You are on page 1of 16

Polymer 211 (2020) 123146

Contents lists available at ScienceDirect

Polymer
journal homepage: http://www.elsevier.com/locate/polymer

LCST polymers: Thermoresponsive nanostructured assemblies


towards bioapplications☆
George Pasparakis, Constantinos Tsitsilianis *
Department of Chemical Engineering, University of Patras, 26500, Patras, Greece

A R T I C L E I N F O A B S T R A C T

Keywords: Thermoresponsive polymers constitute an important class of materials for biomedical application due to their
LCST thermally reversible coil-to-globule transition which can be exploited in a multitude of biomedical applications
Thermoresponsive polymer spanning from triggered drug delivery systems in the form of bulk hydrogels and nanoparticles, “smart” cell
Assembly
culture setups, sensors and actuators, and separation technologies. In this perspective article we present the basic
Bioapplication
physicochemical properties of poly (N-isopropylacrylamide) (PNIPAM), which is the most widely studied ther­
moresponsive polymer with a lower critical solution temperature. We describe the basic thermodynamic and
physicochemical parameters that affect the LCST and present selective applications that utilize thermoresponsive
polymers in the form of nano-assemblies including micelles, polymersomes, microcapsules and microgels, as well
as injectable hydrogels for biomedical applications.

1. Introduction that exhibit large thermally-controlled volume changes, thermores­


ponsive micelles that release anticancer drugs in an on-demand manner,
Polymeric materials find numerous applications in the biomedical “smart” PNIPAM-coated surfaces for cell-sheet engineering applications
field as biomaterials, fillers, structural components, sensors and actua­ [11,12], and PNIPAM-protein conjugates with thermally-controlled
tors, tissue engineering matrices and as drug delivery systems. A sub- bioactivity [13–15]. The advent of reversible-deactivation radical
class of polymers for biomedical applications, the so called “smart” polymerization (RDRP) routes [16,17] and conjugation protocols via
polymers, can change their physicochemical properties in response to a “click” chemistries [18,19] further boosted the field as it enabled for
stimulus such as application of a pH gradient, the presence of a specific robust control on the macromolecular architecture on surfaces and in­
analyte, irradiation, or heat, enabling their use as dynamic modulators terfaces as well as the precise attachment on biomolecules and nano­
of biochemical cues at the interface of biological and biomedical sci­ particles which was previously not possible [20].
ences [1–3]. Thermoresponsive polymers constitute an interesting class The possibility to combine PNIPAM with additional stimuli cues
of materials that can change their physicochemical properties by further expanded the potential uses for biomedical applications
application of a thermal stimulus within a narrow temperature window. requiring even subtler polymer multi-stimuli responses in biological
The field of thermoresponsive polymers has grown exponentially over milieu even at isothermal conditions [21,22]. In this perspective article
the last 50 years with important contributions in the fields of controlled we describe the key properties of PNIPAM and other LCST polymers, we
drug delivery, cell culture technologies and tissue engineering, separa­ describe the thermodynamics of LCST and the factors that influence it,
tion analytical biosciences, as well as in sensors and actuators technol­ and finally we present characteristic example applications utilizing
ogies [4–8]. The field was sparked by the seminal work of Heskins and mainly PNIPAM as a responsive macromolecular element. Although
Guillet who first reported the exact thermal transition mechanism of there are several review articles summarizing the concepts of responsive
poly (N-isopropylacrylamide) (PNIPAM) [9], the first and most studied polymers and their applications from various viewpoints, We regard the
synthetic thermoresponsive polymer. Soon after, landmark works fol­ present manuscript not as a classical review but as an introductory
lowed conceptualizing PNIPAM-based thermoresponsive hydrogels [10] reference describing key concepts of LCST-polymers, mostly in a


This perspective article has been submitted for consideration in the special Issue of Polymer entitled: From “Makromolekel” to POLYMER: A Centennial Cele­
bration of Staudinger’s “On Polymerization”
* Corresponding author.
E-mail address: ct@chemeng.upatras.gr (C. Tsitsilianis).

https://doi.org/10.1016/j.polymer.2020.123146
Received 13 July 2020; Received in revised form 8 September 2020; Accepted 13 October 2020
Available online 17 October 2020
0032-3861/© 2020 Published by Elsevier Ltd.
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

Scheme 1. Schematic showing the coil-to-globule transition with well polymer solvation below the LCST and domination of the hydrophobic interactions above the
LCST in a), binodal diagrams for U/LCST polymers depending on polymer fraction ϕ in b) and a typical diagram of transmittance vs. temperature of an aqueous
PNIPAM sample with full transmittance below the LCST followed by complete suppression above the LCST in c).

nano-assembled form, combined with selective recent applications the amide segments conveying a more ordered state of negative ΔS. It is
related to the biomedical field in a wider context. the enthalpic term at this phase that drives mixing as ΔGmix must be
negative. An increase of the temperature triggers perturbation and ul­
1.1. Thermodynamics of LCST timately disruption of the hydrogen bonds at a temperature near the
LCST. At T > LCST, endothermic breaking of hydrogen bonding makes
Polymer solvent interaction roots back to the seminal work by Flory the enthalpic term less negative and the entropy gain renders the
[23] and Huggins [24] who first described the polymer-solvent inter­ entropic term ΤΔS predominant leading to demixing and phase separa­
action parameter and explained the deviation of polymer solubility from tion. At T > LCST, the hydrophobic interactions of the backbone and the
classic solubilization of small molecules, followed by the concept of free isopropyl segments dominate and induce polymer globule formation
volume [25] which was later used to explain the lower/upper critical which at higher concentrations can be observed as flocculation and
solution temperature phenomenon. Generally, thermoresponsive poly­ precipitation in aqueous media. Since the entropic coil-to-globule
mers are categorized in two categories depending on whether they transition affects mostly the re-orientation of water molecules during
exhibit a lower critical solution temperature (LCST) or an upper critical de-mixing, the transitions tend to be quite faster compared to UCST
solution temperature (UCST) (Scheme 1a) [26,27]. Both classes polymers which partly explains why LCST polymers are more widely
comprise polymers that change their physicochemical properties above exploited in the biomedical field compared to their UCST counterparts.
a certain temperature. UCST polymers are non-soluble at low tempera­ It should be noted that PNIPAM in its collapsed state is not fully de-
tures and become soluble above the UCST by an enthalpic-driven pro­ hydrated, hydrophobic interactions are dominant above the LCST
cess [28]. Conversely, LCST-type polymers interact well with their however the globules still contain a substantial amount of water
solvent at low temperatures but undergo sharp coil-to-globule transition (Scheme 1) [36–38]. The thermal transition process is fully reversible
above their LCST and drop out of solution by an entropy increase. Both with minor hysteresis taking place within a few microseconds upon
mechanisms are rooted to the so-called hydrophobic effect [29] of water application of a temperature jump above the LCST [39] depending on
and can be described as binary systems composed of the solvent (or a the polymer concentration the rate of heating/cooling of the experi­
mixture of solvents) and the polymer, with their mixability being mental, the polymer topology and confinement [40] and the presence of
dependent by temperature and the polymer fraction. The phase hydrophobic moieties [41]. The LCST is the lowest temperature of the
boundary is denoted as the binodal and corresponds to the temperature binodal at which demixing is observed and should not be confused with
where de-mixing occurs at each corresponding volume fraction of the the cloud point temperature (Tcp) [42]. The latter is defined as the
polymer (Scheme 1b) [30]. temperature at which phase-separation occurs at a specific polymer
In the context of biomedical applications, by far the most studied concentration which should be given and is usually observed by a
polymers are those interacting with water and in particular polymers characteristic cloudiness of the solution (Scheme 1c). Turbidimetry
that exhibit LCST. For UCST-type polymers the reader can refer to measurements with UV/Vis spectrometry coupled with temperature
excellent review articles here [31,32]. control (i.e. a Peltier unit) are most commonly used to determine the
As previously mentioned, PNIPAM is the most studied thermores­ LCST, although differential scanning calorimetry, light scattering, and
ponsive polymer mainly because it has an LCST onset close to physio­ nuclear magnetic resonance can also be used to determine the thermal
logical temperature (32 ◦ C) and it resembles the structure of isoleucine behavior of these materials [42,43].
and therefore can be regarded as a polypeptide mimic [33]. PNIPAM
undergoes an entropic coil-to-globule transition in aqueous media 1.2. Factors affecting the LCST
(Scheme 1a) [34]. The thermally controlled mixing/de-mixing transi­
tion can be understood by the Gibbs free energy equation ΔG = ΔH-TΔS The LCST of PNIPAM can be affected by many parameters depending
[35]; at low temperatures (Τ<LCST) the enthalpy of mixing is negative on the molecular structure and the aqueous environment. The end-group
due to the formation of hydrogen bonds between water molecules and was shown to minimally affect the LCST when preparing polymers with

2
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

azo-initiators such as the commonly used azoisobutyronitrile (AIBN) Interestingly the interplay of confined polymer topology in combi­
[44]. Minor influence of the LCST was also observed with the use of nation with co-monomers that elicit co-operative interactions with
potassium persulfate initiators which was significant only at lower NIPAM monomer (i.e. via intramolecular H-bonding) can often lead to
molecular weight polymer samples (ca. 2 kDa) [45]. Expectedly, soluble unexpected manipulation motifs of the LCST as elegantly shown in a
end-groups such as -OH, -NH2 tend to increase the LCST while more series of studies by Richtering and co-workers [57–59].
hydrophobic end-groups seem to decrease the LCST [44,46]. These ef­ The LCST is greatly affected by the presence of salts in aqueous
fects have been observed by controlled polymerization routes including media, the so called Hofmeister salt series which are categorized as
atom transfer radical polymerization (ATRP) and reversible cosmotropes and chaotropes, salts that either “make” (stabilize) or
addition-fragmentation chain transfer (RAFT) polymerizations and “break” (de-stabilize) the structure of water molecules and were origi­
generally are more pronounced at samples below ca. 50 kDa [44,47]. nally ranked according to their capacity to “salt-out” (precipitate) pro­
Polymer tacticity seems to affect the solubility as isotactic and syndio­ teins [60]. Cosmotropic salt addition leads to polarization of the water
tactic PNIPAM are insoluble in water; insertion of isotactic blocks across molecules adjacent to the amide units and increase the surface tension of
a random PNIPAM polymer chain was found to suppress the cloud point the hydrophobic segments resulting in suppression of the LCST. Chaot­
[48,49]; ultimately, the monomer sequence [50] across the polymer ropes seem to effect on the surface tension by direct ion binding to the
chain seems to play a critical role on the physicochemical properties in amide moieties and removal of the hydration layer of the hydrophobic
responsive polymers and still poses a synthetic challenge that needs isopropyl and backbone segments, resulting in salting-in of the polymer
further development. [61].
The most effective way to control the exact LCST onset is the use of a Finally, another factor affecting the LCST, albeit less relevant to
comonomer. Water soluble co-monomers contribute to the energy bioapplications, is the so called cononsolvency effect [62] (usually the
required to disrupt H-bonding, hence shifting the LCST to higher tem­ use of a mixture of water with a good organic solvent such as water/­
peratures, while hydrophobic co-monomers tend to have a decremental MeOH) which tends to decrease and increase the LCST onset at moderate
effect. Careful selection of the co-monomers should be considered with and high co-solvent fractions respectively [63,64].
similar reactivity ratios during the polymerization reactions in order to
achieve homogeneous incorporation of the co-monomer across the 2. Other LCST polymers
PNIPAM polymer chain and finely control the LCST shifting with
adequate molecular fidelity. A classic example of co-monomer feed Many polymers with H-bonding sites have a LCST (Scheme 2);
include the use of ionizable monomers such as methacrylic acid [51], notable polymers of interest in the biomedical field include poly ((2-
and diethylaminoethyl methacrylate [52,53]. Since the solubility of dimethylamino)ethyl methacrylate)) (PDMAEMA) (pH-dependent LCST
these co monomers can be tuned by their respective pKa, it is possible to range from 32oC to 53 ◦ C) [65,66], some poly (ethylene oxide)s (PEOs,
precisely adjust the hydrophilic-hydrophobic balance of the final LCST range from 100oC to 150 ◦ C) [67,68] and poly (propylene oxide)s
copolymer to exhibit coil-to-globule transition isothermally. This is (PPO, LCST 10oC–45 ◦ C) [69] within a certain molecular weight range,
particularly useful for applications in vivo where large differences in the hydroxypropyl cellulose (LCST, 45 ◦ C) [70], poly (N-vinyl caprolactam)
pH of different tissues and intracellular compartments can be harnessed (PNVCL) (LCST at 32oC–34 ◦ C) [71], poly (ethylene glycol methacry­
to elicit triggered molecular cargo release (i.e. drugs, nucleic acids, etc.) late)s (LCST range from ca. 10–100 ◦ C) [72] and poly (2-oxazoline)s
in a pH-dependent manner. This approach greatly expand the scope of (pOxs, LCST range from ca. 25–100 ◦ C) [73].
PNIPAM as LCST can be used as an optical readout mechanism [54], as a PDMAEMA has pH dependent LCST due to the ionizable tertiary
reporter of ligand-receptor interactions and biomolecular events [55], as amine moiety which however raises toxicity issues for in vivo applica­
well as hide-and-reveal [56] strategies by simple rational copolymer tions [74,75]. PEOs and PPOs have both been mostly exploited as tri­
design principles. block copolymers with the structure PEO-PPO-PEO which are

Scheme 2. Alterative to PNIPAM polymers with LCST in aqueous media.

3
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

Fig. 1. Schematic representation of various topologies bearing LCST polymeric blocks utilized to fabricate thermoresponsive assemblies: a) A-b-B, b) (A-co-B)-b-(C-
co-D), c) B-b-A-b-B, d) A-b-B-b-C, e) A-b-(C-co-D), f) (C-co-D)-b-A-b-(C-co-D), g) protein-g-A conjugate, h) A-g-B, i) A-g-B-co-C, d) An [B-b-(C-g-D)]n (bold denotes
LCST blocks).

commercially available as Poloxamers, and Pluronics [76] and find uses 3. LCST-based segmented macromolecules
as pharmaceutical excipients. (PNVCL) in an interesting polymer as it
shows similar thermal properties with PNIPAM, and was relatively Stimuli responsive polymeric systems have attracted tremendous
recently polymerized by controlled polymerization methods (see for attention due to a vast variety of potential applications. Temperature is
example [77]), and also it does not degrade under in vivo conditionς one of the stimuli that can be used in the field of biomedicine as the
which is appealing for certain in vivo applications. human body is thermo-regulated at about 37 ◦ C, under physiological
PEGMAs have been proposed as PNIPAM alternatives as they conditions. Moreover, hyperthermia can be applied to solid tumors and
resemble the structure of poly (ethylene glycol) which is currently can synergistically enhance tumor cytotoxicity when combined with
considered as the gold standard biocompatible polymer, and they endow other treatments; hyperthermia preferentially increases cell membrane
chemical versatility owing to their polymerization via controlled poly­ fluidity as well as the permeability of tumor vasculature, facilitating the
merization routes [78]. One of their drawbacks is that they can coor­ delivery of drugs. The potential to harness the benefits of locally applied
dinate heavy metal ions which could pose certain limitations for their hyperthermia motivated the design of associative macromolecules with
use. pOxs have recently emerged as a promising PNIPAM alternative, thermo-responsiveness in the form of block copolymers. In recent years,
they are synthesized by living cationic ring opening polymerization, the enormous growth of Macromolecular Engineering through the
they do not suffer hysteresis issues during recovery (i.e. from heating to development of controlled/living polymerization methods [91], has
cooling) and more importantly they exhibit protein repulsion properties offered numerous possibilities in designing tailor-made thermo-res­
similar to poly (ethylene glycol) (PEG) which renders them attractive for ponsive block copolymers of various topologies (Fig. 1). As it is
the biomedical field [79]. well-known, block copolymers self-assemble in selective media through.
Elastin-like-peptides (ELPs) constitute an interesting alternative to Distinct interactions, forming nanostructured assemblies of various
responsive polymers that exhibit unique physicochemical properties that morphologies, i.e. micellar assemblies and/or 3D networks, depending
could render them viable candidates as responsive nanocarriers of on their macromolecular Architecture (Fig. 1). Hence, the integration of
translational potential. ELPs are oligo-/poly-amino acid sequences with blocks exhibiting LCST type of behavior, endow these self-assemblies
the general repeating unit [aPGbG]n [80] where a = I or V, b is any amino with reversible thermo-responsiveness. For instance, in block co­
acid except P, and n is the number of repeating units [81–83]. ELPs un­ polymers/aqueous media systems, the LCST-type block can be trans­
dergo an inverse temperature responsive transition (Tt [84], unlike the formed from hydrophilic to hydrophobic upon heating as passing
LCST in thermoresponsive polymers) where they exist in a fully soluble through its LCST, inducing reversible hydrophobic association. This
state below Tt and collapse in the form of aggregates/coacervates above kind of effect has inspired the design of “smart” thermo-responsive
the Tt. Phase separation is a reversible phenomenon and the precise Tt can polymeric and/or polymer/inorganic hybrid nanostructured assem­
be tuned by varying n, a, and b, and is also dependent on the ELP con­ blies comprising micelles, nanogels micro/nano-capsules as well as 3D
centration and the ionic strength of the medium. networks, targeting biomedical applications i.e. as drug delivery systems
In a seminal research article, Rees et al. [85] demonstrated that a (DDS). Bellow we demonstrate selective examples of these kinds of
single repeating octapeptide unit exhibited unique entropy-driven ther­ thermo-responsive systems.
mally induced transitions in a fully reversible manner similar to its
polymeric analogue. This striking finding implies that the Tt is not a 4. Nanostructured assemblies
collective phenomenon derived by the cooperative interaction between
the repeating units, but that each separate peptide unit exhibits indi­ 4.1. Micelles
vidual thermoresponsive properties; also, it was shown that the config­
urational entropy (ΔS) for the formation of type-II turns is independent of The simplest thermoresponsive block copolymer is of the A-b-B
the peptide sequence length. Hathorne and Bermudez [86], and Nuhn diblock topology where a LCST-type A block is joined together with
and Klok [87] confirmed these findings by the systematic analysis of the another B block by their end-groups by a single covalent bond. Two
conformational properties of short ELPs (n = 1–8) as a function of tem­ kinds of micelles can be formed depending on the nature of the B block,
perature changes around Tt by circular dichroism and molecular namely amphiphilic or double hydrophilic if the B block is hydrophobic,
modelling. Further studies showed that ELPs constitute excellent build­ or hydrophilic respectively. Concerning the first class, a PNIPAM-b-
ing blocks to form T-responsive micelles [88,89] by simple coupling with PBMA (poly (butyl methacrylate)) diblock copolymer self-assembles in
hydrophilic polypeptide chains or water soluble polymers (i.e. PEG) with aqueous media at room temperature through hydrophobic association,
applications in precision oncology [90] and hyperthermia therapeutics forming core-shell micelles with a PBMA hydrophobic core, stabilized by
[82]. the thermosensitive PNIPAM shell [92]. These micelles were tested as

4
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

drug carriers by loading doxorubicin in their cores. By increasing tem­ thermoresponsive micelles self-assembled from (c-PNIPAM)-b-PCL,
perature above the PNIPAM’s LCST, the outer hydrophilic shell, that exhibited lower LCST and improved drug loading and releasing capac­
prevents inner core interaction with biocomponents and other micelles, ity. These results indicate the importance of macromolecular architec­
can be rapidly switched to hydrophobic, allowing drug release due to ture of thermoresponsive block copolymers, on the self-assembling,
colloidal destabilization of the micelles. The dramatic thermal phase transition properties and their functions as drug nano­
thermo-responsive on/off switching behavior of drug release correlated carriers for controlled release.
with the in vitro cytotoxicity opened opportunities to construct novel Double hydrophilic block copolymers constituted of PEG (hydro­
drug delivery systems in combination with localized hyperthermia. The philic and biocompatible with “stealth” properties) and thermosensitive
same group expanded their work by replacing the hydrophobic block N-(2-hydroxyethyl)methacrylamide-oligolactates, P(HEMAm-Lacn)
with the biodegradable poly (D,L-lactide) (PLA), while the LCST of were designed, as efficient drug carrier systems for therapeutic appli­
PNIPAAm was adjusted to 39.4 ◦ C, slightly above the physiological cations [98]; conveniently, the LCST of the thermosensitive block can be
temperature, by incorporating N,N-dimethylacrylamide moieties. The tuned by the number of lactate moieties thus a block copolymer PEG-b-P
as-designed P(NIPAM-co-DMAAm)-b-PLA copolymer was (80%HEMAm-Lac2)-co-(20%HEMAm-Lac4), exhibiting low LCST,
end-functionalized by a fluorescent molecule to fabricate self-assembled in room temperature forming micellar carriers with fast
self-assembling labeled thermoresponsive micelles [93]. These micelles degradable core and a PEG stabilized shell. Under physiological condi­
showed time-dependent intracellular uptake without exhibiting cyto­ tions (pH 7.4 and 37 ◦ C), the micelles started to swell after 4 h and were
toxicity. In another study [94] it was shown that the intracellular uptake fully destabilized within 8 h due to hydrolysis of the lactate side chains.
is significantly limited at a temperature below the micellar LCST at 37 This destabilization profile seems advantageous for in vivo use because

C, whereas the micelles were internalized into the cytoplasm above the the observed induction period is just long enough to allow accumulation
micellar LCST (e.g. 42 ◦ C), being dependent on polymer concentration, of the micelles after intravenous administration at their site of action, e.
time, and temperature. Importantly, no intracellular uptake of g., a tumor via the enhanced permeation and retention (EPR) effect [99,
PEG-b-PLA micelles was observed, regardless of temperature changes 100].
showing the role of the LCST-block. The enhanced intracellular micelle An interesting double hydrophilic diblock copolymer exhibiting two
uptake was attributed to the enhanced interactions between the micelles LCSTs in distinct blocks were designed, based on PNVCL (Fig. 3) [101].
and cell membranes through the dehydration of corona-forming ther­ The first block is a statistical copolymer of VC with the hydrophobic
moresponsive polymer chains. This temperature dependence of the monomer 3-methyl-N-vinylcaprolactam P(MVC-co-VC) with LCST1
micelle outer shell with the cell membranes was further corroborated by around room temperature (19–27 ◦ C). The second block is a statistical
our group with the use of P(PEGMA)s as the T-responsive element which copolymer of VC with the hydrophilic monomer N-vinylpyrrolidone
enabled the co-delivery of multiple drug combinations both in the form (VP) P(VC-co-VP) with adjusted LCST2 slightly above physiological
of micelles [95] or polymer-coated liposomes [96] (Fig. 2). temperature (41–42 ◦ C). The diblock copolymer P(MVC-co-VC)-b-P
Another PNIPAM-based copolymer with sophisticated macromolec­ (VC-co-VP) self-assemble above the LCST1 forming micelles, enabling
ular architecture, was designed as DDS based on a tadpole-shaped, drug encapsulation in their hydrophobic P (MVC-co-VC) cores. Upon
linear-cyclic diblock copolymer, consisting of the biodegradable hy­ increasing temperature above LCST2, the dehydrated P(VC-co-VP)
drophobic linear pol (ε-caprolactone) (PCL) and the thermoresponsive corona blocks associate leading to colloidally unstable aggregates, fol­
macrocyclic PNIPAM [(c-PNIPAM)-b-PCL] [97]. Compared with the lowed by drug release. Moreover, the elevated temperature can trigger
linear diblock copolymer counterpart, (l-PNIPAM)-b-PCL, the loaded micelle aggregation/accumulation within the tumor

Fig. 2. P(PEGMA) based thermoresponsive block copolymers synthesized by RAFT polymerization and their applications as standalone micellar and liposome
dispersions for hydrophobic and hydrophilic drug loading, respectively. LCST-driven disruption leads to augmented drug release rates as well as higher cellular
uptake due to augmented cell membrane interactions (adapted with permission from references [95,96]).

5
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

Fig. 3. Schematic representation of a double hydrophilic diblock with double LCST response, associating upon heating, (from ref. [102] by permission of ACS).

resulting in enhanced passive targeting. Besides passive targeting DDS, attempts have been undertaken to
Triple thermoresponsive triblock terpolymers, comprising three design active targeting thermoresponsive micellar nanocarriers (Fig. 4).
different blocks exhibiting resolved thermal transitions of the LCST type Thermoresponsive (PEO-PPG-PEO (Pluronic F127, commercially avail­
[poly (N-n-propylacrylamide) (LCST ~22 ◦ C), poly (methoxydiethylene able) covalently bonded with poly (d,l-lactic acid) (F127-PLA, abbrevi­
glycol acrylate) (LCST ~39 ◦ C), and poly (N-ethylacrylamide) (LCST ated as FP) copolymer micelles were developed and decorated with
~73 ◦ C)] have also been synthesized and their self-assembly, by step­ folate (FA) acting as the targeting functional moiety. The system
wise thermal switching, was explored in respect of the block sequence exhibited LCST at 39.2 ◦ C. Thus at 37 ◦ C, a small amount of encapsulated
(ABC, BAC, and ACB) [103]. The self-organization (micellization) of anticancer drug DOX was released from the self-assembling micelles,
such terpolymers and their molecular cargo ability of hydrophobic while at a temperature slightly above LCST (40 ◦ C), the shrinkage of
agents (Nile Red probe) in the collapsed lower LCST block forming the thermoresponsive PEO-PPG-PEO segments induced rapid release of DOX
hydrophobic core of the micelles, were found to depend strongly on the and rapid increase of the drug concentration locally [104]. As demon­
position of the blocks across the macromolecular chain. strated, this copolymer has excellent cytocompatibility, and the

Fig. 4. Schematic representation of thermosensitive nanocarrier working as a targeted drug delivery system, (from ref. [107] by permission of ACS).

6
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

FA-decorated FP micelles exhibited improved cellular uptake and higher as a triggerable DDS. Magnetic heating of the embedded SPIONs led to
cytotoxicity against folate receptor (FR)-overexpressed HeLa cells. increased bilayer permeability through dehydration of the thermores­
Importantly, the cytotoxicity of DOX-loaded FA-FP micelles against ponsive PNIPAM block. Thus, calcein could be released in controlled
HeLa cells was significantly more pronounced under hyperthermia (40 doses solely via exposure to pulses of an alternating magnetic field. This

C) compared to physiological temperature, showing promising poten­ work demonstrates controlled drug release potential applications that
tial as a drug nanocarrier for anticancer applications. combines rational polymersome design with addressed external mag­
Non-linear segmented copolymers have also been employed as DDS. netic field-triggered release [110].
Folate-decorated thermoresponsive micelles based on the star-shaped The coexistence of thermo- and pH-responsive (weak poly­
amphiphilic block copolymer bearing four [poly (ε-caprolactone)-b-2s electrolytes) blocks in the macromolecular chain endows polymers with
(poly (N-isopropylacrylamide-co-acrylamide)-b-methoxy poly (ethylene dual responsiveness. To this direction, multicompartmental polymer­
glycol)/poly (ethylene glycol)-folate)] arms (i.e., 4s [PCL-b-2s (P somes were fabricated by star-graft quarterpolymers, composed of hy­
(NIPAМ-co-AAm)-b-MPEG/PEG-FA)] were developed for the tumor drophobic PS and pH-sensitive P2VP-b-PAA arms, emanated from a
targeted delivery and temperature-induced controlled release of pacli­ common nodule, enriched by thermo-sensitive PNIPAM grafts, cova­
taxel (PTX) [105]. The cytotoxicity studies showed that the transport of lently bonded on the PAA block-arms. These pH/thermo-sensitive pol­
PTX by these micelles into cancer cells was higher than that by the ymersomes were evaluated as nanocarriers for co-delivery of
commercial PTX formulation Tarvexol. More thermoresponsive nano­ doxorubicin (hydrophilic) and paclitaxel (hydrophobic) anticancer
systems based on polymeric micelles can be viewed in previous reviews chemotherapeutic agents, encapsulated in different compartments (i.e.
[27,106]. lumen and hydrophobic membrane, respectively). The loaded poly­
mersomes were found to be internalized efficiently in human lung
adenocarcinoma epithelial cells. Enhanced release was observed for
4.2. Polymersomes
both payloads at pH 6.0 and physiological temperature (37 ◦ C). At the
same total drug level, co-delivery of these drugs with the polymersomes
Polymersomes constitute a special class of polymeric nanostructured
had enhanced cytotoxicity and induced significantly higher cell
assemblies with vesicular morphology, capable to carry hydrophilic and
apoptosis in the cancer cell line, compared to the polymersomes loaded
hydrophobic therapeutic agents in their lumen and membrane respec­
with each of the two drugs separately [111].
tively, rendering them promising DDSs. Thermoresponsive polymer­
The combination of T-responsive polymers with polyelectrolytes
somes have been designed using segmented copolymers and/or
constitutes a powerful approach as it harnesses strong Coulombic in­
terpolymers incorporating thermoresponsive blocks [108]. Most of these
teractions and the thermally driven transition properties; for example,
systems are focused on heating induced polymersome formation and
McCormick et al. [112] coupled PNIPAM with poly (N-(3-aminopropyl)
some with the effect of temperature on the controlled delivery of ther­
methacrylamide P (AMPA) which was crosslinked with poly (sodium
apeutics, for example polymersomes formed by self-assembling of
2-acrylamido-2-meth-ylpropanesulfonate) (PAMPS) via interpolyelec­
methoxy-poly (ethylene oxide)-b-poly (e-caprolactone)-b-poly (N-iso­
trolyte interactions to “lock” polymersomes (Fig. 5.); the same group
propylacrylamide) (mPEO-b-PCL-b-PNIPAM) triblock terpolymers
further extended their approach to crosslink PNIPAM-rich block copol­
[109]. These polymersomes can effectively act as a drug release carrier
ymer micelles [113].
of indomethacin at physiological conditions, that is, store the drug at a
In a series of studies, Plamper et al. [114–116] explored the combi­
lower temperature and release it at a higher temperature.
nation of LCST triggering as a means to kinetically control the shape of
Polymersomes endowed with magneto-thermal ability (by inducing
polyelectrolyte nanoassemblies and their compartmentalization; this
hyperthermia), were prepared by the self-assembly of poly (isoprene)-b-
approach opens up the opportunity to develop multi-stimuli responsive
PNIPAM amphiphilic block copolymer, encapsulating monodisperse
nanocontainers capable to carry and deliver multiple molecular cargos
hydrophobic superparamagnetic iron oxide nanoparticles (SPION)
of different solubility. In a similar context Sumerlin et al. [117] proposed
within their hydrophobic compartment. A water-soluble dye, calcein,
the polymerization-induced thermal self-assembly (a variation of the so
entrapped in the lumen of the vesicles was used to evaluate this system

Fig. 5. Formation of T-responsive block copolymersomes and their polyelectrolyte crosslinking conferring additional colloidal stability (reproduced from ref. [112]
with permission).

7
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

called polymerization induced self-assembly PISA [118]) where the Particularly, star-graft quarterpolymers (SGQP) PSn [P2VP-b-(PAA-g-
formation of nano-objects (i.e. spherical or worm-like nanoassemblies) PNIPAM)]n (PS: polystyrene, P2VP: poly (2-vinylpyridine), PAA: poly
with tunable shape can be finely controlled via in situ control of the (acrylic acid)) containing two kinds of arms, one is a shorter hydro­
degree of polymerization and the application of thermal stimulation that phobic PS arm, the other is a pH responsive hydrophilic P2VP-b-PAA
modulates the packing parameter in aqueous media. block copolymer arm with grafted PNIPAM chains on the outer PAA
PNIPAM conjugated to proteins (BSA-NH2) to obtain thermally block were assembled with tannic acid through H-bonding on a sacri­
responsive microcompartments called “proteinsomes”, by analogy to ficial silica microparticle, forming multilayered nanostructured micro­
polymersomes, have also been reported, expanding the design capabil­ capsules. Due to the presence of weak polyelectrolytes and PNIPAM in
ities towards thermoresponsive biological hybrid vesicular systems. By the layers, this microcapsule exhibited pH and thermosensitivity. It was
grafting PNIPAM chains onto the protein surface, temperature- shown that both hydrophobic (small molecules) and hydrophilic (mac­
dependent chain conformational transition can be exploited to pro­ romolecules) cargos can be encapsulated in the shell or the lumen of the
duce stimulus-responsive proteinosomes that exhibit self-mediated microcapsules, concurrently. A programmable and sequential release of
gated permeability at water/water phase boundaries. In principle, this hydrophobic and hydrophilic molecules can be triggered independently
offers the ability not only to produce targeted microcontainers but also by temperature and pH variations. These stimuli affect the hydropho­
to use localized cooling of targeted cells and tissues for the temperature- bicity and chain conformation of the star block copolymers to initiate
mediated release of bioactive payloads [119]. out-of-shell release (elevated temperature), or change the overall star
conformation and interlayer interactions to trigger increased perme­
ability of the shell and out-of-core release (pH). Reversing of the stim­
4.3. Microcapsules ulus order, completely alters the release process allowing pre-
programmed, on-demand, release opposite sequences without mutual
Polymeric vesicular carriers, namely, microcapsules and/or nano­ interference [121].
capsules have emerged for potential applications in biomedicine. Among A different strategy to fabricate pH/thermo responsive organic/
them, systems incorporating LCST type thermoresponsive polymers inorganic hybrid microcapsules, enabling multiple loading and pro­
have been designed and prepared. An interesting example is the fabri­ grammable release was reported recently. PNIPAM surface modified
cation of ultrathin thermoresponsive microcapsules through direct co­ stellate mesoporous silica nanoparticles were used as nanocarriers for
valent layer-by-layer (LbL) assembly by “click” chemistry using azido- one guest compound (e.g. Nile red), serving simultaneously as Pickering
and acetylene-functionalized poly (N-isopropylacrylamide) (PNIPAM) emulsifiers to stabilize oil-in-water droplets. These droplets contain pH-
“clickable” random copolymers [120]. The synthesized poly [N-iso­ responsive monomers and another guest molecule, 5 (6)-carboxy­
propylacrylamide-co-(trimethylsilyl)propargylacrylamide] (PNIPA­ fluorescein diacetate (CFDA). Upon UV irradiation polymerization, the
M-Ace) and poly (N-isoropylacrylamide-co-3-azideopropylacrylamide) pH-responsive monomers were converted into polymer microcapsules.
(PNIPAM-Az) after removing the protective trimethylsilyl groups, were The release of both cargos could be selectively activated by changing the
assembled alternately onto azido-modified silica particles in aqueous temperature or the pH value [122].
media through “click” reactions. After removing the silica template, the Partially biodegradable thermoresponsive self-folding capsules,
microcapsules remained stable because of the presence of the covalently capable of controlled capture and release of cells (e.g. yeast cells) in
bonded triazole units (Fig. 6). response to a temperature have been demonstrated by using star-like
The microcapsules exhibited swelling/deswelling behavior by patterned polycaprolactone and PNIPAM crosslinked bilayers. Backing
responding reversibly to temperature changes of the medium. Evalua­ yeast cells were deposited on the patterned polymer bilayer from buffer
tion study on the permeability of microcapsules revealed that the mi­ dispersion at elevated temperature when PNIPAM is in the collapsed
crocapsules with tighter packing wall are selectively permeable to state. Upon decreasing temperature, due to PNIPAM swelling, folding of
molecules and show potential for applications for the encapsulation of a the polymer capsules occurs, encapsulating the cells which can be
variety of materials. delivered upon heating, inducing deswelling of PNIPAM and unfolding
Dual pH/thermo responsive microcapsules fabricated by the LbL of the capsule (Fig. 7) [123].
technique through non-covalent interactions have also been designed.

Fig. 6. Schematic Representation of the Preparation of Covalently Stabilized Thermoresponsive Microcapsules Through Layer-by-Layer Assembly Using Click
Chemistry, (from ref. [120] by permission of ACS).

8
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

Fig. 7. Schematic and bright field optical microscopy images of the folding star-shaped polymer bilayer. Swelling of the thermoresponsive hydrogel layer at lower
temperature increases stress in the film that results in bending of the star arms and folding (from ref. [123] by permission of RSC).

4.4. Microgels integrate living cell populations during fabrication which enables the in
situ generation of cell encapsulating microparticles in the form of par­
T-responsive gels harness the large volume changes driven by tem­ ticle dispersions, or as injectable soft biomaterials [130,131].
perature stimuli at the micro- or nano-scale finding interesting appli­ Generally, T-responsive microgels retain their temperature respon­
cations mostly as triggerable molecular cargo release systems [124]. Of sive properties and in analogy with bulk hydrogels exhibit large changes
particular interest for bioapplications are micro-/nano-sized PNIPAM of their volume albeit in a more confined space and hence the transition
gels owing to their relative ease of fabrication, their inherent capability temperature is often referred to as volume phase transition temperature
to be combined with different co-monomers and as recently reported (VPTT). In addition to cell delivery systems, they have found numerous
their excellent protein repellent properties [125] which render them applications as stimuli responsive drug delivery systems in precision
suitable drug carriers for in vivo delivery applications. medicine, as T-recoverable catalysts for homogeneous and heterogenous
Initial reports on PNIPAM microgels were based on surfactant-free catalysis applications, fluorescent reporters and as building blocks for
free radical polymerization in presence of crosslinkers with the use of photonic crystals with T-switchable optical properties [132,133].
persulfate initiators at 60–70 ◦ C. This protocol is sometimes referred to Microgels’ confined topology allows for facile integration with micro­
as “precipitation polymerization” as the high temperature induces fluidic devices [134] owing to their large surface to volume areas which,
instant collapsing of NIPAM oligomers during the propagation step and in addition to their T-induced volume changes, can be exploited for
produces well dispersed microgel particles [126]. Variations of this efficient surface modification of electrodes and further loading of en­
protocol include the incorporation of surfactants such as sodium dodecyl zymes [135,136] for biosensing applications that exploit and also their
sulfate, or the addition of macro-stabilizers in the form of macro­ T-induced volume changes of the microgels as readout mechanisms with
monomers or macro initiators in the case of CRP methodologies. Pre­ very low detection limitscan be exploited for efficient surface modifi­
cipitation polymerization is somewhat advantageous to classic emulsion cation of electrodes and further loading of enzymes [135,136].
polymerization as only a single aqueous reaction phase is needed which
simplifies reaction setups and purification steps. Lyon et al. [127] fully 4.5. Injectable hydrogels
exploited the robustness of precipitation polymerization to construct
core-shell microgel particles with fully detached compartments; first Hydrogels are three-dimensional (3D) soft materials consisting usu­
they grew seed PNIPAM particles with the use of N,N′ -methylenebis ally of a polymeric network capable to entrap high amounts of water,
(acrylamide) (BIS) (Fig. 8) followed by the growth of a sacrificial shell making them suitable for medicinal applications as DDS and/or scaffolds
by replacing BIS with N,N′ -(1,2-dihydroxyethylene)bisacrylamide. A for tissue engineering. Two types of hydrogels can be distinguished,
final precipitation polymerization step by using BIS again as crosslinker depending on the chemical or physical crosslinking of the polymeric
leads to the formation of a final outer shell. Treatment of the particles gelator, leading to permanent or reversible networks respectively.
with sodium periodate results in hydrolysis of the middle sacrificial shell Hydrogels can be classified according to the origin of the gelator in
affording detached core-shell T-responsive microgel particles. natural (e.g. polysaccharides), synthetic (copolymers) and semi­
For more complex molecular architectures, emulsion polymerization synthetic (natural/synthetic combinations); by incorporating LCST
remains the method of choice, especially, for protocols requiring mul­ macromolecular chains in the gelator, thermoresponsive hydrogels can
tiple monomers [128,129]. In addition, core-shell microgels can be be achieved.
obtained by the sequential addition of a second monomer either in situ A special category, namely injectable hydrogels, are those gels that
or by using an additional seed-polymerization step. In recent years, the are formed in situ after injection of a polymeric solution of freely moving
fabrication of microgel particles with the use of microfluidic devices entities (either free or associated chains) capable to form a network by
have also been employed as they allow for superior homogeneity and crosslinking. In the case of thermoresponsive hydrogels, two strategies
control on the microscale architecture and more importantly, they can have been developed so far. 1) chemical crosslinking comprising LCST

9
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

Fig. 8. Core-shell PNIPAM microgels with fully detached compartments synthesized by three consecutive precipitation polymerization steps (core (C), core-shell
(CS), and core-double-shell (CDS)) followed by periodate hydrolysis of the middle shell. Under the atomic force microscope, the outer shells look almost
completely flat due to the complete absence of structural rigidity (adapted with permission from ref. [127]).

polymers bearing functions reacting in situ to form a network and 2) 137–144]. In the following paragraphs, some new trends are briefly
physical crosslinking of block copolymer type macromolecules which demonstrated.
relies on the hydrophilic-to-hydrophobic transition above the LCST of Polysaccharides are very attractive biocompatible and biodegrad­
the thermosensitive blocks of the gelator, inducing self-association of able natural biopolymers, bearing functional groups susceptible to easy
the freely moving entities into a spanning network. In the second modification toward designing thermo-induced injectable hydrogels
strategy we can distinguish two main trends. 1) use of amphiphilic co­ [145,146]. Grafting methodologies allow attachment of LCST polymers
polymers that form micellar self-assemblies at room temperature before as pendant chains, endowing the resulting semisynthetic graft co­
injection and 2) double hydrophilic copolymers that self-assemble in the polymers with thermoresponsiveness. Efforts have been made to fine
physiological temperature after injection. In the former case immobili­ tune the rheological properties including injectability. For instance so­
zation of the micelles occurs either by jamming or by secondary thermo- dium alginate, grafted by a thermo-responsive random copolymer of
association (Fig. 9). PNIPAM, enriched with 10% mol of the hydrophobic N-tert-butylacry­
The research activities in the area of thermogels is very extensive due lamide, Alg-g-P(NIPAM90-co-NtBAM10) showed sol-gel transition (Tgel)
to potential bioapplications, namely: 1) minimally invasive drug de­ at about 32 ◦ C and excellent shear- and thermo-induced injectability
livery; 2) injectable tissue engineering (3D cell/stem cell culture); 3) [147]. The presence of the hydrophobic moieties in the thermores­
post-surgical treatments for adhesion prevention; 4) long-term magnetic ponsive PNIPAM grafting chains (stickers) regulate their LCST as well as
resonance imaging (MRI) contrasting; 5) transcatheterarterial emboli­ the strength of hydrophobic association, in terms of exchange dynamics
zation; 6) 3D live cell imaging for cellular processes; and 7) wound of the stickers of the network that control the rheological properties.
dressings (Fig. 10) [137]. However, the Tgel and the gel properties are influenced by a number of
Several examples of thermoresponsive injectable hydrogels can be factors, namely, gelator concentration, grafting density, nature and
viewed in previous reviews and under different contexts [10,27, length of grafts as well as by the environmental conditions like

Fig. 9. Schematic representation of gelation above the LCST (red color illustrates the LCST blocks): A) jamming of micelles formed upon heating of double hy­
drophilic diblock copolymers; B) double hydrophilic triblocks self-assemble in a micellar network by bridging; C) amphiphilic triblock micelles, with LCST chains in
the corona, associate.

10
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

Fig. 10. Potential applications of a thermogelling material, (from ref. [137] by permission from RSC).

temperature, pH, salinity and other solutes, which all have to be versatility and the possibility to be combined with other monomers to
co-evaluated to target tailor-made hydrogel properties for specific confer additional isothermal responsiveness towards other stimuli such
applications. as pH gradients, irradiation, or the presence analytes. PNVCL-based and
The use of PNIPAM-based random copolymers instead of the ho­ pOx-based polymers are also emerging especially for applications where
mopolymer PNIPAM can be utilized to introduce biodegradability. An non-degradability is a prerequisite while PEGMAs constitute an attrac­
interesting example of this concept is the bioresorbable polymer tive alternative owing to their resemblance of PEG and their capacity to
injectable hydrogels formed by poly (N-isopropylacrylamide-co- be easily polymerized via controlled/living polymerization reactions.
dimethyl-c-butyrolactone acrylate-co-acrylic acid) containing 7.00 mol Arguably ELPs will continue to be developed and show the highest po­
% of the hydrolyzable dimethyl-c-butyrolactone acrylate monomer tential to for direct in vivo applications owing to their degradability and
[148]. their natural polypeptide structure.
Stem cell transplantation using injectable hydrogel is an application T-responsive materials have now been diffused into niche areas
requiring a two-step gelation before and after injection in order to beyond classic polymer chemistry and materials science and they are
protect the cells during injection and to provide adequate cell immobi­ invaluable laboratory tools albeit mostly for proof-of-concept applica­
lization in the target tissue. Shear/thermo induced injectability fulfill tions. The field has matured enough to expect practical applications to
these requirements. For this purpose, a two-component system has been continue to emerge not only in the biomedical field but also in the fields
developed comprising aldehyde-modified hyaluronic acid (HA-ALD) of catalysis, separation and purification technologies and biosensing
and the thermoresponsive hydrazine-modified elastin-like protein (ELP- applications; indeed patent filings including the word “poly (N-iso­
HY). By mixing the polymeric components at room temperature a shear- propylacrylamide)” have linearly increased in the last 10–20 years and
induced injectable hydrogel is formed through reversible dynamic co­ therefore it remains to be seen whether these fascinating polymers will
valent hydrazone bonds. Upon heating at 37 ◦ C a secondary physical be integrated in out-of-the-lab applications. Of note, there are still
crosslinking occurs, due to ELP LCST type thermal phase transition (see concerns about the toxicity of PNIPAM [151–153] and hence certain in
for example [149]), mechanically reinforcing the hydrogel (Fig. 11). vivo applications may be hampered which potentially leaves room for
This system enables. other alternative LCST polymers to be used for clinical practice. More
Cell culture for three weeks post injection. Yet, the encapsulated cells realistic expectations should lie on in vitro and/or ex vivo applications
maintain their ability to differentiate into multiple lineages, including such as molecular diagnostics, biosensing and integration with micro­
chondrogenic, adipogenic, and osteogenic cell types [150]. fluidic devices.
Finally, there is a clear trend towards more refined polymer topol­
5. Conclusions and outlook ogies and architectures owing to the rapid development of controlled
polymerization methods and conjugation protocols (i.e. “click” chem­
PNIPAM remains the polymer of choice for applications requiring istries, bio-orthogonal coupling reactions, etc.); these advances have
materials with stimuli responsive polymers due to its synthetic allowed for unprecedented molecular fidelity which was not possible to

11
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

Fig. 11. Injectable ELP–HA hydrogels. a) ELP–HA is composed of hydrazine-modified elastin-like protein (ELP-HYD) and aldehyde-modified hyaluronic acid (HA-
ALD). b) Schematic of ELP–HA hydrogel formation. c) Photographs demonstrating the injectability and rapid self-healing of ELP–HA hydrogels, (from ref. [150] by
permission of Wiley VCH).

achieve 20 years ago (a classic example-testament of these advances is properties, which we expect to continue to be developed in the near
the so-called “grafting from” approach directly from protein molecules future.
[154]); such synthetic precision constitutes a major synthetic milestone
enabling the potential use of LCST polymers in clinical areas requiring Declaration of competing interest
pharmaceutical grade materials and good manufacturing protocols.
More robust chemistries have also lead to the easier coupling of The authors declare that they have no known competing financial
T-responsive polymers with additional stimuli-responsive components interests or personal relationships that could have appeared to influence
(i.e. pH-responsive polymers, magnetic/plasmonic nanoparticles, etc.) the work reported in this paper.
[21,155] which have already led to the synthesis of multi-stimuli
responsive materials with complex behavior and often exotic

12
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

References [28] L.D. Taylor, L.D. Cerankowski, Preparation of films exhibiting a balanced
temperature dependence to permeation by aqueous solutions—a study of lower
consolute behavior, J. Polym. Sci. Polym. Chem. Ed. 13 (1975) 2551–2570,
[1] M.A.C. Stuart, W.T.S. Huck, J. Genzer, M. Müller, C. Ober, M. Stamm, G.
https://doi.org/10.1002/pol.1975.170131113.
B. Sukhorukov, I. Szleifer, V. V Tsukruk, M. Urban, F. Winnik, S. Zauscher,
[29] N.T. Southall, K.A. Dill, A.D.J. Haymet, A view of the hydrophobic effect, J. Phys.
I. Luzinov, S. Minko, Emerging applications of stimuli-responsive polymer
Chem. B 106 (2002) 521–533, https://doi.org/10.1021/jp015514e.
materials, Nat. Mater. 9 (2010) 101–113, https://doi.org/10.1038/nmat2614.
[30] A. Halperin, M. Kröger, F.M. Winnik, Poly(N-isopropylacrylamide) phase
[2] C. de las H. Alarcón, S. Pennadam, C. Alexander, Stimuli responsive polymers for
diagrams: fifty years of research, Angew. Chem. Int. Ed. 54 (2015) 15342–15367,
biomedical applications, Chem. Soc. Rev. 34 (2005) 276–285, https://doi.org/
https://doi.org/10.1002/anie.201506663.
10.1039/B406727D.
[31] J. Seuring, S. Agarwal, Polymers with upper critical solution temperature in
[3] A.P. Blum, J.K. Kammeyer, A.M. Rush, C.E. Callmann, M.E. Hahn, N.
aqueous solution, Macromol. Rapid Commun. 33 (2012) 1898–1920, https://doi.
C. Gianneschi, Stimuli-responsive nanomaterials for biomedical applications,
org/10.1002/marc.201200433.
J. Am. Chem. Soc. 137 (2015) 2140–2154, https://doi.org/10.1021/ja510147n.
[32] J. Seuring, S. Agarwal, Polymers with upper critical solution temperature in
[4] A.S. Hoffman, Stimuli-responsive polymers: biomedical applications and
aqueous solution: unexpected properties from known building blocks, ACS Macro
challenges for clinical translation, Adv. Drug Deliv. Rev. 65 (2013) 10–16,
Lett. 2 (2013) 597–600, https://doi.org/10.1021/mz400227y.
https://doi.org/10.1016/j.addr.2012.11.004.
[33] H.G. Schild, Poly(N-isopropylacrylamide): experiment, theory and application,
[5] L.A. Lyon, Z. Meng, N. Singh, C.D. Sorrell, A. St John, Thermoresponsive
Prog. Polym. Sci. 17 (1992) 163–249, https://doi.org/10.1016/0079-6700(92)
microgel-based materials, Chem. Soc. Rev. 38 (2009) 865–874, https://doi.org/
90023-R.
10.1039/B715522K.
[34] M. Heskins, J.E. Guillet, Solution properties of poly(N-isopropylacrylamide),
[6] Y. Gao, X. Li, M.J. Serpe, Stimuli-responsive microgel-based etalons for optical
J. Macromol. Sci. Part A - Chem. 2 (1968) 1441–1455, https://doi.org/10.1080/
sensing, RSC Adv. 5 (2015) 44074–44087, https://doi.org/10.1039/
10601326808051910.
C5RA02306H.
[35] C. Zhao, Z. Ma, X.X. Zhu, Rational design of thermoresponsive polymers in
[7] L. Hu, Y. Wan, Q. Zhang, M.J. Serpe, Harnessing the power of stimuli-responsive
aqueous solutions: a thermodynamics map, Prog. Polym. Sci. 90 (2019) 269–291,
polymers for actuation, Adv. Funct. Mater. 30 (2020) 1903471, https://doi.org/
https://doi.org/10.1016/j.progpolymsci.2019.01.001.
10.1002/adfm.201903471.
[36] R. Pelton, Poly(N-isopropylacrylamide) (PNIPAM) is never hydrophobic,
[8] H. Sun, Z. Zhong, 100th anniversary of macromolecular science viewpoint:
J. Colloid Interface Sci. 348 (2010) 673–674, https://doi.org/10.1016/j.
biological stimuli-sensitive polymer prodrugs and nanoparticles for tumor-
jcis.2010.05.034.
specific drug delivery, ACS Macro Lett. (2020) 1292–1302, https://doi.org/
[37] B. Jean, L.-T. Lee, B. Cabane, Interactions of sodium dodecyl sulfate with
10.1021/acsmacrolett.0c00488.
acrylamide –N-isopropylacrylamide) statistical copolymer, Colloid Polym. Sci.
[9] M. Heskins, J.E. Guillet, Solution properties of poly(N-isopropylacrylamide),
278 (2000) 764–770, https://doi.org/10.1007/s003960000310.
J. Macromol. Sci. Part A - Chem. 2 (1968) 1441–1455, https://doi.org/10.1080/
[38] M. Stieger, W. Richtering, J.S. Pedersen, P. Lindner, Small-angle neutron
10601326808051910.
scattering study of structural changes in temperature sensitive microgel colloids,
[10] L. Klouda, A.G. Mikos, Thermoresponsive hydrogels in biomedical applications,
J. Chem. Phys. 120 (2004) 6197–6206, https://doi.org/10.1063/1.1665752.
Eur. J. Pharm. Biopharm. 68 (2008) 34–45, https://doi.org/10.1016/j.
[39] J. Adelsberger, I. Grillo, A. Kulkarni, M. Sharp, A.M. Bivigou-Koumba,
ejpb.2007.02.025.
A. Laschewsky, P. Müller-Buschbaum, C.M. Papadakis, Kinetics of aggregation in
[11] K. Nagase, J. Kobayashi, T. Okano, Temperature-responsive intelligent interfaces
micellar solutions of thermoresponsive triblock copolymers – influence of
for biomolecular separation and cell sheet engineering, J. R. Soc. Interface 6
concentration, start and target temperatures, Soft Matter 9 (2013) 1685–1699,
(2009) S293–S309, https://doi.org/10.1098/rsif.2008.0499.focus.
https://doi.org/10.1039/C2SM27152D.
[12] K. Nagase, M. Yamato, H. Kanazawa, T. Okano, Poly(N-isopropylacrylamide)-
[40] R. Keidel, A. Ghavami, D.M. Lugo, G. Lotze, O. Virtanen, P. Beumers, J.
based thermoresponsive surfaces provide new types of biomedical applications,
S. Pedersen, A. Bardow, R.G. Winkler, W. Richtering, Time-resolved structural
Biomaterials 153 (2018) 27–48, https://doi.org/10.1016/j.
evolution during the collapse of responsive hydrogels: the microgel-to-particle
biomaterials.2017.10.026.
transition, Sci. Adv. 4 (2018), eaao7086, https://doi.org/10.1126/sciadv.
[13] P.S. Stayton, T. Shimoboji, C. Long, A. Chilkoti, G. Ghen, J.M. Harris, A.
aao7086.
S. Hoffman, Control of protein–ligand recognition using a stimuli-responsive
[41] J. Wang, D. Gan, L.A. Lyon, M.A. El-Sayed, Temperature-jump investigations of
polymer, Nature 378 (1995) 472–474, https://doi.org/10.1038/378472a0.
the kinetics of hydrogel nanoparticle volume phase transitions, J. Am. Chem. Soc.
[14] M.S. Messina, K.M.M. Messina, A. Bhattacharya, H.R. Montgomery, H.
123 (2001) 11284–11289, https://doi.org/10.1021/ja016610w.
D. Maynard, Preparation of biomolecule-polymer conjugates by grafting-from
[42] Q. Zhang, C. Weber, U.S. Schubert, R. Hoogenboom, Thermoresponsive polymers
using ATRP, RAFT, or ROMP, Prog. Polym. Sci. 100 (2020) 101186, https://doi.
with lower critical solution temperature: from fundamental aspects and
org/10.1016/j.progpolymsci.2019.101186.
measuring techniques to recommended turbidimetry conditions, Mater. Horizons.
[15] E.S. Gil, S.M. Hudson, Stimuli-reponsive polymers and their bioconjugates, Prog.
4 (2017) 109–116, https://doi.org/10.1039/C7MH00016B.
Polym. Sci. 29 (2004) 1173–1222, https://doi.org/10.1016/j.
[43] C. Boutris, E.G. Chatzi, C. Kiparissides, Characterization of the LCST behaviour of
progpolymsci.2004.08.003.
aqueous poly(N-isopropylacrylamide) solutions by thermal and cloud point
[16] D.J. Siegwart, J.K. Oh, K. Matyjaszewski, ATRP in the design of functional
techniques, Polymer (Guildf) 38 (1997) 2567–2570, https://doi.org/10.1016/
materials for biomedical applications, Prog. Polym. Sci. 37 (2012) 18–37, https://
S0032-3861(97)01024-0.
doi.org/10.1016/j.progpolymsci.2011.08.001.
[44] S. Furyk, Y. Zhang, D. Ortiz-Acosta, P.S. Cremer, D.E. Bergbreiter, Effects of end
[17] C. Boyer, V. Bulmus, T.P. Davis, V. Ladmiral, J. Liu, S. Perrier, Bioapplications of
group polarity and molecular weight on the lower critical solution temperature of
RAFT polymerization, Chem. Rev. 109 (2009) 5402–5436, https://doi.org/
poly(N-isopropylacrylamide), J. Polym. Sci. Part A Polym. Chem. 44 (2006)
10.1021/cr9001403.
1492–1501, https://doi.org/10.1002/pola.21256.
[18] I. Cobo, M. Li, B.S. Sumerlin, S. Perrier, Smart hybrid materials by conjugation of
[45] K. Otake, H. Inomata, M. Konno, S. Saito, Thermal analysis of the volume phase
responsive polymers to biomacromolecules, Nat. Mater. 14 (2015) 143–159,
transition with N-isopropylacrylamide gels, Macromolecules 23 (1990) 283–289,
https://doi.org/10.1038/nmat4106.
https://doi.org/10.1021/ma00203a049.
[19] H.C. Kolb, M.G. Finn, K.B. Sharpless, Click chemistry: diverse chemical function
[46] J.E. Chung, M. Yokoyama, T. Aoyagi, Y. Sakurai, T. Okano, Effect of molecular
from a few good reactions, Angew. Chem. Int. Ed. 40 (2001) 2, https://doi.org/
architecture of hydrophobically modified poly(N-isopropylacrylamide) on the
10.1002/1521-3773(20010601)40:11<2004::AID-ANIE2004>3.0.CO,
formation of thermoresponsive core-shell micellar drug carriers, J. Contr. Release
2004–2021.
53 (1998) 119–130, https://doi.org/10.1016/S0168-3659(97)00244-7.
[20] S. Qiao, H. Wang, Temperature-responsive polymers: synthesis, properties, and
[47] X. Qiu, T. Koga, F. Tanaka, F.M. Winnik, New insights into the effects of
biomedical applications, Nano Res 11 (2018) 5400–5423, https://doi.org/
molecular weight and end group on the temperature-induced phase transition of
10.1007/s12274-018-2121-x.
poly(N-isopropylacrylamide) in water, Sci. China Chem. 56 (2013) 56–64,
[21] G. Pasparakis, M. Vamvakaki, Multiresponsive polymers: nano-sized assemblies,
https://doi.org/10.1007/s11426-012-4781-9.
stimuli-sensitive gels and smart surfaces, Polym. Chem. 2 (2011) 1234–1248,
[48] B. Ray, Y. Okamoto, M. Kamigaito, M. Sawamoto, K. Seno, S. Kanaoka,
https://doi.org/10.1039/C0PY00424C.
S. Aoshima, Effect of tacticity of poly(N-isopropylacrylamide) on the phase
[22] X. Fu, L. Hosta-Rigau, R. Chandrawati, J. Cui, Multi-stimuli-responsive polymer
separation temperature of its aqueous solutions, Polym. J. 37 (2005) 234–237,
particles, films, and hydrogels for drug delivery, Chem 4 (2018) 2084–2107,
https://doi.org/10.1295/polymj.37.234.
https://doi.org/10.1016/j.chempr.2018.07.002.
[49] S. Habaue, Y. Isobe, Y. Okamoto, Stereocontrolled radical polymerization of
[23] P.J. Flory, Thermodynamics of high polymer solutions, J. Chem. Phys. 10 (1942)
acrylamides and methacrylamides using Lewis acids, Tetrahedron 58 (2002)
51–61, https://doi.org/10.1063/1.1723621.
8205–8209, https://doi.org/10.1016/S0040-4020(02)00969-9.
[24] M.L. Huggins, Thermodynamic properties OF solutions OF long-chain
[50] Y. Kametani, F. Tournilhac, M. Sawamoto, M. Ouchi, Unprecedented sequence
compounds, Ann. N. Y. Acad. Sci. 43 (1942) 1–32, https://doi.org/10.1111/
control and sequence-driven properties in a series of AB-alternating copolymers
j.1749-6632.1942.tb47940.x.
consisting solely of acrylamide units, Angew. Chem. Int. Ed. 59 (2020)
[25] D. Patterson, Free volume and polymer solubility. A qualitative view,
5193–5201, https://doi.org/10.1002/anie.201915075.
Macromolecules 2 (1969) 672–677, https://doi.org/10.1021/ma60012a021.
[51] M. Shibayama, F. Ikkai, S. Inamoto, S. Nomura, C.C. Han, pH and salt
[26] A. Bordat, T. Boissenot, J. Nicolas, N. Tsapis, Thermoresponsive polymer
concentration dependence of the microstructure of poly(N-isopropylacrylamide-
nanocarriers for biomedical applications, Adv. Drug Deliv. Rev. (2018), https://
co-acrylic acid) gels, J. Chem. Phys. 105 (1996) 4358–4366, https://doi.org/
doi.org/10.1016/j.addr.2018.10.005.
10.1063/1.472252.
[27] M.A. Ward, T.K. Georgiou, Thermoresponsive Polymers for Biomedical
Applications, Polymer 3 (2011), https://doi.org/10.3390/polym3031215.

13
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

[52] D. Schmaljohann, Thermo- and pH-responsive polymers in drug delivery, Adv. [77] D. Wan, Q. Zhou, H. Pu, G. Yang, Controlled radical polymerization of N-
Drug Deliv. Rev. 58 (2006) 1655–1670, https://doi.org/10.1016/j. vinylcaprolactam mediated by xanthate or dithiocarbamate, J. Polym. Sci. Part A
addr.2006.09.020. Polym. Chem. 46 (2008) 3756–3765, https://doi.org/10.1002/pola.22722.
[53] W.L.J. Hinrichs, N.M.E. Schuurmans-Nieuwenbroek, P. van de Wetering, W. [78] J.-F. Lutz, Ö. Akdemir, A. Hoth, Point by point comparison of two thermosensitive
E. Hennink, Thermosensitive polymers as carriers for DNA delivery, J. Contr. polymers exhibiting a similar LCST: is the age of poly(NIPAM) over? J. Am.
Release 60 (1999) 249–259, https://doi.org/10.1016/S0168-3659(99)00075-9. Chem. Soc. 128 (2006) 13046–13047, https://doi.org/10.1021/ja065324n.
[54] G. Pasparakis, M. Vamvakaki, N. Krasnogor, C. Alexander, Diol–boronic acid [79] T.X. Viegas, M.D. Bentley, J.M. Harris, Z. Fang, K. Yoon, B. Dizman, R. Weimer,
complexes integrated by responsive polymers—a route to chemical sensing and A. Mero, G. Pasut, F.M. Veronese, Polyoxazoline: chemistry, properties, and
logic operations, Soft Matter 5 (2009) 3839–3841. applications in drug delivery, Bioconjugate Chem. 22 (2011) 976–986, https://
[55] G. Pasparakis, A. Cockayne, C. Alexander, Control of bacterial aggregation by doi.org/10.1021/bc200049d.
thermoresponsive glycopolymers, J. Am. Chem. Soc. 129 (2007) 11014–11015. [80] S. Aluri, S.M. Janib, J.A. Mackay, Environmentally responsive peptides as
[56] F. Mastrotto, P. Caliceti, V. Amendola, S. Bersani, J.P. Magnusson, M. Meneghetti, anticancer drug carriers, Adv. Drug Deliv. Rev. 61 (2009) 940–952, https://doi.
G. Mantovani, C. Alexander, S. Salmaso, Polymer control of ligand display on org/10.1016/j.addr.2009.07.002.
gold nanoparticles for multimodal switchable cell targeting, Chem. Commun. 47 [81] J. Andrew Mackay, A. Chilkoti, Temperature sensitive peptides: engineering
(2011) 9846–9848, https://doi.org/10.1039/C1CC12654G. hyperthermia-directed therapeutics, Int. J. Hyperther. 24 (2008) 483–495,
[57] F.A. Plamper, A.A. Steinschulte, C.H. Hofmann, N. Drude, O. Mergel, C. Herbert, https://doi.org/10.1080/02656730802149570.
M. Erberich, B. Schulte, R. Winter, W. Richtering, Toward copolymers with ideal [82] D.E. Meyer, B.C. Shin, G.A. Kong, M.W. Dewhirst, A. Chilkoti, Drug targeting
thermosensitivity: solution properties of linear, well-defined polymers of N- using thermally responsive polymers and local hyperthermia, J. Contr. Release 74
isopropyl acrylamide and N,N-diethyl acrylamide, Macromolecules 45 (2012) (2001) 213–224, https://doi.org/10.1016/S0168-3659(01)00319-4.
8021–8026, https://doi.org/10.1021/ma301606c. [83] I.L. Karle, D.W. Urry, Crystal structure of cyclic (APGVGV)2, an analog of elastin,
[58] M. Keerl, V. Smirnovas, R. Winter, W. Richtering, Interplay between hydrogen and a suggested mechanism for elongation/contraction of the molecule,
bonding and macromolecular architecture leading to unusual phase behavior in Biopolymers 77 (2005) 198–204, https://doi.org/10.1002/bip.20214.
thermosensitive microgels, Angew. Chem. Int. Ed. 47 (2008) 338–341, https:// [84] D.W. Urry, Free energy transduction in polypeptides and proteins based on
doi.org/10.1002/anie.200703728. inverse temperature transitions, Prog. Biophys. Mol. Biol. 57 (1992) 23–57,
[59] M. Keerl, J.S. Pedersen, W. Richtering, Temperature sensitive copolymer https://doi.org/10.1016/0079-6107(92)90003-O.
microgels with nanophase separated structure, J. Am. Chem. Soc. 131 (2009) [85] H. Reiersen, A.R. Clarke, A.R. Rees, Short elastin-like peptides exhibit the same
3093–3097, https://doi.org/10.1021/ja807367p. temperature-induced structural transitions as elastin polymers: implications for
[60] K.D. Collins, M.W. Washabaugh, The Hofmeister effect and the behaviour of protein engineering11Edited by A. R. Fersht, J. Mol. Biol. 283 (1998) 255–264,
water at interfaces, Q. Rev. Biophys. 18 (1985) 323–422, https://doi.org/ https://doi.org/10.1006/jmbi.1998.2067.
10.1017/S0033583500005369. [86] A.P. Hathorne, H. Bermudez, Effects of short elastin-like peptides on filamentous
[61] Y. Zhang, S. Furyk, D.E. Bergbreiter, P.S. Cremer, Specific ion effects on the water particles and their transition behavior, Biotechnol. Bioeng. 110 (2013)
solubility of Macromolecules: PNIPAM and the hofmeister series, J. Am. Chem. 1822–1830, https://doi.org/10.1002/bit.24854.
Soc. 127 (2005) 14505–14510, https://doi.org/10.1021/ja0546424. [87] H. Nuhn, H.-A. Klok, Secondary structure formation and LCST behavior of short
[62] D. Mukherji, C.M. Marques, K. Kremer, Polymer collapse in miscible good elastin-like peptides, Biomacromolecules 9 (2008) 2755–2763, https://doi.org/
solvents is a generic phenomenon driven by preferential adsorption, Nat. 10.1021/bm800784y.
Commun. 5 (2014) 4882, https://doi.org/10.1038/ncomms5882. [88] R.E. Sallach, M. Wei, N. Biswas, V.P. Conticello, S. Lecommandoux, R.A. Dluhy, E.
[63] F.M. Winnik, H. Ringsdorf, J. Venzmer, Methanol-water as a co-nonsolvent L. Chaikof, Micelle density regulated by a reversible switch of protein secondary
system for poly(N-isopropylacrylamide), Macromolecules 23 (1990) 2415–2416, structure, J. Am. Chem. Soc. 128 (2006) 12014–12019, https://doi.org/10.1021/
https://doi.org/10.1021/ma00210a048. ja0638509.
[64] C. Scherzinger, A. Schwarz, A. Bardow, K. Leonhard, W. Richtering, [89] W. Kim, J. Thévenot, E. Ibarboure, S. Lecommandoux, E.L. Chaikof, Self-assembly
Cononsolvency of poly-N-isopropyl acrylamide (PNIPAM): microgels versus of thermally responsive amphiphilic diblock copolypeptides into spherical
linear chains and macrogels, Curr. Opin. Colloid Interface Sci. 19 (2014) 84–94, micellar nanoparticles, Angew. Chem. Int. Ed. 49 (2010) 4257–4260, https://doi.
https://doi.org/10.1016/j.cocis.2014.03.011. org/10.1002/anie.201001356.
[65] T. Manouras, E. Koufakis, S.H. Anastasiadis, M. Vamvakaki, A facile route [90] J. Andrew MacKay, M. Chen, J.R. McDaniel, W. Liu, A.J. Simnick, A. Chilkoti,
towards PDMAEMA homopolymer amphiphiles, Soft Matter 13 (2017) Self-assembling chimeric polypeptide–doxorubicin conjugate nanoparticles that
3777–3782, https://doi.org/10.1039/C7SM00365J. abolish tumours after a single injection, Nat. Mater. 8 (2009) 993–999, https://
[66] W. Agut, A. Brûlet, C. Schatz, D. Taton, S. Lecommandoux, pH and temperature doi.org/10.1038/nmat2569.
responsive polymeric micelles and polymersomes by self-assembly of poly[2- [91] M.U. Kahveci, Y. Yagci, A. Avgeropoulos, M.E. Tsitsilianis, Polymeric materials –
(dimethylamino)ethyl methacrylate]-b-Poly(glutamic acid) double hydrophilic well defined block copolymers, in: Reference Module in Materials Science and
block copolymers, Langmuir 26 (2010) 10546–10554, https://doi.org/10.1021/ Materials Engineering, Elsevier, 2016. https://doi.org/10.1016/B978-0-12-80
la1005693. 3581-8.01447-8.
[67] A. Saraiva, O. Persson, A. Fredenslund, An experimental investigation of cloud- [92] J.E. Chung, M. Yokoyama, M. Yamato, T. Aoyagi, Y. Sakurai, T. Okano, Thermo-
point curves for the poly(ethylene glycol)/water system at varying molecular responsive drug delivery from polymeric micelles constructed using block
weight distributions, Fluid Phase Equil. 91 (1993) 291–311, https://doi.org/ copolymers of poly(N-isopropylacrylamide) and poly(butylmethacrylate),
10.1016/0378-3812(93)85105-U. J. Contr. Release 62 (1999) 115–127, https://doi.org/10.1016/S0168-3659(99)
[68] M. Rackaitis, K. Strawhecker, E. Manias, Water-soluble polymers with tunable 00029-2.
temperature sensitivity: solution behavior, J. Polym. Sci. Part B Polym. Phys. 40 [93] J. Akimoto, M. Nakayama, K. Sakai, T. Okano, Temperature-induced intracellular
(2002) 2339–2342, https://doi.org/10.1002/polb.10284. uptake of thermoresponsive polymeric micelles, Biomacromolecules 10 (2009)
[69] H.G. Schild, D.A. Tirrell, Microcalorimetric detection of lower critical solution 1331–1336, https://doi.org/10.1021/bm900032r.
temperatures in aqueous polymer solutions, J. Phys. Chem. 94 (1990) [94] J. Akimoto, M. Nakayama, K. Sakai, T. Okano, Thermally controlled intracellular
4352–4356, https://doi.org/10.1021/j100373a088. uptake system of polymeric micelles possessing poly(N-isopropylacrylamide)-
[70] X. Xia, S. Tang, X. Lu, Z. Hu, formation and volume phase transition of Based outer coronas, Mol. Pharm. 7 (2010) 926–935, https://doi.org/10.1021/
hydroxypropyl cellulose microgels in salt solution, Macromolecules 36 (2003) mp100021c.
3695–3698, https://doi.org/10.1021/ma0216728. [95] M. Emamzadeh, D. Desmaële, P. Couvreur, G. Pasparakis, Dual controlled
[71] N.A. Cortez-Lemus, A. Licea-Claverie, Poly(N-vinylcaprolactam), a delivery of squalenoyl-gemcitabine and paclitaxel using thermo-responsive
comprehensive review on a thermoresponsive polymer becoming popular, Prog. polymeric micelles for pancreatic cancer, J. Mater. Chem. B. 6 (2018) 2230–2239,
Polym. Sci. 53 (2016) 1–51, https://doi.org/10.1016/j. https://doi.org/10.1039/C7TB02899G.
progpolymsci.2015.08.001. [96] M. Emamzadeh, M. Emamzadeh, G. Pasparakis, Dual controlled delivery of
[72] J.-F. Lutz, Polymerization of oligo(ethylene glycol) (meth)acrylates: toward new gemcitabine and cisplatin using polymer-modified thermosensitive liposomes for
generations of smart biocompatible materials, J. Polym. Sci. Part A Polym. Chem. pancreatic cancer, ACS Appl. Bio Mater. 2 (n.d.) 1298–1309. https://doi.org/10.
46 (2008) 3459–3470, https://doi.org/10.1002/pola.22706. 1021/acsabm.9b00007.
[73] R. Hoogenboom, H.M.L. Thijs, M.J.H.C. Jochems, B.M. van Lankvelt, M.W. [97] X. Wan, T. Liu, S. Liu, Synthesis of amphiphilic tadpole-shaped linear-cyclic
M. Fijten, U.S. Schubert, Tuning the LCST of poly(2-oxazoline)s by varying diblock copolymers via ring-opening polymerization directly initiating from
composition and molecular weight: alternatives to poly(N-isopropylacrylamide)? cyclic precursors and their application as drug nanocarriers, Biomacromolecules
Chem. Commun. (2008) 5758–5760, https://doi.org/10.1039/B813140F. 12 (2011) 1146–1154, https://doi.org/10.1021/bm101463d.
[74] Y.-Z. You, D.S. Manickam, Q.-H. Zhou, D. Oupický, Reducible poly(2- [98] C.J.F. Rijcken, T.F.J. Veldhuis, A. Ramzi, J.D. Meeldijk, C.F. van Nostrum, W.
dimethylaminoethyl methacrylate): synthesis, cytotoxicity, and gene delivery E. Hennink, Novel fast degradable thermosensitive polymeric micelles based on
activity, J. Contr. Release 122 (2007) 217–225, https://doi.org/10.1016/j. PEG-block-poly(N-(2-hydroxyethyl)methacrylamide-oligolactates),
jconrel.2007.04.020. Biomacromolecules 6 (2005) 2343–2351, https://doi.org/10.1021/bm0502720.
[75] S. Agarwal, Y. Zhang, S. Maji, A. Greiner, PDMAEMA based gene delivery [99] S. Wilhelm, A.J. Tavares, Q. Dai, S. Ohta, J. Audet, H.F. Dvorak, W.C.W. Chan,
materials, Mater. Today 15 (2012) 388–393, https://doi.org/10.1016/S1369- Analysis of nanoparticle delivery to tumours, Nat. Rev. Mater. 1 (2016) 16014,
7021(12)70165-7. https://doi.org/10.1038/natrevmats.2016.14. http://www.nature.com/articl
[76] G. Dumortier, J.L. Grossiord, F. Agnely, J.C. Chaumeil, A review of poloxamer es/natrevmats201614. #supplementary-information.
407 pharmaceutical and pharmacological characteristics, Pharm. Res. 23 (2006) [100] J. Park, Y. Choi, H. Chang, W. Um, J.H. Ryu, I.C. Kwon, Alliance with EPR effect:
2709–2728, https://doi.org/10.1007/s11095-006-9104-4. combined strategies to improve the EPR effect in the tumor microenvironment,
Theranostics 9 (2019) 8073–8090, https://doi.org/10.7150/thno.37198.

14
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

[101] X. Liang, F. Liu, V. Kozlovskaya, Z. Palchak, E. Kharlampieva, Thermoresponsive [125] E. Miceli, B. Kuropka, C. Rosenauer, E.R. Osorio Blanco, L.E. Theune, M. Kar,
micelles from double LCST-poly(3-methyl-N-vinylcaprolactam) block copolymers C. Weise, S. Morsbach, C. Freund, M. Calderón, Understanding the elusive protein
for cancer therapy, ACS Macro Lett. 4 (2015) 308–311, https://doi.org/10.1021/ corona of thermoresponsive nanogels, Nanomedicine 13 (2018) 2657–2668,
mz500832a. https://doi.org/10.2217/nnm-2018-0217.
[102] X. Liang, F. Liu, V. Kozlovskaya, Z. Palchak, E. Kharlampieva, Thermoresponsive [126] R. Pelton, Temperature-sensitive aqueous microgels, Adv. Colloid Interface Sci.
micelles from double LCST-poly(3-methyl-N-vinylcaprolactam) block copolymers 85 (2000) 1–33, https://doi.org/10.1016/S0001-8686(99)00023-8.
for cancer therapy, ACS Macro Lett. 4 (2015) 308–311, https://doi.org/10.1021/ [127] X. Hu, Z. Tong, L.A. Lyon, Multicompartment core/shell microgels, J. Am. Chem.
mz500832a. Soc. 132 (2010) 11470–11472, https://doi.org/10.1021/ja105616v.
[103] J. Weiss, A. Laschewsky, Temperature-induced self-assembly of triple-responsive [128] J.M. Asua, Emulsion polymerization: from fundamental mechanisms to process
triblock copolymers in aqueous solutions, Langmuir 27 (2011) 4465–4473, developments, J. Polym. Sci. Part A Polym. Chem. 42 (2004) 1025–1041, https://
https://doi.org/10.1021/la200115p. doi.org/10.1002/pola.11096.
[104] X. Guo, D. Li, G. Yang, C. Shi, Z. Tang, J. Wang, S. Zhou, Thermo-triggered drug [129] C.S. Chern, Emulsion polymerization mechanisms and kinetics, Prog. Polym. Sci.
release from actively targeting polymer micelles, ACS Appl. Mater. Interfaces 6 31 (2006) 443–486, https://doi.org/10.1016/j.progpolymsci.2006.02.001.
(2014) 8549–8559, https://doi.org/10.1021/am501422r. [130] S. Seiffert, D.A. Weitz, Microfluidic fabrication of smart microgels from
[105] S.J.T. Rezaei, M.R. Nabid, H. Niknejad, A.A. Entezami, Folate-decorated macromolecular precursors, Polymer (Guildf) 51 (2010) 5883–5889, https://doi.
thermoresponsive micelles based on star-shaped amphiphilic block copolymers org/10.1016/j.polymer.2010.10.034.
for efficient intracellular release of anticancer drugs, Int. J. Pharm. 437 (2012) [131] R. Riahi, A. Tamayol, S.A.M. Shaegh, A.M. Ghaemmaghami, M.R. Dokmeci,
70–79, https://doi.org/10.1016/j.ijpharm.2012.07.069. A. Khademhosseini, Microfluidics for advanced drug delivery systems, Curr. Opin.
[106] J. Bergueiro, M. Calderón, Thermoresponsive nanodevices in biomedical Chem. Eng. 7 (2015) 101–112, https://doi.org/10.1016/j.coche.2014.12.001.
applications, Macromol. Biosci. 15 (2015) 183–199, https://doi.org/10.1002/ [132] F.A. Plamper, W. Richtering, Functional microgels and microgel systems, Acc.
mabi.201400362. Chem. Res. 50 (2017) 131–140, https://doi.org/10.1021/acs.accounts.6b00544.
[107] X. Guo, D. Li, G. Yang, C. Shi, Z. Tang, J. Wang, S. Zhou, Thermo-triggered drug [133] G. Agrawal, R. Agrawal, Functional microgels: recent advances in their
release from actively targeting polymer micelles, ACS Appl. Mater. Interfaces 6 biomedical applications, Small 14 (2018) 1801724, https://doi.org/10.1002/
(2014) 8549–8559, https://doi.org/10.1021/am501422r. smll.201801724.
[108] X. Hu, Y. Zhang, Z. Xie, X. Jing, A. Bellotti, Z. Gu, Stimuli-responsive [134] R.L. Srinivas, S.C. Chapin, P.S. Doyle, Aptamer-functionalized microgel particles
polymersomes for biomedical applications, Biomacromolecules 18 (2017) for protein detection, Anal. Chem. 83 (2011) 9138–9145, https://doi.org/
649–673, https://doi.org/10.1021/acs.biomac.6b01704. 10.1021/ac202335u.
[109] X. Cao, Y. Chen, W. Chai, W. Zhang, Y. Wang, P.-F. Fu, Thermoresponsive self- [135] L. Sigolaeva, D. Pergushov, M. Oelmann, S. Schwarz, M. Brugnoni, I. Kurochkin,
assembled nanovesicles based on amphiphilic triblock copolymers and their F. Plamper, A. Fery, W. Richtering, Surface functionalization by stimuli-sensitive
potential applications as smart drug release carriers, J. Appl. Polym. Sci. 132 microgels for effective enzyme uptake and rational design of biosensor setups,
(2015), https://doi.org/10.1002/app.41361. Polymers (Basel) 10 (2018) 791, https://doi.org/10.3390/polym10070791.
[110] O. Bixner, S. Kurzhals, M. Virk, E. Reimhult, Triggered release from [136] L. V Sigolaeva, S.Y. Gladyr, A.P.H. Gelissen, O. Mergel, D. V Pergushov, I.
thermoresponsive polymersomes with superparamagnetic membranes, Mater 9 N. Kurochkin, F.A. Plamper, W. Richtering, Dual-stimuli-sensitive microgels as a
(2016), https://doi.org/10.3390/ma9010029. tool for stimulated spongelike adsorption of biomaterials for biosensor
[111] Z. Iatridi, A. Angelopoulou, E. Voulgari, K. Avgoustakis, C. Tsitsilianis, Star-graft applications, Biomacromolecules 15 (2014) 3735–3745, https://doi.org/
quarterpolymer-based polymersomes as nanocarriers for Co-delivery of 10.1021/bm5010349.
hydrophilic/hydrophobic chemotherapeutic agents, ACS Omega 3 (2018) [137] H.J. Moon, D.Y. Ko, M.H. Park, M.K. Joo, B. Jeong, Temperature-responsive
11896–11908, https://doi.org/10.1021/acsomega.8b01437. compounds as in situ gelling biomedical materials, Chem. Soc. Rev. 41 (2012)
[112] Y. Li, B.S. Lokitz, C.L. McCormick, Thermally responsive vesicles and their 4860–4883, https://doi.org/10.1039/C2CS35078E.
structural “locking” through polyelectrolyte complex formation, Angew. Chem. [138] B. Jeong, S.W. Kim, Y.H. Bae, Thermosensitive sol–gel reversible hydrogels, Adv.
Int. Ed. 45 (2006) 5792–5795, https://doi.org/10.1002/anie.200602168. Drug Deliv. Rev. 54 (2002) 37–51, https://doi.org/10.1016/S0169-409X(01)
[113] B.S. Lokitz, A.J. Convertine, R.G. Ezell, A. Heidenreich, Y. Li, C.L. McCormick, 00242-3.
Responsive nanoassemblies via interpolyelectrolyte complexation of amphiphilic [139] M.H. Park, M.K. Joo, B.G. Choi, B. Jeong, Biodegradable thermogels, Acc. Chem.
block copolymer micelles, Macromolecules 39 (2006) 8594–8602, https://doi. Res. 45 (2012) 424–433, https://doi.org/10.1021/ar200162j.
org/10.1021/ma061672y. [140] A.P. Constantinou, T.K. Georgiou, Thermoresponsive gels based on ABC triblock
[114] C. Dähling, J.E. Houston, A. Radulescu, M. Drechsler, M. Brugnoni, H. Mori, D. V copolymers: effect of the length of the PEG side group, Polym. Chem. 7 (2016)
Pergushov, F.A. Plamper, Self-templated generation of triggerable and restorable 2045–2056, https://doi.org/10.1039/C5PY02072G.
nonequilibrium micelles, ACS Macro Lett. 7 (2018) 341–346, https://doi.org/ [141] M.C. Koetting, J.T. Peters, S.D. Steichen, N.A. Peppas, Stimulus-responsive
10.1021/acsmacrolett.8b00096. hydrogels: theory, modern advances, and applications, Mater. Sci. Eng. R Rep. 93
[115] C. Dähling, G. Lotze, H. Mori, D. V Pergushov, F.A. Plamper, Thermoresponsive (2015) 1–49, https://doi.org/10.1016/j.mser.2015.04.001.
segments retard the formation of equilibrium micellar interpolyelectrolyte [142] C. Chassenieux, C. Tsitsilianis, Recent trends in pH/thermo-responsive self-
complexes by detouring to various intermediate structures, J. Phys. Chem. B 121 assembling hydrogels: from polyions to peptide-based polymeric gelators, Soft
(2017) 6739–6748, https://doi.org/10.1021/acs.jpcb.7b04238. Matter 12 (2016) 1344–1359, https://doi.org/10.1039/C5SM02710A.
[116] C. Dähling, G. Lotze, M. Drechsler, H. Mori, D. V Pergushov, F.A. Plamper, [143] A.P. Constantinou, H. Zhao, C.M. McGilvery, A.E. Porter, T.K. Georgiou,
Temperature-induced structure switch in thermo-responsive micellar A comprehensive systematic study on thermoresponsive gels: beyond the common
interpolyelectrolyte complexes: toward core–shell–corona and worm-like architectures of linear terpolymers, Polymer 9 (2017), https://doi.org/10.3390/
morphologies, Soft Matter 12 (2016) 5127–5137, https://doi.org/10.1039/ polym9010031.
C6SM00757K. [144] X. Xu, Y. Liu, W. Fu, M. Yao, Z. Ding, J. Xuan, D. Li, S. Wang, Y. Xia, M. Cao, Poly
[117] C.A. Figg, A. Simula, K.A. Gebre, B.S. Tucker, D.M. Haddleton, B.S. Sumerlin, (N-isopropylacrylamide)-Based thermoresponsive composite hydrogels for
Polymerization-induced thermal self-assembly (PITSA), Chem. Sci. 6 (2015) biomedical applications, Polymer 12 (2020), https://doi.org/10.3390/
1230–1236, https://doi.org/10.1039/C4SC03334E. polym12030580.
[118] N.J.W. Penfold, J. Yeow, C. Boyer, S.P. Armes, Emerging trends in [145] H.F. Darge, A.T. Andrgie, H.-C. Tsai, J.-Y. Lai, Polysaccharide and polypeptide
polymerization-induced self-assembly, ACS Macro Lett. 8 (2019) 1029–1054, based injectable thermo-sensitive hydrogels for local biomedical applications, Int.
https://doi.org/10.1021/acsmacrolett.9b00464. J. Biol. Macromol. 133 (2019) 545–563, https://doi.org/10.1016/j.
[119] X. Huang, M. Li, D.C. Green, D.S. Williams, A.J. Patil, S. Mann, Interfacial ijbiomac.2019.04.131.
assembly of protein–polymer nano-conjugates into stimulus-responsive [146] S. Graham, P.F. Marina, A. Blencowe, Thermoresponsive polysaccharides and
biomimetic protocells, Nat. Commun. 4 (2013) 2239, https://doi.org/10.1038/ their thermoreversible physical hydrogel networks, Carbohydr. Polym. 207
ncomms3239. (2019) 143–159, https://doi.org/10.1016/j.carbpol.2018.11.053.
[120] C.-J. Huang, F.-C. Chang, Using click chemistry to fabricate ultrathin [147] Z. Iatridi, S.-F. Saravanou, C. Tsitsilianis, Injectable self-assembling hydrogel from
thermoresponsive microcapsules through direct covalent layer-by-layer assembly, alginate grafted by P(N-isopropylacrylamide-co-N-tert-butylacrylamide) random
Macromolecules 42 (2009) 5155–5166, https://doi.org/10.1021/ma900478n. copolymers, Carbohydr. Polym. 219 (2019) 344–352, https://doi.org/10.1016/j.
[121] W. Xu, P.A. Ledin, Z. Iatridi, C. Tsitsilianis, V. V Tsukruk, Multicompartmental carbpol.2019.05.045.
microcapsules with orthogonal programmable two-way sequencing of [148] Z. Cui, B.H. Lee, C. Pauken, B.L. Vernon, Degradation, cytotoxicity, and
hydrophobic and hydrophilic cargo release, Angew. Chem. Int. Ed. 55 (2016) biocompatibility of NIPAAm-based thermosensitive, injectable, and bioresorbable
4908–4913, https://doi.org/10.1002/anie.201600383. polymer hydrogels, J. Biomed. Mater. Res. 98A (2011) 159–166, https://doi.org/
[122] Y. Cong, Q. Li, M. Chen, L. Wu, Synthesis of dual-stimuli-responsive 10.1002/jbm.a.33093.
microcontainers with two payloads in different storage spaces for [149] Y. Xiao, Z.S. Chinoy, G. Pecastaings, K. Bathany, E. Garanger, S. Lecommandoux,
preprogrammable release, Angew. Chem. Int. Ed. 56 (2017) 3552–3556, https:// Design of polysaccharide-b-elastin-like polypeptide bioconjugates and their
doi.org/10.1002/anie.201612291. thermoresponsive self-assembly, Biomacromolecules 21 (2020) 114–125, https://
[123] G. Stoychev, N. Puretskiy, L. Ionov, Self-folding all-polymer thermoresponsive doi.org/10.1021/acs.biomac.9b01058.
microcapsules, Soft Matter 7 (2011) 3277–3279, https://doi.org/10.1039/ [150] H. Wang, D. Zhu, A. Paul, L. Cai, A. Enejder, F. Yang, S.C. Heilshorn, Covalently
C1SM05109A. adaptable elastin-like protein–hyaluronic acid (ELP–HA) hybrid hydrogels with
[124] T.R. Hoare, D.S. Kohane, Hydrogels in drug delivery: progress and challenges, secondary thermoresponsive crosslinking for injectable stem cell delivery, Adv.
Polymer (Guildf) 49 (2008), https://doi.org/10.1016/j.polymer.2008.01.027, Funct. Mater. 27 (2017) 1605609, https://doi.org/10.1002/adfm.201605609.
1993–2007. [151] A. Mellati, M. Valizadeh Kiamahalleh, S. Dai, J. Bi, B. Jin, H. Zhang, Influence of
polymer molecular weight on the in vitro cytotoxicity of poly (N-

15
G. Pasparakis and C. Tsitsilianis Polymer 211 (2020) 123146

isopropylacrylamide), Mater. Sci. Eng. C 59 (2016) 509–513, https://doi.org/ bioadhesive properties of poly-N-isopropylacrylamide hydrogel, Heliyon 5
10.1016/j.msec.2015.10.043. (2019), e01474, https://doi.org/10.1016/j.heliyon.2019.e01474.
[152] H. Vihola, A. Laukkanen, L. Valtola, H. Tenhu, J. Hirvonen, Cytotoxicity of [154] K.L. Heredia, D. Bontempo, T. Ly, J.T. Byers, S. Halstenberg, H.D. Maynard, In
thermosensitive polymers poly(N-isopropylacrylamide), poly(N- situ preparation of Protein− “Smart” polymer conjugates with retention of
vinylcaprolactam) and amphiphilically modified poly(N-vinylcaprolactam), bioactivity, J. Am. Chem. Soc. 127 (2005) 16955–16960, https://doi.org/
Biomaterials 26 (2005) 3055–3064, https://doi.org/10.1016/j. 10.1021/ja054482w.
biomaterials.2004.09.008. [155] T. Manouras, M. Vamvakaki, Field responsive materials: photo-, electro-,
[153] V. Capella, R.E. Rivero, A.C. Liaudat, L.E. Ibarra, D.A. Roma, F. Alustiza, magnetic- and ultrasound-sensitive polymers, Polym. Chem. 8 (2017) 74–96,
F. Mañas, C.A. Barbero, P. Bosch, C.R. Rivarola, N. Rodriguez, Cytotoxicity and https://doi.org/10.1039/C6PY01455K.

16

You might also like