/kda). For example, catalyst activity was found to increase in the order 2,2’-bipyridine (BPY) <

N,N,N’,N’,N’’-pentamethyldiethylenetriamine (PMDETA) < tris(2-pyridylmethyl)amine (TPMA) < tris(2-

(dimethylamino)ethyl)amine (Me6TREN) [198]. Generally, the reducing power of CuIL activator complexes

with nitrogen-based ligands have been correlated to the denticity (4 > 3 > 2), nature of the nitrogen atoms (alkyl

amine ≈ pyridine > imine > aromatic amine), distance between nitrogen atoms (C2 > C3 > C4), ligand topology

(branched > cyclic > linear), and electronic substituent (donating > withdrawing) [152, 199, 200].

Cyclic voltammetry also provides useful information on the catalyst’s reactivity toward alkyl halides.

Addition of an alkyl halide initiator to a solution of Cu IIL during voltammetric measurements causes a loss of

reversibility and, in most cases, a concomitant increase of the cathodic current because of regenerated Cu IIL via

the previously described catalytic EC’ process (Figure 4). Further evidence of this is provided by the decrease in

the anodic current from electrochemically generated CuIL near the electrode surface being oxidized by its

reaction with RX [119, 122]. A clear trend is observed with cathodic current enhancement increasing with the

following order of increasing activity: PMDETA < TPMA < Me6TREN (Figure 4). These results show that

more active catalysts react at a faster rate with RX to form XCuIIL and R.

2.6. Electrochemical Determination of Activation and Deactivation Kinetics

Recently, electrochemical techniques were developed to accurately evaluate activation and deactivation

rate constants of various catalyst-initiator systems [201-206]. Three different techniques were used:

16
chronoamperometry at a rotating disk electrode, comparison of the degree of catalysis in CV to theoretical

working curves, and comparison between experimental and simulated CVs.

To determine ka, experiments were typically carried out in the presence of a stable radical, i.e. 2,2,6,6-

tetramethyl-1-piperidinyloxy (TEMPO), which effectively trapped radicals generated through the ATRP

activation step. In the presence of TEMPO, the ATRP activation reaction becomes irreversible, allowing to

extract reliable kinetic information from the activation step.s

_ENREF_72 Chronoamperometry at a Rotating Disk Electrode

Chronoamperometry on a rotating disk electrode (RDE) was employed to determine ka in various solvents

[195, 203, 205]. At a RDE, both binary Csu/L and ternary XCu/L complexes exhibit a well-defined wave in

linear sweep voltammetry (see example in Figure 5). These experiments were conducted starting with a solution

of CuIL or XCuIL (and TEMPO); then, the potential was set to a value in the plateau region for the oxidation

of CuI to CuII (e.g. 0.1 V vs. SCE) and the current was monitored as a function of time (first data points in

Figure 5b). The initiator is then injected into the cell and the current decay is monitored as a function of time

(Figure 5b). Since the current is proportional to the concentration of CuI, the data are elaborated, using directly

the current or after conversion to concentration. The experiment is performed under pseudo-first-order ([RX]

>> [CuIL]) or second-order conditions ([RX] = [CuIL]) depending on the rate of the activation reaction.

The limit of detection of this technique is around ka = 5 × 103 M-1 s-1, which is at least one order of

magnitude higher then values obtained using spectrophotometric or NMR methods. The same technique was

used also for the determination of disproportionation rate constants of CuIL complexes [193].

Comparison of the Current Enchancement in CV to Theoretical Working Curves

CV measurements of X–CuIIL were used to determine ka of very active ATRP systems in aqueous media,

exploiting the current enhancement measured in the presence of RX (Figure 6) [90]. First, the ratio between
17
the cathodic peak currents (Ip,c/Ip,d) for the reduction of XCuIIL in the absence and presence of an initiator RX

was recorded. The process was repeated for different scan rates and different RX concentrations (excess ratio γ

= [RX]/[XCuIIL]). Then, theoretical working curves relating Ip,c/Ip,d with a kinetic parameter λ = RTkaCCuII/Fv

were constructed by digital simulation of cyclic voltammetry under the same experimental conditions. Last, the

rate coefficient ka was determined by comparing the experimental Ip,c/Ip,d to the simulated curve. Exceptionally

high ka values of up to 2.7×107 M-1 s-1 were measured for the activation of HEBiB by CuIMe6TREN in water.

Simulation of the Complete CV Curve

Both ka and kda were determined by CV measurements and simulations of the whole voltammetric curves

[201, 204]. When CV of CuIIL is run in organic solvents in the presence of an alkyl halide initiator, a new

reduction peak appears at potentials more negative than the reversible peak of Cu IIL. This new peak, associated

to the reduction of XCuIIL, is catalytic and its intensity depend on ka. Also the reversible peak of CuIIL loses

(or partially loses) its anodic partner [206]. In this case, digital simulation of the CV pattern can be used to

obtain kinetic information. Indeed, simulation of the entire catalytic mechanism, under the exact experimental

conditions, provided a theoretical CV that was superimposed to the experimental one (Figure 7). This

comparison provided a direct measurement of the overall activation and possibly also deactivation rates. The

direct determination of very high ka and kda is an advantage of this technique. In particular, the accurate

measurement of kda is usually difficult, because the bimolecular deactivation reaction approaches diffusion-

controlled limits (kda ∼107−108 M−1 s−1) and the propagating radical (Pn•) cannot be isolated [207]. Using

simulation of the CV patterns of BrCuIIPMDETA/RBr in DMSO, both ka and kda were determined in the

same experiment [204].

18
2.7. eATRP Parameters

Applied Potential

When conducting polymerizations under potentiostatic conditions, the choice of the applied potential

(Eapp) is of primary importance. In particular, two different conditions may arise. The first case is when Eapp is

close to the E1/2 of the catalyst complex. It must be noted that in normal experimental conditions, X- ions are

present in solution, therefore the redox potential of reference is E1/2 of the XCuIIL/X–CuIL couple. In this

regime, the polymerization rate can be modulated by changing the overpotential, η (η = Eapp – E1/2). Changing

Eapp causes an alteration in the [XCuIIL]/[X–CuIL] ratio at the electrode surface, which is closely related to the

Nernst equation:

RT [X - Cu II L]
Eapp  E1 / 2  ln (15)
F [X - Cu I L]

Changing the [CuII]/[CuI] ratio modulates the reaction rate, because, according to eq. 1, Rp is directly

proportional to [CuIL]/[XCuIIL]. Polymerization rates are correlated with η, whereby more negative potentials

provide larger cathodic currents and faster rates of polymerization (kpapp) until mass transfer limits are reached

(Figure 8) [122].

Once a sufficiently negative η is applied, the rate of XCuIIL reduction becomes limited by mass transport

of the reagents to and from the electrode surface. Under such conditions, Rp reaches a plateau and becomes

independent of η, i.e. progressively more negative Eapp values do not further enhance the polymerization rate.

This Rp limit is proportional to the square root of [XCuIIL] (see Section 2.3, eq. 5).

The rate limit observed at sufficiently negative η values is sometimes referred to as “diffusion limit.”

However, a more correct definition is “mass-transport limit”, a term that includes all modes of mass transport

and cannot be confused with the kinetic “diffusion limit” of a bimolecular reaction [122].

19
Regardless of the applied potential, fast decay of the current was observed in the early stages of

polymerization under potentiostatic condition. This indicated efficient conversion of XCuIIL to XCuIL in the

bulk solution until the [CuII]/[CuI] ratio reached the value dictated by eq. 15. More negative η produced larger

initial currents and faster reduction rates. Afterward, however, steady-state conditions were reached, where

[PnX], [Pn•], [CuIL], and [XCuIIL] are at equilibrium, and fairly constant currents were observed [122, 208].

Beside these considerations, the choice of Eapp should be judicious to avoid other undesired electron

transfer reactions. In ATRP systems, one of these is the reduction of CuIL to metallic Cu0 and free ligand. This

usually occurs at potentials more negative than that of the XCuIIL/X–CuIL redox couple [122, 129, 130, 133,

143, 190].

For the preparation of star polymers, the polymerization rate was usually controlled by applying less

negative Eapp values, thereby decreasing radical concentrations and suppressing termination reactions between

growing arms (Figure 9) [130, 133, 137, 143]. Termination between growing arms can be intermolecular or

intramolecular, whereby the former leads to star-star coupling and the latter to uneven arm growth. Both types

of coupling were diminished by applying less negative Eapp.

An on-demand polymerization was realized by modulating Eapp (and therefore [XCuIIL]/[CuIL]) in-situ

[42, 119]. Once the activator CuIL was generated by applying a negative overpotential (η < 0), polymerization

started and proceeded smoothly, but the application of a positive η quickly oxidized the catalyst to the

deactivator XCuIIL state stopping polymerization. By cycling the potential from negative to sufficiently

positive values of η, the polymerization could be stopped and reinitiated several times without deleterious

effects on the polymerization. An example of this electrochemical switching is illustrated in Figure 10. All

chains were effectively reinitiated after each “ON/OFF” step proving the livingness of the process. Alterations

of polymerization rate through switching the catalyst oxidation state enabled precise modification of

macromolecular species at any specified time. Moreover, since the system can be started and stopped by

20
appropriately adjusting η, auto acceleration reactions can be easily prevented by limiting the polymerization rate

through the applied potential when employing exothermic polymerizations.

_ENREF_72 Effect of Stirring Rate

As discussed previously, the polymerization kinetics depend on [XCuIIL]/[CuIL]. Since the reduction of

the catalyst occurs only at the WE surface, efficient stirring is necessary to establish a fixed and homogeneous

[XCuIIL]/[CuIL] ratio in the bulk solution. Efficient agitation homogeneously distributes the reduced activators

in solution, and prevents deposition of the growing chains at the WE surface, which causes lower Rp and

broader molecular-weight distributions. In the polymerization of BA in DMF, it was found that a stirring rate

greater than 450 rpm was sufficient to homogenously mix the reactants (Figure 11). Polymerizations conducted

without sufficient stirring suffered from sluggish polymerization rates, high dispersities, and limited conversion

of monomer.

The [XCuIIL]/[CuIL] ratio was also controlled by adjusting the distance between WE and CE, when an

eATRP was conducted without stirring [122]. The influence of distance between electrodes is addressed in

Section 3.2 for the preparation of polymeric surfaces through surface-initiated eATRP.

Influence of Catalyst Structure and Concentration

The polymerization rate strongly depends on the catalyst’s structure [122]. When the same overpotential η

was applied, i.e. fixing the [XCuIIL]/[CuIL] ratio, complexes with higher reducing power provided faster

polymerization rates due to a larger KATRP, i.e. Cu/Me6TREN > Cu/TPMA > Cu/PMDETA (Figure 12).

Uniquely with eATRP, when employing low activity catalysts (e.g., Cu/PMDETA), the Mw/Mn can be

reduced by using a less reducing η to increase [XCuIIL] as shown by the blue and magenta points in Figure

12b. In this instance, the dispersity of the polymer was reduced from ca. 2.2 to 1.5 at a similar monomer

conversion.
21
In contrast, a relatively positive Eapp (η > 0) was used to diminish [CuIL] and obtain controlled

polymerization in water, where catalyst activity is typically very high. A ratio [XCuIIL]/[CuIL] ~ 100 was

selected for the very active Cu/Me6TREN in 10% OEGMA (v/v) in H2O using 0.1 M Et4NBr [193].

According to eqs. 1 and 5, Rp is expected to be proportional to the square root of [XCuIIL]. Figure 13

shows that kpapp, inferred from the slope of pseudo-first order plots, increased with the initial concentration of

deactivating catalyst [122, 128-130, 188, 189]. In addition, higher catalyst loadings showed narrower molecular

weight distributions, as the polymer’s dispersity is inversely proportional to [XCuIIL] in agreement with eq. 2.

Recently, poly(butyl acrylate) homopolymers were prepared via eATRP using only 10 ppm of XCuIIL [189],

much lower than commonly reported levels at 50–400 ppm [122, 126, 128, 130, 132, 133, 137, 143, 188].

Electrolyte Effects

eATRP is also affected by the supporting electrolyte. A decrease in the rate of polymerization was noted

when the concentration of supporting electrolyte was decreased in potentiostatic eATRP [190]. The decrease in

the rate of polymerization can be explained as the consequence of an increased solution resistance. Addition of

an electrolyte to the polymerization system with bulky non-coordinating counter ions, e.g. tetrabutylammonium

hexaflourophosphate (TBAPF6), has little, or no, adverse effects on eATRP. Besides lowering the resistance,

adding an electrolyte provides a means to control the concentration of a desired counterion in the reaction

medium. Switching the electrolyte anion from PF6 to bromide (TBABr) changed the behavior of the system

and resulted in slower polymerization rates. An excess of bromide ions can form an appreciable quantity of the

inactive BrCuIL complex [193, 195], or can even displace the amine ligand from the metal complex [120,

196]. Contrary to organic solvents, ATRP in water benefits from the presence of relatively high concentrations

(0.03-0.3 M) of chloride of bromide salts [42], to increase the concentration of XCuIIL (as further explained in

the next paragraph). Consequently, it is usually not necessary to add additional supporting electrolyte in

aqueous eATRP.
22
Solvent Selection: from Organic to Aqueous Media

The choice of the solvent plays a crucial role in eATRP. Typically, eATRP reactions have been carried

out in DMF and MeCN as solvents. The polymerization in DMF was faster in comparison to reactions

conducted in MeCN because of its higher KATRP [25, 143]. An aqueous environment further enhanced reaction

rates, since the KATRP in aqueous solution is ca. 103 times higher than in MeCN [127, 129, 193, 209]. The

activation rate constant, ka, is similarly very high in aqueous solutions [91, 120].

In pure water, the activity of the catalyst is so high that the CV pattern of Cu I/TPMA and HEBiB (2-

hydroxyethyl bromoisobutyrate) was observed to exhibit “total catalysis” behavior (Figure 14) [120]. This

occurs when the reaction between catalyst and alkyl halide is so fast that only a tiny fraction of the catalyst is

sufficient to quickly react with the totality of initiator. In such a case, the catalytic process occurs at a potential

more positive than the reduction peak of the catalyst and involves only a small fraction of the catalyst.

Therefore, the system shows a catalytic pre-peak, besides the normal reversible peak couple of the catalyst. In

contrast, less polar solvents, such as anisole, showed much lower catalytic activity and required higher

concentration of the supporting electrolyte to achieve suitable conductivity [190].

Aqueous eATRP is very attractive from both economic and environmental points of view. During the past

few years, water soluble poly(meth)acrylates and polyacrylamides were successfully prepared using eATRP.

Relatively narrow Ð and experimental Mn close to the theoretical values were obtained [120, 127, 129, 139,

142, 193]. However, an aqueous eATRP entails several challenges. The XCuIIL bond can easily dissociate due

to halidophilicity constant between 1 < KX < 14 [193]. Therefore, either high concentrations of catalysts or

addition of salts (halide anions, X–) was required [120, 210]. The addition of excess X– promoted the formation

of XCuIIL and a noticeable improvement in the molecular weight distributions.

Reaction temperature was found to have an important effect in aqueous ATRP. The concurrent

nucleophilic substitution (solvolysis) of alkyl halides in water was suppressed at lower temperatures. Thus,

23
aqueous eATRPs conducted at relatively low temperature (0 – 25 °C) and high concentration of Cu catalysts

were found most effective [94, 127, 211-213].

The effect of pH on eATRP of OEGMA and MAA was investigated. Stable complexes such as Cu/TPMA

allowed the controlled synthesis of OEGMA in water down to pH 1.5 whereas pH > 7 should be avoided [193].

Using ClCuIITPMA as catalyst, the aqueous eATRP of methacrylic acid was recently obtained under strongly

acidic conditions (pH < 1) [42].

Figure 15 displays the necessary conditions to achieve a well-controlled eATRP of OEGMA when using

Cu/TPMA, Cu/Me6TREN and Cu/PMDETA complexes. Optimal results were found when using a high

concentration of halide salt, an appropriate pH value, and a sufficiently positive Eapp to ensure [CuII]/[CuI] ≥ 10.

ATRP in water is also problematic due to fast disproportionation of Cu IL and high KATRP. A high ATRP

equilibrium constant leads to high radical concentrations and consequently to more frequent chain termination

events. Both of these negative aspects, fast disproportionation and high KATRP, could be circumvented by

eATRP. This technique easily allowed working conditions to be established that generated a high and well-

controlled [XCuIIL]/[CuIL] ratio (η > 0), suppressing CuIL disproportionation and also maintaining a low

radical concentration [94, 120, 127, 193].

Aqueous media were used for the eATRP of nonpolar monomers in miniemulsion [190, 214, 215]. These

systems were characterized by the presence of two heterogeneous interfaces, i.e., an oil/water and an

electrode/water interface. The formulation of a satisfactory eATRP in miniemulsion required reduced reaction

temperatures (60-65 °C) and the use of a dual catalytic system, composed of an aqueous phase catalyst (Cu/Laq)

and an organic phase catalyst (Cu/Lorg). Among the tested catalysts, the best results for n-butyl acrylate (BA)

polymerization were obtained with Laq = N,N-bis(2-pyridylmethyl)-2-hydroxyethylamine (BPMEA) and Lorg =

bis[2-(4-methoxy-3,5-dimethyl)pyridylmethyl]octadecylamine (BPMODA*), which produced stable latexes and

well-defined polymers with different DP.

24
The addition of a second hydrophobic copper complex allowed electron transfer from the aqueous to the

organic phase (droplet). Cu/Laq complex played the role of an electron messenger, transporting electrons from

the WE surface to the organic phase, via two possible mechanisms (Scheme 4): (A) direct reduction of CuII/Lorg

or (B) activation of the hydrophobic chain end followed by deactivation by the Cu II/Lorg [214]. Differently from

traditional ATRP systems, the reactivity of the dual catalytic system did not depend on the redox potential of

the catalysts, but instead depended on the “hydrophobicity” and partition coefficient of the aqueous phase

catalyst. In miniemulsion eATRP, catalyst partition and interfacial dynamics are new important parameters to

regulate the process.

Efficient electropolymerizations of BA and tert-butyl acrylate (tBA) in miniemulsion were also conducted

with a single catalytic system, composed only of an aqueous phase catalyst (CuIIBr2/2TPMA) and an active

surfactant (sodium dodecyl sulfate). The proper balance between SDSCuIIL interaction and CuIIL redox

potential provided efficient electrocatalysis and well-defined polymers by eATRP in miniemulsion [215].

Polymerization Under Potentiostatic versus Galvanostatic Conditions

Up to now, most eATRPs were conducted at a fixed Eapp (potentiostatic condition) in a three-electrode

cell. On the other hand, electrolysis under galvanostatic conditions is attractive from an industrial and

experimental standpoint, as it requires less expensive instrumentation (current generator vs potentiostat) and a

simpler two-electrode system (WE and CE). Galvanostatic conditions were successfully applied for the

synthesis of homopolymers and copolymers using eATRP [122, 126, 129-133, 136, 143, 215, 216].

The current profile (chronoamperometry) recorded during potentiostatic electrolysis was used to select the

appropriate current program for galvanostatic mode (Figure 16a). Therefore, the total charge passed was nearly

the same in both experiments. This multiple-stage current program was chosen to rapidly convert a significant

amount of XCuIIL to XCuIL at the beginning of the polymerization to quickly initiate eATRP. Next, the

lower current steps allowed CuIL to be continuously (re)generated from XCuIIL produced from unavoidable

25
radical termination reactions. Concerning polymerization control, both polymerizations displayed a linear

increase in molecular weight with conversion and low Mw/Mn values during the course of the reaction (Figure

16b). When less current steps were used in the galvanostatic experiment, the potential was not appropriately

controlled, and a slower reaction and/or copper deposition on the electrode surface were observed [122, 126,

129, 130, 133, 143, 215, 216].

_ENREF_79s

2.8. eATRP: Advantages and Limitations

The application of electrochemical techniques to ATRP introduced a unique and novel strategy to

dynamically control this redox mediated radical polymerization process. The extent and rate of reduction of the

catalytic complex (XCuIIL) is dictated by the Eapp, allowing fine-tuning of the polymerization rate [190].

Cycling the Eapp between appropriate values higher and lower than E1/2, allowed to realize an on-demand

polymerization, by triggering active or inactive catalytic (electrochemical switch, Scheme 5) [42, 119]. On the

other hand, mass transport also plays an important role, and the spatial control of eATRP can be achieved using

catalyst concentration gradient in solution by controlling the distance between the traditional WE [123] or

bipolar electrodes (BPE) [142] and initiator-modified substrate [123] (more details in section 3.2).

Electrochemical methods provide significant improvements to, and control over, ATRP by offering

adjustable “dials” (e.g., current, potential, total charge passed) to change catalyst oxidation states,

polymerization rates, and active/inactive chain growth stages during polymerization. In addition,

electrochemistry offers an environmentally friendly alternative to current ATRP techniques because chemical

reducing agents, such as those used in ARGET ATRP or ICAR ATRP, are not required. Moreover, the

procedure is tolerant to small amounts of oxygen and allows a new and straightforward approach to catalyst

removal/recycle through electrodeposition (see Section 4) [24, 51, 52, 122, 217-

219]._ENREF_133_ENREF_124

26
Furthermore, eATRP can be conducted with very low concentrations of activator and in a full range of

solvents, including water, thereby including bio-active agents to the range of possible initiators or monomers.

Increasing the polymerization rate by applying a negative Eapp allowed lowering the temperature at which the

process was performed [94, 127]. eATRP, relying on an external stimulus, is less sensitive to changes in the

polymerization conditions, compared to traditional chemical techniques such as ARGET [42] or ICAR.

Nevertheless, there are some limitations associated with eATRP, especially in the complexity of reaction

setup, relatively to normal ATRP or other low ppm ATRP methods. Moreover, in eATRP the (re)generation of

the active catalyst occurs only at the WE surface, not in the bulk of the solution. A supporting electrolyte was

always added to the polymerization medium. Typically, eATRP has been carried out in a three-electrode cell by

a potentiostatic process, i.e., at constant Eapp. The role of the anode is to close the electric circuit and once Eapp

(or current) is fixed, the most favorable reaction occurring at the anode is the one with the lowest oxidation

potential. If an undivided cell is used, this could be oxidation of Cu I back to CuII, which is highly undesirable.

Another equally undesired possibility is oxidation of halide ions, which are always present in the reaction. Also

oxidation of the solvent and/or supporting electrolyte cannot be excluded. In general, the CE is separated from

the reaction mixture to avoid undesired side reactions and side effects due to anodic reactions. In fact, if the

anodic and cathodic compartments are properly separated (e.g. by a glass frit and methylcellulose gel), the

products generated at the anode, whatever the anodic process, do not contaminate the polymerization process.

2.9. Simplified eATRP (seATRP)

eATRP Using Sacrificial Counter Electrode

The complexity of the electrochemical reaction setup can be reduced by using a sacrificial counter

electrode. This procedure, referred to as simplified eATRP (seATRP), does not require the separation of

cathodic and anodic compartments, and the CE can be directly inserted into the reaction mixture (Figure 17).

The use of an undivided cell allowed for a simpler and lower cost process. Moreover, the ohmic drop caused by

27
the separator is avoided bringing a beneficial energy saving. This system was further simplified by working

under galvanostatic conditions using a two-electrode system directly powered by a current generator. Both

advantages should benefit commercial and academic efforts in development of eATRP [126, 129, 130, 133,

136, 137, 143, 215, 216].

In the ideal case of seATRP, the sacrificial counter electrode as well as the products of its oxidation

should not react with the CuL catalysts. Oxidation of CuIL to CuIIL did not occur at the Al anode because

anodic dissolution of the anode itself was more favourable. Al3+ ions were therefore generated in solution while

the CuII/CuI redox system was not affected by the Al anode. Nevertheless, an excess of amine ligand was

required to complex the Al3+ ions [126, 132].

A series of seATRP polymerizations of BA in DMF were carried out at different Eapp, with a Pt mesh WE,

an Al CE, and Ag/AgI/I- or saturated calomel electrode (SCE) RE [126, 130, 133, 136, 137, 143, 216]. The

polymerizations showed MW evolution and narrow Ð throughout the reactions similar to conventional three-

electrode eATRP (Figure 18). Well-controlled seATRP polymerizations were also achieved with other

monomers, such as MMA [129, 137], tBA [130, 136, 137, 216], 2-(dimethylamino)ethyl methacrylate

(DMAEMA) [143, 216], and oligo (ethylene glycol) acrylate (OEGA) [133, 136].

eATRP Using Non Platinum Electrodes

eATRP was further developed and simplified by replacing Pt cathode with cheaper, more easily available,

and easily functionalizable materials [131]. Indeed, the role of the cathodic material in eATRP is to act solely as

an electron source for catalyst reduction, and, in principle, any electronic conductor can accomplish this task.

However, an electrode material is also required to be stable (e.g. no corrosion) and inert with a wide

electrochemical potential window (it should not directly activate CX bonds).

A series of polymerizations of OEGMA in aqueous solution were performed on Au, GC, Ti, NiCr (80/20,

wt%) alloy, and 304 stainless steel (SS304) electrodes. The reactions proceeded with fast kinetics, high

28
conversions, low dispersity, and good agreement between theoretical and experimental Mn (Figure 19). While

noble metals and GC worked well under potentiostatic mode, in aqueous environment non-noble metals

required galvanostatic conditions with pre-electrolysis of XCuIIL before the addition of initiator. In DMF,

working electrodes made of GC, Au, Fe, NiCr and SS304 were used for the polymerization of n-BuA without

any substantial modification of the electropolymerization procedure [132].

In each solvent, release of metal ions from the WE into the polymerization solution was negligible. These

cathodes were also used in combination with a sacrificial Al anode in an undivided cell, completely removing Pt

from the reaction setup. These achievements represent a significant step toward the scale-up and the

commercialization of this polymerization technique.

3. Polymer Architectures by eATRP

3.1. Homopolymers, Block Copolymers and Star Polymers

Table 1 lists homopolymers and block copolymers prepared by eATRP [42, 94, 119-137, 139, 140, 142,

143, 188, 189, 193, 206, 215, 216, 220-223]. However, eATRP was also used to prepare copolymers with more

complex architectures [124, 130, 133, 136, 137, 143]. Indeed, eATRP could have several advantages for the

synthesis of polymers with complex architectures. For example, eATRP was used in the preparation of star

polymers, which combine interesting properties of a branched architecture, globular shape, and chemically

cross-linked structure [224-228]. In this case, eATRP was applied to favor cross-linking of the star cores but

disfavor coupling between stars.

One approach to prepare star polymers by eATRP is the macroinitiator (MI) method, in which the chain-

end functionalized linear MIs can react with multifunctional crosslinkers to form a dense core (from chain

extension with a cross-linker) with multiple peripheral arms. The MI method can provide star polymers with

various chemical composition, stimuli-responsiveness, and site specific functionalities, while maintaining high

star yields with low Cu catalyst loadings (∼100 ppm, w/w).


29
The first eATRP synthesis of star macromolecules was attempted with PBA MIs under potentiostatic

conditions. The final MW and dispersity of the star polymer were 60,000 and 1.40, respectively [190]. Another

successful example of eATRP for star polymer synthesis was demonstrated with poly(ethylene oxide) (PEO)

(Figure 20a) [124]. PEO ATRP MIs were used for star synthesis by reacting them with ethylene glycol

diacrylate cross-linkers. Various experimental conditions were examined to optimize star formation, including

MI concentration, MI to cross-linker molar ratio, and Eapp. High star yields of up to 95% were obtained under

the optimized conditions. The use of a multistep potentiostatic technique allowed for a gradual increase in the

Rp, thus reducing star−star coupling and providing higher MW stars in higher yields (Figure 20b).

Recently, the first example of both 14-arm and 21-arm -cyclodextrin (-CD)-based PBA stars was

demonstrated by eATRP under both potentiostatic and galvanostatic conditions (Figure 21) [130].

Polymerization was fast (> 80% conversion in 3 h) with low catalyst concentrations (50 ppm) and produced star

polymers with narrow molecular weight distributions. The rate of the polymerizations (Rp) was controlled by

applying different potentials Eapp, with lower Rp observed using more positive Eapp values, thereby, suppressing

intermolecular termination reactions between growing arms and subsequent star-star coupling. In addition, by

exploiting the high ω-chain end functionality on the periphery of star polymers, well-defined block star

copolymers (β-CD-PBA-PtBA) were successfully prepared by seATRP.

Other 21-arm -CD-based star polymers were also prepared by seATRP: β-CD-PDMAEMA-b-PBA

[143], β-CD-PMMA-b-PtBA [137], and β-CD-PMMA-b-PtBA-b-PBA [137]. Another example is a glucose-

based star-like block copolymer with a hydrophilic α-D-glucose core and amphiphilic block copolymer arms

composed of BA and OEGA, synthesized also via seATRP (Figure 22) [133]. These new cyclodextrin/glucose-

based polymer materials have potential applications as biomaterials with increased hydrophobicity, as well as

antifouling coatings, in biological media [133, 137].

30
3.2. Modification of Surfaces with Polymeric Brushes

An excellent example of the application of eATRP is the direct growth of polymer brushes from surfaces

functionalized with ATRP initiators. Several polymers including POEGMA, PHEMA, PDMAEMA, PSPMA,

PSBMA, PGMA, PMMA, PNIPAM, PMETAC, PAM, poly(N,N-methylene-bis-acrylamide) (PMBA), PVP,

PBA, and PtBA were successfully grafted from solid surfaces, such as gold surfaces with ATRP initiators [121,

123, 125, 134, 136, 139, 140, 142, 216, 220-223,

229]._ENREF_128_ENREF_128_ENREF_77_ENREF_140_ENREF_80_ENREF_73

Typically, the initiator was grafted on Au via a self-assembled monolayer (SAM), i.e. anchoring a linker

with a thiol (–SH) group on one end and an ATRP initiator (–Br) on the other

end._ENREF_128_ENREF_128_ENREF_77_ENREF_140_ENREF_80_ENREF_73 The length of linkers

between the –SH and –Br groups influenced surface conductivity and grafting density [121]. Short linkers (HS–

C6–Br) provided good electrical conductivity, but poor quality initiator monolayers. Long linkers (HS–C15–Br)

provided good quality initiator films, but in this case a fraction of 2-naphtothiol was required to provide

sufficient electron transfer pathways, at the expense of graft density. Microcontact printing technique allowed

preparation of thiol-based ATRP initiator monolayers at certain location of the WE. eATRP was carried out by

reduction of CuIICl2/BPY on initiator-decorated Au electrodes to prepare well-defined polymeric surfaces

(Figure 23). The thickness of polymer brushes was controlled through adjusting the fraction of initiator and

applied potentials.

Similarly, surfaces were modified with water soluble polymers (PHEMA or POEGMA) by eATRP in an

aqueous media without additional supporting electrolyte (Figure 24) [139]. The use of an electrochemical quartz

crystal microbalance displayed a highly regular increase in surface confined mass only after the addition of the

copper catalyst which was reduced in situ and transformed into its active form. After isolation and washing of

the modified electrode surface, re-initiation of polymerization process was achieved, with retention of the

controlled electrochemical ATRP. This approach has interesting implications for the synthesis of smart thin film

31
and also offers the possibility of post-modification via additional electrochemically induced reactions. When

combined with atomic force microscopy (AFM), the direct growth of water soluble polymeric chains, i.e.

PSPMA or PMETAC, was directly visualized [140]. Surfaces modified with polyelectrolyte brushes have

potential applications in areas such as biosensors and “smart” coatings.

Another advantage of surface initiated eATRP (SI-eATRP) is the ability to synthesize gradient polymer-

brush surfaces by controlling the distance between the WE and initiator-modified substrate. CuIL diffusion from

the WE generated catalyst’s concentration gradients, which affected polymer growth on the nearby substrate

(Figure 25) [123]. Initiator moieties close to the WE were exposed to a higher concentration of CuIL and

therefore exhibited higher polymerization rates, leading to increased brush lengths. For example, when the

silicon (Si) wafer, modified with ATRP initiators, was placed at a titled angle with respect to the WE, the

XCuIIL/CuIL ratio was different depending on the proximity between the WE and substrate surface

(designated as “d” in Figure 25). Therefore, polymerization rates were different and resulted in surface brushes

with a MW gradient. In addition, the steepness of the gradient could be easily tuned by changing the tilt angle

between WE and substrate. A linear continuous increase in the thickness of the PGMA gradient layer along one

direction (i.e., x-axis) was achieved (Figure 26). Moreover, a “wedge” or “stair” shaped gradient pattern was

formed by pre-patterning the initiator areas (Figure 26). This approach of using the [XCuIIL]/[CuIL]

concentration gradients for tailored gradient slopes can be potentially used to prepare very complex surface

topographies in a straightforward fashion.

Gradient polymeric surfaces were also fabricated using bipolar electrolysis. A bipolar electrode (BPE) is a

wireless electrodes produced by the polarization of a conducting material immersed in a solution with low

concentration of a supporting electrolyte in the presence of an external electric field, which is generated by two

“driving electrodes” (Figure 27). BPEs have unique characteristics, including the ability to generate a gradient

potential [230-236]. A potential gradient generated on a BPE allowed the formation of a concentration gradient

of a CuIL polymerization catalyst through the one-electron reduction of XCuIIL, resulting in the gradient

32
growth of PNIPAM and PMMA brushes from an initiator-modified substrate surface set close to a BPE (Figure

27) [142]. These polymer brushes could be fabricated in three-dimensional gradient shapes with control over

thickness, steepness, and modified area by varying the electrolytic conditions. Moreover, a polymer brush with

a circular pattern was successfully formed by using a Pt disk as driving electrode, in conjunction with an

insulating cylinder and a second toroidal electrode, which resulted in a site-selective reduction of CuII to CuI

(Figure 28) [142]. Because the BPE system requires low concentration of supporting salts, it allows flexible

selection of the electrolytic media and monomers.

Densely grafted (co)polymer brushes, including PBA, PDMAEMA, PBA-b-PtBA, and PDMAEMA-b-

PtBA, can be achieved from silicon wafer surfaces by SI-seATRP under both constant potential and constant

current electrolysis conditions (Figure 29) [216]. The rate of the polymerization was controlled by applying

more negative potential values, which allowed to increase the surface thickness (up to 118 nm and grafting

densities up to 0.88 nm–2). These brushes are promising candidates for applications in microelectronics,

nanotechnology, and bioresponsive materials.

One interesting example of the application of eATRP is grafting PGMA from carbon fibers (CFs) and

post-modification to introduce iminodiacetic acid functionalities. The modified CFs can be used for nano-nickel

recovery from spent electroless nickel plating (EN) baths (Figure 30) [125]. In this case, the first stage is

electro-grafting of aryl layers from a diazonium salt bearing ATRP initiating groups. The second step is surface-

initiated eATRP followed by post-functionalization of PGMA grafts by reaction of iminodiacetic acid with the

epoxy groups in the presence of sodium hydroxide. The last step is adsorption of Ni2+ at CFs from spent

electroless nickel plating baths. The CFs with absorbed Ni2+ were then used as a WE for the electrodeposition

of Ni at -0.7 V in sulfuric acid to obtain nano-nickel coated CFs. These chelating carbon fibers may also be used

in the treatment of wastewater containing heavy metals.

Another approach to the electrochemical modification of surfaces was the grafting of N-vinylpyrrolidone

from a ordered mesoporous carbons (OMCs) containing 2-bromoisobutyryl ATRP initiating sites (Figure 31)

33
[222]. The introduction of hydrophilic poly(N-vinylpyrrolidone) PVP brushes on the surface of OMCs results in

the improvement of wettability and an increase in the specific capacitance. The SI-eATRP method opens the

way for the design of improved carbon-based electrodes for electrochemical energy storage devices.

Super-low protein absorption polymeric surfaces can be produced by SI-eATRP [220]. A well-defined

zwitterionic polymer hydrogel (PSBMA) coating was prepared from gold substrate with immobilized ω-

mercaptoundecyl bromoisobutyrate initiator under potentiostatic condition (Figure 32). In vitro and in vivo

impedance-time scans showed that the PSBMA coating effectively delayed the sensitivity decay of the

electrodes. Preparation of zwitterionic polymer surfaces provides the possibility of multiple applications,

especially in the development of biocompatible materials for implantable devices such as neural and biosensor

electrodes [220].

Recently, new synthetic route for the preparation of antifouling zwitterionic PSBMA coating on enzyme-

based glucose (GOx) sensor (deposited on Pt wire) by SI-eATRP was presented (Figure 33) [223]. The

preparation of the enzyme sensor was performed by coating Pt wire with a PANI film, followed by physical

absorption of protein initiator (GOx-Br). This highly controllable electrodeposition procedure could be a

promising method to develop implantable biosensors with long-time stability.

Another successful example of SI-eATRP is the growth of PAM polymer brushes on the surface of a Au

electrode modified by SAM of initiator (Figure 34) [221]. Polymerization was performed on the modified

Br/Au electrode immersed in a solution of CuIIBr2/BPY by applying a potential Eapp = E1/2 – 120 mV. The

PAM/Au electrode can be used for the determination of Pb2+ ion by differential pulse voltammetry (DPV).

Recently, a synthetic route for the preparation of a molecularly imprinted polymer (MIP) on the surface of

Au electrode by eATRP with hemoglobin (Hb) both as catalyst and template molecule was presented (Figure

35) [229, 237]. In this example, mixed SAMs of 2-bromoisobutyryl bromide and 4-hydroxythiophenol were

self-assembled on the surface of Au electrode. Then, a Au working electrode was inserted into a protein

solution, containing functional monomer and cross-linker, and then polarized to reduce Hb–FeIII to Hb–FeII

34
triggering polymerization. After the removal of Hb, a MIP was successfully prepared on the electrode surface

(MIP/Au). The modified MIP/Au electrode was used as a biosensor and applied for the determination of Hb by

DPV measurements.

Another approach to the electrochemical modification of surfaces was the grafting of hydrophilic

monomers from a conducting polymer (CP) containing ATRP initiating sites (poly(2-(2,5-di(pyrrol-2-yl)

thiophen-3-yl) ethyl 2-bromopropanoate) [134]. A gold surface acted first as a support for the CP cathodic

deposition and then as a WE to reduce CuIIBr2/TPMA and control HEMA ‘grafting from’ the CP (Figure 36).

However, competing electrografting mechanisms occurred: HEMA partially polymerized in solution, followed

by ‘grafting onto’ reactions. Additionally, the mechanism of initiation differed from previously reported eATRP

systems in that the conducting polymer itself acted as a co-reductant for the copper catalyst.

4. Other Related Electrochemical Techniques

Other electro-polymerization methods have been used in polymer chemistry in addition to eATRP.

Typically, monomers react at an electrode often forming insoluble electrode coatings, without good control on

molecular weight or dispersity of the produced chains. In this case polymerization occurs by oxidation (or

reduction) of monomers which can cross-link upon application of various electrochemical reactions. Thus, the

field of classical electropolymerization is a very active area of research [238-242]. Polymerizations controlled

by electrochemical means were mostly applied to the modification of surfaces and electrodes (section 4.1).

Nevertheless, a few reactions that are conceptually very close to homogeneous eATRP were reported in solution

(section 4.2). Electrochemistry was also efficiently used for polymer purification, through electrodeposition of

several metal catalysts (section 4.3).

35
4.1. Surface Modification

Recently, an electroactive initiator to grow polymer brushes from Pt, GC, and Au surfaces was utilized

[243]. These custom initiator molecules contained two reactive groups, one suited for surface attachment and

the other functioning as a latent radical initiator for subsequent polymerization (Figure 37). In the first step, an

electrochemical oxidation was performed to anchor 4-phenylacetic diazonium salts to the electrode surface

generating a diazonium-terminated surface. Then, this diazonium functionality was electrochemically reduced

to irreversibly generate aryl radicals via the Meerwein reaction, which is an activation by CuIL like ATRP.

Conducting the process in the presence of monomer and ATRP catalysts then allowed for the formation of

polymer brushes from a variety of substrates. Grafted diazonium salts acted as dormant radical sources; the

subsequent reduction of the diazonium functionality was capable of controlling film thicknesses and was

conducted in a simple and versatile fashion.

Surface-initiated polymerization was also used to prepare thin hydrogel films (Figure 38) [180, 181, 244].

Electron transfer from the WE surface to a redox-active radical initiator (e.g. ammonium persulfate) promoted

polymerization close to the electrode, either by a free radical process or by RAFT. The prepared polymer chains

were immobilized at the electrode in the presence of a cross-linking agent, leading to hydrogel-modified

electrodes. Two different RAFT agents were employed in either acidic or basic solution. In both cases, addition

of RAFT agents had an influence on the thickness and the surface morphology of the films. Neither RAFT

agent displayed electrochemical activity on its own, but they modified the electrochemical behavior of the

initiator. The addition of a CTA strongly enhanced the homogeneity of the hydrogel surfaces, which

presumably was due to a reduced amount of microgel formation [181].

Recently, a synthetic route for the preparation of a functionalizable conducting poly(3,4-(1-

azidomethylethylene) dioxythiophene) (PEDOT-N3) was presented [245]. Poly(3,4-ethylenedioxythiophene)

(PEDOT) films were functionalized selectively with alkyne-modified moieties in an analogous

electrochemically controlled polymer click reaction [246, 247]. An array of PEDOT-N3 covered electrodes was

36
used, which could be modified only where the potential was applied (Figure 39). This allowed advanced

patterning of PEDOT films. Gradients of these “electro-clicked” groups along the substrate were achieved by

the electrochemical generation of the active CuI species within a dedicated electrode setup [141]. Separated

microelectrode patterns were functionalized with different alkyne species in turn, by controlling the potential

applied to the different electrodes. The method is fast and versatile, and facilitates easy post-functionalization of

an electrically conducting polymers [246].

An approach to pattern fluorescent molecules was proposed, which do not participate by themselves in an

electrochemical reaction, onto an azido-functionalized glass substrate using “click” reactions spatially guided

by using scanning electrochemical microscope (SECM) [248-250]. This technique extends the surface

patterning ability of SECM to the level at which any inert small molecule or macromolecule can be grown on an

insulating substrate through a click chemistry reaction to form a covalent bond directly under a microelectrode

[248]. The setup was typically constituted as follows: first, an azido-terminated monolayer was self-assembled

onto a glass substrate; then, in an electrochemical cell an acetylene-functionalized fluorophore was reacted to

form a triazole ring with the azide moieties through a “click reaction” (Figure 40). A gold microelectrode was

used to locally form CuI in the small gap between a tip and a substrate to catalyze the click reaction in a small

volume. The amount of deposited molecules was controlled by adjusting the tip-substrate distance, the amount

of CuI, and the reaction time. This approach can also be used on a larger scale and at higher density to pattern

protein arrays.

Recently, a new electrochemical sensor electrode was constructed with an original electrode modification

technique, termed “click electrochemistry” (CEC) (Figure 41) [251]. First, 4-azidoaniline was

electropolymerized on a GC electrode; then a terminal-alkynyl cobalt phthalocyanine (CoPc-TA) was

electrochemically bonded to the azido end of electro-polymerized 4-azidoaniline (PANI-N3). In both water and

organic solvent (DMSO), three different redox couples were present on the modified electrode: Co II/CoI,

37
CoIII/CoII and oxidized/reduced PANI-N3. The presence of versatile redox active sides allowed usage of the

electrode for catalysis or sensing of various target species (e.g. pesticides).

Micropatterning of PEDOT on an agarose hydrogel was also reported, to provide an organic, moist, and

flexible electrode [252]. The PEDOT/agarose electrodes were prepared through two electrochemical processes:

electropolymerization of PEDOT into the hydrogel and electrochemical-actuation-assisted peeling (Figure 42).

The melted 2.8 wt% agarose solution was poured over a Pt microelectrode fabricated on a glass plate. After

gelation of the agarose, with typical thickness of 2 mm, electropolymerization was conducted on the gel-

covered electrode in the aqueous monomer solution. Successively, redox cycles caused a volume change of the

PEDOT hydrogel, which was effective for the nondestructive peeling of the soft gel film from the master

electrode.

The process, used to prepare PEDOT micropatterned films, was versatile and successful with pre-cured

films of other natural hydrogels, e.g. collagen or glucomannan, and synthetic hydrogels, e.g. PAM or PHEMA,

as shown in Figure 43 [253]. These hydrogel-based patterns represent an electrode that can be combined with

cells and tissues without disturbing the physiological conditions, including the continuous supply of O2 and

nutrients. Such properties are ideal for use as in vivo and in vitro electrodes for stimulation and recording. In

addition to such cellular applications, these electrodes with improved conductivity should be applicable to a

variety of hydrogel-based bio-electronic applications such as a transdermal iontophoresis patch.

4.2. Controlled Polymerization in Solution

One example of an electrochemically controlled “click reaction” was demonstrated for protein

modification [254]. This conjugation could be performed in air, as the concentration of the reactive catalyst

species was maintained by electrochemical means. This concept, based on Cu I-catalyzed azide–alkyne

cycloaddition (CuAAC), is similar to the eATRP procedure [255-260]. CuAAC requires that the copper

catalyst, which is usually prepared with an appropriate chelating ligand, be maintained in the air-sensitive CuI

38
oxidation state, usually by the use of an in situ reducing agent such as an ascorbate or tris(2-carboxy-

ethyl)phosphine (TCEP) [261]. In electrochemically mediated CuAAC the active Cu I species is (re)generated at

the WE.

Recently, electrochemically assisted ethylene polymerization and ethylene-propylene copolymerization

were conducted with the vanadium-based Ziegler-Natta catalyst VOCl3/ethylaluminum sesquichloride (EASC)

(Scheme 6) [262]. The poorly active V2+ was oxidized to the active V3+ at a WE. The catalytic activity of the

catalyst was largely improved by changing its oxidation state: the electrochemically-controlled system

displayed higher olefin (co-)polymerization activities compared to the non-electrochemically controlled system,

and produced higher MW (co-)polymers. This can improve the production of polyolefin and synthetic rubber.

4.3. Catalyst Removal

Copper removal from the solution was accomplished by electrodeposition [124, 219]. Polymers with high

affinity for Cu complexes, such as 2-ureido-4H-pyrimidone (UPy)-POEGMA could not be efficiently purified

by dialysis, extraction, precipitation, or passing through alumina columns. More efficient removal of Cu from

reaction mixtures was achieved when Cu species were electrochemically reduced to Cu0 and deposited on the

electrode surface (Figure 44). Typically higher applied potentials and longer electrolysis periods (e.g., 12 V for

44 h) showed higher efficiency of Cu removal (> 98% removal) without any damage or degradation to the

polymers.

Removal of copper improved the stability of polymeric poly(n-butyl methacrylate) (PBMA) latexes

prepared by emulsion ATRP [263]. Higher voltage (9 V) provided higher copper removal (>99% in 7 h).

Copper electrodeposition did not influence particle distribution or polymer molecular weight (Figure 45).

Instead, polymers produced from latexes subsequent to Cu removal showed significantly greater aging and

weathering stability as compared to those containing residual Cu (Figure 46).

39
5. Summary

Electropolymerization methods are useful for the synthesis of new types of polymer materials. Among

these polymerization methods, electrochemically mediated ATRP is a rapidly developing area of polymer

science. This article summarizes the detailed experimental set-up, advantages and limitations as well as

mechanistic details of eATRP. In eATRP, the ratio of the concentration of activator to deactivator is precisely

controlled by electrochemical means. Several parameters, such as applied current or potential, and total charge

passed, can be externally tuned in eATRP to control and program the desired concentration of the redox-active

catalytic species.

This review aims also to provide an overview of selected, but representative recent developments, trends,

and emerging applications of polymer materials obtained by eATRP and other electropolymerization techniques

conceptually similar to eATRP. These include electrochemical copper-catalyzed azide-alkyne cycloaddition

(click chemistry), electrochemical micropatterning, and RAFT polymerization combined with redox-sensitive

initiators. The rapid increase of research activity in this field within the last decade indicates that

electrochemically controlled methods have become an essential tool to design and synthetize advanced and

novel polymer materials. The potential future application of this technique include polymerization from protein,

DNA and nanoparticles.

Acknowledgments

Financial support from Rzeszow University of Technology (U-553/DS, U-718/DS, and U-771/DS/M), National

Science Foundation (CHE 1400052) and University of Padova (CPDA150001) is gratefully acknowledged.

P.C. acknowledges Kosciuszko Foundation Fellowship.

40
References

[1] Wang JS, Matyjaszewski K.;1; Controlled living radical polymerization - Atom-transfer radical

polymerization in the presence of transition-metal complexes. J Am Chem Soc 1995;117:5614-5.

[2] Matyjaszewski K, Patten TE, Xia JH.;1; Controlled/''living'' radical polymerization. Kinetics of the

homogeneous atom transfer radical polymerization of styrene. J Am Chem Soc 1997;119:674-80.

[3] Patten TE, Matyjaszewski K.;1; Atom transfer radical polymerization and the synthesis of polymeric

materials. Adv Mater 1998;10:901-15.

[4] Matyjaszewski K, Miller PJ, Shukla N, Immaraporn B, Gelman A, Luokala BB, Siclovan TM,

Kickelbick G, Vallant T, Hoffmann H, Pakula T.;1; Polymers at interfaces: Using atom transfer radical

polymerization in the controlled growth of homopolymers and block copolymers from silicon surfaces in the

absence of untethered sacrificial initiator. Macromolecules 1999;32:8716-24.

[5] Shipp DA, Matyjaszewski K.;1; Kinetic analysis of controlled/“living” radical polymerizations by

simulations. 1. The importance of diffusion-controlled reactions. Macromolecules 1999;32:2948-55.

[6] Shah RR, Merreceyes D, Husemann M, Rees I, Abbott NL, Hawker CJ, Hedrick JL.;1; Using atom

transfer radical polymerization to amplify monolayers of initiators patterned by microcontact printing into

polymer brushes for pattern transfer. Macromolecules 2000;33:597-605.

41
[7] Matyjaszewski K, Shipp DA, McMurtry GP, Gaynor SG, Pakula T.;1; Simple and effective one-pot

synthesis of (meth)acrylic block copolymers through atom transfer radical polymerization. J Polym Sci Part A

Polym Chem 2000;38:2023-31.

[8] Coessens V, Pintauer T, Matyjaszewski K.;1; Functional polymers by atom transfer radical

polymerization. Prog Polym Sci 2001;26:337-77.

[9] Cheng GL, Boker A, Zhang MF, Krausch G, Muller AHE.;1; Amphiphilic cylindrical core-shell

brushes via a "grafting from" process using ATRP. Macromolecules 2001;34:6883-8.

[10] Percec V, Popov AV, Ramirez-Castillo E, Monteiro M, Barboiu B, Weichold O, Asandei AD,

Mitchell CM.;1; Aqueous room temperature metal-catalyzed living radical polymerization of vinyl chloride. J

Am Chem Soc 2002;124:4940-1.

[11] Pyun J, Kowalewski T, Matyjaszewski K.;1; Synthesis of polymer brushes using atom transfer

radical polymerization. Macromol Rapid Commun 2003;24:1043-59.

[12] Qin SH, Oin DQ, Ford WT, Resasco DE, Herrera JE.;1; Polymer brushes on single-walled carbon

nanotubes by atom transfer radical polymerization of n-butyl methacrylate. J Am Chem Soc 2004;126:170-6.

[13] Kong H, Gao C, Yan DY.;1; Controlled functionalization of multiwalled carbon nanotubes by in situ

atom transfer radical polymerization. J Am Chem Soc 2004;126:412-3.

42
[14] Min K, Gao HF, Matyjaszewski K.;1; Preparation of homopolymers and block copolymers in

miniemulsion by ATRP using activators generated by electron transfer (AGET). J Am Chem Soc

2005;127:3825-30.

[15] Braunecker WA, Matyjaszewski K.;1; Controlled/living radical polymerization: Features,

developments, and perspectives. Prog Polym Sci 2007;32:93-146.

[16] Tsarevsky NV, Matyjaszewski K.;1; \"Green\" atom transfer radical polymerization: From process

design to preparation of well-defined environmentally friendly polymeric materials. Chem Rev 2007;107:2270-

99.

[17] Nakatani K, Ouchi M, Sawamoto M.;1; Terminal umpolung in metal-catalyzed living radical

polymerization: Quantitative end-capping of carbon−halogen bond via a modifier monomer. Macromolecules

2008;41:4579-81.

[18] Matyjaszewski K, Tsarevsky NV.;1; Nanostructured functional materials prepared by atom transfer

radical polymerization. Nat Chem 2009;1:276-88.

[19] Lee Hl, Pietrasik J, Sheiko SS, Matyjaszewski K.;1; Stimuli-responsive molecular brushes. Prog

Polym Sci 2010;35:24-44.

[20] Boyer C, Soeriyadi AH, Roth PJ, Whittaker MR, Davis TP.;1; Post-functionalization of ATRP

polymers using both thiol/ene and thiol/disulfide exchange chemistry. Chem Commun 2011;47:1318-20.

43
[21] Matyjaszewski K.;1; Atom transfer radical polymerization: From mechanisms to applications. Isr J

Chem 2012;52:206-20.

[22] Matyjaszewski K.;1; Atom transfer radical polymerization (ATRP): Current status and future

perspectives. Macromolecules 2012;45:4015-39.

[23] Siegwart DJ, Oh JK, Matyjaszewski K.;1; ATRP in the design of functional materials for biomedical

applications. Prog Polym Sci 2012;37:18-37.

[24] Król P, Chmielarz P.;1; Recent advances in ATRP methods in relation to the synthesis of copolymer

coating materials. Prog Org Coat 2014;77:913-48.

[25] Matyjaszewski K, Tsarevsky NV.;1; Macromolecular engineering by atom transfer radical

polymerization. J Am Chem Soc 2014;136:6513-33

[26] Adali-Kaya Z,;1; Tse Sum Bui B, Falcimaigne-Cordin A, Haupt K. Molecularly imprinted polymer

nanomaterials and nanocomposites: Atom-transfer radical polymerization with acidic monomers. Angew Chem

Int Ed 2015;54:5192-5.

[27] Gao C, Yan D.;1; Hyperbranched polymers: from synthesis to applications. Prog Polym Sci

2004;29:183-275.

[28] Wang X, Graff RW, Shi Y, Gao H.;1; One-pot synthesis of hyperstar polymers via sequential ATRP

of inimers and functional monomers in aqueous dispersed media. Polym Chem 2015;6:6739-45.

44
[29] Graff RW, Wang X, Gao H.;1; Exploring self-condensing vinyl polymerization of inimers in

microemulsion to regulate the structures of hyperbranched polymers. Macromolecules 2015;48:2118-26.

[30] Qiu J, Matyjaszewski K.;1; Polymerization of substituted styrenes by atom transfer radical

polymerization. Macromolecules 1997;30:5643-8.

[31] Teodorescu M, Matyjaszewski K.;1; Atom transfer radical polymerization of (meth)acrylamides.

Macromolecules 1999;32:4826-31.

[32] Destarac M, Pees B, Boutevin B.;1; Radical telomerization of vinyl acetate with chloroform.

Application to the synthesis of poly(vinyl acetate)-block-polystyrene copolymers by consecutive telomerization

and atom transfer radical polymerization. Macromol Chem Phys 2000;201:1189-99.

[33] Kowalewski T, Tsarevsky NV, Matyjaszewski K.;1; Nanostructured carbon arrays from block

copolymers of polyacrylonitrile. J Am Chem Soc 2002;124:10632-3.

[34] Lu X, Gong S, Meng L, Li C, Yang S, Zhang L.;1; Controllable synthesis of poly(N-

vinylpyrrolidone) and its block copolymers by atom transfer radical polymerization. Polymer 2007;48:2835-42.

[35] Nicolay R, Kwak Y, Matyjaszewski K.;1; Synthesis of poly(vinyl acetate) block copolymers by

successive RAFT and ATRP with a bromoxanthate iniferter. Chem Commun 2008 :5336-8.

45
[36] Huang CF, Nicolaÿ R, Kwak Y, Chang FC, Matyjaszewski K.;1; Homopolymerization and block

copolymerization of N-vinylpyrrolidone by ATRP and RAFT with haloxanthate inifers. Macromolecules

2009;42:8198-210.

[37] Wu Z, Chen H, Liu X, Zhang Y, Li D, Huang H.;1; Protein adsorption on poly(N-vinylpyrrolidone)-

modified silicon surfaces prepared by surface-initiated atom transfer radical polymerization. Langmuir

2009;25:2900-6.

[38] Tang H, Radosz M, Shen Y.;1; Atom transfer radical polymerization and copolymerization of vinyl

acetate catalyzed by copper halide/terpyridine. AIChE J 2009;55:737-46.

[39] Wever DAZ, Raffa P, Picchioni F, Broekhuis AA.;1; Acrylamide homopolymers and acrylamide-N-

isopropylacrylamide block copolymers by atomic transfer radical polymerization in water. Macromolecules

2012;45:4040-5.

[40] Liu X, Sun K, Wu Z, Lu J, Song B, Tong W, Shi X, Chen H.;1; Facile synthesis of thermally stable

poly(N-vinylpyrrolidone)-modified gold surfaces by surface-initiated atom transfer radical polymerization.

Langmuir 2012;28:9451-9.

[41] Au-Duong AN, Vo DT, Lee CK.;1; Bactericidal magnetic nanoparticles with iodine loaded on

surface grafted poly(N-vinylpyrrolidone). J Mater Chem B 2015;3:840-8.

[42] Fantin M, Isse AA, Venzo A, Gennaro A, Matyjaszewski K.;1; Atom transfer radical polymerization

of methacrylic acid: A won challenge. J Am Chem Soc 2016;138:7216-9.

46
[43] Messndes JP, Mendonca PV, Maximiano P, Abreu CMR, Guliashvili T, Serra AC, Coelho JFJ.;1;

Getting faster: low temperature copper-mediated SARA ATRP of methacrylates, acrylates, styrene and vinyl

chloride in polar media using sulfolane/water mixtures. RSC Adv 2016;6:9598-603.

[44] Mazzotti G, Benelli T, Lanzi M, Mazzocchetti L, Giorgini L.;1; Straightforward synthesis of well-

defined poly(vinyl acetate) and its block copolymers by atom transfer radical polymerization. Eur Polym J

2016;77:75-87.

[45] Coca S, Matyjaszewski K.;1; Block copolymers by transformation of “living” carbocationic into

“living” radical polymerization. Macromolecules 1997;30:2808-10.

[46] Coca S, Matyjaszewski K.;1; Block copolymers by transformation of “living” carbocationic into

“living” radical polymerization. II. ABA-type block copolymers comprising rubbery polyisobutene middle

segment. J Polym Sci Part A Polym Chem 1997;35:3595-601.

[47] Coca S, Paik HJ, Matyjaszewski K.;1; Block copolymers by transformation of living ring-opening

metathesis polymerization into controlled/“living” atom transfer radical polymerization. Macromolecules

1997;30:6513-6.

[48] Paik HJ, Teodorescu M, Xia J, Matyjaszewski K.;1; Block copolymerizations of vinyl acetate by

combination of conventional and atom transfer radical polymerization. Macromolecules 1999;32:7023-31.

47
[49] Acar MH, Matyjaszewski K.;1; Block copolymers by transformation of living anionic polymerization

into controlled/“living” atom transfer radical polymerization. Macromol Chem Phys 1999;200:1094-100.

[50] Kajiwara A, Matyjaszewski K.;1; Formation of block copolymers by transformation of cationic ring-

opening polymerization to atom transfer radical polymerization (ATRP). Macromolecules 1998;31:3489-93.

[51] Matyjaszewski K, Jakubowski W, Min K, Tang W, Huang JY, Braunecker WA, Tsarevsky NV.;1;

Diminishing catalyst concentration in atom transfer radical polymerization with reducing agents. P Natl Acad

Sci USA 2006;103:15309-14.

[52] Matyjaszewski K, Dong H, Jakubowski W, Pietrasik J, Kusumo A.;1; Grafting from surfaces for

"Everyone": ARGET ATRP in the presence of air. Langmuir 2007;23:4528-31.

[53] Dong H, Matyjaszewski K.;1; ARGET ATRP of 2-(dimethylamino)ethyl methacrylate as an intrinsic

reducing agent. Macromolecules 2008;41:6868-70.

[54] Stoffelbach F, Griffete N, Bui C, Charleux B.;1; Use of a simple surface-active initiator in

controlled/living free-radical miniemulsion polymerization under AGET and ARGET ATRP conditions. Chem

Commun 2008 :4807-9.

[55] Chan N, Cunningham MF, Hutchinson RA.;1; ARGET ATRP of methacrylates and acrylates with

stoichiometric ratios of ligand to copper. Macromol Chem Phys 2008;209:1797-805.

48
[56] Jakubowski W, Kirci-Denizli B, Gil RR, Matyjaszewski K.;1; Polystyrene with improved chain-end

functionality and higher molecular weight by ARGET ATRP. Macromol Chem Phys 2008;209:32-9.

[57] Kwak Y, Matyjaszewski K.;1; ARGET ATRP of methyl methacrylate in the presence of nitrogen-

based ligands as reducing agents. Polym Int 2009;58:242-7.

[58] Jonsson M, Nystrom D, Nordin O, Malmstrom E.;1; Surface modification of thermally expandable

microspheres by grafting poly(glycidyl methacrylate) using ARGET ATRP. Eur Polym J 2009;45:2374-82.

[59] Hansson S, Ostmark E, Carlmark A, Malmstrom E.;1; ARGET ATRP for versatile grafting of

cellulose using various monomers. ACS Appl Mater Interfaces 2009;1:2651-9.

[60] Chen H, Yang L, Liang Y, Hao Z, Lu Z.;1; ARGET ATRP of acrylonitrile catalyzed by

FeCl3/isophthalic acid in the presence of air. J Polym Sci Part A Polym Chem 2009;47:3202-7.

[61] Nicolay R, Kwak Y, Matyjaszewski K.;1; A green route to well-defined high-molecular-weight

(co)polymers using ARGET ATRP with alkyl pseudohalides and copper catalysis. Angew Chem Int Ed

2010;49:541-4.

[62] Magenau AJD, Kwak Y, Matyjaszewski K.;1; ATRP of methacrylates utilizing CuIIX2/L and copper

wire. Macromolecules 2010;43:9682-9.

[63] Kwak Y, Magenau AJD, Matyjaszewski K.;1; ARGET ATRP of methyl acrylate with inexpensive

ligands and ppm concentrations of catalyst. Macromolecules 2011;44:811-9.

49
[64] Simakova A, Averick SE, Konkolewicz D, Matyjaszewski K.;1; Aqueous ARGET ATRP.

Macromolecules 2012;45:6371-9.

[65] Cheesman BT, Willott JD, Webber GB, Edmondson S,;1; Wanless EJ. pH-Responsive brush-

modified silica hybrids synthesized by surface-initiated ARGET ATRP. ACS Macro Lett 2012;1:1161-5.

[66] Król P, Chmielarz P.;1; Synthesis of PMMA-b-PU-b-PMMA tri-block copolymers through ARGET

ATRP in the presence of air. Express Polym Lett 2013;7:249-60.

[67] Payne KA, D’hooge DR, Van Steenberge PHM, Reyniers MF, Cunningham MF, Hutchinson RA,

Marin GB.;1; ARGET ATRP of butyl methacrylate: Utilizing kinetic modeling to understand experimental

trends. Macromolecules 2013;46:3828-40.

[68] Król P, Chmielarz P.;1; Synthesis of poly(urethane-methacrylate) copolymers using

tetraphenylethane-urethane macroinitiator by ARGET ATRP controlled polymerization method. Polimery

2014;59:279-92.

[69] Williams VA, Ribelli TG, Chmielarz P, Park S, Matyjaszewski K.;1; A silver bullet: Elemental silver

as an efficient reducing agent for atom transfer radical polymerization of acrylates. J Am Chem Soc

2015;137:1428-31.

50
[70] Król P, Chmielarz P.;1; Synthesis of polystyrene-b-polyurethane-b-polystyrene tri-block copolymers

through ARGET ATRP polymerization method. Part I. Synthesis of macroinitiator and polymerization

mechanism. Polimery 2015;60:316-21.

[71] Lan T, Torkelson JM.;1; Substantial spatial heterogeneity and tunability of glass transition

temperature observed with dense polymer brushes prepared by ARGET ATRP. Polymer 2015;64:183-92.

[72] Mueller L, Jakubowski W, Tang W, Matyjaszewski K.;1; Successful chain extension of polyacrylate

and polystyrene macroinitiators with methacrylates in an ARGET and ICAR ATRP. Macromolecules

2007;40:6464-72.

[73] Plichta A, Li WW, Matyjaszewski K.;1; ICAR ATRP of styrene and methyl methacrylate with

Ru(Cp*)Cl(PPh3)(2). Macromolecules 2009;42:2330-2.

[74] Zhang L, Miao J, Cheng Z, Zhu X.;1; Iron-mediated ICAR ATRP of styrene and methyl

methacrylate in the absence of thermal radical initiator. Macromol Rapid Commun 2010;31:275-80.

[75] Zhu G, Zhang L, Zhang Z, Zhu J, Tu Y, Cheng Z, Zhu X.;1; Iron-mediated ICAR ATRP of methyl

methacrylate. Macromolecules 2011;44:3233-9.

[76] D'hooge DR, Konkolewicz D, Reyniers MF, Marin GB, Matyjaszewski K.;1; Kinetic modeling of

ICAR ATRP. Macromol Theor Simul 2012;21:52-69.

51
[77] Toloza Porras C, D'Hooge DR, Reyniers MF, Marin GB.;1; Computer-aided optimization of

conditions for fast and controlled ICAR ATRP of n-butyl acrylate. Macromol Theor Simul 2013;22:136-49.

[78] Jiang X, Wu J, Zhang L, Cheng Z, Zhu X.;1; Highly active ppm level organic copper catalyzed

photo-induced ICAR ATRP of methyl methacrylate. Macromol Rapid Commun 2014;35:1879-85.

[79] Matyjaszewski K, Coca S, Gaynor SG, Wei M, Woodworth BE.;1; Zerovalent metals in

controlled/\"living\" radical polymerization. Macromolecules 1997;30:7348-50.

[80] Matyjaszewski K, Tsarevsky NV, Braunecker WA, Dong H, Huang J, Jakubowski W, Kwak Y,

Nicolay R, Tang W, Yoon JA.;1; Role of Cu-0 in controlled/"living" radical polymerization. Macromolecules

2007;40:7795-806.

[81] Zhang Y, Zhang W, Chen X, Cheng Z, Wu J, Zhu J, Zhu X.;1; Synthesis of novel three-arm star azo

side-chain liquid crystalline polymer via ATRP and photoinduced surface relief gratings. J Polym Sci Part A

Polym Chem 2008;46:777-89.

[82] Wang Y, Zhang Y, Parker B, Matyjaszewski K.;1; ATRP of MMA with ppm levels of iron catalyst.

Macromolecules 2011;44:4022-5.

[83] Abreu CMR, Serra AC, Popov AV, Matyjaszewski K, Guliashvili T, Coelho JFJ.;1; Ambient

temperature rapid SARA ATRP of acrylates and methacrylates in alcohol-water solutions mediated by a mixed

sulfite/Cu(II)Br-2 catalytic system. Polym Chem 2013;4:5629-36.

52
[84] Konkolewicz D, Wang Y, Zhong MJ, Krys P, Isse AA, Gennaro A, Matyjaszewski K.;1; Reversible-

deactivation radical polymerization in the presence of metallic copper. A critical assessment of the SARA

ATRP and SET-LRP mechanisms. Macromolecules 2013;46:8749-72.

[85] Cordeiro RA, Rocha N, Mendes JP, Matyjaszewski K, Guliashvili T, Serra AC, Coelho JFJ.;1;

Synthesis of well-defined poly(2-(dimethylamino)ethyl methacrylate) under mild conditions and its co-

polymers with cholesterol and PEG using Fe(0)/Cu(II) based SARA ATRP. Polym Chem 2013;4:3088-97.

[86] Wang Y, Zhong M, Zhu W, Peng CH, Zhang Y, Konkolewicz D, Bortolamei N, Isse AA, Gennaro

A, Matyjaszewski K.;1; Reversible-deactivation radical polymerization in the presence of metallic copper.

Comproportionation–disproportionation equilibria and kinetics. Macromolecules 2013;46:3793-802.

[87] Peng CH, Zhong M, Wang Y, Kwak Y, Zhang Y, Zhu W, Tonge M, Buback J, Park S, Krys P,

Konkolewicz D, Gennaro A, Matyjaszewski K.;1; Reversible-deactivation radical polymerization in the

presence of metallic copper. Activation of alkyl halides by Cu0. Macromolecules 2013;46:3803-15.

[88] Zhou YN, Luo ZH.;1; Facile synthesis of gradient copolymers via semi-batch copper(0)-mediated

living radical copolymerization at ambient temperature Polym Chem 2013;4:76-84.

[89] Gois JR, Konkolewicz D, Popov AV, Guliashvili T, Matyjaszewski K, Serra AC, Coelho JFJ.;1;

Improvement of the control over SARA ATRP of 2-(diisopropylamino)ethyl methacrylate by slow and

continuous addition of sodium dithionite. Polym Chem 2014;5:4617-26.

53
[90] Konkolewicz D, Krys P, Gois JR, Mendonca PV, Zhong MJ, Wang Y, Gennaro A, Isse AA, Fantin

M, Matyjaszewski K.;1; Aqueous RDRP in the presence of Cu-0: The exceptional activity of Cu-I confirms the

SARA ATRP mechanism. Macromolecules 2014;47:560-70.

[91] Konkolewicz D, Wang Y, Krys P, Zhong MJ, Isse AA, Gennaro A, Matyjaszewski K.;1; SARA

ATRP or SET-LRP. End of controversy? Polym Chem 2014;5:4396-417.

[92] Mendonca PV, Konkolewicz D, Averick SE, Serra AC, Popov AV, Guliashvili T, Matyjaszewski K,

Coelho JFJ.;1; Synthesis of cationic poly((3-acrylamidopropyl)-trimethylammonium chloride) by SARA ATRP

in ecofriendly solvent mixtures. Polym Chem 2014;5:5829-36.

[93] Zhou YN, Luo ZH.;1; Copper(0)-mediated reversible-deactivation radical polymerization: Kinetics

insight and experimental study. Macromolecules 2014;47:6218-29.

[94] Chmielarz P, Krys P, Park S, Matyjaszewski K.;1; PEO-b-PNIPAM copolymers via SARA ATRP

and eATRP in aqueous media. Polymer 2015;71:143-7.

[95] Costa JRC, Mendonça PV, Maximiano P, Serra AC, Guliashvili T, Coelho JFJ.;1; Ambient

temperature “flash” SARA ATRP of methyl acrylate in water/ionic liquid/glycol mixtures. Macromolecules

2015;48:6810-5.

[96] Zhou YN, Luo ZH.;1; An old kinetic method for a new polymerization mechanism: Toward

photochemically mediated ATRP. AIChE J 2015;61:1947-58.

54
[97] Chmielarz P, Krol P.;1; PSt-b-PU-b-PSt copolymers using tetraphenylethane-urethane macroinitiator

through SARA ATRP. Express Polym Lett 2016;10:302-10.

[98] Erel I, Cianga I, Serhatli E, Yagci Y.;1; Synthesis of block copolymers by combination of

photoinduced and atom transfer radical polymerization routes. Eur Polym J 2002;38:1409-15.

[99] Ishizu K, Kakinuma H.;1; Synthesis of nanocylinders consisting of graft block copolymers by the

photo-induced ATRP technique. J Polym Sci Part A Polym Chem 2005;43:63-70.

[100] Wang G, Zhu X, Wu J, Zhu J, Chen X, Cheng Z.;1; Synthesis and photoinduced surface-relief

grating of well-defined azo-containing polymethacrylates via atom transfer radical polymerization. J Appl

Polym Sci 2007;106:1234-42.

[101] Ishizu K, Kayakawa N, Takano S, Tokuno Y, Ozawa M.;1; Synthesis of polymer particles

possessing radical initiating sites on the surface by emulsion copolymerization and construction of core-shell

structures by a photoinduced atom transfer radical polymerization approach. J Polym Sci Part A Polym Chem

2007;45:1771-7.

[102] Ishizu K, Murakami T, Takano S.;1; Architecture of brush-on-brush copolymers by photoinduced

ATRP approach. J Colloid Interface Sci 2008;322:59-64.

[103] Tasdelen MA, Uygun M, Yagci Y.;1; Studies on photoinduced ATRP in the presence of

photoinitiator. Macromol Chem Phys 2011;212:2036-42.

55
[104] Tasdelen MA, Ciftci M, Yagci Y.;1; Visible light-induced atom transfer radical polymerization.

Macromol Chem Phys 2012;213:1391-6.

[105] Yan J, Li B, Zhou F, Liu W.;1; Ultraviolet light-induced surface-initiated atom-transfer radical

polymerization. ACS Macro Lett 2013;2:592-6.

[106] Taskin OS, Yilmaz G, Tasdelen MA, Yagci Y.;1; Photoinduced reverse atom transfer radical

polymerization of methyl methacrylate using camphorquinone/benzhydrol system. Polym Int 2014;63:902-7.

[107] Dadashi-Silab S, Tasdelen MA, Asiri AM, Khan SB, Yagci Y.;1; Photoinduced atom transfer

radical polymerization using semiconductor nanoparticles. Macromol Rapid Commun 2014;35:454-9.

[108] Ribelli TG, Konkolewicz D, Pan XC, Matyjaszewski K.;1; Contribution of photochemistry to

activator regeneration in ATRP. Macromolecules 2014;47:6316-21.

[109] Zhang T, Chen T, Amin I, Jordan R.;1; ATRP with a light switch: photoinduced ATRP using a

household fluorescent lamp. Polym Chem 2014;5:4790-6.

[110] Ribelli TG, Konkolewicz D, Bernhard S, Matyjaszewski K.;1; How are radicals (re)generated in

photochemical ATRP? J Am Chem Soc 2014;136:13303-12.

[111] Pan X, Lamson M, Yan J, Matyjaszewski K.;1; Photoinduced metal-free atom transfer radical

polymerization of acrylonitrile. ACS Macro Lett 2015;4:192-6.

56
[112] Yang Q, Dumur F, Morlet-Savary F, Poly J, Lalevée J.;1; Photocatalyzed Cu-based ATRP

involving an oxidative quenching mechanism under visible light. Macromolecules 2015;48:1972-80.

[113] Doran S, Yagci Y.;1; Graft polymer growth using tandem photoinduced photoinitiator-free

CuAAC/ATRP. Polym Chem 2015;6:946-52.

[114] Liu X, Zhang L, Cheng Z, Zhu X.;1; Metal-free photoinduced electron transfer-atom transfer radical

polymerization (PET-ATRP) via a visible light organic photocatalyst. Polym Chem 2016;7:689-700.

[115] Pan X, Fang C, Fantin M, Malhotra N, So WY, Peteanu LA, Isse AA, Gennaro A, Liu P,

Matyjaszewski K.;1; Mechanism of photoinduced metal-free atom transfer radical polymerization:

Experimental and computational studies. J Am Chem Soc 2016;138:2411-25.

[116] Amado E, Kressler J.;1; Reversible complexation of iminophenylboronates with mono- and

dihydroxy methacrylate monomers and their polymerization at low temperature by photoinduced ATRP in one

pot. Macromolecules 2016;49:1532-44.

[117] Huang Z, Feng C, Guo H, Huang X.;1; Direct functionalization of poly(vinyl chloride) by photo-

mediated ATRP without a deoxygenation procedure. Polym Chem 2016;7:3034-45.

[118] Zhou YN, Guo JK, Li JJ, Luo ZH.;1; Photoinduced iron(III)-mediated atom transfer radical

polymerization with in situ generated initiator: Mechanism and kinetics studies. Ind Eng Chem Res

2016;55:10235-42.

57
[119] Magenau AJD, Strandwitz NC, Gennaro A, Matyjaszewski K.;1; Electrochemically mediated atom

transfer radical polymerization. Science 2011;332:81-4.

[120] Bortolamei N, Isse AA, Magenau AJD, Gennaro A, Matyjaszewski K.;1; Controlled aqueous atom

transfer radical polymerization with electrochemical generation of the active catalyst. Angew Chem Int Ed

2011;50:11391-4.

[121] Li B, Yu B, Huck WTS, Zhou F, Liu W.;1; Electrochemically induced surface-initiated atom-

transfer radical polymerization. Angew Chem Int Ed 2012;51:5092-5.

[122] Magenau AJD, Bortolamei N, Frick E, Park S, Gennaro A, Matyjaszewski K.;1; Investigation of

electrochemically mediated atom transfer radical polymerization. Macromolecules 2013;46:4346-53.

[123] Li B, Yu B, Huck WTS, Liu W, Zhou F.;1; Electrochemically mediated atom transfer radical

polymerization on nonconducting substrates: Controlled brush growth through catalyst diffusion. J Am Chem

Soc 2013;135:1708-10.

[124] Park S, Cho HY, Wegner KB, Burdynska J, Magenau AJD, Paik HJ, Jurga S, Matyjaszewski K.;1;

Star synthesis using macroinitiators via electrochemically mediated atom transfer radical polymerization.

Macromolecules 2013;46:5856-60.

[125] Jin GP, Fu Y, Bao XC, Feng XS, Wang Y, Liu WH.;1; Electrochemically mediated atom transfer

radical polymerization of iminodiacetic acid-functionalized poly(glycidyl methacrylate)grafted at carbon fibers

for nano-nickel recovery from spent electroless nickel plating baths. J Appl Electrochem 2014;44:621-9.

58
[126] Park S, Chmielarz P, Gennaro A, Matyjaszewski K.;1; Simplified electrochemically mediated atom

transfer radical polymerization using a sacrificial anode. Angew Chem Int Ed 2015;54:2388-92.

[127] Chmielarz P, Park S, Simakova A, Matyjaszewski K.;1; Electrochemically mediated ATRP of

acrylamides in water. Polymer 2015;60:302-7.

[128] Guo JK, Zhou YN, Luo ZH.;1; Kinetic insight into electrochemically mediated ATRP gained

through modeling. AIChE J 2015;61:4347-57.

[129] Chmielarz P, Sobkowiak A, Matyjaszewski K.;1; A simplified electrochemically mediated ATRP

synthesis of PEO-b-PMMA copolymers. Polymer 2015;77:266-71.

[130] Chmielarz P, Park S, Sobkowiak A, Matyjaszewski K.;1; Synthesis of β-cyclodextrin-based star

polymers via a simplified electrochemically mediated ATRP. Polymer 2016;88:36-42.

[131] Fantin M, Lorandi F, Isse AA, Gennaro A.;1; Sustainable electrochemically-mediated atom transfer

radical polymerization with inexpensive non-platinum electrodes. Macromol Rapid Commun 2016;37:1318-22.

[132] Lorandi F, Fantin M, Isse AA, Gennaro A.;1; Electrochemically mediated atom transfer radical

polymerization of n-butyl acrylate on non-platinum cathodes. Polym Chem 2016;7:5357-65.

[133] Chmielarz P.;1; Synthesis of α-D-glucose-based star polymers through simplified electrochemically

mediated ATRP. Polymer 2016;102:192-8.

59
[134] Strover LT, Malmström J, Stubbing LA, Brimble MA, Travas-Sejdic J.;1; Electrochemically-

controlled grafting of hydrophilic brushes from conducting polymer substrates. Electrochim Acta 2016;188:57-

70.

[135] Guo JK, Zhou YN, Luo ZH.;1; Kinetic insights into the iron-based electrochemically mediated atom

transfer radical polymerization of methyl methacrylate. Macromolecules 2016;49:4038-46.

[136] Chmielarz P.;1; Cellulose-based graft copolymers prepared by simplified electrochemically

mediated ATRP. Express Polym Lett 2017;11:140-51.

[137] Chmielarz P.;1; Synthesis of multiarm star block copolymers via simplified electrochemically

mediated ATRP. Chem Pap 2017;71:161-70.

[138] Listak J, Jakubowski W, Mueller L, Plichta A, Matyjaszewski K, Bockstaller MR.;1; Effect of

symmetry of molecular weight distribution in block copolymers on formation of “metastable” morphologies.

Macromolecules 2008;41:5919-27.

[139] Hosseiny SS, van Rijn P.;1; Surface initiated polymerizations via e-ATRP in pure water. Polymers

2013;5:1229-40.

[140] Li B, Yu B, Zhou F.;1; In situ AFM investigation of electrochemically induced surface-initiated

atom-transfer radical polymerization. Macromol Rapid Commun 2013;34:246-50.

60
[141] Plamper F.;1; Polymerizations under electrochemical control. Colloid Polym Sci 2014;292:777-83.

[142] Shida N, Koizumi Y, Nishiyama H, Tomita I, Inagi S.;1; Electrochemically mediated atom transfer

radical polymerization from a substrate surface manipulated by bipolar electrolysis: Fabrication of gradient and

patterned polymer brushes. Angew Chem Int Ed 2015;54:3922-6.

[143] Chmielarz P.;1; Synthesis of cationic star polymers by simplified electrochemically mediated

ATRP. Express Polym Lett 2016;10:810-21.

[144] Lin CY, Coote ML, Gennaro A, Matyjaszewski K.;1; 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;130:12762-74.

[145] Krys P, Ribelli TG, Matyjaszewski K, Gennaro A.;1; Relation between overall rate of ATRP and

rates of activation of dormant species. Macromolecules 2016;49:2467-76.

[146] Bortolamei N, Isse AA, Gennaro A.;1; Estimation of standard reduction potentials of alkyl radicals

involved in atom transfer radical polymerization. Electrochim Acta 2010;55:8312-8.

[147] Isse AA, Lin CY, Coote ML, Gennaro A.;1; Estimation of standard reduction potentials of halogen

atoms and alkyl halides. J Phys Chem B 2011;115:678-84.

[148] Kamigaito M, Ando T, Sawamoto M.;1; Metal-catalyzed living radical polymerization. Chem Rev

2001;101:3689-746.

61
[149] Ouchi M, Terashima T, Sawamoto M.;1; Transition metal-catalyzed living radical polymerization:

Toward perfection in catalysis and precision polymer synthesis. Chem Rev 2009;109:4963-5050.

[150] di Lena F, Matyjaszewski K.;1; Transition metal catalysts for controlled radical polymerization.

Prog Polym Sci 2010;35:959-1021.

[151] Kwiatkowski P, Jurczak J, Pietrasik J, Jakubowski W, Mueller L, Matyjaszewski K.;1; High

molecular weight polymethacrylates by AGET ATRP under high pressure. Macromolecules 2008;41:1067-9.

[152] Magenau AJD, Kwak Y, Schröder K, Matyjaszewski K.;1; Highly active bipyridine-based ligands

for atom transfer radical polymerization. ACS Macro Lett 2012;1:508-12.

[153] Horn M, Matyjaszewski K.;1; Solvent effects on the activation rate constant in atom transfer radical

polymerization. Macromolecules 2013;46:3350-7.

[154] Fischer H.;1; The persistent radical effect: a principle for selective radical reactions and living

radical polymerizations. Chem Rev 2001;101:3581-610.

[155] Tang W, Tsarevsky NV, Matyjaszewski K.;1; Determination of equilibrium constants for atom

transfer radical polymerization. J Am Chem Soc 2006;128:1598-604.

[156] Shen Y, Tang H, Ding S.;1; Catalyst separation in atom transfer radical polymerization. Prog Polym

Sci 2004;29:1053-78.

62
[157] Matyjaszewski K, Miller PJ, Pyun J, Kickelbick G, Diamanti S.;1; Synthesis and characterization of

star polymers with varying arm number, length, and composition from organic and hybrid inorganic/organic

multifunctional initiators. Macromolecules 1999;32:6526-35.

[158] Kickelbick G, Paik HJ, Matyjaszewski K.;1; Immobilization of the copper catalyst in atom transfer

radical polymerization. Macromolecules 1999;32:2941-7.

[159] Pruitt JD, Husseini G, Rapoport N, Pitt WG.;1; Stabilization of pluronic P-105 micelles with an

interpenetrating network of N,N-diethylacrylamide. Macromolecules 2000;33:9306-9.

[160] Honigfort ME, Brittain WJ, Bosanac T, Wilcox CS.;1; Use of precipitons for copper removal in

atom transfer radical polymerization. Macromolecules 2002;35:4849-51.

[161] Hong SC, Matyjaszewski K.;1; Fundamentals of supported catalysts for atom transfer radical

polymerization (ATRP) and application of an immobilized/soluble hybrid catalyst system to ATRP.

Macromolecules 2002;35:7592-605.

[162] Sarbu T, Pintauer T, McKenzie B, Matyjaszewski K.;1; Atom transfer radical polymerization of

styrene in toluene/water mixtures. J Polym Sci Part A Polym Chem 2002;40:3153-60.

[163] Yang J, Ding S, Radosz M, Shen Y.;1; Reversible catalyst supporting via hydrogen-bonding-

mediated self-assembly for atom transfer radical polymerization of MMA. Macromolecules 2004;37:1728-34.

63
[164] Barré G, Taton D, Lastécouères D, Vincent JM.;1; Closer to the “ideal recoverable catalyst” for

atom transfer radical polymerization using a molecular non-fluorous thermomorphic system. J Am Chem Soc

2004;126:7764-5.

[165] Faucher S, Okrutny P, Zhu S.;1; Facile and effective purification of polymers produced by atom

transfer radical polymerization via simple catalyst precipitation and microfiltration. Macromolecules 2006;39:3-

5.

[166] Duquesne E, Labruyère C, Habimana J, Degée P, Dubois P.;1; Copper-based supported catalysts for

the atom transfer radical polymerization of methyl methacrylate: How can activity and control be tuned up? J

Polym Sci Part A Polym Chem 2006;44:744-56.

[167] Munirasu S, Deshpande A, Baskaran D.;1; Hydrated clay for catalyst removal in copper mediated

atom transfer radical polymerization. Macromol Rapid Commun 2008;29:1538-43.

[168] Jakubowski W, Min K, Matyjaszewski K.;1; Activators regenerated by electron transfer for atom

transfer radical polymerization of styrene. Macromolecules 2006;39:39-45.

[169] Jakubowski W, Matyjaszewski K.;1; Activators regenerated by electron transfer for atom-transfer

radical polymerization of (meth)acrylates and related block copolymers. Angew Chem Int Ed 2006;45:4482-6.

[170] Min K, Gao H, Matyjaszewski K.;1; Use of ascorbic acid as reducing agent for synthesis of well-

defined polymers by ARGET ATRP. Macromolecules 2007;40:1789-91.

64
[171] Abreu CMR, Mendonca PV, Serra AC, Popov AV, Matyjaszewski K, Guliashvili T, Coelho JFJ.;1;

Inorganic sulfites: Efficient reducing agents and supplemental activators for atom transfer radical

polymerization. ACS Macro Lett 2012;1:1308−11.

[172] Woodruff SR, Davis BJ, Tsarevsky NV.;1; Epoxides as reducing agents for low-catalyst-

concentration atom transfer radical polymerization. Macromol Rapid Commun 2014;35:186-92.

[173] Zhang Y, Wang Y, Matyjaszewski K.;1; ATRP of methyl acrylate with metallic zinc, magnesium,

and iron as reducing agents and supplemental activators. Macromolecules 2011;44:683-5.

[174] Mueller L, Matyjaszewski K.;1; Reducing copper concentration in polymers prepared via atom

transfer radical polymerization. Macromol React Eng 2010;4:180-5.

[175] Saraç AS, Erbil C, Soydan AB.;1; Polymerization of acrylamide initiated with electrogenerated

cerium (IV) in the presence of EDTA. J Appl Polym Sci 1992;44:877-81.

[176] Saraç AS, Soydan AB, Coka V.;1; Aqueous polymerization of acrylamide by electrolitically

generated KMnO4 organic acid redox systems. J Appl Polym Sci 1996;62:111-6.

[177] Saraç AS.;1; Redox polymerization. Prog Polym Sci 1999;24:1149-204.

[178] Jenkins DW, Hudson SM.;1; Review of vinyl graft copolymerization featuring recent advances

toward controlled radical-based reactions and illustrated with chitin/chitosan trunk polymers. Chem Rev

2001;101:3245-74.

65
[179] Cosnier S, Karyakin A.;1; Electropolymerization: concepts, materials and applications. New York:

John Wiley & Sons Inc, 2011. p. 1-26.

[180] Reuber J, Reinhardt H, Johannsmann D.;1; Formation of surface-attached responsive gel layers via

electrochemically induced free-radical polymerization. Langmuir 2006;22:3362-7.

[181] Bünsow J, Mänz M, Vana P, Johannsmann D.;1; Electrochemically induced RAFT polymerization

of thermoresponsive hydrogel films: Impact on film thickness and surface morphology. Macromol Chem Phys

2010;211:761-7.

[182] Bonometti V, Labbe E, Buriez O, Mussini P, Amatore C.;1; Exploring the first steps of an

electrochemically-triggered controlled polymerization sequence: Activation of alkyl- and benzyl halide

initiators by an electrogenerated Fe(II)Salen complex. J Electroanal Chem 2009;633:99-105.

[183] Testa AC, Reinmuth WH.;1; Chronopotentiometry with step-functional changes in current. Anal

Chem 1961;33:1324-8.

[184] Testa AC, Reinmuth WH.;1; Stepwise reactions in chronopotentiometry. Anal Chem 1961;33:1320-

4.

[185] Bard AJ, Faulkner LR.;1; Electrochemical methods: fundamentals and applications 2nd ed. New

York: John Wiley & Sons Inc, 2001. p. 87-155.

66
[186] Wang Y, Zhong MJ, Zhang YZ, Magenau AJD, Matyjaszewski K.;1; Halogen conservation in atom

transfer radical polymerization. Macromolecules 2012;45:8929-32.

[187] Inzelt G, Lewenstam A, Scholz F, Baucke FG.;1; Handbook of reference electrodes.

Berlin/Heidelberg: Springer, 2013. p. 77-188.

[188] Chmielarz P, Sobkowaik A.;1; Synthesis of urethane-acrylic multiblock copolymers via

electrochemically mediated ATRP. Chem Pap 2016;70:1228-37.

[189] Chmielarz P, Sobkowiak A.;1; Synthesis of poly(butyl acrylate) using a electrochemically mediated

ATRP. Polimery 2016;61:585-90.

[190] Matyjaszewski K, Bortolamei N, Magenau A, Gennaro A, Isse AA.;1; Electrochemically mediated

tom transfer radical polymerization. US 2014/0183055, 2014.

[191] Qiu J, Matyjaszewski K, Thouin L, Amatore C.;1; Cyclic voltammetric studies of copper complexes

catalyzing atom transfer radical polymerization. Macromol Chem Phys 2000;201:1625-31.

[192] Isse AA, Gennaro A, Lin CY, Hodgson JL, Coote ML, Guliashvili T.;1; Mechanism of

carbon−halogen bond reductive cleavage in activated alkyl halide initiators relevant to living radical

polymerization: Theoretical and experimental study. J Am Chem Soc 2011;133:6254-64.

[193] Fantin M, Isse AA, Gennaro A, Matyjaszewski K.;1; Understanding the fundamentals of aqueous

ATRP and defining conditions for better control. Macromolecules 2015;48:6862-75.

67
[194] Braunecker WA, Brown WC, Morelli BC, Tang W, Poli R, Matyjaszewski K.;1; Origin of activity

in Cu-, Ru-, and Os-mediated radical polymerization. Macromolecules 2007;40:8576-85.

[195] De Paoli P, Isse AA, Bortolamei N, Gennaro A.;1; New insights into the mechanism of activation of

atom transfer radical polymerization by Cu(I) complexes. Chem Commun 2011;47:3580-2.

[196] Bortolamei N, Isse AA, Di Marco VB, Gennaro A, Matyjaszewski K.;1; Thermodynamic properties

of copper complexes used as catalysts in atom transfer radical polymerization. Macromolecules 2010;43:9257-

67.

[197] Lanzalaco S, Fantin M, Scialdone O, Galia A, Isse AA, Gennaro A, Matyjaszewski K.;1; Atom

transfer radical polymerization with different halides (F, Cl, Br, and I): Is the process “living” in the presence of

fluorinated initiators? Macromolecules 2017;50:192-202.

[198] Tang W, Matyjaszewski K.;1; Effect of ligand structure on activation rate constants in ATRP.

Macromolecules 2006;39:4953-9.

[199] Pintauer T, Matyjaszewski K.;1; Structural aspects of copper catalyzed atom transfer radical

polymerization. Coord Chem Rev 2005;249:1155-84.

[200] Schröder K, Mathers RT, Buback J, Konkolewicz D, Magenau AJD, Matyjaszewski K.;1;

Substituted tris(2-pyridylmethyl)amine ligands for highly active ATRP catalysts. ACS Macro Lett 2012;1:1037-

40.

68
[201] Bell CA, Bernhardt PV, Monteiro MJ.;1; A rapid electrochemical method for determining rate

coefficients for copper-catalyzed polymerizations. J Am Chem Soc 2011;133:11944-7.

[202] Zerk TJ, Bernhardt PV.;1; Solvent dependent anion dissociation limits copper(i) catalysed atom

transfer reactions. Dalton T 2013;42:11683-94.

[203] Isse AA, Bortolamei N, De Paoli P, Gennaro A.;1; On the mechanism of activation of copper-

catalyzed atom transfer radical polymerization. Electrochim Acta 2013;110:655-62.

[204] Zerk TJ, Bernhardt PV.;1; New method for exploring deactivation kinetics in copper-catalyzed

atom-transfer-radical reactions. Inorg Chem 2014;53:11351-3.

[205] Lorandi F, Fantin M, Isse AA, Gennaro A.;1; RDRP in the presence of Cu0: The fate of Cu(I)

proves the inconsistency of SET-LRP mechanism. Polymer 2015;72:238-45.

[206] Fantin M, Isse AA, Bortolamei N, Matyjaszewski K, Gennaro A.;1; Electrochemical approaches to

the determination of rate constants for the activation step in atom transfer radical polymerization. Electrochim

Acta 2016;220:393-401.

[207] Soerensen N, Barth J, Buback M, Morick J, Schroeder H, Matyjaszewski K.;1; SP-PLP-EPR

measurement of ATRP deactivation rate. Macromolecules 2012;45:3797-801.

69
[208] Tobias C, Eisenberg M, Wilke C.;1; Fiftieth anniversary: diffusion and convection in electrolysis - a

theoretical review. J Electrochem Soc 1952;99:359C-65C.

[209] Braunecker WA, Tsarevsky NV, Gennaro A, Matyjaszewski K.;1; Thermodynamic components of

the atom transfer radical polymerization equilibrium: quantifying solvent effects. Macromolecules

2009;42:6348-60.

[210] Tsarevsky NV, Pintauer T, Matyjaszewski K.;1; Deactivation efficiency and degree of control over

polymerization in ATRP in protic solvents. Macromolecules 2004;37:9768-78.

[211] Millard PE, Mougin NC, Boeker A, Mueller AHE.;1; Controlling the fast ATRP of N-

isopropylacrylamide in water. ACS Symp Ser 2009;1023:127-37.

[212] Zhang Q, Wilson P, Li ZD, McHale R, Godfrey J, Anastasaki A, Waldron C, Haddleton DM.;1;

Aqueous copper-mediated living polymerization: Exploiting rapid disproportionation of CuBr with Me6TREN.

J Am Chem Soc 2013;135:7355-63.

[213] Anastasaki A, Haddleton AJ, Zhang Q, Simula A, Droesbeke M, Wilson P, Haddleton DM.;1;

Aqueous copper-mediated living radical polymerisation of N-acryloylmorpholine, SET- LRP in water.

Macromol Rapid Commun 2014;35:965-70.

[214] Fantin M, Park S, Wang Y, Matyjaszewski K.;1; Electrochemical atom transfer radical

polymerization in miniemulsion with a dual catalytic system. Macromolecules 2016;49:8838-47.

70
[215] Fantin M, Chmielarz P, Wang Y, Lorandi F, Matyjaszewski K.;1; Surfactant as active reagent in

miniemulsion simplified electrochemically mediated atom transfer radical polymerization. Macromolecules

2017;submitted.

[216] Chmielarz P, Krys P, Wang Z, Wang Y, Matyjaszewski K.;1; Synthesis of well-defined polymer

brushes from silicon wafers via surface-initiated seATRP. Macromol Chem Phys 2017;submitted.

[217] Nasser-Eddine M, Delaite C, Dumas P, Vataj R, Louati A.;1; Copper removal in atom transfer

radical polymerization through electrodeposition. Macromol Mater Eng 2004;289:204-7.

[218] Pintauer T, Matyjaszewski K.;1; Atom transfer radical addition and polymerization reactions

catalyzed by ppm amounts of copper complexes. Chem Soc Rev 2008;37:1087-97.

[219] Jasinski N, Lauer A, Stals PJM, Behrens S, Essig S, Walther A, Goldmann AS, Barner-Kowollik

C.;1; Cleaning the click: A simple electrochemical avenue for copper removal from strongly coordinating

macromolecules. ACS Macro Lett 2015;4:298-301.

[220] Hu Y, Yang G, Liang B, Fang L, Ma G, Zhu Q, Chen S, Ye X.;1; The fabrication of superlow

protein absorption zwitterionic coating by electrochemically mediated atom transfer radical polymerization and

its application. Acta Biomater 2015;13:142-9.

[221] Sun Y, Du H, Deng Y, Lan Y, Feng C.;1; Preparation of polyacrylamide via surface-initiated

electrochemical-mediated atom transfer radical polymerization (SI-eATRP) for Pb2+ sensing. J Solid State

Electr 2016;20:105-13.

71
[222] Zhang J, Yi XB, Ju W, Fan HL, Wang QC, Liu BX, Liu S.;1; Hydrophilic modification of ordered

mesoporous carbons for supercapacitor via electrochemically induced surface-initiated atom-transfer radical

polymerization. Electrochem Commun 2016;74:19-23.

[223] Hu Y, Liang B, Fang L, Ma G, Yang G, Zhu Q, Chen S, Ye X.;1; Antifouling zwitterionic coating

via electrochemically mediated atom transfer radical polymerization on enzyme-based glucose sensors for long-

time stability in 37 °C serum. Langmuir 2016;32:11763-70.

[224] Hadjichristidis N, Pitsikalis M, Pispas S, Iatrou H.;1; Polymers with complex architecture by living

anionic polymerization. Chem Rev 2001;101:3747-92.

[225] Vamvakaki M, Hadjiyannakou SC, Loizidou E, Patrickios CS, Armes SP, Billingham NC.;1;

Synthesis and characterization of novel networks with nano-engineered structures:  Cross-linked star

homopolymers. Chem Mater 2001;13:4738-44.

[226] Gao H, Matyjaszewski K.;1; Synthesis of functional polymers with controlled architecture by CRP

of monomers in the presence of cross-linkers: From stars to gels. Prog Polym Sci 2009;34:317-50.

[227] Blencowe A, Tan JF, Goh TK, Qiao GG.;1; Core cross-linked star polymers via controlled radical

polymerisation. Polymer 2009;50:5-32.

72
[228] Gregory A, Stenzel MH.;1; Complex polymer architectures via RAFT polymerization: From

fundamental process to extending the scope using click chemistry and nature's building blocks. Prog Polym Sci

2012;37:38-105.

[229] Sun Y, Du H, Lan Y, Wang W, Liang Y, Feng C, Yang M.;1; Preparation of hemoglobin (Hb)

imprinted polymer by Hb catalyzed eATRP and its application in biosensor. Biosens Bioelectron 2016;77:894-

900.

[230] Mavré F, Anand RK, Laws DR, Chow KF, Chang BY, Crooks JA, Crooks RM.;1; Bipolar

electrodes: A eseful tool for concentration, separation, and detection of analytes in microelectrochemical

systems. Anal Chem 2010;82:8766-74.

[231] Ramakrishnan S, Shannon C.;1; Display of solid-state materials using bipolar electrochemistry.

Langmuir 2010;26:4602-6.

[232] Loget G, Kuhn A.;1; Shaping and exploring the micro- and nanoworld using bipolar

electrochemistry. Anal Bioanal Chem 2011;400:1691-704.

[233] Fattah Z, Loget G, Lapeyre V, Garrigue P, Warakulwit C, Limtrakul J, Bouffier L, Kuhn A.;1;

Straightforward single-step generation of microswimmers by bipolar electrochemistry. Electrochim Acta

2011;56:10562-6.

[234] Ishiguro Y, Inagi S, Fuchigami T.;1; Gradient doping of conducting polymer films by means of

bipolar electrochemistry. Langmuir 2011;27:7158-62.

73
[235] Fosdick SE, Knust KN, Scida K, Crooks RM.;1; Bipolar electrochemistry. Angew Chem Int Ed

2013;52:10438-56.

[236] Loget G, Zigah D, Bouffier L, Sojic N, Kuhn A.;1; Bipolar electrochemistry: From materials

science to motion and beyond. Acc Chem Res 2013;46:2513-23.

[237] Sun Y, Lan Y, Yang L, Kong F, Du H, Feng C.;1; Preparation of hemoglobin imprinted polymers

based on graphene and protein removal assisted by electric potential. RSC Adv 2016;6:61897-905.

[238] Patton DL, Taranekar P, Fulghum T, Advincula R.;1; Electrochemically active dendritic−linear

block copolymers via RAFT polymerization: Synthesis, characterization, and electrodeposition properties.

Macromolecules 2008;41:6703-13.

[239] Wessling B.;1; New insight into organic metal polyaniline morphology and structure. Polymers

2010;2:786-98.

[240] Friebe C, Hager MD, Winter A, Schubert US.;1; Metal-containing polymers via

electropolymerization. Adv Mater 2012;24:332-45.

[241] Li M, Zhang J, Nie HJ, Liao M, Sang L, Qiao W, Wang ZY, Ma Y, Zhong YW, Ariga K.;1; In situ

switching layer-by-layer assembly: one-pot rapid layer assembly via alternation of reductive and oxidative

electropolymerization. Chem Commun 2013;49:6879-81.

74
[242] Bardini L, Ceccato M, Hinge M, Pedersen SU, Daasbjerg K, Marcaccio M, Paolucci F.;1;

Electrochemical polymerization of allylamine copolymers. Langmuir 2013;29:3791-6.

[243] Hazimeh H, Piogé S, Pantoustier N, Combellas C, Podvorica FI, Kanoufi F.;1; Radical chemistry

from diazonium-terminated surfaces. Chem Mater 2013;25:605-12.

[244] Bünsow J, Johannsmann D.;1; Patterned hydrogel layers produced by electrochemically triggered

polymerization. Macromol Rapid Commun 2009;30:858-63.

[245] Daugaard AE, Hvilsted S, Hansen TS, Larsen NB.;1; Conductive polymer functionalization by click

chemistry. Macromolecules 2008;41:4321-7.

[246] Hansen TS, Daugaard AE, Hvilsted S, Larsen NB.;1; Spatially selective functionalization of

conducting polymers by “electroclick” chemistry. Adv Mater 2009;21:4483-6.

[247] Hansen TS, Lind JU, Daugaard AE, Hvilsted S, Andresen TL, Larsen NB.;1; Complex surface

concentration gradients by stenciled ”electro click chemistry”. Langmuir 2010;26:16171-7.

[248] Ku SY, Wong KT, Bard AJ.;1; Surface patterning with fluorescent molecules using click chemistry

directed by scanning electrochemical microscopy. J Am Chem Soc 2008;130:2392-3.

[249] Amemiya S, Bard AJ, Fan FRF, Mirkin MV, Unwin PR.;1; Scanning electrochemical microscopy.

Annu Rev Anal Chem 2008;1:95-131.

75
[250] Ye H, Lee J, Jang JS, Bard AJ.;1; Rapid screening of BiVO4-based photocatalysts by scanning

electrochemical microscopy (SECM) and studies of their photoelectrochemical properties. J Phys Chem C

2010;114:13322-8.

[251] İpek Y, Dinçer H, Koca A.;1; Electrode modification based on “click electrochemistry” between

terminal-alkynyl substituted cobalt phthalocyanine and 4-azidoaniline. Sensor Actuat B 2014;193:830-7.

[252] Sekine S, Ido Y, Miyake T, Nagamine K, Nishizawa M.;1; Conducting polymer electrodes printed

on hydrogel. J Am Chem Soc 2010;132:13174-5.

[253] Ido Y, Takahashi D, Sasaki M, Nagamine K, Miyake T, Jasinski P, Nishizawa M.;1; Conducting

polymer microelectrodes anchored to hydrogel films. ACS Macro Lett 2012;1:400-3.

[254] Hong V, Udit AK, Evans RA, Finn MG.;1; Electrochemically protected copper(I)-catalyzed azide–

alkyne cycloaddition. ChemBioChem 2008;9:1481-6.

[255] Meldal M. Polymer;1; “clicking” by CuAAC reactions. Macromol Rapid Commun 2008;29:1016-

51.

[256] Aragão-Leoneti V, Campo VL, Gomes AS, Field RA, Carvalho I.;1; Application of copper(I)-

catalysed azide/alkyne cycloaddition (CuAAC) ‘click chemistry’ in carbohydrate drug and neoglycopolymer

synthesis. Tetrahedron 2010;66:9475-92.

76
[257] Antoni P, Robb MJ, Campos L, Montanez M, Hult A, Malmström E, Malkoch M, Hawker CJ.;1;

Pushing the limits for thiol−ene and CuAAC reactions: Synthesis of a 6th generation dendrimer in a single day.

Macromolecules 2010;43:6625-31.

[258] Hein JE, Fokin VV.;1; Copper-catalyzed azide-alkyne cycloaddition (CuAAC) and beyond: new

reactivity of copper(i) acetylides. Chem Soc Rev 2010;39:1302-15.

[259] Crowley JD, Bandeen PH, Hanton LR.;1; A one pot multi-component CuAAC “click” approach to

bidentate and tridentate pyridyl-1,2,3-triazole ligands: Synthesis, X-ray structures and copper(II) and silver(I)

complexes. Polyhedron 2010;29:70-83.

[260] Liang L, Astruc D.;1; The copper(I)-catalyzed alkyne-azide cycloaddition (CuAAC) “click”

reaction and its applications. An overview. Coord Chem Rev 2011;255:2933-45.

[261] Chan TR, Hilgraf R, Sharpless KB, Fokin VV.;1; Polytriazoles as copper(I)-stabilizing ligands in

catalysis. Org Lett 2004;6:2853-5.

[262] Zhang HX, Hu YM, Zhang CY, Lee DH, Yoon KB, Zhang XQ.;1; Electrochemically assisted

ethylene (co-)polymerization with a vanadium-based Ziegler-Natta catalyst. Catal Commun 2016;83:39-42.

[263] Wei Y, Zhang Q, Wang WJ, Li BG, Zhu S.;1; Improvement on stability of polymeric latexes

prepared by emulsion ATRP through copper removal using electrolysis. Polymer 2016;106:261-6.

77
Figure 1. Chronoamperometry recorded during potentiostatic eATRP of 50% (v/v) n-BuA in DMF + 0.1 M

Et4NBF4 on Pt at Eapp − E1/2 = − 60 mV with initial 1.0 mM BrCuIIMe6TREN (adapted from [132]). Copyright

2016. Reproduced with permission from the Royal Society of Chemistry.

Figure 2. Schematic of an eATRP setup and reactor [119]. Copyright 2011. Reproduced with permission from

Advancing Science, Serving Society.

Figure 3. (a) CV of ClCuIL in MeCN + 0.1 M Bu4NBF4 recorded on Pt at 0.5 Vs-1. (b) Correlation between

apparent KATRP and E1/2 of BrCuIIL/BrCuIL; Keqapp measured in methyl acrylate with ethyl 2-bromopropionate

as initiator [191]. Copyright 2000. Reproduced with permission from John Wiley & Sons Inc.

Figure 4. Cyclic voltammetry of BrCuIIL (L = Me6TREN (a), TPMA (b) or PMDETA (c)) in the absence (black

line) and presence (red line) of ethyl 2-bromoisobutyrate (EBiB) [122]. Copyright 2013. Reproduced with

permission from the American Chemical Society.

Figure 5. (a) Linear sweep voltammetry at RDE showing oxidation of Cu IMe6TREN and reduction of

ClCuIIMe6TREN at rotating disk electrode in MeCN + 0.1 M Et4NBF4. (b) Chronoamperometry of CuIL in the

presence of excess TEMPO, with injection of benzyl chloride at t = 0 s. Inset: kinetic analysis according to a

pseudo-first-order rate law [195]. Copyright 2011. Reproduced with permission from the Royal Society of

Chemistry.

Figure 6. Cyclic voltammetry recorded at v = 0.2 V s−1 for 1.0 mM CuIIMe6TREN2+: (a) in water or in 18 wt%

OEGA in water; (b) in water in the absence and presence of oligo(ethylene oxide) bromopropionate (OEOBrP);

(c) in 18 wt% OEGA in water in the absence and presence of OEOBrP. Determination of ka for the reaction of

CuIMe6TREN with (d) HEBiB in 18 wt% OEGA in water, (e) OEOBrP in pure water, and (f) OEOBrP in 18

wt% OEGA in water, by fitting of the experimental data on theoretical working curves at 25 °C (excess ratio γ =

78
[RX]/[XCuIIL], kinetic parameter λ = RTkaCCuII/Fv) [90]. Copyright 2014. Reproduced with permission from

the Royal Society of Chemistry.

Figure 7. Experimental and simulated CVs of CuIIMe6TREN in MeCN + 0.1 M Et4NBF4 in the presence of

bromoacetonitrile recorded at γ = [RX]/[CuIIL] = 0.5 [206]._ENREF_135 Copyright 2016. Reproduced with

permission from Elsevier Ltd.

Figure 8. (a) Voltammetry of BrCuIITPMA in 56% (v/v) BA in DMF + 0.2 M Bu4NClO4, recorded at 0.05 V

s-1 in the absence (dotted line) and presence of excess EBiB without stirring (black solid line) or under vigorous

stirring (solid red line). (b) Effect of the Eapp on the polymerization rate (Rp) [122]. Copyright 2013. Reproduced

with permission from the American Chemical Society.

Figure 9. Synthesis of glucose-based star homopolymers as a function of applied potential. (a) First-order

kinetic plots for different Eapp values, and (b) Mn and Mw/Mn versus monomer conversion [133]. Copyright

2016. Reproduced with permission from Elsevier Ltd.

Figure 10. eATRP of 10% MAA (v/v) in water at pH = 0.9; [M]:[RX]:[CuCl2]:[TPMA]:[NaCl] =

200:1:0.1:0.4:29. (a) Variation of conversion (full circles) and applied potential (dashed lines) with time during

eATRP. No Eapp was employed during the first 20 min. (b) Evolution of molecular weight with monomer

conversion during the active steps [42]. Copyright 2016. Reproduced with permission from the American

Chemical Society.

Figure 11. Effect of stirring on eATRP of 56% BA (v/v) in DMF + 0.2 M Bu4NClO4 at Eapp – E1/2 = -125 mV

and [M]:[EBiB]:[TPMA]:[CuII(OTf)2]:[n-Bu4NBr] = 300:1:0.03:0.03:0.03. (a) First-order kinetic plots and (b)

variations of molecular weight and Mw/Mn with monomer conversion at different stirring rates [122]. Copyright

2013. Reproduced with permission from theAmerican Chemical Society.

79
Figure 12. eATRP of 56% BA in DMF with a variety of ligands (tris(2-(dimethylamino)ethyl)amine

(Me6TREN), tris(2-pyridylmethyl)amine (TPMA), and N,N,N’,N’,N’’-pentamethyldiethylenetriamine

(PMDETA)); Eapp – E1/2 = -0.180 V (circles) or -0.125 V (squares). (a) First-order kinetic plot. (b) Molecular

weight and Mw/Mn versus monomer conversion. The arrows indicate the scale used [122]. Copyright 2013.

Reproduced with permission from the American Chemical Society.

Figure 13. Effect of the initial catalyst concentration on eATRP with 56% BA in DMF. (A) First-order kinetic

plot and (B) the evolution of the number-average molecular weight and Mw/Mn versus conversion [122].

Copyright 2013. Reproduced with permission from the American Chemical Society.

Figure 14. Cyclic voltammetry of 1.01 mM CuIITPMA in H2O + 0.1 M Et4NBF4 recorded at 0.1 V s-1 in the

absence and presence of different amounts HEBiB [120]. Copyright 2011. Reproduced with permission from

John Wiley & Sons Inc.

Figure 15. Schematic representation of the optimal conditions for a well-controlled aqueous eATRP. ΔE = Eapp

– E1/2. Controlled polymerizations (Mw/Mn < 1.3 and Mn in good agreement with theoretical values) were found

for the experimental conditions that are inside the colored zones [193]. Copyright 2015. Reproduced with

permission from the American Chemical Society.

Figure 16. Galvanostatic versus potentiostatic (Eapp = Epc - 25 mV) seATRP of 50 % BA (v/v) in DMF + 0.2 M

TBAP. (a) Current versus time for a potentiostatic (green) and galvanostatic polymerization (black). (b) First-

order kinetic plot of a potentiostatic and galvanostatic eATRP [133]. Copyright 2016. Reproduced with

permission from Elsevier Ltd.

Figure 17. Comparison of eATRP and seATRP reaction setup [126]. Copyright 2015. Reproduced with

permission from Gesellschaft Deutscher Chemiker.

80
Figure 18. Multi-step chronoamperometry for galvanostatic seATRP. (a) First-order kinetic plots by

potentiostatic and galvanostatic conditions. (b) MW and MWD results from GPC with PMMA standards [126].

Copyright 2015. Reproduced with permission from Gesellschaft Deutscher Chemiker.

Figure 19. Potentiostatic eATRP of OEGMA in water + 0.1 M Et4NBr performed on different cathodes. (a)

First-order kinetic plots and (b) evolution of molecular weight and dispersity with conversion [131]. Copyright

2016. Reproduced with permission from John Wiley & Sons Inc.

Figure 20. Synthesis of PEO star polymers using multi-step Eapp (a), and cross-linker conversion and Eapp

versus time (b) [124]. Copyright 2013. Reproduced with permission from the American Chemical Society.

Figure 21. Strategy for the preparation of star block copolymers having cyclodextrin cores [130]. Copyright

2016. Reproduced with permission from Elsevier Ltd.

Figure 22. Strategy for the preparation of star block copolymers having glucose cores [133]. Copyright 2016.

Reproduced with permission from Elsevier Ltd.

Figure 23. PSPMA brush generation on an electrode by eATRP using a grafting-from approach [121].

Copyright 2012. Reproduced with permission from Gesellschaft Deutscher Chemiker.

Figure 24. A monolayer of ATRP-initiator (bis[2-(2′-bromoisobutyryloxy)ethyl]disulfide) on a flat gold

substrate, which is used as the WE in eATRP. XCuIIL was reduced to CuIL that initiated the reaction on the

electrode surface [139]. Copyright 2013. Reproduced with permission from Multidisciplinary Digital Publishing

Institute.

Figure 25. Illustration of fabrication of gradient polymer brushes by SI-eATRP [123]. Copyright 2013.

Reproduced with permission from the American Chemical Society.

81
Figure 26. Thickness gradient brushes on homogeneous and patterned surfaces. (a) Ellipsometry thickness

gradients versus position and time. (b) AFM image of a part of patterned gradients grown on gold surfaces

[123]. Copyright 2013. Reproduced with permission from the American Chemical Society.

Figure 27. Illustration of the electrochemical apparatus used for eATRP with GC as the BPE, positioned

between Pt driving electrodes [142]. Copyright 2015. Reproduced with permission from Gesellschaft Deutscher

Chemiker.

Figure 28. Illustration of the apparatus employed for polymer brush patterning. (a) Side view of the system in

which a Pt mesh electrode was placed under the cylinder without a gap, and a glass substrate was placed 130

mm apart from the Pt mesh. (b) SI-eATRP of NIPAM generating a pattern by bipolar electrochemistry. (c)

Optical micrograph of a water droplet adsorbed on the patterned PNIPAM brush [142]. Copyright 2015.

Reproduced with permission from Gesellschaft Deutscher Chemiker.

Figure 29. Schematic illustration of the grafted process on a silicon wafer surfaces via SI-seATRP [216].

Copyright 2017. Reproduced with permission from John Wiley & Sons Inc.

Figure 30. Strategy for the preparation of nano-nickel coated carbon fibers (CFs) [125]. Copyright 2014.

Reproduced with permission from Springer Berlin Heidelberg.

Figure 31. Schematic illustration of the grafted process of PVP on an OMCs surface via SI-eATRP [222].

Copyright 2016. Reproduced with permission from Elsevier Ltd.

82
Figure 32. Antifouling surfaces prepared by SI-eATRP [220]. Copyright 2015. Reproduced with permission

from Elsevier Ltd.

Figure 33. Preparation of the Pt-PANI-GOx-PSBMA sensor. Red arrows indicate additional bromination sites

of primary amino groups on PANI [223]. Copyright 2016. Reproduced with permission from the American

Chemical Society.

Figure 34. Synthesis of PAM brushes on the Au electrode surface via SI-eATRP [221]. Copyright 2016.

Reproduced with permission from Springer Berlin Heidelberg.

Figure 35. Synthesis of MIP on the Au electrode surface via eATRP with Hb both as catalyst and template

[229]. Copyright 2016. Reproduced with permission from Elsevier Ltd.

Figure 36. Electrografting of HEMA from a conducting polymer macroinitiator via an SI-eATRP mechanism

[134]. Copyright 2016. Reproduced with permission from Elsevier Ltd.

Figure 37. Electrochemical attachment of 4-phenylacetic diazonium salt and its subsequent use as initiator for

generating polymer brushes [243]. Copyright 2013. Reproduced with permission from the American Chemical

Society.

Figure 38. Structure formation by surface-initiated polymerizations using ammonium persulfate as redox-

sensitive initiator (a) without RAFT agent and (b) with RAFT agent, leading to a homogeneous hydrogel

PNIPAM film [181]. Copyright 2010. Reproduced with permission from John Wiley & Sons Inc.

83
Figure 39. Functionalization of electrodes of azide-modified electrical conducting polymer PEDOT-N3. (a)

Dried electrodes after immersion at appropriate potential in a solution of alkynated fluorescein (yellow) and

catalytically inactive CuII. (b) Application of a cathodic potential to electrode set 1 reduces Cu II to CuI in its

immediate vicinity and initiates the coupling of alkyne reagent 1 to form a triazole link. (c) Replacing

fluorescein with alkynated rhodamine (purple) in solution and reversing the electrode potentials leads to

selective coupling of the second reagent to the azides of electrode set 2 [246]. Copyright 2009. Reproduced with

permission from John Wiley & Sons Inc.

Figure 40. Local reduction of CuII to CuI at a gold microelectrode and immobilization of acetylene fluorophore

derivatives onto a glass substrate through “click chemistry”. Reprinted with permission from [248]. Copyright

2008. Reproduced with permission from the American Chemical Society.

Figure 41. Electrode modification via “click chemistry” (CC) and CEC to use of electrochemical pesticide

sensors [251]. Copyright 2014. Reproduced with permission from Elsevier Ltd.

Figure 42. Schematic illustration of the fabrication of a conducting polymer/hydrogel electrode. (a) A melted

agarose solution is poured onto a Pt microelectrode substrate. (b) Electropolymerization of the PEDOT into the

gel. (c) Electrochemical actuation causes peeling of the gel. (d) Photograph of the PEDOT microelectrode array

on the gel sheet [252]. Copyright 2010. Reproduced with permission from the American Chemical Society.

Figure 43. Photographs of the PEDOT microelectrodes anchored to the preliminarily molded hydrogel films of

collagen (0.3 mm thick), glucomannan (1 mm thick), PAM (1 mm thick), and a commercial soft contact lens

made of PHEMA. Scale bar: 5 mm _ENREF_191[253]. Copyright 2012. Reproduced with permission from the

American Chemical Society.

84
Figure 44. Electrochemical removal of Cu from polymers with high Cu II affinity [219]. Copyright 2015.

Reproduced with permission from the American Chemical Society.

Figure 45. TEM images of Sample 2 (A) prior and (B) subsequent to electrolysis [263]. Copyright 2016.

Reproduced with permission from Elsevier Ltd.

Figure 46. Tensile curves for PBMA samples with Cu residual prior (A) and subsequent to aging (A’), and after

Cu removal before (B) and after aging (B’). The latex was electrolyzed at 25 °C and 9 V for 500 min to remove

Cu [263]. Copyright 2016. Reproduced with permission from Elsevier Ltd.

Scheme 1. ATRP processes (ka = activation rate constant, kda = deactivation rate constant, kp = propagation rate

constant, kt = termination rate constant, M = monomer, L = ligand, X = halogen, and Pn−X = alkyl halide).

Scheme 2. ATRP processes with (re)generation of the catalyst (kred = reduction rate constant).

Scheme 3. The electrochemical catalytic process involved in eATRP [122].

Scheme 4. Proposed mechanism of miniemulsion polymerization by eATRP (coordination of X– to CuII/L was

omitted for simplicity) [214]. Copyright 2016. Reproduced with permission from the American Chemical

Society.

Scheme 5. On-demand polymerization by eATRP. Polymerization can be (re)initiated or stopped by changing

Eapp [119]. Copyright 2011. Reproduced with permission from Advancing Science, Serving Society.

85
Scheme 6. Mechanism of electrochemically mediated olefin polymerization [262]. Copyright 2016. Reproduced

with permission from Elsevier Ltd.

86
Table 1. Examples of eATRP polymerizations of various monomers.

Polymer Catalyst system Polymer architecture Mn Đ [ref]

poly(n-butyl acrylate) CuIIBr2/TPMA homopolymers 25 100 1.08 [122]

(PBA) CuIIBr2/TPMA homopolymers 18 150 1.06 [126]

CuIIBr2/TPMA homopolymers 27 500 1.10 [128]

CuIIBr2/TPMA homopolymers 19 400 1.08 [189]

CuIIBr2/TPMA block copolymers 25 000 1.12 [124]

(PEO-b-PBA)

CuIIBr2/TPMA block copolymers 16 200 1.19 [188]

(PBA-b-PU-b-PBA)

CuIIBr2/TPMA star polymers 64 200 1.07 [130]

(-CD-PBA)

CuIIBr2/TPMA star polymers 90 700 1.07 [133]

(glucose-PBA)

CuIIBr2/TPMA star block copolymers 165 100 1.14 [143]

(-CD-(PDMAEMA-b-PBA))

CuIIBr2/TPMA star block copolymers 108 700 1.17 [137]

(-CD-(PMMA-b-PtBA-b-PBA))
CuIIBr2/TPMA graft polymers 54 900 1.10 [216]

(Si-g-PBA)
CuIIBr2/Me6TREN homopolymers 35 500 1.16 [132]

CuIIBr2/BPMEA + homopolymers 27 400 1.19 [214]

CuIIBr2/BPMODA*

poly(tert-butyl acrylate) CuIIBr2/TPMA block copolymers 19 900 1.22 [188]

(PtBA) (PtBA-b-PBA-b-PU-b-PBA-b-

PtBA)

87
CuIIBr2/TPMA star block copolymers 140 200 1.23 [130]

(-CD-(PBA-b-PtBA))

CuIIBr2/TPMA star block copolymers 74 100 1.18 [137]

(-CD-(PMMA-b-PtBA))

CuIIBr2/TPMA graft polymers 38 200 1.14 [136]

(cellulose-g-PtBA)
CuIIBr2/TPMA graft copolymers 145 000 1.17 [216]

(Si-g-(PBA-b-PtBA))

CuIIBr2/TPMA graft copolymers 57 600 1.11


[216]
(Si-g-(PDMAEMA-b-PtBA))

poly(methyl acrylate) CuIIBr2/Me6TREN homopolymers 36 000 1.06 [119]

(PMA)

poly(oligo (ethylene glycol) CuIIBr2/TPMA star block copolymers 182 200 1.18 [133]

acrylate) (glucose-(PBA-b-POEGA))

(POEGA) CuIIBr2/TPMA graft block copolymers 78 800 1.11 [136]

(cellulose-g-(PtBA-POEGA))

poly(methyl methacrylate) FeIIIBr3/TMPP homopolymers 10 400 1.29 [135]

(PMMA) CuIIBr2/TPMA

CuIIBr2/TPMA block copolymers 41 800 1.12 [129]

(PEO-b-PMMA)

CuIICl2/PMDETA star polymers 38 00 1.23 [137]

(-CD-PMMA)

graft polymers   [142]

(glass-g-PMMA)

88
poly(tert-butyl CuIICl2/PMDETA graft polymers   [142]

methacrylate) (glass-g-PtBMA)

(PtBMA)

poly(glycidyl methacrylate) CuIIBr2/BPY graft polymers   [125]

(PGMA) (CFs-g-PGMA)

poly(potassium 3- CuIIBr2/BPY graft polymers   [123]

sulfopropyl methacrylate) (Si-g-PSPMA)

(PSPMA) CuIIBr2/BPY graft polymers   [140]

(Au-g-PSPMA)

CuIIBr2/BPY graft copolymers   [121]

(Au-g-PSPMA)

poly(hydroxyethyl CuIIBr2/Me6TREN graft polymers   [139]

methacrylate) (Au-g-PMEMA)

(PHEMA) CuIIBr2/BPY graft polymers   [134]

(Au-g-PMEMA)

CuIIBr2/BPY graft copolymers   [121]

(Au-g-PMEMA)

CuIICl2/PMDETA graft polymers [142]


 
(glass-g-PNIPAM)

poly(sulfobetaine CuIICl2/BPY graft polymers   [220]

methacrylate) (Au-g-PSBMA)

(PSBMA) CuIICl2/TPMA graft polymers   [223]

(Pt-g-PSBMA)

poly(sodium methacrylate) CuIICl2/PMDETA graft polymers   [142]

(PMA-Na) (glass-g-PMA-Na)

89
poly(oligo(ethylene oxide) CuIIBr2/TPMA homopolymers 106 000 1.19 [120]

monomethyl ether CuIIBr2/TPMA homopolymers 62 000 1.17 [131]

methacrylate) CuIIBr2/TPMA homopolymers 55 000 1.17 [193]

(POEGMA) CuIIBr2/Me6TREN homopolymers 58 000 1.28 [193]

CuIIBr2/Me6TREN graft polymers   [139]

(Au-g-POEGMA)

CuIIBr2/PMDETA homopolymers 54 000 1.23 [193]

poly(2- CuIIBr2/TPMA star polymers 78 500 1.08 [143]

(dimethylamino)ethyl (-CD-PDMAEMA)

methacrylate) CuIIBr2/TPMA graft polymers 31 600 1.31 [216]

(PDMAEMA) (Si-g-PDMAEMA)

poly(methacrylic acid) CuIIBr2/TPMA homopolymers 87 600 1.33 [42]

(PMAA)

poly(acrylamide) CuIIBr2/TPMA block copolymers 5 800 1.15 [127]

(PAM) (PEO-b-PAM)

CuIIBr2/bis(4-methoxy- block copolymers 14 400 1.10 [127]

3,5-dimethyl-pyridin-2- (PEO-b-PAM)

ylmethyl)-pyridin-

2-ylmethyl-amine

(TPMA*2)

CuIIBr2/Me6TREN block copolymers 9 000 1.09 [127]

(PEO-b-PAM)

CuIIBr2/BPY graft polymers   [221]

(Au-g-PAM)

poly(N- CuIIBr2/Me6TREN block copolymers 24 500 1.18 [94]

90
isopropylacrylamide) (PEO-b-PNIPAM)

(PNIPAM) CuIIBr2/Me6TREN block copolymers 36 700 1.25 [127]

(PEO-b-PAM-b-P(NIPAM-stat-

AM))

CuIICl2/PMDETA graft polymers   [142]

(glass-g-PNIPAM)

poly(N-vinylpyrrolidone) CuIICl2/N,N,N,N- graft polymers   [222]


tetramethylethylenediam
PVP ine (TMEDA) (OMC-g-PVP)

poly(2- CuIIBr2/Me6TREN graft polymers   [139]

(methacryloyloxy)ethyl (Au-g-PMETAC)

trimethylammonium CuIIBr2/BPY graft polymers   [140]

chloride) (Au-g-PMETAC)

(PMETAC)

91

You might also like