Strandman 2015

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 72

Accepted Manuscript

Title: Thermo-responsive block copolymers with multiple


phase transition temperatures in aqueous solutions

Author: S. Strandman X.X. Zhu

PII: S0079-6700(14)00120-8
DOI: http://dx.doi.org/doi:10.1016/j.progpolymsci.2014.10.008
Reference: JPPS 902

To appear in: Progress in Polymer Science

Please cite this article as: Strandman S, Zhu XX, Thermo-responsive block copolymers
with multiple phase transition temperatures in aqueous solutions, Progress in Polymer
Science (2014), http://dx.doi.org/10.1016/j.progpolymsci.2014.10.008

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.
Thermo-responsive block copolymers with multiple phase transition
temperatures in aqueous solutions
S. Strandman and X. X. Zhu

Department of Chemistry, University of Montreal,


CP 6128, Succursale Centre-ville, Montreal, QC, H3C3J7, Canada.

t
ip
Abstract

cr
Stimuli-responsive block copolymers that exhibit multiple thermal transitions are an emerging topic,
important for understanding thermo-sensitive self-assembling processes and furthermore, for

us
developing interesting molecular devices. This review describes synthetic strategies for building di-
and multiblock copolymers composed of several thermo-responsive segments that provide multiple

an
transitions. The structural and compositional factors that affect the transition temperatures (upper or
lower critical solution temperatures, UCST and LCST) are discussed. The aggregation behavior is
M
highlighted through several examples of diblock and multiblock copolymers showing sequential
thermal transitions.
d

Keywords
te

Thermo-responsiveness, multiple transitions, UCST, LCST, block copolymers, aggregation

List of abbreviations
p
ce

AA acrylamide DMA N,N-dimethylacrylamide


AAc acrylic acid DMP 2-dodecylsulfanylthiocarbonylsulfanyl-
ATRP atom transfer radical polymerization 2-methyl propionic acid
Ac

CA cholic acid DMSO dimethyl sulfoxide


CP cloud point DP degree of polymerization
CTA chain transfer agent DSC differential scanning calorimetry
dNbpy 4,4′-dinonyl-2,2′-dipyridyl EA N-ethylacrylamide
DEA N,N-diethylacrylamide EEGE ethoxyethyl glycidyl ether
DEAEMA 2-(diethylamino)ethyl methacrylate EMA N,N-ethylmethacrylamide
DEGEA ethoxydi(ethylene glycol) acrylate HEA N-hydroxyethyl acrylamide
DEGMA di(ethyleneglycol) methyl ether HMTETA 1,1,4,7,10,10-hexamethyltriethylene-
methacrylate tetramine
DLS dynamic light scattering
1
Page 1 of 71
HS-DSC high-sensitivity differential scanning PEOEOVE poly(2-(2-ethoxy)ethoxyethyl vinyl
calorimetry ether)
iBA N-isobutylacrylamide PEtOx poly(2-ethyl-2-oxazoline)
iBMA N-(isobutoxymethyl)acrylamide PHEMA poly(2-hydroxyethyl methacrylate)
iPA N-isopropylacrylamide (also known as PiPrOx poly(2-isopropyl-2-oxazoline)
NIPA or NIPAm) PMDETA N,N,N′,N′′,N′′-

t
iPMA N-isopropylmethacrylamide pentamethyldiethylenetriamine

ip
I-S insolubility-solubility transition PMEMA poly(2-(N-morpholino)ethyl
LCST lower critical solution temperature methacrylate

cr
MAPTAC 3-(methacryloylamino)propyl]- PMPC poly(2-methacroyloxyethyl phosphoryl
trimethylammonium chloride choline)

us
MDEGA methoxydiethyleneglycol acrylate PVCL poly(N-vinylcaprolactam)
MEO2MA 2-(2-methoxyethoxy)ethyl methacrylate POEG(M)A poly(oligo ethylene glycol)
Me6TREN tris [2-(dimethylamino)ethyl]amine (meth)acrylate

an
Mn,theor theoretical number-average molar mass PPO poly(propylene oxide)
MOTAC [2-(methacryloyloxy)ethyl]trimethyl- PPropOx poly(2-n-propyl-2-oxazoline)
ammonium chloride PTEGSt poly(4-vinylbenzyl methoxytris-
M
mPEG methoxypoly(ethylene glycol) (oxyethylene) ether)
MVA N-methylvinylacetamide RAFT reversible addition-fragmentation chain
d

nBA N-n-butylacrylamide transfer (polymerization)


NMP nitroxide-mediated polymerization sBA N-sec-butylacrylamide
te

NMR nuclear magnetic resonance S-I solubility-insolubility transition


spectroscopy SLS static light scattering
p

nPA N-n-propylacrylamide SS styrene sulfonic acid


ce

OEGMA oligo(ethylene glycol) methacrylate tBA N-tert-butylacrylamide


P4VP poly(4-vinylpyridine) Tdem demixing temperature
PAGE poly(allyl glycidyl ether) TEGMA methoxytri(ethylene glycol)
Ac

P(A-Hyp-OH) poly(N-acryloyl-4-trans-hydroxy-L- methacrylate


proline) Tmix mixing temperature
P(A-Pro-OMe) poly(N-acryloyl-L-proline methyl TMS trimethylsilyl
ester) UCST upper critical solution temperature
PDI polydispersity index UV-vis ultraviolet-visible light spectroscopy
PDMAEMA poly(dimethylamino)ethyl VA N-vinylacetamide
methacrylate)) VAc vinyl acetate
PEG poly(ethylene glycol) VBA vinylbenzoic acid
PEOVE poly(2-ethoxyethyl vinyl ether) VPi vinyl pivalate

2
Page 2 of 71
1. Introduction

Stimuli-sensitive or so-called “smart” polymers that are capable of undergoing a transition from
hydrophilic to hydrophobic in response to an external stimulus have been investigated and reviewed
extensively [1–6]. While most of the fundamental studies have focused on thermo-responsive systems
with a single phase transition temperature, more complex tailored polymers with multiple transitions

t
between soluble and insoluble states have recently been designed and studied for better understanding

ip
of the structural effects on phase transitions, necessary for developing interesting molecular devices [7,
8]. Recent reviews have highlighted the advances in multi-stimuli responsive systems capable of

cr
multiple phase transitions in response to other types of stimuli, such as pH, ionic strength, light,
magnetic fields, enzymatic action or redox agent [9–14], but less emphasis has been put on the

us
emerging systems that can exhibit stepwise temperature-sensitive phase transitions. Therefore, these
systems will be the focus of this review. Some of the suggested applications of stimuli-responsive

an
materials lie in biomedical and pharmaceutical fields, such as gene or drug delivery, cell carrier
systems and imaging [15–21], analytical field, such as chromatography and biosensing [22], and
M
membrane and separation technology [23–25]. Many of these applications require controlled
interactions with (bio)molecules and a precise non-linear response to the stimuli. Advanced synthetic
techniques have given researchers a toolkit to design copolymers with pre-defined distributions of
d

responsive and functional segments that could eventually fulfill such requirements.
te

The thermo-responsiveness of polymers may be generally described by two threshold


temperatures. The polymers are soluble at all compositions above the upper critical solution
p

temperature (UCST) and below the lower critical solution temperature (LCST). There are considerably
ce

fewer reports on the UCST-type systems in aqueous solutions [26] than on polymers exhibiting the
LCST transition. The latter type of behavior is attributed to a local structural transition involving water
Ac

molecules that surround specific segments of the polymer, which depends on the delicate balance
between polymer-solvent hydrogen bonding and hydrophobic and hydrophilic polymer-polymer
interactions [Error! Bookmark not defined.]. However, simply analyzing the balance of hydrophobic
and hydrophilic molecular fragments does not always predict the exact LCST [27], as it is affected also
by other factors, such as molar mass, polymer architecture, end groups, neighboring blocks and
possible linkers in the case of block copolymers, concentration, spatial separation between the
comonomers, and the presence of salts or additives [Error! Bookmark not defined., 28, 29]. The
order of hydrophilic and hydrophobic segments in multiblock systems can be controlled by block
copolymerization, and amphiphilic or double-hydrophilic thermo-responsive diblock copolymers
3
Page 3 of 71
exhibiting a single thermal transition have been widely studied, particularly for their self-assembling
characteristics [30–34]. Di- or multiblock copolymers with at least two blocks that show well
distinguishable LCSTs enable sequential changes in amphiphilicity and thus, graded temperature-
controlled self-assembling behavior in aqueous solutions.
In this review, we will first present the synthetic strategies for building thermo-responsive di- and
multiblock copolymers, focusing especially on the controlled syntheses of systems that consist of

t
ip
several thermo-responsive segments in the same polymer. We will then review the factors that
influence the phase transition temperatures and how these can be used in tailoring the thermo-

cr
responsive self-assembling behavior. Finally, the aggregation of thermo-responsive multiblock
copolymers will be discussed through examples of systems showing sequential thermal transitions.

us
an
2. Syntheses of di- and multiblock copolymers

2.1. Background
M
Reversible deactivation or so-called “living” or controlled radical polymerization techniques are widely
used in the preparation of block copolymers with well-defined structures and architectures owing to
their applicability to a broad range of monomers and synthetic conditions [35–39]. They are more
d

robust than ionic polymerization techniques. In the case of thermo-responsive polymers, the low
te

polydispersity of well-defined polymers allows precise examination of thermal transitions. A key


requirement for building block copolymers is the presence of living or functional group-bearing chain
p

ends to further continue the polymerization (chain extension) or to attach a preformed polymer chain.
ce

The latter method is useful when the blocks are prepared by different polymerization mechanisms or
from monomers with different reactivities, or when finding a suitable copolymerization solvent is
Ac

difficult [40, 41]. Currently, the most common controlled radical polymerization techniques used for
building thermo-responsive block copolymers are reversible addition-fragmentation chain transfer
(RAFT) polymerization and atom transfer radical polymerization (ATRP). These methods have been
the subject of several reviews [42–47]. Although nitroxide-mediated polymerization (NMP) [48] has
also gained popularity in the controlled polymerization of many important monomers, our focus will
mainly be on these two techniques, and their principles will be discussed briefly as a background for
block copolymerizations.

4
Page 4 of 71
RAFT is a radical polymerization method mediated by a chain transfer agent (CTA), which can be
thiocarbonylthio compounds, such as dithioesters, dithiocarbamates, trithiocarbonates, and xanthates,
or more recently, organometallic compounds [49]. Scheme 1 presents the principle of RAFT through its
main equilibrium and the structures of common chain transfer agents. Owing to its robustness and
versatility, RAFT is popular in the synthesis of thermo-responsive block copolymers, particularly those
composed of (meth)acrylamides. In addition, the presence of end group from the CTA may allow

t
ip
binding a functional moiety, such as fluorescence probe, to a suitable polymer chain end. Often the end
group can be modified in a one-pot procedure without isolating the intermediate polymer [50].

cr
Although the direct RAFT polymerization of (meth)acrylamides proceeds with ease, sometimes an
activated ester monomer was preferred and was allowed to react with an amine after the polymerization

us
step to yield poly(meth)acrylamides [51, 52]. This strategy is useful in introducing randomly
distributed functional comonomers or synthesizing a library of polymers with identical degree of

an
polymerization (DP) but different substituents, and it can be used with various polymerization methods
[Error! Bookmark not defined., 53, 54, 55]. The presence of unprotected primary or secondary amine
groups may lead to the aminolysis of the CTA and loss of control during the polymerization in aqueous
M
medium, although this can be minimized at acidic pH [56]. Loss of chain end functionality may also
occur because of degradation during the polymerization, workup, and storage steps, for example due to
d

the presence of peroxides in the solvent (such as in the case of cyclic ethers) or due to slow thermal
te

elimination [57].The stoichiometric ratios of monomer, chain transfer agent and initiator concentrations
determine the theoretical molar mass, the chain end functionalities (dormant vs. dead chains), and the
p

rate of polymerization [58], and allow the design of polymers with desired block lengths. More reactive
CTAs with high chain transfer coefficients provide narrower molar mass distributions and better
ce

control over the molar mass with time [59]


Ac

Scheme 1. Schematic of the main RAFT equilibrium [Error! Bookmark not defined.]. R and Z
represent the radical leaving group and the C=S bond activating group in a thiocarbonylthio chain
transfer agent, respectively, Pn is the growing polymer chain and M is the monomer. The lower part
shows the structures of common CTAs: (a) dithiobenzoates, (b) trithiocarbonates, (c) dithiocarbamates,
(d) xanthates, (e) switchable CTAs (N-methyl-N-(4-pyridyl)dithiocarbamates).

5
Page 5 of 71
ATRP and its variations are based on a transition metal (Cu, Fe, Ru, etc.) complex that participates
in the reversible end-capping of propagating polymer chain through oxidation-reduction mechanism
[Error! Bookmark not defined.,Error! Bookmark not defined.]. The general mechanism of ATRP
is presented in Scheme 2. This method has become popular because of a wide range of polymerizable
functional monomers and commercially available catalysts, ligands and initiators [60]. As the rates of
activation and deactivation of the growing chain depend on the catalyst system used, particularly the

t
ip
choice of ligand, some of the initial attempts to polymerize (meth)acrylamides by ATRP to produce
thermo-responsive polymers had poor control over the polymerization because of the deactivation of

cr
the copper catalyst due to binding to amide, cyclization of the halogen end with penultimate amide
nitrogen, and inadequate redox potential of catalysts leading to increased radical concentration and thus

us
termination reactions [61, 62]. Therefore, for monomers containing strongly coordinating groups, such
as amine, amide, carboxylate or pyridine groups, ligands forming very stable metal complexes (Scheme

an
2) need to be used [Error! Bookmark not defined.]. Polymerizing monomers with bulky substituents
in the ester or amide group requires careful selection of the catalyst and initiator, particularly when
multifunctional sterically-hindered initiators are used [63]. The combination of suitable
M
catalyst/initiator pairs and polar solvents allows the ATRP of (meth)acrylamides [64, 65]. In
conventional ATRP, the concentration of transition metal/ligand complex acting as an activator (metal
d

in reduced form) and deactivator (oxidized form) needs to be sufficiently high to provide a controlled
te

polymerization process with a reasonable rate but at the same time as low as possible to reduce the
metal residues in the final product and to ease the post-polymerization purification process.
p
ce

Scheme 2. Schematic of ATRP equilibrium [Error! Bookmark not defined.]. Here R-X is a halogen
end-capped polymer chain, R• is the corresponding radical, and Mt-Y/ligand is the transition metal
Ac

complex (catalyst). The lower part shows the structures of common ligands: (a) HMTETA
(1,1,4,7,10,10-hexamethyltriethylene-tetramine), (b) PMDETA (N,N,N′,N′′,N′′-
pentamethyldiethylenetriamine), (c) dNbpy (4,4′-dinonyl-2,2′-dipyridyl), (d) Me6TREN (tris[2-
(dimethylamino)ethyl]amine).

2.2. Block copolymerization

The relative reactivities of propagating radicals depend on the monomer structure. Bulky groups in the
vicinity of vinyl functionality prevent the propagating radical from reacting effectively with the
6
Page 6 of 71
monomer, which makes controlling the polymerization of sterically hindered monomers, such as
itaconate-based monomers, a challenging task [66, 67]. In the RAFT polymerization of a series of N-
alkyl-substituted acrylamides, the number and nature of the substituents in the nitrogen of amide group
influences the monomer reactivity and thus, the controllability of the polymerization [68]. The
monomer structure has important consequences for the design and preparation of di- and multiblock
copolymers.

t
ip
The polymerization kinetics of both N-alkyl- and N,N-dialkyl-substituted acrylamides (iPA, nPA,
DMA and DEA, see Table 1 for full names) were studied under identical conditions (temperatures [M]/

cr
[CTA]/[I] ratios, solvent) using a trithiocarbonate CTA with high chain transfer efficiency, 2-
dodecylsulfanylthiocarbonylsulfanyl-2-methyl propionic acid (DMP) [Error! Bookmark not

us
defined.]. While all polymerizations showed negative deviation from the theoretical number-average
molar mass (Mn,theor) at high conversions leading to decreased control particularly at higher

an
temperatures, the polymerizations of N,N-dialkyl-substituted monomers were faster and better
controlled with diminished influence of temperature, less deviation from Mn,theor and lower
M
polydispersity indexes (PDI) than those of N-alkyl-substituted monomers. This could stem from the
higher reactivity and more stable radicals of N,N-disubstituted monomers because of the stronger
electron-donating conjugative effect in the molecular structure [Error! Bookmark not defined.]. The
d

result is in accordance with an earlier study on the reactivity ratios of N-alkylacrylamides [69].
te

In order to produce well-defined block copolymers by ATRP, the rate of initiation of the block
should be higher than the rate of propagation, and thus, the order of blocks should follow the
p

decreasing order of monomer reactivity (e.g. acrylonitrile > methacrylates > styrenes ~ acrylates)
ce

[Error! Bookmark not defined.]. When the reactivities of monomers are similar, it is possible to
prepare multiblock copolymers sequentially. If the polymer architecture requires adding monomers in
Ac

the order of increasing reactivity, halogen exchange from Br to Cl can be used for the initiating chain
end, as C-Cl bond has higher dissociation energy than C-Br leading to more efficient initiation. One
method for preparing multi-block copolymers with symmetrical order of blocks (e.g. ABA) is using
difunctional initiators or CTAs, particularly when the reactivity of the monomers is similar. In RAFT
polymerization, the monomers that have greater steric stabilization and hence, better leaving group
ability (methacrylates, methacrylamides) should be polymerized before monomers which produce
stabilized secondary propagating radicals, such as styrene, acrylates and acrylamides. These in turn
should be polymerized before vinyl esters and vinylamides [Error! Bookmark not defined.]. In

7
Page 7 of 71
RAFT, switchable CTAs that can be suitably protonated or deprotonated (Scheme 1) may be used for
the block copolymerization of monomers with a large difference in reactivity [70].

Table 1. Chemical structures and aqueous solution properties of poly(N-alkylacrylamide)s. [Error!


Bookmark not defined.], Copyright 2007. Adapted from with permission from the American

t
Chemical Society.

ip
cr
CH2 CH LCST

us
n
O R1 R2 abbrev Ref.
(°C)
N
R1 R2

Polyacrylamide

Poly(N,N-dimethylacrylamide)
H

Me
H

Me an PAA

PDMA
soluble

soluble
85
Error!
M
Bookmark not
defined.
Error!
Bookmark not
d

defined.
(Error!
Poly(N-ethylacrylamide) H Et PEA >85 (82,70)
te

Bookmark not
defined.,Error!
Bookmark not
p

defined.)
Error!
ce

Poly(N,N-ethylmethacrylamide) Me Et PEMA 70 Bookmark not


defined.
Error!
Ac

Poly(N,N-diethylacrylamide) Et Et PDEA 32 Bookmark not


defined.
Error!
Bookmark not
defined.
Poly(N-isopropylacrylamide) H iPr PiPA 32 (36)
(Error!
Bookmark not
defined.)
Error!
Bookmark not
Poly(N-n-propylacrylamide) H nPr PnPA 25 (20) defined.
(Error!
Bookmark not
defined. Error!
8
Page 8 of 71
defined.,Error!
Bookmark not
defined.)
Error!
Poly(N-n-butylacrylamide) H nBu PnBA insoluble Bookmark not
defined.
Error!
Poly(N-isobutylacrylamide) H iBu PiBA insoluble Bookmark not

t
defined.

ip
Error!
Poly(N-sec-butylacrylamide) H sBu PsBA insoluble Bookmark not
defined.

cr
Error!
Poly(N-tert-butylacrylamide) H tBu PtBA insoluble Bookmark not

us
defined.

The above-mentioned difference in the reactivities of mono- and disubstituted acrylamides led

an
to testing the capacity of their corresponding homo- and copolymers as macro-CTAs. Monomers with
similar reactivity are expected to produce random copolymers, which was verified by the kinetics and
M
monomer ratios throughout the copolymerization of two mono-substituted N-alkylacrylamides, nPA
and N-ethylacrylamide, EA [71]. While all tested poly(N-alkylacrylamide)s showed good performance
as macro-CTAs in the block copolymerization of both mono- and dialkyl-substituted acrylamides, their
d

N,N-disubstituted counterparts could provide chain extension but often yielded a mixture of unreacted
te

macro-CTA and block copolymer under the same conditions (Table 2) [Error! Bookmark not
defined.]. Therefore, the less reactive N-alkylacrylamides should be polymerized before the more
p

reactive N,N-dialkylacrylamides in the block copolymerization.


ce

a
Table 2. RAFT block copolymerization of N-alkyl-substituted acrylamides [Error! Bookmark not
Ac

defined.], Copyright 2007. Reproduced with permission from the American Chemical Society.

Second comonomer
Macro-CTA iPA nPA nBA EMA DMA DEA iBA sBA tBA

PiPA √ √ √ √ √ √ √ √ √
PnPA √ √ M √ √ √ √ √ √
PnBA √ M √ √ √ √ √ √ √
PEMA M M M N/A √ √ M M M

9
Page 9 of 71
PDMA M M M M M √ M M M
PDEA M M M M M M M M M

a
Polymerizations at 70 °C in DMSO at constant [M]/ [CTA]/ [I]; √ = successful block
copolymerization; M = mixture of macro-CTA and block copolymer; N/A = data not available.

t
ip
This sequential polymerization strategy has successfully been used in the synthesis of well-defined

cr
thermo-responsive di-, tri- and tetrablock copolymers. As multiblock copolymers exhibiting multiple
phase transition temperatures are less common, some examples of such polymers synthesized by (i)

us
sequential RAFT polymerization include triblock and tetrablock copolymers PnPA-b-PiPA-b-PEMA
and PnPA-b-PiPA-b-PEMA-b-PDMA (Scheme 3a) [Error! Bookmark not defined.], where the name

an
indicates the order of block addition in the copolymerization, triblock copolymers of a sulfobetaine and
vinylbenzoic acid (VBA) with poly(oligo (ethylene glycol) acrylate) (POEGA), DMA, or styrene
sulfonate (SS) (Scheme 3d) [72], triblock copolymers of nPA, methoxydiethyleneglycol acrylate
M
(MDEGA), and EA (Scheme 3c) [73], and triblock copolymers of iPA, DMA and N-
acryloylpyrrolidine (Scheme 3f) [74], both with different orders of blocks. Diblock and ABA-type
d

thermo-responsive triblock copolymers composed of PiPA and poly(2-methoxyethylvinyl ether)


te

(PMOVE) (Scheme 3g), were synthesized by (ii) the combination of RAFT polymerization and living
cationic polymerization [75, 76], and an ABA-type triblock copolymer of N-(4-vinylbenzyl)-N,N-
p

dialkyldimethylamines was synthesized by RAFT (Scheme 3e) [77], both using a dicarboxylic acid-
ce

functional CTA. In the case of cationic polymerization, the polymerization was initiated from a proton
derived from the carboxylic RAFT agent. After quenching the polymerization, the synthesis was
continued by RAFT process [Error! Bookmark not defined.]. Ethylvinyl ether-based thermo-
Ac

responsive diblock and triblock copolymers composed of poly(2-ethoxyethyl vinyl ether) (PEOVE),
PMOVE, and poly(2-(2-ethoxy)ethoxyethyl vinyl ether) (PEOEOVE) blocks (Scheme 3h) were
synthesized by (iii) living cationic polymerization [78, 79]. Finally, examples of thermo-responsive
multiblock copolymers synthesized by (iv) ATRP include triblock copolymers composed of
poly(ethylene glycol), PiPA and P(iPA-co-HEA) blocks, where HEA is N-hydroxyethyl acrylamide
(Scheme 3b) [80], ABA-type block copolymers of a sulfobetaine and 2-(2-methoxyethoxy)ethyl
methacrylate (MEO2MA) (Scheme 3i) [81] and triblock copolymers of poly(propylene oxide) (PPO),

10
Page 10 of 71
poly(2-methacroyloxyethyl phosphoryl choline) (PMPC) and PiPA (Scheme 3j) [82].The range of
different diblock copolymers is wider because of fewer polymerization steps.

Scheme 3. Structures of multiblock copolymers discussed in this review: (a) PnPA-b-PiPA-b-PEMA-


b-PDMA [Error! Bookmark not defined.], (b) PEG-b-PiPA-b-P(iPA-co-HEA) [Error! Bookmark

t
ip
not defined.], (c) triblock copolymer of EA, nPA and MDEGA (x = 2) [Error! Bookmark not
defined.], (d) triblock copolymer of POEGA (x = 8), sulfobetaine and VBA [Error! Bookmark not

cr
defined.], (e) PVMA-b-PVEA-b-PVMA (R1 = Me, R2 = Et) [Error! Bookmark not defined.], (f)
triblock copolymer of iPA, N-acryloylpyrrolidine and DMA [Error! Bookmark not defined.], (g)

us
PMOVE-b-PiPA-b-PMOVE [Error! Bookmark not defined.], (h) PEOVE-b-PMOVE-b-PEOEOVE
[Error! Bookmark not defined.], (i) ABA-type triblock copolymer of a sulfobetaine and MEO2MA

an
[Error! Bookmark not defined.], (j) PPO-b-PMPC-b-PiPA [Error! Bookmark not defined.], and (k)
PEEGE-b-PPO-b-PEEGE [Error! Bookmark not defined.]. Note that the letters indicating the
numbers of repeating units do not represent the order of the blocks added, and in some cases, different
M
orders of blocks have been investigated. The end groups or groups linking the blocks have been
omitted in most structures for simplicity.
d
te

3. Multiple phase transitions


p

3.1. Determination methods


ce

The most common techniques used for detecting the phase transition temperatures in aqueous solutions
include turbidimetric detection and differential scanning calorimetry (DSC, Fig. 1) [Error! Bookmark
Ac

not defined.]. The former method gives a cloud point (CP), which has various experimental definitions
when determined from the transmittance curve obtained by UV-vis spectrometry, for example from the
first appearance of cloudiness or from different stages of absorbance increase or transmittance decrease
(e.g. 10% or 50% drop in transmittance). The turbidity changes with temperature can be slow or abrupt
and can depend on the polymer concentration and heating/cooling rate [83]. Because of the popularity
of this method, the term LCST is often used as a synonym for CP or in conjunction with it, although
LCST actually refers to a specific point in the phase diagram, below which the polymer is soluble at all
concentrations. The same applies to the UCST and the turbidity changes are discussed in terms of a
clearing point. Calorimetric techniques, particularly microcalorimetry, give a demixing (Tdem) or
11
Page 11 of 71
mixing temperature (Tmix) either from the onset or the maximum of the endotherm [Error! Bookmark
not defined.]. Other methods, such as static and dynamic light scattering (SLS, DLS), nuclear
magnetic resonance spectroscopy (NMR), or fluorescence spectroscopic and X-ray scattering methods
are often used in combination with the methods above to gain detailed information on the nature of the
transition and the structure of aggregates [Error! Bookmark not defined.].

t
ip
Fig. 1 UV-absorbance and DSC endotherm (∆Cp = difference in the heat capacity between the sample

cr
and the reference) as a function of temperature for a PiPA solution [84]. Copyright 1991. Reproduced
with permission from the American Chemical Society.

us
Detecting the LCSTs of individual blocks of a thermo-responsive di- or multiblock copolymer is

an
not trivial, as the transition temperatures may be too close to each other to be detected separately, the
enthalpy change associated to the collapse of individual blocks may be small, or the collapse may be
M
quick [85]. Also the heating rate and concentration may affect the transitions [86, 87]. For example, the
CP of 0.2 wt-% aqueous solution of PDEA shifted with 8 °C (from 33 to 41 °C) when the heating rate
was increased from 0.06 to 5 °C/min [88]. The effect of concentration depends on the determination
d

method, as optical methods are sensitive to particle size, and hence, to concentration [Error!
te

Bookmark not defined.]. When using UV-vis spectrophotometer for turbidity detection, the
temperature-dependent transmittance and the visibility of stepwise transitions depend on the detection
p

wavelength. This is demonstrated in Fig. 2 for a PnPA-b-PiPA-b-PEMA block copolymer (structures of


ce

blocks given in Table 1) [89]. The scattering depends strongly on the wavelength of light (~λ-4) [90].
Therefore, the light of the shorter wavelength is more scattered than that of the longer wavelength, and
Ac

as a result, the transmittance at longer wavelengths will be higher. Moreover, only the particles larger
than the wavelength of light are visible at a certain observation wavelength [Error! Bookmark not
defined.,Error! Bookmark not defined.,Error! Bookmark not defined.]. Winnik and coworkers
detected the phase transitions of individual blocks of a diblock copolymer composed of poly(2-
isopropyl-2-oxazoline) (PiPrOx) and poly(2-ethyl-2-oxazoline) (PEtOx) blocks with the help of high-
sensitivity differential scanning calorimetry (HS-DSC) when UV-vis spectrophotometry at 550 nm
showed only a single CP [91]. Hence, although the system would consist of multiple thermo-responsive
segments, only one transition may be observed. The CPs of individual blocks should have enough
difference so that stepwise phase transitions could be observed instead of one broad transition.
12
Page 12 of 71
Fig. 2. Transmittance curves at different detection wavelengths for a 1.0 mg/mL aqueous solution of
PnPA-b-PiPA-b-PEMA block copolymer as a function of temperature with heating rate of 0.1 °C/min
(structures of blocks given in Table 1) [Error! Bookmark not defined.]. Copyright 2009. Reproduced
with permission from the American Chemical Society.

t
ip
3.2. Tailoring LCSTs by copolymerization

cr
The modification of the LCST of thermo-responsive polymers can be generally achieved by
copolymerization or by external conditions; by the addition of salts, cosolvents, or surfactants in the

us
aqueous solution, which affect the whole polymer [Error! Bookmark not defined.,Error! Bookmark
not defined., 92, 93, 94, 95, 96]. In general, polymerization with a more hydrophobic monomer will

an
decrease the LCST and more hydrophilic comonomer will increase it. The effect of comonomers has
been demonstrated for several series of copolymers based on N-substituted acrylamides with varying
M
degrees of hydrophilicity and hydrophobicity [Error! Bookmark not defined.,Error! Bookmark not
defined., 97, 98]. Some of the results are shown in Fig. 3 presenting the ranges of LCSTs for different
copolymers. The solubility of the corresponding homopolymers vary from soluble to insoluble below
d

the boiling point of water (Table 1), and the solubilities and LCSTs of the copolymers lie between
te

those of homopolymers. The LCST data (Fig. 3) can be fitted to


p

(1)
ce

where μ1 and μ2 are the molar fractions of monomers 1 and 2, respectively, T1 and T2 are the LCSTs of
corresponding homopolymers, and K is a weighting parameter deduced from the fitting of the curve to
Ac

experimental results. Ideally, the K of 1 would give a linear relationship between the composition and
LCST [Error! Bookmark not defined.]. However, all the curves for polymers depicted in Fig. 3, as
well as for poly(nPA-co-EA) [Error! Bookmark not defined.], showed a concave shape. The
copolymerization allowed adjusting the LCSTs in the range from 14 °C to 87 °C (or their solubility
also increases in this range) [Error! Bookmark not defined.]. Tuning the LCST of poly(N-
isopropylacrylamide) in the range from < 16 to 100 °C by copolymerization has been reviewed
elsewhere [Error! Bookmark not defined.]. Also other subtle structural changes, for example, the

13
Page 13 of 71
nature of the end group or the oxidation state of a reactive comonomer, can have a significant effect on
the LCST and solubility [99, 100].

Fig. 3. LCSTs of different series of poly(DEA-co-X) as a function of their chemical compositions


(molar fractions μ); X = AA, DMA or EA. Dashed lines represent fittings to Eq. (1) [Error!

t
Bookmark not defined.]. Copyright 1999. Reproduced with permission from Elsevier.

ip
cr
The LCSTs of poly((meth)acrylamide)s could also be adjusted between 0 and 100 °C by the post-
modification of water-soluble hydroxyl group-bearing homopolymer precursors by introducing

us
hydrophobic moieties by acetylation, benzoylation or cinnamoylation [Error! Bookmark not
defined.]. In another post-modification approach, poly(pentafluorophenyl (meth)acrylate)s were

an
modified by acyl substitution with oligo(ethylene glycol) methyl ether amines and hexylamine, which
resulted in (meth)acrylamides with tuneable solubilities (from insoluble to fully soluble in the range
from 0 to 93 °C) and CPs ranging from 26 to 78 °C [101]. Interestingly, the (meth)acrylamide
M
homopolymers without hexyl substituents showed side chain-length dependent UCST in 2-propanol
and 1-octanol [Error! Bookmark not defined.]. This synthetic strategy was used to construct
d

POEGMA-b-PiPA block copolymers exhibiting double thermo-responsive LCST behavior in aqueous


te

solution and one UCST in the above-mentioned alcohols [102].

Lutz and coworkers demonstrated the tuning of the thermo-responsiveness of poly(oligo(ethylene


p

glycol) methacrylate) (POEGMA) by the copolymerization with more hydrophobic 2-(2-


ce

methoxyethoxy)ethyl methacrylate (MEO2MA) (Scheme 4e) [103]. The LCSTs were in the range of
26−90 °C [104]. In another example, Laschewsky and coworkers showed that the random
Ac

copolymerization of 4-vinylbenzyl tri- and tetra(oxyethylene) ethers yielded polymers with CPs
between 16 and 61 °C in pure water, and the alternating copolymerization of 4-vinylbenzyl
methoxytetra(oxyethylene) ether with a series of N-substituted maleimides (Scheme 4g) resulted in
polymers with LCSTs covering the full range of liquid water at ambient pressure from insoluble (LCST
< 0 °C) to fully soluble (LCST > 100 °C) [105].

The copolymerization strategy has been successfully used for tuning the LCSTs of individual
blocks to introduce multiple phase transitions. The LCST of the second block in PiPA-based diblock
copolymer could be adjusted above or below the LCST o PiPA by the amount of suitable comonomers

14
Page 14 of 71
[106]. In PiPA-b-P(iPA-co-HMA) (Scheme 4a, R = H), the more hydrophilic N-
(hydroxymethyl)acrylamide (HMA) increased the LCST of PiPA up to 80 °C and two phase transitions
were visible in the UV-vis transmittance curve when the HMA content was between 10 and 34 mol-%
[107, 108]. In PiPA-b-P(iPA-co-iBMA) (Scheme 4a, R = CH2CH(CH3)2), the more hydrophobic N-
(isobutoxymethyl)acrylamide (iBMA) decreased the LCST of PiPA and two CPs were observed with
iBMA contents of 5−15 mol-% [109]. Both series of polymers were synthesized by one-pot ATRP.

t
ip
This method was also used in the synthesis of a double thermo-responsive PEG-b-PiPA-b-P(iPA-co-
HEA) (Scheme 3b), which showed two CPs at 37 and 48 °C when the ratio of iPA and N-hydroxyethyl

cr
acrylamide (HEA) in the copolymer block was 10:1 [Error! Bookmark not defined.].

Debuigne and coworkers [110] found that the CP of poly(N-vinylcaprolactam) (PVCL) could be

us
tuned from 36 °C down to 19 °C by the cobalt-mediated radical copolymerization with more
hydrophobic vinyl acetate (VAc) or vinyl pivalate (VPi), or up to 81 °C with more hydrophilic N-

an
methyl-N-vinylacetamide (MVA) or N-vinylacetamide (VA), depending on the comonomer ratio. This
inspired the preparation of double thermo-responsive PVCL-b-P(VCL-stat-MVA) block copolymers
M
(Scheme 4d) with the first CP at 35−37 °C and the second at 40 or 70 °C [Error! Bookmark not
defined.]. Weiss and Laschewsky prepared double-thermoresponsive tapered copolymers by one-pot
ATRP or RAFT polymerization of maleimide-based monomers and the excess of 4-vinylbenzyl
d

methoxytetrakis(oxyethylene) ether (Scheme 4h) [111]. Because of the alternating nature of the
te

copolymerization of maleimides with 4-vinylbenzyl monomer, the resulting copolymer was roughly
composed of two blocks, one alternating copolymer and one homopolymer. The CPs of the alternating
p

block were in the range of 20−38 °C and the transition temperatures of the homopolymer block were
ce

31−51 °C, depending on the relative block lengths and the composition of the maleimide-based
monomer [Error! Bookmark not defined.].
Ac

Scheme 4. Structures of copolymers and block copolymers with random copolymer segments: (a)
PiPA-b-P(iPA-co-HMA) (R = H) and PiPA-b-P(iPA-co-iBMA) (R = CH2CH(CH3)2) [Error!
Bookmark not defined.,Error! Bookmark not defined.,Error! Bookmark not defined.], (b)
P(nPAxEA(1-x))-b-P(nPAyEA(1-y)) [Error! Bookmark not defined.], (c) P(nPAxDEAEMA(1-x))-b-
P(nPAyEA(1-y)) [Error! Bookmark not defined.], (d) PVCL-b-P(VCL-stat-MVA) [Error! Bookmark
not defined.], (e) P(OEGMA-co-MEO2MA), x = 8-9 [Error! Bookmark not defined.], (f)
P(TEGMA-co-AAc)-b-P(DEGEA-co-AAc) [Error! Bookmark not defined.], (g) alternating

15
Page 15 of 71
copolymers of 4-vinylbenzyl methoxyoligo(oxyethylene) ether with N-substituted maleimides [Error!
Bookmark not defined.], (h) tapered copolymers of 4-vinylbenzyl methoxytetra(oxyethylene) ether
and N-substituted maleimides [Error! Bookmark not defined.]. Note that the letters indicating the
numbers of repeating units do not represent the order of the blocks added. The end groups or groups
linking the blocks have been omitted in most structures for simplicity.

t
ip
To take the copolymerization approach even further, a diblock copolymer, where both blocks

cr
consist of random P(nPA-co-EA) copolymers with tailored CPs, was synthesized by Zhu group using
sequential RAFT and the ratio of blocks was suited to ensure the colloidal stability of the aggregates.

us
The compositions selected to obtain two well separable transition temperatures were P(nPA0.7EA0.3)-b-
P(nPA0.4EA0.6) (Scheme 4b), where the subscripts refer to the molar fraction of the monomer in each
block [Error! Bookmark not defined.]. The CPs of the individual random copolymers were 37 °C

an
(P(nPA0.7EA0.3)) and 52 °C (P(nPA0.4EA0.6)), and two separate CPs could be observed by UV-vis
spectroscopy and DLS (Fig. 4) [Error! Bookmark not defined.]. Later the same approach was used
M
by the same group to prepare a double stimuli-responsive diblock copolymer, where P(nPA0.8EA0.2)
block responded only to temperature (CP = 33 °C) and P(nPA0.8DEAEMA0.2) block (DEAEMA = 2-
(diethylamino)ethyl methacrylate, Scheme 4c) shows pH-dependent CP in the range of 68 and 13 °C
d

(pH 6 and >9, respectively). This polymer showed two separate CPs both at pH 7 and 10, where
te

DEAEMA units were partially protonated and deprotonated, respectively [112]. In the same way,
Maric and coworkers prepared a series of double thermo-responsive diblock random copolymers by
p

nitroxide-mediated polymerization (NMP), where the blocks consisted of OEGMA (with 8-9 EG units)
ce

and MEO2MA, and the corresponding CPs were between 31 and 72 °C [113].
Ac

Fig. 4. Temperature-dependent transmittance curves 1.0 mg/mL aqueous solutions of (A)


P(nPA0.7EA0.3) (1) and P(nPA0.4EA0.6) (2), observed at a wavelength of 350 nm, and (B)
P(nPA0.7EA0.3)-b-P(nPA0.4EA0.6) (Scheme 4b) at two wavelengths, 350 and 250 nm [Error!
Bookmark not defined.]. Copyright 2012. Reproduced with permission from the American Chemical
Society.

3.3. Effect of neighboring blocks on LCSTs

16
Page 16 of 71
The LCSTs of individual blocks in thermo-reponsive multiblock copolymers are not affected only by
their composition but also by their neighboring blocks and end groups, which needs to be taken into
consideration when designing copolymers with stepwise phase transitions at precise temperatures. The
more hydrophilic neighboring block is expected to slightly increase the LCST of the less hydrophilic
block and vice versa, and this effect depends on the relative lengths of the blocks. The change in LCST
may be minor if the blocks are well separated in the aggregation process, and therefore for many block

t
ip
and graft copolymers no changes were reported [Error! Bookmark not defined.]. In some cases, a
single CP located between those of individual blocks of a double thermo-sensitive system was reported

cr
on the basis of the transmittance curve, but other methods, such as DLS, revealed more complex
aggregation behavior for the same polymer under the same conditions [Error! Bookmark not

us
defined.]. The effect of the neighboring block is easiest to observe in systems consisting only of a
single thermo-responsive block. For example, the cloud point-increasing effect of the more hydrophilic

an
PEG on the block and graft copolymers of PiPA has been well demonstrated [114–119], although the
demixing temperature and kinetics have been reported to strongly depend on the polymer architecture
and concentration, as Tdem decreases significantly at high concentrations [120]. The effect of the
M
relative block lengths on the LCSTs has also been demonstrated for other thermo-responsive block
copolymers with a single LCST [121–123].
d

Laschewsky and coworkers showed a shift in the CPs of dually thermo-responsive PnPA-b-PEA
te

diblock copolymers according to the relative block lengths [124].When the relative length of the more
hydrophilic PEA block increased upon changing the nPA:EA ratio from 2:1 to 1:1 and further to 1:3,
p

the CP of the more hydrophobic PnPA block increased from 17 to 23 and then to 27 °C, respectively.
ce

The less pronounced CPs of the PEA block were in the range of 60−70 °C. The hydrophobic end group
from CTA participated the aggregation process and its immobilization upon aggregation depended on
Ac

whether it was attached to PnPA or PEA block. This further depended on the order of blocks in the
copolymerization [Error! Bookmark not defined.]. In another example, for dually thermo-responsive
PiPAx-b-PEO20-b-PPO70-b-PEO20-b-PiPAx pentablock copolymers, the first Tdem arising from the PPO
block increased from 24.4 to 29.0 °C when the length of PiPA block (x) increased from 10 to 97 units.
At this length range, Tdem of the PiPA block remained nearly constant (~ 35 °C) [125].The mutual
effect of neighboring blocks was also observed for a 1:1 POEGMA-b-PiPMA (number of EG units = 4)
block copolymer bearing biotin at a chain end [Error! Bookmark not defined.]. The CPs of
POEGMA and poly(N-isopropylmethacrylamide) blocks shifted upon block copolymerization from 64
[126] to 57 °C and from 45 [127] to 47°C, respectively [Error! Bookmark not defined.]. The double
17
Page 17 of 71
thermo-sensitivity allowed stimulus-controlled adsorption of the block copolymer on protein-
immobilized surface: as the biotin was attached to the end of the more hydrophobic PiPMA block,
according to surface plasmon resonance (SPR) measurements, the immobilization of biotin to
streptavidin-functionalized surface was possible below the LCST of the first block, but inhibited above
it (Fig. 5) [Error! Bookmark not defined.].

t
ip
Fig. 5. Schematic of streptavidin-functionalized surface (A), immobilization of POEGMA-b-PiPMA-

cr
biotin (structure above) on it below the LCST of PiPMA (46.5 °C) (B), and the absence of adsorption
when the surface is exposed to the block copolymer above the LCST of PiPMA (C) [Error!

us
Bookmark not defined.]. Copyright 2010. Reproduced with permission from the American Chemical
Society.

an
Similar to diblock copolymers, the multiple stepwise LCSTs of multiblock copolymers were
M
affected by the lengths and hydrophilicity/hydrophobicity of the neighboring blocks. For two P(nPA-b-
iPA-b-EMA) triblock copolymers [Error! Bookmark not defined.,Error! Bookmark not defined.],
the phase transition temperatures of PnPA and PiPA blocks were higher than those of their
d

homopolymers, but the largest shift towards lower temperatures was observed for the most hydrophilic
te

PEMA block (from 70 °C down to 58 °C). The stability of the aggregates depended on the length of the
PEMA block that collapsed last upon heating [Error! Bookmark not defined.]. In the case of random
p

diblock copolymers, both the monomer composition and block ratio influenced the CPs of neighboring
ce

blocks, such as in the copolymers composed of P(OEGMA-co-MEO2MA) blocks (number of EG units


= 8-9, Scheme 4e) [Error! Bookmark not defined.] or P(nPA-co-EA) blocks [Error! Bookmark not
defined.]. The protonation of a pH-responsive comonomer, such as DEAEMA in P(nPA0.8EA0.2)-b-
Ac

P(nPA0.8DEAEMA0.2) (Scheme 4c), shifted the CP of the corresponding block but also that of the
neighboring block towards higher temperatures [Error! Bookmark not defined.]. The effect of pH
and the two-step thermo-responsiveness of the block copolymer are demonstrated in Fig. 6.

Fig. 6. Dual stimuli-responsive behavior and the two-step transition observed for aqueous solutions of
diblock random copolymer P(nPA0.8EA0.2)-b-P(nPA0.8DEAEMA0.2) (Scheme 4c) at (A) pH 7
(DEAEMA partially protonated) and (B) at pH 10 (DEAEMA deprotonated), measured by UV-vis

18
Page 18 of 71
transmittance at different wavelengths [Error! Bookmark not defined.]. Copyright 2013. Reproduced
with permission from the American Chemical Society.

4. Aggregation behavior

4.1. Aggregation of double thermo-responsive diblock copolymers

t
ip
The effects of relative block lengths and end group on the aggregation were demonstrated for aqueous
solutions of double thermo-responsive R-PnPA137-b-PEA67, PnPA133-b-PEA133-R and PnPA34-b-

cr
PEA94-R diblock copolymers, where R is an aryl-TMS group from the CTA and the block ratios are
2:1, 1:1 and 1:3, respectively [Error! Bookmark not defined.]. All the polymers showed two-step

us
self-assembling behavior in aqueous solutions. While the first polymer with long core-forming block
(2:1 block ratio) gave micellar clusters above the LCST of PnPA, which shrank above the LCST of

an
PEA, the two other polymers formed star-flower-like micelles, where the hydrophobic end groups were
partially incorporated in the hydrophobic cores. Further heating above the LCST of the outer block led
M
to a reversible partial collapse of the shell (1:1 block ratio) or reversible clustering into well-defined
aggregates (1:3 block ratio) [Error! Bookmark not defined.]. The shrinkage above the first CP and
subsequent clustering was also observed by our group for P(nPA0.7EA0.3)-b-P(nPA0.4EA0.6) with a
d

block ratio of 1:2.4 (Scheme 4b, Fig. 7) [Error! Bookmark not defined.]. Notably, the transition from
te

free polymer chains to micelles is usually accompanied by an initial, sometimes very strong increase in
the hydrodynamic diameter, followed by shrinkage upon the dehydration and/or reorganization of core-
p

forming blocks. After this, the micelles either remain stable or aggregate further.
ce

Fig. 7. Schematic illustration of the aggregation behavior of P(nPA0.7EA0.3)-b-P(nPA0.4EA0.6) (Scheme


Ac

4b) with a block ratio of 1:2.4 in water upon heating: (a) molecularly soluble polymers and small
micelles below the first CP; (b) dehydration of the P(nPA0.7EA0.3) core at 40 °C; (c) collapsed
aggregates at 44−53 °C; (d) clusters above the second CP at 53 °C [Error! Bookmark not defined.].
Copyright 2012. Reproduced with permission from the American Chemical Society.

For comparison, a series of PiPA-b-P(iPA-co-HMA) diblock copolymers (Scheme 4a) with block
ratios of 1.5:1 to 1.3:1 and HMA contents of 10−34 mol-% [Error! Bookmark not defined.] formed
stable micelles similar to the 1:1 polymer above, while clustering tendency above the second LCST
19
Page 19 of 71
was observed for a P(nPA0.7EA0.3)-b-P(nPA0.4EA0.6) diblock random copolymer with a block ratio of
1:2.4 (Scheme 4b) [Error! Bookmark not defined.]. P(iPA-co-iBMA)-b-PiPA block copolymers with
a block ratio of 1:2 and hydrophobic iBMA contents of 5-15 mol-% also showed tendency towards
reversible clustering above the second LCST, but the polymer precipitated upon further heating
[Error! Bookmark not defined.]. However, in the case of PVCL-b-P(MVA-stat-VCL) diblock
copolymers (Scheme 4d) with a block ratio of ~1:1, clusters were obtained at 1.0 g/L after the initial

t
ip
micellization and dehydration of the micellar core. Further aggregation above the second LCST was
reversible upon cooling [Error! Bookmark not defined.]. In a dendritic PiPA-b-PDMAEMA system

cr
(PDMAEMA = poly(dimethylamino)ethyl methacrylate)), where double thermo-responsive diblock
copolymers at block ratios 1:2 and 1:3 were grafted from a hyperbranched polyester core, the clustering

us
above the second LCST was prevented by the high branching density, and unimolecular micelles with
collapsed coronas were obtained [128].

an
Thermal transitions may also lead to different aggregate morphologies. Our group observed that
both temperature- and pH-responsive P(nPA0.8EA0.2)-b-P(nPA0.8DEAEMA0.2) diblock random
M
copolymer (Scheme 4c) with a block ratio of 1:1.25 (45:55 wt-%) showed double thermo-
responsiveness both at pH 7 and pH 10 (Fig. 5), where the pH-sensitive DEAEMA-containing block
was partially charged and uncharged, respectively. Invertible vesicles (Fig. 8) were obtained above the
d

first LCST assigned to P(nPA0.8EA0.2) block at pH 7 and P(nPA0.8DEAEMA0.2) block at pH 10 [Error!


te

Bookmark not defined.]. These collapsed blocks formed the core of the vesicles at 37 and 25 °C.
Increasing the temperature above the second LCST led to clustering of the aggregates [Error!
p

Bookmark not defined.]. Further study on these diblock random copolymers with different block
ce

ratios suggested that the polymer composition may be quite critical for the self-assembly process and
thus, for aggregate morphologies [129]. For comparison, in the work of McCormick and coworkers,
Ac

PiPA-b-PDEAEMA diblock copolymers with a block ratio of 50:50 wt-% self-assembled into micelles
both below the CP of PiPA at high pH (pH 9) and above the CP of PiPA at low pH (pH 5, PDMAEMA
fully protonated), while at 29:71 wt-%, the polymer formed micelles at low temperature and high pH,
but vesicles at high temperature and low pH [130]. Note that the PDEAEMA homopolymer block
responds only to pH, and thus, the pH-dependent aggregation of the dually thermo-responsive diblock
random copolymers may be more complex because of the compositional similarity and hence possible
poorer separation of the blocks.

In another example on the aggregate morphologies of diblock copolymers, a double temperature-


responsive poly(2-(dimethylamino)ethyl methacrylate)-b-poly(di(ethyleneglycol) methyl ether
20
Page 20 of 71
methacrylate) (PDMAEMA-b-PDEGMA)) at ∼50:50 (wt %) formed multilamellar vesicles above the

LCST of PDEGMA and unilamellar vesicles above the LCST of partially protonated PDMAEMA

t
ip
block in deionized water [131]. PDMAEMA is known to show pH-, molar mass-, and ionic strength-
dependent thermo-responsiveness in aqueous solutions. In the case of its block copolymers above, two

cr
separate CPs could only be observed at a certain composition (~50:50 wt-%) in deionized water
[Error! Bookmark not defined.]. McCormick and coworkers showed the change in the aggregate

us
morphologies of partially protonated PiPA-b-PDMAEMA above their critical aggregation temperatures
[132]. The polymers self-assembled into spherical micelles at block ratio 32:68 wt-%, into a mixture of

an
worm-like and spherical micelles at 52:48 wt-%, and into vesicles at 64:36 wt-%. Predicted aggregate
morphologies based on the mass fraction of the more hydrophilic PDMAEMA block (f) are presented
in Fig. 9 [Error! Bookmark not defined.].
M
Fig. 8. Illustration of the self-assembly of P(nPA0.8EA0.2)-b-P(nPA0.8DEAEMA0.2) into vesicles and the
d

formation of aggregates upon heating. The insets show representative TEM images of the vesicles and
te

AFM images of the aggregates [Error! Bookmark not defined.]. Copyright 2013. Reproduced with
permission from the American Chemical Society.
p
ce

Fig. 9. Aggregate morphologies of PDMAEMA-b-PiPA according to the mass fraction of


PDMAEMA block (f). Here red domains represent PDMAEMA and blue ones PiPA [Error!
Ac

Bookmark not defined.]. Copyright 2009. Reproduced with permission from the American
Chemical Society.

4.2. Aggregation of triblock copolymers

The aqueous aggregation behavior of multiblock copolymers composed solely of thermo-responsive


blocks depends on the distribution of the blocks in the polymer chain and their relative lengths. The
possible triblock copolymer systems (ABC, BAC, ACB) are depicted in Fig. 10, where A represents

21
Page 21 of 71
the block with the lowest LCST and C the block with the highest LCST. In the illustrated system, the
triblock copolymers are composed of PnPA (LCST ~ 22 °C), PMDEGA (LCST ~ 39 °C), and PEA
blocks (LCST ~ 73 °C, structures shown in Scheme 3c) [Error! Bookmark not defined.]. The
molecular level transitions convert the initially triple hydrophilic polymer into amphiphilic one with
high hydrophilic/hydrophobic balance, then into one with low hydrophilic/hydrophobic balance, and
finally into a triple hydrophobic polymer.

t
ip
cr
Fig. 10. Schematic stepwise thermal switching of the amphiphilicity of triblock copolymers exhibiting
three LCST transition, in dependence on the block sequence. Here A represents the block with lowest

us
LCST and C the block with the highest LCST [Error! Bookmark not defined.]. Copyright 2011.
Reproduced with permission from the American Chemical Society.

an
Fig. 11 presents the schematic illustration of the micellization of two N-alkylacrylamide-based
M
ABC triblock copolymers, where the only difference is the length of the outermost C block [Error!
Bookmark not defined.]. According to DLS and SLS results on heating and cooling of the aqueous
0.1 g/L polymer solutions, heating above the LCST of the A block (PnPA in the case of ref [Error!
d

Bookmark not defined.]) led to the formation of polymeric micelles with collapsed hydrophobic A
te

core and hydrophilic swollen B-C shell (PiPA-b-PEMA) [Error! Bookmark not defined.]. These
blocks will further collapse upon their dehydration above the respective LCSTs. When the C block is
p

too short, it cannot stabilize the collapsed A-B core completely, which leads to the clustering of the
ce

micelles. A longer C block can prevent or delay the formation of micellar clusters. This behavior is in
accordance with the observations by Weiss and Laschewsky on a PiPA52-b-PMDEGA32-b-PEA90 block
copolymer [Error! Bookmark not defined.]. Upon cooling, hysteresis was observed for the N-
Ac

alkylacrylamide-based systems due to the formation of additional inter- and intrachain hydrogen bonds
in the collapsed state [Error! Bookmark not defined.].

Fig. 11. Schematic illustration of the temperature-dependent aggregation behavior of two ABC-type
triblock copolymers with a short (Tri1) or long (Tri2) outermost block. The compositions of the
depicted polymers are P(nPA124-b-iPA80-b-EMA44) (Tri1) and P(nPA124-b-iPA80-b-EMA160) (Tri2)

22
Page 22 of 71
[Error! Bookmark not defined.]. Copyright 2009. Reproduced with permission from the American
Chemical Society.

In the case of BAC block copolymer, when the B and C blocks are longer than A, the initial
collapse of the central A block above its LCST leads to unimolecular micellization rather than

t
aggregation, which requires the collapse of the second block (B). This was shown for a 1.0 g/L solution

ip
of PMDEGA54-b-PiPA34-b-PEA94 block copolymer [Error! Bookmark not defined.], and it resembles

cr
the solution behavior of amphiphilic ABA block copolymers which do not aggregate if the length of
hydrophilic A blocks exceeds the length of hydrophobic B block [133]. Longer hydrophobic or initially

us
collapsing block would lead to increased aggregation tendency, which was demonstrated for instance
for a series of double thermo-responsive PMOVE-b-PiPA-b-PMOVE ABA-type polymers (Scheme 3g)
[Error! Bookmark not defined.]. The continued heating above the LCST of the B block of the ABC

an
polymer enforced the clustering of aggregates and increase of the Dh above the LCST of the third block
[Error! Bookmark not defined.]. Interestingly, Dehghani and coworkers reported a change in the
M
aggregate morphologies of double thermo-responsive PEG45-b-PiPA380-b-P(iPA423-co-HEA42) (Scheme
3b) with temperature, where the polymer formed micelles above the LCST of the PiPA middle block
and vesicles upon further heating above the LCST of the terminal P(iPA-co-HEA) block due to the
d

significant change in the hydrophilic/hydrophobic balance and short shielding PEG block [Error!
te

Bookmark not defined.]. This system could also be considered as an example of a BAC-type block
copolymer where B block is longer than A, although here the PEG block is not thermo-responsive
p

under the studied conditions, and it demonstrates the complexity of the aggregation behavior of multi-
ce

block copolymers.

The ACB-type block copolymers with the composition of PiPA120-b-PEA60-b-PMDEGA13 showed


Ac

relatively similar aggregation behavior to the ABC-type block copolymer, forming aggregates above
the thermal collapse of the terminal A block [Error! Bookmark not defined.]. Heating above the
LCST of the B block in the other end of the chain made the aggregates shrink, although at higher
temperatures than in the case of ABC polymer. Further heating above the collapse of C block made the
aggregates shrink even more, but without clustering [Error! Bookmark not defined.].

Tsvetanov and coworkers [134] studied the aggregation of double thermo-responsive ABA and
BAB triblock copolymers of ethoxyethyl glycidyl ether (EEGE) and PPO (Scheme 3k), which were
synthesized by sequential anionic polymerization. Only single cloud points were obtained, indicating
intermolecular association due to the collapse of the more hydrophobic PEEGE block, and the CPs of
23
Page 23 of 71
ABA copolymers, where A = PPO and B = PEEGE, varied between 7 and 23 °C, while those of BAB
copolymers were lower, between 7 and 18 °C. In both cases the terminal blocks were considerably
shorter than the middle block. The ABA copolymers formed stable micelles but underwent secondary
aggregation above the certain composition-dependent temperature upon the dehydration of PPO block.
On the other hand, the aggregates of BAB copolymers showed continuous increase in size with
temperature owing to the interchain association of the terminal PEEGE blocks. The associated block

t
ip
copolymers coexisted with unimers which gradually joined the aggregated structures with heating (Fig.
12) [Error! Bookmark not defined.].

cr
us
Fig. 12. Schematic representation of the possible states of chain association of PEEGE-b-PPO-b-
PEEGE block copolymers (Scheme 3k) at (a) low and (b) high temperatures [Error! Bookmark not
defined.]. Copyright 2004. Reproduced with permission from the American Chemical Society.

4.3. Association and gelation at high concentrations an


M
The interaction of collapsed blocks above their LCSTs may lead to the hydrophobic association and
subsequent gelation at high concentrations. This brings about interesting gelation patterns that in the
d

case of triblock copolymers depend on the order of blocks. Such behavior has been observed both N-
te

alkylacrylamide- and oligo(ethylene glycol)-based block copolymers.


a) Association of ACB- and ABA-type triblock copolymers. Aoshima and coworkers studied the
p

aggregation of thermo-responsive ABC- and ACB-type triblock copolymers with identical block
ce

lengths (DP = 200), where the LCSTs of PEOVE, PEOEOVE, and PMOVE blocks composed of
oxyethylene units were 20, 41 and 70 °C, respectively [Error! Bookmark not defined.]. The
association behavior of ABC-type block copolymers will be discussed in the section below. The
Ac

heating of 6.0 wt-% aqueous solution of ACB-type PEOVE-b-PMOVE-b-PEOEOVE (Scheme 3h) led
to micellization, gelation, and finally precipitation above the respective LCSTs of the segments. Above
the second LCST, i.e., after the PEOVE and PEOEOVE blocks have collapsed, the polymer resembles
amphiphilic ABA triblock copolymers, where two not necessarily thermo-sensitive hydrophobic
segments are bridged by a soluble segment. Interestingly, the authors noticed that the gel formed by the
ACB-type block polymer was stiffer than the one by ABA-type PEOVE-b-PMOVE-b-PEOVE block
copolymer. The difference stems from the balance between inter- and intramolecular association,
favoring respectively the gelation and flowerlike micellization of ABA polymers [Error! Bookmark

24
Page 24 of 71
not defined.]. The properties of ABA polymers and other types of stimuli-responsive gelators have
been reviewed elsewhere [135].
Armes and coworkers reported the gelation behavior of a double thermo-responsive ACB-type
PPO43-b-PMPC160-b-PiPA81 block terpolymer (Scheme 3j), where two of the blocks were thermo-
responsive; the LCSTs of PPO and PiPA blocks were 15 and 37 °C, respectively [Error! Bookmark
not defined.]. In this case, the gelation efficiency of the terpolymer was lower than that observed for a

t
ip
series of ABA-type PiPA-b-PMPC-b-PiPA triblock copolymers. This was explained by the above-
mentioned micellization above the first LCST and gelation through the intermicellar interactions after

cr
the collapse of the second block above its LCST, as opposed to the simultaneous processes in ABA
polymers. The stepwise process was supported by the viscosity profile of the ACB polymer, which

us
revealed a local viscosity minimum prior to the gelation, arising from the intramicellar collapse of the
outer B block [Error! Bookmark not defined.].

an
b) Association of ABC-type triblock copolymers and AB diblock copolymers. The ABC-type
PEOVE-b-PEOEOVE-b-PMOVE triblock copolymer of Aoshima group formed a gel above the LCST
of PEOVE block, but no increase in viscosity was observed above the second LCST [Error!
M
Bookmark not defined.]. This indicates that the gelation of ABC-type polymers would occur through
the interaction of micellar cores and the second aggregation step only adds the second segment to them.
In this sense, the gelation of ABC-type polymers through intermicellar weak interactions or micellar
d

packing resembles that of thermo-responsive AB diblock copolymers. However, in the case of EOVE-
te

b-EOEOVE, EOVE-b-MOVE, and EOEOVE-b-MOVE diblock copolymers with 1:2 block ratios, the
double thermo-sensitivity resulted in a ‘‘gelation window’’: when the temperature exceeded the first
p

LCST, gelation occurred through the micellization and packing of micelles at sufficiently high
ce

concentrations instead of bridging of the polymer chains in a network. Further increase in the
temperature led to a gel-to-sol transition and finally phase separation at the LCST of the second block
Ac

[Error! Bookmark not defined.]. The process is illustrated in Fig. 13. The block length-dependent
packing of micelles leading to gelation has been described in detail for example for thermo-responsive
poly(oxyethylene)-b-poly(oxybutylene) diblock copolymers in water, although such systems exhibit
only a single sol-to-gel transition that depends on polymer concentration [136]. The packing of
micelles is analogous to the arrangement of sterically stabilized or charge-stabilized colloidal particles
(hard or soft spheres) into colloidal crystalline arrays [137]

25
Page 25 of 71
Fig. 13. Schematic illustration of the micellization and related gelation of a double thermo-responsive
diblock copolymer composed of oligo(ethylene glycol)-based blocks [Error! Bookmark not
defined.]. Copyright 2005. Reproduced with permission from the American Chemical Society.

c) Tuning the gelation behavior. The temperature where the gel-to-sol transition takes place depends

t
on the polymer structure (sequence, composition, molar mass distribution), concentration, and the

ip
presence of additives, such as salt and nonsolvents [Error! Bookmark not defined.]. The shifting

cr
effect of salt (NaCl) on both LCSTs is expected for blocks composed of oxyethylene-based units owing
to their salt-sensitivity. The gelation window was also observed for PTEGMA-b-P(DEGEA-co-AAc)

us
[138], P(TEGMA-co-AAc)-b-PDEGEA [139] and P(TEGMA-co-AAc)-b-P(DEGEA-co-AAc)
(Scheme 4f) [140] diblock copolymers, where TEGMA = methoxytri(ethylene glycol) acrylate and
DEGEA = ethoxydi(ethylene glycol) acrylate. The presence of acrylic acid (AAc) rendered the

an
corresponding blocks both pH- and thermo-responsive, and thus, either only the sol-to-gel, gel-to-sol,
or both transitions could be adjusted with pH (Fig. 14) [Error! Bookmark not defined.,Error!
M
Bookmark not defined.,Error! Bookmark not defined.]. The gelation window was also observed for
PTEGMA-b-PDEGEA without AAc, but no changes according to pH were observed [Error!
Bookmark not defined.].
d
te

Fig. 14. Sol-gel phase diagrams determined by the vial inversion method for P(TEGMA-co-AAc)-b-
P(DEGEA-co-AAc) (Scheme 4f) in a 30 mM potassium hydrogen phthalate buffer at different pH
p

values: pH = 3.29 (solid black square), pH = 5.10 (solid red circle), and pH = 5.79 (solid blue triangle).
ce

The green hollow circles show the phase diagram at pH = 5.10 obtained after decreasing the pH again
from 5.79: the slightly lower values arise from the higher ionic strength after readjusting the pH
Ac

[Error! Bookmark not defined.]. Copyright 2012. Reproduced with permission from the American
Chemical Society.

4.4. Other transitions in aqueous solutions

Certain block copolymers with interacting segments show complex thermo-responsive aggregation
behavior, leading to multiple transitions between soluble (S) and insoluble (I) states. Even if these are
often referred to as LCST (S-I transition) or UCST (I-S transition) phase transitions, the solubility
changes can sometimes be very subtle, observed through the changes in transmittance by UV-vis
26
Page 26 of 71
spectroscopy or in hydrodynamic diameter by DLS. Typically, these polymers possess a combination
of suitable hydrophilic/hydrophobic balance and specific interactions, such as hydrogen bonding or
electrostatic interactions that can be perturbed by heating. The electrostatic interactions are strongly
sensitive to electrolyte and polymer concentrations, while hydrogen bonds are less sensitive, except for
specific hydrogen bond breaking agents [141]. With a proper selection of the interacting comonomer, a
polymer normally exhibiting only a single thermal transition can be transformed into a dually

t
ip
responsive one and the transition temperatures can be controlled with the monomer ratio. For example,
poly(2-hydroxyethyl methacrylate) (PHEMA) exhibits a CP when DP = 20−45 owing to intra- and

cr
interchain H-bonding, and can be made into a dually responsive, highly pH- and salt-sensitive system
exhibiting both S-I and I-S transitions by copolymerization with DMAEMA, 3-

us
(methacryloylamino)propyl]trimethylammonium chloride (MAPTAC), or [2-
(methacryloyloxy)ethyl]trimethylammonium chloride (MOTAC) [142]. The UCST-type behavior may

an
also be induced by multivalent counterions, such as [Co(CN6)]3- for PDMAEMA and its block
copolymers [143–145] and 2,6-naphthalene-disulfonate for P(iPA-co-DMA)-b-P4VP multiblock
copolymer (P4VP = poly(4-vinylpyridine)) (Scheme 5a) [146]. Aliphatic alcohols induced the UCST-
M
type transitions of poly(N-(4-vinylbenzyl)-N,N-dialkylamine)s [Error! Bookmark not defined.] and
various copolymers containing PEG, POEGMA or poly(2-oxazoline) segments [Error! Bookmark not
d

defined.,Error! Bookmark not defined., 147, 148, 149, 150]. The selection of homopolymers and
te

copolymers exhibiting UCST-type transitions is wider than that of block copolymers [Error!
Bookmark not defined.]. Recently, the thermal transitions of thermo-responsive polymers and block
p

copolymers have also been discovered in ionic liquids or ionic liquid/water mixtures [151–155].
Among the block copolymers with thermo-responsive segments, I-S-I (UCST-LCST) transitions in
ce

water have been shown for P(iPA-co-DMA)-b-P4VP multiblock copolymer (Scheme 5a) [Error!
Bookmark not defined.] for sulfobetaine diblock copolymers with PiPA [156, 157], PMEMA (poly(2-
Ac

(N-morpholino)ethyl methacrylate), (Scheme 5b) [158], PDEA [159], or P(OEGMA-co-MEO2MA)


[160] blocks, and POEGMA-b-PDMAEMA diblock copolymers in the presence of counterions [Error!
Bookmark not defined.], while S-I-S (LCST-UCST) transitions have been reported for PTEGMA-b-
PTEGSt (PTEGSt = poly(4-vinylbenzyl methoxytris(oxyethylene) ether)) [161], proline-based block
copolymers [162, 163], and PDMAEMA-based systems [164].

Scheme 5. Examples of block copolymers exhibiting multiple transitions between soluble and
insoluble states: (a) multiblock P(iPA-co-DMA)-b-P4VP [Error! Bookmark not defined.], (b)
27
Page 27 of 71
diblock copolymer of sulfobetaine and MEMA [Error! Bookmark not defined.], (c) P(A-Pro-OMe)-
b-P(A-Hyp-OH) [139], (d) PEtOx-b-P(EtOx-stat-PropOx) [Error! Bookmark not defined.], (e)
CA(PEG-b-PAGE)4 [144].

The thermal transitions that involve ionic and/or hydrogen bonding are often pH-dependent. For

t
example, a proline-based block copolymer (Scheme 5c) composed of a thermo-responsive poly(N-

ip
acryloyl-L-proline methyl ester) block, P(A-Pro-OMe), and a water-soluble poly(N-acryloyl-4-trans-

cr
hydroxy-L-proline) block, P(A-Hyp-OH) exhibited S-I-S transition at pH 2, where initially soluble
micelles with hydrogen-bonded P(A-Pro-OMe) were formed (Fig. 15) [Error! Bookmark not

us
defined.]. Upon heating above the first transition temperature (corresponding to the LCST of P(A-Pro-
OMe)), the polymer formed intra- and interchain hydrogen bonds between the blocks. Further heating
increased the motion of the polymer chains and disruption of hydrogen bonds, observed as second

an
transition temperature (I-S transition). As a result, micelles with dehydrated P(A-Pro-OMe) core were
formed. At pH 7, no detectable transition was observed [Error! Bookmark not defined.]
M
Fig. 15. Postulated mechanism of the S-I-S transition of a P(A-Pro-OMe)-b-P(A-Hyp-OH) block
d

copolymer (blocks represented orange and blue segments, respectively) via self-assembly and hydrogen
te

bonding at low pH (pH = 2) [Error! Bookmark not defined.]. Copyright 2010. Reproduced with
permission from the American Chemical Society.
p
ce

MPEG-b-PDMAEMA diblock copolymers (Fig. 16) with block ratios ~ 1:2 and ~ 1:4 showed an
S-I transition at neutral pH (PDMAEMA partially protonated) and an S-I-S transition at pH 12
Ac

(PDMAEMA deprotonated) [Error! Bookmark not defined.]. The pH dependence of the S-I
transition has earlier been shown for PDMAEMA homopolymers with different architectures [165].
The addition of NaCl had only a minor effect on the S-I transition at neutral pH and at pH 12, but it
gradually increased the I-S transition at pH 12. It was suggested that at neutral pH the polymer formed
micelles with PDMAEMA core and PEG shell, which then aggregated with further heating, while the
S-I-S transition at high pH arised from micellization (PEG core), formation of phase-separated
aggregates, and finally inverse micellization (PDMAEMA core) upon heating [Error! Bookmark not
defined.]. For comparison, Au nanoparticles decorated with PDMAEMA-b-PEG (inner and outer

28
Page 28 of 71
block, respectively) showed LCST behavior in their native state and UCST behavior after the
quaternization into a sulfobetaine, both in aqueous solutions [166].

Fig. 16. Temperature-dependent transmittance curves for mPEG-b-PDMAEMA at pH 3 (black


squares), pH 8 (red circles), pH 10 (blue upward triangles), and pH 12 (green downward triangles)

t
[Error! Bookmark not defined.]. Copyright 2013. Reproduced with permission from Wiley-VCH

ip
Verlag GmbH & Co. KGaA, Weinheim.

cr
Temperature-induced changes in the aggregation or in the supramolecular self-assembly of

us
amphiphilic block copolymers in aqueous solutions may appear as multiple transitions even if only one
block is thermo-responsive. For example, amphiphilic four-arm star-like PEG-based block copolymers,

an
CA(PAGE-b-PEG)4 and CA(PEG-b-PAGE)4 (Scheme 5e) where CA = cholic acid core and PAGE =
poly(allyl glycidyl ether), showed double LCST-type thermo-responsiveness at certain block ratios,
M
owing to the stepwise micellization, aggregation, and dehydration processes as well as clustering of
dehydrated aggregates [167, 168]. While PEG and its derivatives are thermo-responsive and their
LCST decreases with hydrophobic substituents, PAGE itself is not known to be thermo-sensitive.
d

However, after introducing –NH2 grops into the allyl moieties, pH-sensitive LCST-type behavior could
te

be observed for CA(PAGE)4 and the CP could be tuned down controllably by further acetylation of the
–NH2 groups [169]. The amino-functionalized CA(PEG-b-PAGE)4 block copolymers with block ratios
p

2:1 and 3:1 showed S-I-S-type transitions at pH 12, where –NH2 groups are not charged [Error!
ce

Bookmark not defined.], similar to mPEG-b-PDMAEMA above. The S-I-S transition may originate
from the restructuring and fragmentation of aggregates upon heating prior to reaggregation and
clustering. Such behavior was also
Ac

demonstrated by Kjøniksen, Hoogenboom and coworkers for the aqueous solutions of poly(2-
oxazoline) block copolymers consisting of a poly(2-ethyl-2-oxazoline) (PEtOx) block and a random
copolymer block of EtOx and 2-n-propyl-2-oxazoline (PropOx) (PEtOx-block-P(EtOx-stat-PropOx),
Scheme 5d) [170]. In this case, the CP of PEtOx is significantly higher than that of PPropOx, but at
room temperature, the block copolymer is below the CPs of individual blocks and the polymer is thus
hydrophilic. Recently, our group explored the concentration and block length dependence of the S-I-S
transition for a series of dually thermo-responsive PnPA-b-PEMA block copolymers. The results gave a
phase diagram for the temperature-dependent self-assembly of PnPA84-b-PEMA42 (Fig. 17), which
29
Page 29 of 71
shows how the clustering, fragmentation and aggregation processes, and thus, the existence of two
clearly observable phase transitions depends on polymer concentration [171]. Such behavior adds to the
complexity of thermo-responsive systems and is currently an emerging topic, along with the studies on
other systems exhibiting multiple thermal transitions.

t
Fig. 17. (A) Schematic of the assembled structures of PnPA84-b-PEMA42, the unimers transformed into

ip
micelles, into clusters and into aggregates with increasing temperature. (B) Concentration-temperature

cr
phase diagram of PnPA84-b-PEMA42. Low concentration range a: The unimers transformed into
micelles and to aggregates; Medium concentration range b: The unimers transformed into micelles, into

us
clusters and into aggregates; High concentration range c: The unimers transformed into clusters and
into aggregates with increasing temperature. The concentration and temperature scales are approximate

an
and serve as a general trend [Error! Bookmark not defined.]. Copyright 2014. Reproduced with
permission from the Royal Society of Chemistry.
M
5. Conclusion
It is possible to prepare polymers and gels that show multiple controlled and tunable thermal LCST-
d

type transitions in water, giving promise to sensor, separation, and controlled delivery applications, and
te

already much is understood of the nature of these transitions and the aggregation behavior of these
polymers in water. However, the examples of systems exhibiting multiple different types of (S-I and I-
p

S) transitions show that further studies are needed to understand and eventually control the conditions
ce

that induce such a behavior. Especially the effects of block copolymer composition, polymer
concentration and different stimuli need to be explored further. Although the system would consist of
Ac

multiple thermo-responsive segments, only one transition may be observed, as the transition
temperatures may be too close to each other to be detected separately, the enthalpy change associated
to the collapse of individual blocks may be small, or the collapse may be quick. The heating rate and
concentration may also affect the transitions. The phase transition temperatures of individual blocks
should have enough difference so that stepwise transitions could be observed instead of one broad
transition. This can be achieved by suitable selection of (co)monomers and block lengths, and the
advances in reversible deactivation or so-called “living” or controlled radical polymerization
techniques together with living anionic or cationic polymerization methods have brought us closer to
tailoring of the individual thermal transitions. Even though many experimental data have been
30
Page 30 of 71
accumulated in this area in a diverse collection of research papers, theoretical work seems to be lagging
behind in producing guidelines or predictions in the design and synthesis of these polymers.

Although biomedical applications have been suggested for many of the polymers and many of
them have already been tested for the biocompatibility, further studies are still needed in this front.
Some of the challenges of multi-stimuli responsive polymers lie in the amplification of responses to

t
biochemical stimuli at low concentrations. The current trend in the research of stimulus-responsive

ip
materials is to expand their scope by designing new polymers, for example by using synthetic amino
acid derivatives, natural compound derivatives and polypeptides, and to explore the conditions that can

cr
induce the thermal transitions, for example the UCST-type transitions for common thermo-responsive
polymers that normally show LCST-type behavior in water. Our group is making efforts to overcome

us
the possible biocompatibility issues of common thermo-responsive polymers through the use of
monomers and initiators based on natural compounds, such as bile acids. The bile acid-based polymers

an
containing poly(ethylene glycol) segments showed good biocompatibility and promising results in the
encapsulation and delivery of active compounds as such or as cosurfactants [172, 173] and thermo-
M
responsiveness with different transitions between soluble and insoluble states at certain compositions
and solution conditions [Error! Bookmark not defined.,Error! Bookmark not defined., 174, 175].
The expansion of the repertoire of such multi-responsive polymers may lead to materials with
d

improved biocompatibility and performance in response to the need in pharmaceutical and biomedical
te

areas.
p
ce

Acknowledgements
Financial support from NSERC of Canada, the FQRNT of Quebec, and Université de Montréal is
gratefully acknowledged. The authors are members of CSACS funded by FQRNT and GRSTB funded
Ac

by FRSQ.

References
[1] Schild HG. Poly(N-isopropylacrylamide): experiment, theory and application. Prog Polym Sci
1992;17:163-249.

[2] Wu C. A comparison between the 'coil-to-globule' transition of linear chains and the volume
phase transition of spherical microgels. Polymer 1998;39:4609-4619.

31
Page 31 of 71
[3] Aseyev V, Tenhu H, Winnik FM. Non-ionic thermoresponsive polymers in water. Adv Polym
Sci 2011;242:29-89.

[4] Hocine S, Li MH. Thermoresponsive self-assembled polymer colloids in water. Soft Matter
2013;9:5839-5861.

[5] Roy D, Brooks WLA, Sumerlin BS. New directions in thermoresponsive polymers. Chem Soc

t
ip
Rev 2013;42:7214-7243.

[6] Gil ES, Hudson SM. Stimuli-responsive polymers and their bioconjugates. Prog Polym Sci

cr
2004;29:1173-1222.

us
[7] Rodrıguez-Hernandez J, Checot F, Gnanou Y, Lecommandoux S. Toward ‘smart’ nano-objects
by self-assembly of block copolymers in solution. Prog Polym Sci 2005;30:691-724.

an
[8] Du J, O’Reilly R. Advances and challenges in smart and functional polymer vesicles. Soft
Matter 2009;5:3544-3561.
M
[9] Roy D, Cambre JN, Sumerlin BS. Future perspectives and recent advances in stimuli-responsive
Materials. Prog Polym Sci 2010;35:278-301.
d

[10] Liu F, Urban MW. Recent advances and challenges in designing stimuli-responsive polymers.
te

Prog Polym Sci 2010;35:3-23.

[11] Cheng R, Meng F, Deng C, Klok HA, Zhong Z. Dual and multi-stimuli responsive polymeric
p

nanoparticles for programmed site-specific drug delivery. Biomaterials 2013;34:3647-3657.


ce

[12] Zhuang J, Gordon MR, Ventura J, Li L, Thayumanavan S. Multi-stimuli responsive


macromolecules and their assemblies. Chem Soc Rev 2013;42:7421-7435.
Ac

[13] Gohy JF, Zhao Y. Photo-responsive block copolymer micelles: design and behavior. Chem Soc
Rev 2013;42:7117-7129.

[14] Dimitrov I, Trzebicka B, Müller AHE, Dworak A, Tsvetanov CB. Thermosensitive water-
soluble copolymers with doubly responsive reversibly interacting entities. Prog Polym Sci
2007;32:1275-1343.

[15] Hoffman AS. Stimuli-responsive polymers: Biomedical applications and challenges for clinical
translation. Adv Drug Deliv Rev 2013;65:10-16.
32
Page 32 of 71
[16] Shim MS, Kwon YJ. Stimuli-responsive polymers and nanomaterials for gene delivery and
imaging applications. Adv Drug Deliv Rev 2012;64:1046-1059.

[17] Ganta S, Devalapally H, Shahiwala A, Amiji M. A review of stimuli-responsive nanocarriers for


drug and gene delivery. J Control Release 2008;126:187-204.

[18] Bajpai AK, Shukla SK, Bhanu S, Kankane S. Responsive polymers in controlled drug delivery.

t
ip
Prog Polym Sci 2008;33:1088-1118.

[19] Schmaljohann D. Thermo- and pH-responsive polymers in drug delivery. Drug Deliv Rev

cr
2006;58:1655-1670.

us
[20] Andriola Silva Brun-Graeppi AK, Richard C, Bessodes CM, Scherman D, Merten OW. Cell
microcarriers and microcapsules of stimuli-responsive polymers. J Control Release 2011;149:209-224.

an
[21] Fleige E, Quadir MA, Haag R. Stimuli-responsive polymeric nanocarriers for the controlled
transport of active compounds: Concepts and applications. Adv Drug Deliv Rev 2012;64:866-884.
M
[22] Islam MR, Lu Z, Li X, Sarker AK, Hu L, Choi P, Li X, Hakobyan N, Serpe MJ. Responsive
polymers for analytical applications: A review. Anal Chim Acta 2013;789:17- 32.
d

[23] Kikuchi A, Okano T. Intelligent thermoresponsive polymeric stationary phases for aqueous
te

chromatography of biological compounds. Prog Polym Sci 2002;27:1165-1193.

[24] Wandera D, Wickramasinghe SR, Husson SM. Stimuli-responsive membranes. J Membrane Sci
p

2010;357:6-35.
ce

[25] Kumar A, Srivastava A, Galaev IY, Mattiasson B. Smart polymers: Physical forms and
bioengineering applications. Prog Polym Sci 2007;32:1205-1237.
Ac

[26] Seuring J, Agarwal S. Polymers with upper critical solution temperature in aqueous solution.
Macromol Rapid Commun 2012;33:1898−1920.

[27] Laschewsky A, Rekai ED, Wischerhoff E. Tailoring of stimuli-responsive water soluble


acrylamide and methacrylamide polymers. Macromol Chem Phys 2001;202:276-286.

[28] Liu R, Fraylich M, Saunders BR. Thermoresponsive copolymers: from fundamental studies to
applications. Colloid Polym Sci 2009;287:627-643.

33
Page 33 of 71
[29] Theato P. One is enough: Influencing polymer properties with a single chromophoric unit.
Angew Chem Int Ed 2011;50:5804-5806.

[30] Topp MDC, Dijkstra PJ, Talsma H, Feijen J. Thermosensitive micelle-forming block
copolymers of poly(ethylene glycol) and poly (N-isopropylacrylamide). Macromolecules 1997;30:8518-
8520.

t
ip
[31] Zhu PW, Napper DH. Aggregation of block copolymer microgels of poly (N-
isopropylacrylamide) and poly(ethylene glycol). Macromolecules 1999;32:2068-2070.

cr
[32] Virtanen J, Holappa S, Lemmetyinen H, Tenhu H. Aggregation in aqueous poly (N-
isopropylacrylamide)-block-poly(ethylene oxide) solutions studied by fluorescence spectroscopy and

us
light scattering. Macromolecules 2002;35:4763-4769.

an
[33] Zhang W, Shi L, Wu K, An Y. Thermoresponsive micellization of poly(ethylene glycol)-b-
poly(N-isopropylacrylamide) in water. Macromolecules 2005;38:5743-5747.
M
[34] Weber C, Hoogenboom R, Schubert US. Temperature responsive bio-compatible polymers
based on poly(ethylene oxide) and poly(2-oxazoline)s. Prog Polym Sci 2012;37:686- 714.

[35]
d

Smith AE, Xu X, McCormick CL. Stimuli-responsive amphiphilic (co)polymers via RAFT


polymerization. Prog Polym Sci 2010;35:45-93.
te

[36] Lowe A, McCormick CL. Reversible addition–fragmentation chain transfer (RAFT) radical
p

polymerization and the synthesis of water-soluble (co)polymers under homogeneous conditions in


ce

organic and aqueous media. Prog Polym Sci 2007;32:283-351.

[37] Lee HI, Pietrasik J, Sheiko SS, Matyjaszewski K. Stimuli-responsive molecular brushes. Prog
Ac

Polym Sci 2010;35:24-44.

[38] Aoshima S, Kanaoka S. Synthesis of stimuli-responsive polymers by living polymerization:


Poly(N-isopropylacrylamide) and poly(vinyl ether)s. Adv Polym Sci 2008;210:169-208.

[39] Braunecker WA, Matyjaszewski K. Controlled/living radical polymerization: Features,


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

[40] Golas PL, Matyjaszwski K. Marrying click chemistry with polymerization: expanding the scope
of polymeric materials. Chem Soc Rev 2010;39:1338-1354.

34
Page 34 of 71
[41] Mansfield U, Pietsch C, Hoogenboom R, Becer CR, Schubert US. Clickable initiators,
monomers and polymers in controlled radical polymerizations – a prospective combination in polymer
science. Polym Chem 2010;1:1560-1598.

[42] Moad G, Rizzardo E, Thang SH. Living radical polymerization by the RAFT process. Aust J
Chem 2005;58:379-410.

t
ip
[43] Moad G, Rizzardo E, Thang SH. Toward living radical polymerization. Acc Chem Res
2008;41:1133-1142.

cr
[44] Moad G, Rizzardo E, Thang SH. Radical addition-fragmentation chemistry in polymer
synthesis. Polymer 2008;49:1079-1131.

us
[45] Matyjaszewski K, Xia J. Atom transfer radical polymerization. Chem Rev 2001;101:2921-2990.

an
[46] Matyjaszewski K. Atom transfer radical polymerization (ATRP): Current status and future
perspectives. Macromolecules 2012;45:4015-4039.
M
[47] Kamigaito M, Ando T, Sawamoto M. Metal-catalyzed living radical polymerization. Chem Rev
2001;101:3689-3745.
d

[48] Nicolas J, Guillaneuf Y, Lefay C, Bertin D, Gigmes D, Charleux B. Nitroxide-mediated


te

polymerization. Prog Polym Sci 2013;38:63-325.

[49] Destarac M. On the critical role of RAFT agent design in reversible addition-fragmentation
p

chain transfer (RAFT) Polymerization. Polym Rev 2011;51:163-187.


ce

[50] Xiu XP, Winnik FM Facile and efficient one-pot transformation of RAFT polymer end groups
via a mild aminolysis/Michael addition sequence. Macromol Rapid Commun 2006;27:1648-1653.
Ac

[51] Jochum FD, Roth PJ, Kessler D, Theato P. Double thermoresponsive block copolymers
featuring a biotin end group. Biomacromolecules 2010;11:2432-2439.

[52] Jochum FD, Theato P. Temperature- and light-responsive polyacrylamides prepared by a double
polymer analogous reaction of activated ester polymers. Macromolecules 2009;42:5941-5945.

[53] Bergbreiter DE, Hughes R, Besinaiz J, Li C, Osburn PL. Phase-selective solubility of poly(N-
alkylacrylamide)s. J Am Chem Soc 2003;125:8244-8249.

35
Page 35 of 71
[54] Wong SY, Putnam D. Overcoming limiting side reactions associated with an NHS-activated
precursor of polymethacrylamide-based polymers. Bioconj Chem 2007;18:970-982.

[55] Gauthier MA, Gibson MI, Klok HA. Synthesis of functional polymers by post-polymerization
modification. Angew Chem Int Ed 2009;48:48-58.

[56] Lowe AB, McCormick CL. RAFT Polymerization in Homogenenous Aqueous Media: Initiation

t
ip
Systems, RAFT Agent Stability, Monomers and Polymer Structures. In: Barner-Kowollik C, editor.
Handbook of RAFT Polymerization. Weinheim: Wiley-VCH, 2008. p. 235-284.

cr
[57] Gruendling T, Pickford R, Guilhaus M, Barner-Kowollik C. Degradation of RAFT polymers in a
cyclic ether studied via high resolution ESI-MS: Implications for synthesis, storage, and end-group

us
modification. J Polym Sci Part A Polym Chem 2008;46:7447-7461.

an
[58] Keddie DJ. A guide to the synthesis of block copolymers using reversible-addition
fragmentation chain transfer (RAFT) polymerization. Chem Soc Rev 2014;43:496-505.
M
[59] Donovan MS, Lowe AB, Sumerlin BS, McCormick CL. Raft polymerization of N,N-
dimethylacrylamide utilizing novel chain transfer agents tailored for high reinitiation efficiency and
structural control. Macromolecules 2002;35:4123-4132.
d

[60] Tsarevsky NV, Matyjaszewski K. “Green” atom transfer radical polymerization: From process
te

design to preparation of well-defined environmentally friendly polymeric materials. Chem Rev


2007;107:2270-2299.
p
ce

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


Macromolecules 1999;32:4826-4831.
Ac

[62] Rademacher J, Baum M, Pallack ME, Brittain WJ, Simonsick Jr WJ. Atom transfer radical
polymerization of N,N-dimethylacrylamide. Macromolecules 2000;33:284-288.

[63] Strandman S, Tenhu H. Star polymers synthesised with flexible resorcinarene-derived ATRP
initiators. Polymer 2007;48:3938-3951.

[64] Xia Y, Yin X, Burke NAD, Stöver HDH. Thermal response of narrow-disperse poly(N-
isopropylacrylamide) prepared by atom transfer radical polymerization. Macromolecules 2005;38:5937-
5943.

36
Page 36 of 71
[65] Ye J, Narain R. Water-assisted atom transfer radical polymerization of N-isopropylacrylamide:
Nature of solvent and temperature. J Phys Chem B 2009;113:676-681.

[66] Szablan Z, Toy AA, Davis TP, Hao X, Stenzel MH, Barner-Kowollik C. Reversible addition
fragmentation chain transfer polymerization of sterically hindered monomers: Toward well-defined
rod/coil architectures. J Polym Sci Part A Polym Chem 2004;42:2432-2443.

t
ip
[67] Szablan Z, Toy AA, Terrenoire A, Davis TP, Stenzel MH, Müller AHE, Barner-Kowollik C.
Living free-radical polymerization of sterically hindered monomers: improving the understanding of

cr
1,1-disubstituted monomer systems. J Polym Sci Part A Polym Chem 2006;44:3692-3710.

[68] Cao Y, Zhu XX, Luo J, Liu H. Effects of substitution groups on the RAFT polymerization of N-

us
alkylacrylamides in the preparation of thermosensitive block copolymers. Macromolecules
2007;40 :6481-6488.

[69]
an
Bork JF, Wyman DP, Coleman LE. Nitrogen-containing monomers. III. Reactivity ratios of N-
alkyl acrylamides with vinyl monomers. J Appl Polym Sci 1963;7:451-459.
M
[70] Benaglia M, Chiefari J, Chong YK, Moad G, Rizzardo E, Thang SH. Universal (switchable)
RAFT agents. J Am Chem Soc 2009;131:6914-1915.
d

[71] Savoji M, Strandman S, Zhu XX. Block random copolymers of N-alkyl-substituted acrylamides
te

with double thermosensitivity. Macromolecules 2012;45:2001-2006.


p

[72] Mertoglu M, Garnier S, Laschewsky A, Skrabania K, Storsberg J. Stimuli responsive


ce

amphiphilic block copolymers for aqueous media synthesised via reversible addition fragmentation
chain transfer polymerisation (RAFT). Polymer 2005;46:7726-7740.
Ac

[73] Weiss J, Laschewsky A. Temperature-induced self-assembly of triple-responsive triblock


copolymers in aqueous solutions. Langmuir 2011;27:4465-4473.

[74] Skrabania K, Kristen J, Laschewsky A, Akdemir Ö, Hoth A, Lutz JF. Design, synthesis, and
aqueous aggregation behavior of nonionic single and multiple thermoresponsive polymers. Langmuir
2007;23:84-93.

[75] Sugihara S, Yamashita K, Matsuzuka K, Ikeda I, Maeda Y. Transformation of living cationic


polymerization of vinyl ethers to RAFT polymerization mediated by a carboxylic RAFT agent.
Macromolecules 2012;45:794-804.
37
Page 37 of 71
[76] Sugihara S, Iwata J, Miura S, Ma’Radzi AH, Maeda Y. Synthesis of dual thermoresponsive
ABA triblock copolymers by both living cationic vinyl polymerization and RAFT polymerization using
a dicarboxylic RAFT agent. Polymer 2013;54:1043-1052.

[77] Dan M, Su Y, Xiao X, Li S, Zhang W. A new family of thermo-responsive polymers based on


poly[n(4-vinylbenzyl)n,ndialkylamine]. Macromolecules 2013;46:3137-3146.

t
ip
[78] Sugihara S, Kanaoka S, Aoshima S. Stimuli-responsive ABC triblock copolymers by sequential
living cationic copolymerization: multistage self-assemblies through micellization to open association. J

cr
Polym Sci Part A Polym Chem 2004;42:2601-2611.

[79] Sugihara S, Kanaoka S, Aoshima S. Double thermosensitive diblock copolymers of vinyl ethers

us
with pendant oxyethylene groups: unique physical gelation. Macromolecules 2005;38:1919-1927.

an
[80] Wei H, Perrier S, Dehn S, Ravarian R, Dehghani F. One-pot ATRP synthesis of a triple
hydrophilic block copolymer with dual LCSTs and its thermo-induced association behavior. Soft Matter
2012;8:9526-9528.
M
[81] Dai F, Wang P, Wang Y, Tang L, Yang J, Liu W, Li H, Wang G. Double thermoresponsive
polybetaine-based ABA triblock copolymers with capability to condense DNA. Polymer 2008;49:5322-
d

5328.
te

[82] Li C, Buurma NJ, Haq I, Turner C, Armes SP, Castelletto V, Hamley IW, Lewis AL. Synthesis
and characterization of biocompatible, thermoresponsive ABC and ABA triblock copolymer gelators.
p

Langmuir 2005;21:11026-11033.
ce

[83] Mannella GA, La Carrubba V, Brucato V. Measurement of cloud point temperature in polymer
solutions. Rev Sci Instr 2013;84:075118/1-7.
Ac

[84] Schild HG, Tirrell DA. Interaction of poly(N-isopropylacrylamide) with sodium n-alkyl sulfates
in aqueous solution. Langmuir 1991;7:665-671.

[85] Xie D, Ye X, Ding Y, Zhang G, Zhao N, Wu K, Cao Y, Zhu XX. Multistep thermosensitivity of
poly(N-n-propylacrylamide)-block-poly(N-isopropylacrylamide)-block-poly(N,N-
ethylmethylacrylamide) triblock terpolymers in aqueous solutions as studied by static and dynamic light
scattering. Macromolecules 2009;42:2715-2720.

38
Page 38 of 71
[86] Xia Y, Burke NAD, Stöver HDH. End group effect on the thermal response of narrow-disperse
poly(N-isopropylacrylamide) prepared by atom transfer radical polymerization. Macromolecules
2006;39:2275-2283.

[87] Lessard DG, Ousalem M, Zhu XX. Effect of the molecular weight on the lower critical solution
temperature of poly(N,N-diethylacrylamide) in aqueous solutions. Can J Chem 2001;70:1870-1874.

t
ip
[88] Idziak I, Avoce D, Lessard D, Gravel D, Zhu XX. Thermosensitivity of aqueous solutions of
poly(N,N-diethylacrylamide). Macromolecules 1999;32:1260-1263.

cr
[89] Cao Y, Zhao N, Wu K, Zhu XX. Solution properties of a thermosensitive triblock copolymer of
N-alkyl substituted acrylamides. Langmuir 2009;25:1699-1704.

us
[90] Kerker M. The Scattering of Light, and Other Electromagnetic Radiation. New York: Academic

an
Press, 1969. p. 339.

[91] Takahashi R, Sato T, Terao K, Qiu XP, Winnik FM. Self-association of a thermosensitive
M
poly(alkyl-2-oxazoline) block copolymer in aqueous solution. Macromolecules 2012;45:6111-6119.

[92] Cho CS, Jung JH, Sung YK, Lee YM. Effect of polymeric surfactants on the cloud point of
d

poly(N-isopropylacrylamide). Macromol Rapid Commun 1995;15:727-732.


te

[93] Taylor LD, Cerankowski LD. Preparation of films exhibiting a balanced temperature
dependence to permeation by aqueous solutions—a study of lower consolute behavior. J Polym Sci
p

Polym Chem Ed 1975;13:2551-2570.


ce

[94] Feil H, Bae YH, Feijen J, Kim SW. Effect of comonomer hydrophilicity and ionization on the
lower critical solution temperature of N-isopropylacrylamide copolymers. Macromolecules
Ac

1993;26:2496-2500.

[95] McPhee W, Tam KC, Pelton R. Poly(N-isopropylacrylamide) latices prepared with sodium
dodecyl sulfate. J Colloid Interface Sci 1993;156:24-30.

[96] Mao H, Li C, Zhang Y, Furyk S, Cremer PS, Bergbreiter DE. High-throughput studies of the
effects of polymer structure and solution components on the phase separation of thermoresponsive
Polymers. Macromolecules 2004;37:1031-1036.

39
Page 39 of 71
[97] Liu HY, Zhu XX. Lower critical solution temperatures of N-substituted acrylamide copolymers
in aqueous solutions. Polymer 1999;40:6985-6990.

[98] Bergbreiter DE, Avilés-Ramos NA, Ortiz-Acosta D. A Combinatorial approach to studying the
effects of N-alkyl groups on poly(N-alkyl and N,N-dialkylacrylamide) solubility. J Comb Chem
2007;9:609-617.

t
ip
[99] Furyk S, Zhang Y, Ortiz-Acosta D, Cremer PS, Bergbreiter DE. Effects of end group polarity
and molecular weight on the lower critical solution temperature of poly(N-isopropylacrylamide). J

cr
Polym Sci Part A Polym Chem 2006;44:1492-1501.

[100] Fu H, Policarpio DM, Batteas JD, Bergbreiter DE. Redox-controlled ‘smart’ polyacrylamide

us
solubility. Polym Chem 2010;1:631-633.

an
[101] Chua GBH, Roth PJ, Duong HTT, Davis TP, Lowe AB. Synthesis and thermoresponsive
solution properties of poly[oligo(ethylene glycol) (meth)acrylamide]s: biocompatible PEG analogues.
Macromolecules 2012;45:1362-1374.
M
[102] Roth PJ, Davis TP, Lowe AB. Comparison between the LCST and UCST transitions of double
thermoresponsive diblock copolymers: insights into the behavior of POEGMA in alcohols.
d

Macromolecules 2012;45:3221-3230.
te

[103] Lutz JF, Hoth A. Preparation of ideal PEG analogues with a tunable thermosensitivity by
controlled radical copolymerization of 2-(2-methoxyethoxy)ethyl methacrylate and oligo(ethylene
p

glycol) methacrylate. Macromolecules 2006;39:893-896.


ce

[104] Lutz JF, Akdemir Ö, Hoth A. Point by point comparison of two thermosensitive polymers
exhibiting a similar LCST: Is the age of poly(NIPAM) over? J Am Chem Soc 2006;128:13046-13047.
Ac

[105] Weiss J, Li A, Wischerhoff E, Laschewsky A. Water-soluble random and alternating copolymers


of styrene monomers with adjustable lower critical solution temperature. Polym Chem 2012;3:352-361.

[106] Maeda T, Kanda T, Yonekura Y, Yamamoto K, Aoyagi T. Hydroxylated poly(N-


isopropylacrylamide) as functional thermoresponsive materials. Biomacromolecules 2006;7:545-549.

[107] Kotsuchibashi Y, Kuboshima Y, Yamamoto K, Aoyagi T. Synthesis and characterization of


double thermo-responsive block copolymer consisting N-isopropylacrylamide by atom transfer radical
polymerization. J Polym Sci Part A Polym Chem 2008;46:6142-6150.
40
Page 40 of 71
[108] Kotsuchibashi Y, Yamamoto K, Aoyagi T. Assembly behavior of double thermo-responsive
block copolymers with controlled response temperature in aqueous solution. J Colloid Interface Sci
2009;336:67-72.

[109] Kotsuchibashi Y, Ebara M, Yamamoto K, Aoyagi T. ‘‘On–off’’ switching of dynamically


controllable self-assembly formation of double-responsive block copolymers with tunable LCSTs. J

t
Polym Sci Part A Polym Chem 2010;48:4393-4399.

ip
[110] Kermagoret A, Fustin CA, Bourguignon M, Detrembleur C, Jérôme C, Debuigne A. One-pot

cr
controlled synthesis of double thermoresponsive N-vinylcaprolactam-based copolymers with tunable
LCSTs. Polym Chem 2013;4:2575-2583.

us
[111] Weiss J, Laschewsky A. One-step synthesis of amphiphilic, double thermoresponsive diblock
copolymers. Macromolecules 2012;45:4158-4165.

an
[112] Savoji M, Strandman S, Zhu XX. Switchable vesicles formed by diblock random copolymers
with tunable pH- and thermo-responsiveness. Langmuir 2013;29:6823-6832.
M
[113] Lessard BH, Ling EJY, Maric M. Fluorescent, thermoresponsive oligo(ethylene glycol)
methacrylate/9-(4-vinylbenzyl)-9H-carbazole copolymers designed with multiple LCSTs via nitroxide
d

mediated controlled radical polymerization. Macromolecules 2012;45:1879-1891.


te

[114] Tang X, Liang X, Yang Q, Fan X, Shen Z, Zhou Q. AB2-type amphiphilic block copolymers
composed of poly(ethylene glycol) and poly(N-isopropylacrylamide) via single-electron transfer living
p

radical polymerization: Synthesis and characterization. J Polym Sci Part A Polym Chem 2009;47:4420-
ce

4427.

[115] Qin S, Geng Y, Discher DE, Yang S. Temperature-controlled assembly and release from
Ac

polymer vesicles of poly(ethylene oxide)-block-poly(N-isopropylacrylamide). Adv Mater


2006;18:2905-2909.

[116] Lee KK, Jung JC, Jhon MS. The synthesis and thermal phase transition behavior of poly(N-
isopropylacrylamide)-b-poly(ethylene oxide). Polym Bull 1998;40:455-460.

[117] Zhao J, Zhang G, Pispas S. Morphological transitions in aggregates of thermosensitive


poly(ethylene oxide)-b-poly(N-isopropylacrylamide) block copolymers prepared via RAFT
polymerization. J Polym Sci Part A Polym Chem 2009;47:4099-4110.

41
Page 41 of 71
[118] Zhu K, Pamies R, Kjøniksen AL, Nyström B. Temperature-induced intermicellization of “hairy”
and “crew-cut” micelles in an aqueous solution of a thermoresponsive copolymer. Langmuir
2008;24:14227-14233.

[119] Maleki A, Zhu K, Pamies R, Schmidt RR, Kjøniksen AL, Karlsson G, Cifre JGH, de la Torre
JG, Nyström B. Effect of polyethylene glycol (PEG) length on the association properties of

t
temperature-sensitive amphiphilic triblock copolymers (PNIPAAMm-b-PEGn-b-PNIPAAMm) in

ip
aqueous solution. Soft Matter 2011;7:8111-8119.

cr
[120] Van Durme K, Van Assche G, Aseyev V, Raula J, Tenhu H, Van Mele B. Influence of
macromolecular architecture on the thermal response rate of amphiphilic copolymers, based on poly(N-

us
isopropylacrylamide) and poly(oxyethylene), in water. Macromolecules 2007;40:3765-3772.

[121] Dinçer S, Tuncel A, Piskin E. A potential gene delivery vector: N-isopropylacrylamide-ethylene

an
imine block copolymers. Macromol Chem Phys 2002;203:1460-1465.

[122] Convertine AJ, Lokitz BS, Vasileva Y, Myrick LJ, Scales CW, Lowe AB, McCormick CL.
M
Direct synthesis of thermally responsive DMA/NIPAM diblock and DMA/NIPAM/DMA triblock
copolymers via aqueous, room temperature RAFT polymerization. Macromolecules 2006;39:1724-
d

1730.
te

[123] Chen Y, Sone M, Fuchise K, Sakai R, Kakuchi R, Duan Q, Sun J, Narumi A, Satoh T, Kakuchi
T. Structural effect of a series of block copolymers consisting of poly(N-isopropylacrylamide) and
p

poly(N-hydroxyethylacrylamide) on thermoresponsive behavior. React Funct Polym 2009;69:463-469.


ce

[124] Weiss J, Böttcher C, Laschewsky A. Self-assembly of double thermoresponsive block


copolymers end-capped with complementary trimethylsilyl groups. Soft Matter 2011;7:483-492.
Ac

[125] Lu Y, Chen T, Mei A, Chen T, Ding Y, Zhang X, Xu J, Fan Z, Du B. Solution behaviors and
microstructures of PNIPAm-P123-PNIPAm pentablock terpolymers in dilute and concentrated aqueous
solutions. Phys Chem Chem Phys 2013;15:8276-8286.

[126] Lutz JF. Polymerization of oligo(ethylene glycol) (meth)acrylates: Toward new generations of
smart biocompatible materials. J Polym Sci Part A Polym Chem 2008;46:3459-3470.

[127] Berndt I, Richtering W. Doubly temperature sensitive core-shell microgels. Macromolecules


2003;36:8780-8785.

42
Page 42 of 71
[128] Xu J, Luo S, Shi W, Liu S. Two-stage collapse of unimolecular micelles with double
thermoresponsive coronas. Langmuir 2006;22:989-997.

[129] Savoji M, Strandman S, Zhu XX. Invertible vesicles and micelles formed by dually-responsive
diblock random copolymers in aqueous solutions. Soft Matter 2014;10:5886-5893.

[130] Smith AE, Xu X, Kirkland-York SE, Savin DA, McCormick CL. “Schizophrenic” self-assembly

t
ip
of block copolymers synthesized via aqueous RAFT polymerization: from micelles to vesicles.
Macromolecules 2010;43:1210-1217.

cr
[131] Pietsch C, Mansfield U, Guerrero-Sanchez C, Hoeppener S, Vollrath A, Wagner M,
Hoogenboom R, Saubern S, Thang SH, Becer CR, Chiefari J, Schubert US. Thermo-induced self-

us
assembly of responsive poly(DMAEMAbDEGMA) block copolymers into multi- and unilamellar
vesicles. Macromolecules 2012;45:9292-9302.

an
[132] Smith AE, Xu X, Abell TU, Kirkland SE, Hensarling RM, McCormick CL. Tuning
nanostructure morphology and gold nanoparticle “locking” of multi-responsive amphiphilic diblock
M
copolymers. Macromolecules 2009;42:2958-2964.

[133] Balsara NP, Tirrell M, Lodge TP. Micelle formation of BAB triblock copolymers in solvents
d

that preferentially dissolve the A block. Macromolecules 1991;24:1975-1986.


te

[134] Dimitrov P, Rangelov S, Dworak A, Tsevtanov CB. Synthesis and associating properties of
poly(ethoxyethyl glycidyl ether)/poly(propylene oxide) triblock copolymers. Macromolecules
p

2004;37:1000-1008.
ce

[135] Tsitsilianis C. Responsive reversible hydrogels from associative ‘‘smart’’ macromolecules. Soft
Matter 2010;6:2372-2388.
Ac

[136] Hamley IW, Daniel C, Mingvanish W, Mai SM, Booth C, Messe L, Ryan AJ. From hard spheres
to soft spheres: The effect of copolymer composition on the structure of micellar cubic phases formed
by diblock copolymers in aqueous solution. Langmuir 2000;16:2508-2514.

[137] Bazin G, Zhu XX. Crystalline colloidal arrays from the self-assembly of polymer microspheres.
Prog Polym Sci 2013;38:406-419.

[138] Jin N, Zhang H, Jin S, Dadmun MD, Zhao B. Tuning of thermally induced sol-to-gel transitions
of moderately concentrated aqueous solutions of doubly thermosensitive hydrophilic diblock
43
Page 43 of 71
copolymers poly(methoxytri(ethylene glycol) acrylate)-b-poly(ethoxydi(ethylene glycol) acrylate-co-
acrylic acid). J Phys Chem B 2012;116:3125-3137.

[139] Jin N, Woodcock JW, Xue C, O’Lenick TG, Jiang X, Jin S, Dadmun MD, Zhao B. Tuning of
thermo-triggered gel-to-sol transition of aqueous solution of multi-responsive diblock copolymer
poly(methoxytri(ethylene glycol) acrylate-co-acrylic acid)-b-poly(ethoxydi(ethylene glycol) acrylate).

t
Macromolecules 2011;44:3556-3566.

ip
[140] Jin N, Zhang H, Jin S, Dadmun MD, Zhao B. Shifting sol−gel phase diagram of a doubly

cr
thermosensitive hydrophilic diblock copolymer poly(methoxytri(ethylene glycol) acrylate-co-acrylic
acid)-b-poly(ethoxydi(ethylene glycol) acrylate-co-acrylic acid) in aqueous solution. Macromolecules

us
2012;45:4790-4800.

[141] Seuring J, Agarwal S. Polymers with upper critical solution temperature in aqueous solution:

an
unexpected properties from known building blocks. ACS Macro Lett 2013;2:597-600.

[142] Longenecker R, Mu T, Hanna M, Burke NAD, Stöver HDH. Thermally responsive 2-


M
hydroxyethyl methacrylate polymers: Soluble-insoluble and soluble-insoluble-soluble transitions.
Macromolecules 2011;44:8962-8971.
d

[143] Plamper FA, Schmalz A, Ballauff M, Müller AHE. Tuning the thermoresponsiveness of weak
te

polyelectrolytes by pH and light: Lower and upper critical-solution temperature of poly(N,N-


dimethylaminoethyl methacrylate). J Am Chem Soc 2007;129:14538-14539.
p

[144] Plamper FA, Synatschke CV, Majewski AP, Schmalz A, Schmalz H, Müller AHE. Star-shaped
ce

poly[2-(dimethylamino)ethyl methacrylate] and its derivatives: toward new properties and applications.
Polimery 2014;59:66-73.
Ac

[145] Zhang Q, Hing JD, Hoogenboom R. A triple thermoresponsive schizophrenic diblock


Copolymer. Polym Chem 2013;4:4322-4325.

[146] Hu J, Ge Z, Zhou Y, Zhang Y, Liu S. Unique thermo-induced sequential gel-sol-gel transition of


responsive multiblock copolymer-based hydrogels. Macromolecules 2010;43:5184-5187.

[147] Hoogenboom R, Thijs HML, Wouters D, Hoeppener S, Schubert US. Tuning solution polymer
properties by binary water–ethanol solvent mixtures. Soft Matter 2008;4:103-107.

44
Page 44 of 71
[148] Lambermont-Thijs HML, Hoogenboom R, Fustin CA, Bomal-D’Haese C, Gohy JF, Schubert
US. Solubility behavior of amphiphilic block and random copolymers based on 2-ethyl-2-oxazoline and
2-nonyl-2-oxazoline in binary water–ethanol mixtures. J Polym Sci Part A Polym Chem 2009;47:515-
522.

[149] Su Y, Dan M, Xiao X, Wang X, Zhang W. A new thermo-responsive block copolymer with

t
tunable upper critical solution temperature and lower critical solution temperature in the alcohol/water

ip
mixture. J Polym Sci Part A Polym Chem 2013;51:4399-4412.

cr
[150] Lambermont-Thijs HML, van Kuringen HPC, van der Put JPW, Schubert US, Hoogenboom R.
Temperature induced solubility transitions of various poly(2-oxazoline)s in ethanol-water solvent

us
mixtures. Polymers 2010;2:188-199.

[151] Reddy PM, Venkatesu P. Ionic liquid modifies the lower critical solution temperature (LCST) of

an
poly(N-isopropylacrylamide) in aqueous solution. J Phys Chem B 2011;115:4752-4757.

[152] Lee HN, Newell N, Bai Z, Lodge TP. Unusual lower critical solution temperature phase
M
behavior of poly(ethylene oxide) in ionic liquids. Macromolecules 2012;45:3627-3633.

[153] Asai H, Fujii K, Ueki T, Showamura S, Nakamura Y, Kitazawa Y, Watanabe M, Han YS, Kim
d

TH, Shibayama M. Structural study on the UCST-type phase separation of poly(N-isopropylacrylamide)


te

in ionic liquid. Macromolecules 2013;46:1101-1106.

[154] Guerrero-Sanchez C, Gohy JF, D’Haese C, Thijs H, Hoogenboom R, Schubert US. Controlled
p

thermoreversible transfer of poly(oxazoline) micelles between an ionic liquid and water. Chem
ce

Commun 2008;2753-2755.

[155] Lee HN, Bai Z, Newell N, Lodge TP. Micelle/inverse micelle self-assembly of a PEO−PNIPAm
Ac

block copolymer in ionic liquids with double thermoresponsivity. Macromolecules 2010;43:9522-9528.

[156] Virtanen J, Arotçaréna M, Heise B, Ishaya S, Laschewsky A, Tenhu H. Dissolution and


aggregation of a poly(NIPA-block-sulfobetaine) copolymer in water and saline aqueous solutions.
Langmuir 2002;18:5360-5365.

[157] Arotçaréna M, Heise B, Ishaya S, Laschewsky A. Switching the inside and the outside of
aggregates of water-soluble block copolymers with double thermoresponsivity. J Am Chem Soc
2002;124:3787-3793.

45
Page 45 of 71
[158] Weaver JVM, Armes SP, Bütün V. Synthesis and aqueous solution properties of a well-defined
thermo-responsive schizophrenic diblock copolymer. Chem Commun 2002;2122-2123.

[159] Maeda Y, Mochiduki H, Ikeda I. Hydration changes during thermosensitive association of a


block copolymer consisting of LCST and UCST blocks. Macromol Rapid Commun 2004;25:1330-
1334.

t
ip
[160] Tian HY, Yan JJ, Wang D, Gu C, You YZ, Chen XS. Synthesis of thermo-responsive polymers
with both tunable UCST and LCST. Macromol Rapid Commun 2011;32:660-664.

cr
[161] Hua F, Jiang X, Zhao B. Temperature-induced self-association of doubly thermosensitive
diblock copolymers with pendant methoxytris(oxyethylene) groups in dilute aqueous solutions.

us
Macromolecules 2006;39:3476-3479.

an
[162] Mori H, Kato I, Endo T. Dual-stimuli-responsive block copolymers derived from proline
derivatives. Macromolecules 2009;42:4985-4992.
M
[163] Mori H, Kato I, Saito S, Endo T. Proline-based block copolymers displaying upper and lower
critical solution temperatures. Macromolecules 2010;43:1289-1298.

[164] Han X, Zhang X, Yin Q, Hu J, Liu H, Hu Y. Thermoresponsive diblock copolymer with tunable
d

soluble–insoluble and soluble–insoluble–soluble transitions. Macromol Rapid Commun 2013;34:574-


te

580.
p

[165] Plamper FA, Ruppel M, Schmalz A, Borisov O, Ballauff M, Müller AHE. Tuning the
ce

thermoresponsive properties of weak polyelectrolytes: Aqueous solutions of star-shaped and linear


poly(N,N-dimethylaminoethyl methacrylate). Macromolecules 2007;40:8361-8366.
Ac

[166] Housni A, Zhao Y. Gold nanoparticles functionalized with block copolymers displaying either
LCST or UCST thermosensitivity in aqueous solution. Langmuir 2010;26:12933-12939.

[167] Li C, Lavigueur C, Zhu XX. Aggregation and thermoresponsive properties of new star block
copolymers with a cholic acid core. Langmuir 2011;27:11174-11179.

[168] Le Devedec F, Strandman S, Baille WE, Zhu XX. Functional star block copolymers with a
cholane core: Thermo-responsiveness and aggregation behavior. Polymer 2013;54:3898-3903.

46
Page 46 of 71
[169] Giguère G, Zhu XX. Functional star polymers with a cholic acid core and their thermosensitive
properties. Biomacromolecules 2010;11:201-206.

[170] Trinh LTT, Lambermont-Thijs HML, Schubert US, Hoogenboom R, Kjøniksen AL.
Thermoresponsive poly(2-oxazoline) block copolymers exhibiting two cloud points: complex multistep
assembly behavior. Macromolecules 2012;45:4337-4345.

t
ip
[171] Jia YG, Zhu XX. Complex thermoresponsive behavior of diblock polyacrylamides. Polym Chem
2014;5:4358-4364.

cr
[172] Le Devedec F, Fuentealba D, Strandman S, Bohne C, Zhu XX. Aggregation behavior of
pegylated bile acid derivatives. Langmuir 2012;28:13431-13440.

us
[173] Le Devedec F, Strandman S, Hildgen P, Leclair G, Zhu XX. PEGylated bile acids for use in

an
drug delivery systems: Enhanced solubility and bioavailability of itraconazole. Mol Pharm
2013;10:3057-3066.
M
[174] Strandman S, Le Devedec F, Zhu XX. Thermosensitivity of bile acid-based oligo(ethylene
glycol) stars in aqueous solutions. Macromol Rapid Commun 2011;32:1185-1189.

[175] Strandman S, Le Devedec F, Zhu XX. Self-assembly of bile acid-PEG conjugates in aqueous
d

solutions. J Phys Chem B 2013;117:252-258.


p te
ce
Ac

47
Page 47 of 71
NOTE TO THE TYPE SETTER:

Neither this note nor the following, between "start" and "end" are to be typeset, but must be retained as is
with the .doc version to maintain the correct reference numbers in the text. The reference list above, which

t
is the list to be used in the publication, is not so linked. This is not required with the .pdf version of the

ip
manuscript also provided herewith.

cr
+++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++++

START

us
an
M
d
p te
ce
Ac

48
Page 48 of 71
S S S S
Pn + S S
R Pn C

t
C R Pn C + R

ip
M Z Z Z
S S S S
R Z R Z1 R Z R

cr
S S S N S O S
(a) (b) Z2 (c) (d)

N S

us
N S
R R

N S HN S (e)

an
Scheme 1.
M
d
p te
ce
Ac

54
Page 49 of 71
kact
R-X + M tn-Y/Ligand R + X-M tn+1-Y/Ligand

t
kdeact
kp

ip
kt
monomer termination

cr
N N N
N N N N
(a) (b)

us
H 3C(H2C)8 (CH2)8CH 3 N

N
N N
N (c) N (d)

an
M
d

Scheme 2.
p te
ce
Ac

55
Page 50 of 71
O S COOH O
m n p q O m x y
p n
O O O O O O O O
OH (a) (b)
N N NH NH NH HN NH

t
ip
OH
Z
m p m m
R n n

cr
m n p

O O O O
(c) (d) (e)
O NH NH O

us
CH2 COOH CH2 CH2 CH2
O O
x N N N N
x
R1 R1 R2 R2 R1 R1

an
R= H2 C TMS
S
Z= S C (CH2 )3 TMS
M
O 3S

m n p m m m n p
n
O O O (f) O O O (g) O O O (h)
d

NH N N NH
te

O O O
O O
p

O
O
m n m p m n Br
ce

O O O O O
(i) (j)
HN O HN O HN

O O OCs
CsO O O
Ac

n n
m m
O O (k)
O
N N O P O O
O O
O O

O3S O3 S N

Scheme 3.

56
Page 51 of 71
O
m x y Br
n
O O O O
NH HN NH (a)

R1 = H
RO
R2 = -CH2CH(CH3 )2

t
ip
S

HOOC x 1-x n y 1-y S SC12H25


m
O O O O

cr
(b)
NH NH NH NH

us
S
Ph
x 1-x n y 1-y m S S(CH2 )2COOH

an
O O O O
(c)
NH NH NH NH M
Et2N

m x y
n
d
O O
N N N
(d)
O
te

O
m n Cl
O O O
(e)
p

x 1-x n y 1-y
m
ce

O O O O O O
x
(f)
O OH O OH

O
Ac

O O
3 2

R R
N N
O O O O

n m n

(g)
(h)

CH2 CH2 H2 C

O O O

O O O 57
x 4 4

Page 52 of 71
Scheme 4.

t
ip
cr
us
an
M
d
te
p
ce
Ac

58
Page 53 of 71
S S

t
n/ 2 m/2 n/2
m/2
S

ip
O O O O
p
NH N N NH
N N (a)

cr
us
n m (b)
O O
O O
O3S N N

an
O

O
M
n
m (c)
O N N N
O m x 1-x n
N O O
N
HO (d)
d
O OH
OH
O
O
p te

(e)
ce
Ac

Scheme 5.

59
Page 54 of 71
t
ip
cr
us
an
M
d
te
p
ce

Fig. 1
Ac

60
Page 55 of 71
t
ip
cr
us
an
M
d
te
p
ce

Fig. 2.
Ac

61
Page 56 of 71
t
ip
cr
us
an
M
d
te
p

Fig. 3.
ce
Ac

62
Page 57 of 71
t
ip
cr
us
an
M
d
te
p
ce
Ac

Fig. 4.

63
Page 58 of 71
t
ip
cr
us
an
M
d
te
p
ce

Fig. 5.
Ac

65
Page 59 of 71
t
ip
cr
us
an
M
d
te
p
ce
Ac

Fig. 6

66
Page 60 of 71
t
ip
cr
us
an
M
d
te
p
ce
Ac

Fig. 7.

68
Page 61 of 71
t
ip
cr
us
an
M
d
te
p
ce
Ac

Fig. 8.

69
Page 62 of 71
t
ip
cr
us
an
M
d
te
p

Fig. 9.
ce
Ac

70
Page 63 of 71
\

t
ip
cr
us
an
M
d
te
p

Fig. 10.
ce
Ac

71
Page 64 of 71
t
ip
cr
us
an
M
d
te
p
ce

Fig. 11.
Ac

72
Page 65 of 71
t
ip
cr
us
an
M
d
te
p
ce
Ac

Fig. 12.

73
Page 66 of 71
t
ip
cr
us
an
M
d
te
p
ce
Ac

Fig. 13.

75
Page 67 of 71
t
ip
cr
us
an
M
d
te
p
ce
Ac

Fig. 14.

77
Page 68 of 71
t
ip
cr
us
an
M
d
te
p
ce
Ac

Fig. 15.

79
Page 69 of 71
t
ip
cr
us
an
M
d
te
p
ce
Ac

Fig. 16.

81
Page 70 of 71
t
ip
cr
us
an
M
d
te
p
ce
Ac

Fig. 17.

82
Page 71 of 71

You might also like