Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

European Polymer Journal 124 (2020) 109488

Contents lists available at ScienceDirect

European Polymer Journal


journal homepage: www.elsevier.com/locate/europolj

Thermoresponsive behavior of poly(DEGMA)-based copolymers. NMR and T


dynamic light scattering study of aqueous solutions
⁎ ⁎,1
Rafał Konefał , Jiří Spěváček , Gabriela Mužíková, Richard Laga
Institute of Macromolecular Chemistry, Czech Academy of Sciences, Heyrovsky Sq. 2, 162 06 Prague 6, Czech Republic

A R T I C LE I N FO A B S T R A C T

Keywords: We studied temperature behavior of diblock copolymers of poly[2-(2-methoxyethoxy)ethyl methacrylate]


Thermoresponsive polymer (PDEGMA) with poly(N-(2-hydroxypropyl)methacrylamide and a statistical copolymer PDEGMA-co-N-pro-
Poly(2-(2-methoxyethoxy)ethyl methacrylate) pargylmethacrylamide in aqueous solutions. NMR spectra and dynamic light scattering showed that the copo-
(PDEGMA) lymers exhibit a reversible phase transition which depends on the composition of the block copolymer and
PHPMA
slightly on polymer concentration. In contrast to statistical copolymer, different behavior of proton groups in
Aqueous solution
NMR
polymer backbone (and in the nearest side chain OCH2 group) on the one hand and in other side chain groups of
Spin-spin relaxation times T2 the PDEGMA on the other hand was observed for diblock copolymers. Additionally, two types of water molecules
NOESY were detected in solutions of the diblock copolymers above the phase transition –“free” with long and “bound”
Dynamic light scattering with short spin-spin relaxation times T2. NOESY spectra indicate that a core- to- shell morphology is unlikely in
these block copolymers. This report emphasizes the importance of understanding of the self-association of co-
polymers in solution on the molecular level and provides important information for the design of “smart”
thermoresponsive polymer-based drug delivery systems.

1. Introduction isopropylacrylamide) (PNIPAm) with LCST around 305 K [12–15].


However, in recent years, polymer chemists reported many interesting
Stimuli-responsive polymers have acquired great attention in a field alternatives to PNIPAm [16–19]. One group of such polymers are
of polymer biomaterials. These polymers after small external stimuli, polymethacrylates containing short oligo(ethylene glycol) (OEG) side
such as changes in temperature, pH, ionic strength, light irradiation or chains. Generally, the phase transition temperatures of these polymers
complexation with appropriate molecules, significantly alter their depend namely on the length of the OEG side chain. For instance, poly
physical or chemical properties [1–5]. Among the miscellaneous stimuli (2-(2-methoxyethoxy)ethyl methacrylate) (PDEGMA) with two ethy-
aforementioned, due to non-invasive treatment and a wide range of lene glycol units, and poly(2-[2-(2- methoxyethoxy)ethoxy] ethyl me-
applications such as drug delivery, tissue engineering, bioseparation, thacrylate) (PTEGMA) with three ethylene oxide units shows LCST
thermoresponsive films, oil-gas industry and nanoreactors, thermo- around 299 K and 325 K, respectively. Polymers with longer side chains
responsive polymers are the most extensively investigated [6–10]. In (4–9 ethylene glycol units) exhibit phase transitions between 333 and
the field of thermoresponsive synthetic polymer biomaterials, most 363 K. A further advantage of those polymers is possibility of tuning of
studies focus on materials that exhibit phase separation in water with their LCST to the application requirements. For example, LCST values
the lower critical solution temperature (LCST) ranging from ~293 K to below physiological temperatures (i.e. 310 K) can be achieved by co-
308 K [11]. This can be very useful, for example, in terms of drug de- polymerizing OEGMA with various hydrophobic comonomers. More-
livery purposes, since the polymers having LCST values in this tem- over, studies of the interactions of POEGMAs with a number of cell lines
perature range can be readily and reproducibly processed and dissolved denote that these polymers are non-toxic and biocompatible [20–23].
at room temperature, at which their chains occur in a random coil Taking into account the abovementioned facts, copolymers based on
conformation (soluble form), while their chains collapse under phy- the PDEGMA appear to be suitable candidates for biomedical applica-
siological conditions (e.g., when introduced into the body) to form tions.
compact globuli (insoluble particles form). In particular, the best In aqueous media, block copolymers where one component is hy-
known and most widely studied polymer from this group is poly(N- drophilic and the other thermoresponsive can form various self-


Corresponding authors.
E-mail addresses: konefal@imc.cas.cz (R. Konefał), spevacek@imc.cas.cz, spevac@seznam.cz (J. Spěváček).
1
Emeritus scientist (J.S.)

https://doi.org/10.1016/j.eurpolymj.2020.109488
Received 4 November 2019; Received in revised form 2 January 2020; Accepted 6 January 2020
Available online 07 January 2020
0014-3057/ © 2020 Elsevier Ltd. All rights reserved.
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

assembled structures (micelles, vesicles, microgels or nanoparticles) at checked by a reversed-phase HPLC, showing a single peak with a re-
elevated temperature [24,25]. Examples of all those nanostructures tention time of 4.0 min (UV detection at 230 nm). 1H NMR (300 MHz,
obtained in polymers based on different POEGMAs are listed in an ex- DMSO) δ ppm: 1.85 (s, 3H, eCH3), 3.05 (s, 1H, ^CH), 3.88 (d, 2H,
cellent review recently published by Szweda et al. [20]. eCH2e), 5.37 and 5.68 (s, 1H, ]CH2), 8.37 (s, 1H, eNHe).
Among the methods used to investigate the temperature behavior of 2-[1-Cyano-1-methyl-4-oxo-4-(2-thioxothiazolidin-3-yl)butylazo]-2-
the LCST-type polymer systems such as cloud point measurement, in- methyl-5-oxo-5-(2-thioxothiazoli- din-3-yl)pentanenitrile (ACVA-(TT)2)
frared (IR) spectroscopy, light scattering, small angle X-ray scattering, was prepared by the reaction of 4,4′-azobis(4-cyanovaleric acid)
calorimetry etc. [26–31], NMR spectroscopy can provide quantitative (ACVA) with TT in tetrahydrofuran in the presence of N,N'-dicyclo-
information on the LCST phase separation behavior [32,33]. Relaxation hexylcarbodiimide (DCC) and 4-dimethylaminopyridine (DMAP) as
time and diffusion experiments measurements offer possibility to follow described in [41].
the changes in molecular motions of polymer and water in solution. Dithiobenzoic acid 1-cyano-1-methyl-4-oxo-4-(2-thioxothiazolidin-
Additionally, the 2D 1H–1H NOESY spectra are extremely valuable to 3-yl)butyl ester (CTA-TT) was synthesized by the reaction of ACVA-
clarify conformational problems of macromolecules. In last decade, 1H (TT)2 with bis(thiobenzoyl) disulfide in ethyl acetate at 353 K [41].
NMR spectroscopy, NMR relaxation times and 2D 1H–1H NOESY tech-
niques were successfully applied to study various single- and multi- 2.3. Synthesis of thermoresponsive diblock copolymers
component thermoresponsive polymer systems in aqueous solutions
[32–39]. The thermoresponsive diblock (A-B type) copolymers were synthe-
In the present work we applied 1H NMR spectroscopy, 1H spin-spin sized by RAFT polymerization technique in two synthetic steps [42].
relaxation times (temperature and time dependences) and 2D nuclear First, the hydrophilic block A (PHPMA) was prepared by polymerizing
Overhauser effect spectroscopy (NOESY) at various temperatures (ap- HPMA in the presence of CTA-TT and ACVA-TT. Then, block A was
plied only to block copolymer) in combination with dynamic light subjected to a chain-extension polymerization with DEGMA in the
scattering (DLS) to study temperature-induced phase separation in presence of AIBN to introduce the thermoresponsive block B. Three
aqueous solutions of block copolymers composed of the hydrophilic different molar ratios of PHPMA to DEGMA were used to synthesize the
poly(N-(2-hydroxypropyl)methacrylamide (PHPMA) block and ther- diblock copolymers with variable lengths of the thermo-responsive
moresponsive PDEGMA block (with different molecular weights of blocks (PHPMA)53-b-(PDEGMA)26-78).
PDEGMA block) and P[(DEGMA)–co-N-propargylmethacrylamide Example: A mixture of CTA-TT (22.1 mg, 58.2 μmol) and ACVA-TT
(PGMA)] statistical copolymer. Since both PHPMA and PDEGMA are (14.1 mg, 29.1 μmol) was dissolved in DMSO (388 μL) and added to a
biocompatible, non-toxic and non-immunogenic materials, those co- solution of HPMA (0.5 g, 3.49 mmol) in tert-butanol (3492 μL). The
polymers are suitable for application in biomedicine and/or drug and reaction mixture was thoroughly bubbled with Ar and polymerized in
gene delivery systems. the sealed glass ampoules at 343 K for 16 h. The polymer was pre-
cipitated to a mixture of acetone and diethyl ether (3:1), re-dissolved in
2. Materials and methods methanol and purified by gel filtration using a SephadexTM LH-20 in
methanol. The methanolic solution was precipitated to diethyl ether
2.1. Chemicals yielding 274.6 mg of the PHPMA hydrophilic polymer precursor
(PHPMA) as a pink powder.
(RS)-1-Aminopropan-2-ol; 4,4′-azobis(4-cyanovaleric acid) (ACVA); Next, a mixture of PHPMA (20.0 mg, 2.61 μmol ~DTB groups) and
azobisisobutyronitrile (AIBN); 4-cyano-4-(phenylcarbonothioylthio) AIBN (0.09 mg, 0.52 μmol) was dissolved in 424 μL of dimethylaceta-
pentanoic acid (CTA-ACVA); 2-cyano-2-propyl benzodithioate (CTA- mide (DMAc), mixed with a solution of DEGMA (64.9 mg, 0.35 mmol)
AIBN); N,N'-dicyclohexylcarbodiimide (DCC); 4-dimethylaminopyr- in 424 μL of tert-butanol, thoroughly bubbled with Ar and polymerized
idine (DMAP); N-(3-dimethylaminopropyl)-N′-ethylcarbodiimide hy- in the sealed glass ampoules at 343 K for 16 h. The diblock copolymer
drochloride (EDC); magnesium sulfate; methacrylic acid; methacryloyl was isolated by precipitation to diethyl ether yielding 65.0 mg of the
chloride; 2-(2-methoxyethoxy)ethyl methacrylate (DEGMA); propargy- pale pink amorphous solid.
lamine; sodium carbonate and thiazolidine-2-thione (TT) were pur- The diblock copolymer was dissolved in 650 μL DMAc, AIBN
chased from Sigma-Aldrich, Czech Republic. All solvents used in this (13.0 mg, 79.2 μmol) was added and the mixture was allowed to react
work were of high-purity grade with extremely low water levels pur- 2 h at 353 K. After cooling the reaction mixture down to r.t., 1-ami-
chased from VWR, Czech Republic. nopropan-2-ol (6.5 mg, 86.5 μmol) was added and the mixture was
allowed to react for next 2 h. The resulting diblock copolymer was
2.2. Synthesis of monomer, initiator and RAFT agent precipitated to diethyl ether, re-dissolved in methanol and purified by
gel filtration using the Sephadex™ LH-20 in methanol. The methanolic
N-(2-Hydroxypropyl)methacrylamide (HPMA) was synthesized by solution was precipitated to diethyl ether yielding 52.1 mg of the di-
reacting methacryloyl chloride with (RS)-1-aminopropan-2-ol in di- block copolymer (PHPMA)53-b-(PDEGMA)78 as a white amorphous
chloromethane in the presence of sodium carbonate according to pro- solid. The SEC characteristics for the hydrophilic polymer precursor as
cedure described in [40]. well as for the diblock copolymers are summarized in the Table 1.
N-Propargylmethacrylamide (PGMA) was synthesized by reacting
methacrylic acid with propargylamine in dichloromethane (DCM) in 2.4. Synthesis of thermoresponsive statistical copolymer
the presence of N-(3-dimethylaminopropyl)-N′-ethylcarbodiimide hy-
drochloride (EDC). A mixture of methacrylic acid (1.567 g, 18.2 mmol) The termoresponsive statistical copolymer was prepared by copo-
and EDC (4.525 g, 23.6 mmol) was dissolved in 86 mL of DCM and lymerizing DEGMA with PGMA through the RAFT mechanism. A mix-
cooled to 255 K. Propargylamine (1,0 g, 18,2 mmol) and a few crystals ture of DEGMA (0.5 g, 2.7 mmol), PGMA (36.4 mg, 0.3 mmol), CTA-
of DMAP were added to the cooled solution and the reaction mixture AIBN (2.0 mg, 8.9 μmol) and AIBN (0.3 mg, 1.8 μmol) was dissolved in
was stirred 2 h at 255 K and then overnight at room temperature. The dioxan (738 μL), thoroughly bubbled with Ar and polymerized in the
DCM solution was washed with brine (3 × 100 mL) and the organic sealed glass ampoules at 343 K for 4 h. The copolymer was precipitated
layer was dried over anhydrous MgSO4. After filtration of MgSO4, the to diethyl ether, re-dissolved in methanol and purified by gel filtration
DCM solution was concentrated under the reduced pressure and the using a Sephadex™ LH-20 in methanol. The methanolic solution was
product was isolated by crystallization from DCM/hexane/diethyl ether precipitated to diethyl ether yielding 240.8 mg of the statistical copo-
mixture. The yield was 1.23 g (55%). The purity of the product was lymer as a pink waxy solid.

2
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

Table 1
SEC, NMR and DLS characteristics for the hydrophilic polymer precursor (HP), thermoresponsive diblock copolymers (DB1 – DB3) and thermoresponsive statistical
copolymer (SP).
** ***
Sample (Co)polymer code DEGMA content (mol%) *Mn [g/mol] *Mw/Mn DH310K [nm] ***
Ttr [K]

PHPMA HP 0 7,660 1.01 4.9 n.a.


P[(HPMA)53-b-(DEGMA)26] DB1 33 11,600 1.20 5.6 312
P[(HPMA)53-b-(DEGMA) 53] DB2 50 17,100 1.20 29.5 308
P[(HPMA)53-b-(DEGMA)78] DB3 60 21,500 1.30 33.8 304
P[(DEGMA)–co-(PGMA)] SP 93 37,320 1.09 1540.0 294

* SEC.
** NMR.
*** DLS.

After that, the statistical copolymer was dissolved in 2.4 mL DMAc, S-shaped curve (sigmoidal curve) fit.
AIBN (21.2 mg, 0.13 mmol) was added and the mixture was allowed to
react 2 h at 353 K. The resulting product was precipitated to diethyl 2.8. NMR spectroscopy
ether, re-dissolved in methanol and purified by gel filtration using the
Sephadex™ LH-20 in methanol. The methanolic solution was pre- Temperature dependences of 1H NMR spectra were acquired with
cipitated to diethyl ether yielding 220.1 mg of the P[(DEGMA)–co- Bruker Avance III 600 spectrometer operating at 600.2 MHz. The width
(PGMA)] statistical copolymer as a white amorphous solid. The SEC of 90° pulse was 10 μs, relaxation delay 10 s, acquisition time 2.18 s, 16
characteristics of the copolymer are stated in the Table 1. scans. Each sample was kept for 15 min at the desired temperature
before measurement. The integrated intensities were determined with
2.5. UV–VIS spectrophotometry spectrometer integration software with an accuracy of ± 1%. 2D 1H–1H
NOESY NMR spectra were recorded on the same spectrometer with
The spectrophotometric analyses of the (co)polymers were carried 4098 Hz spectral window in f1 and f2 frequency axes, and mixing times
out in quartz glass cuvettes with an optical path length of 0.1 cm on a in the range 100–600 ms. A total of 16 scans was accumulated over
UV–VIS spectrophotometer Specord Plus (Analytik Jena, Germany). 512 t1 (evolution time) increments with a relaxation delay of 10 s. The
The molar content of dithiobenzoate (DTB) end group in the (co) temperature and time dependences of 1H spin-spin relaxation times T2
polymers were determined at 302 nm in methanol using the molar of HDO and selected proton groups the copolymer were measured using
absorption coefficient of 12,100 L/mol·cm. The molar content of the the CPMG pulse sequence 90°x-(td-180°y-td)n-acquisition [43]. The re-
terminal carbonylthiazolidine-2-thione reactive group in the (co)poly- laxation delay between scans was 100 s, acquisition time 2.84 s with 2
mers was determined at 305 nm in methanol using the molar absorption scans. The relative error for T2 values did not exceed ± 5%. In all
coefficient of 10,300 L/mol·cm. measurements temperature was maintained constant within ± 0.2 K in
the range 290–330 K using a BVT 3000 temperature unit. Temperature
2.6. Size-exclusion chromatography (SEC) was calibrated using a standard 80% ethylene glycol (DMSO‑d6)
sample. All samples in D2O (Euriso-top, 99.9% of deuterium) solutions
The molecular weights and molecular weights distributions of the (polymer concentrations c = 0.67 and 3 wt%) were filled into 5-mm
(co)polymers were determined by SEC on a HPLC system (Shimadzu Norell NMR Tubes ST500-7 HT with ± 0.77 mm wall thickness and
VP, Japan), equipped with internal UV–VIS photodiode array detector, 178 mm length, which were degassed and sealed under nitrogen.
and external differential refractive index and multiangle light scattering
detectors (Wyatt Technologies, USA). The TSK-Gel SuperAW3000 and 3. Results and discussion
SuperAW4000 columns (6.0 × 150 mm, Tosoh Bioscience, Japan)
connected in series and 80% methanol/20% sodium acetate buffer 3.1. Copolymer synthesis
(0.3 M, pH 6.5) mixture as a mobile phase (flow rate 0.6 mL/min) was
used. A method based on the known total injected mass with an as- All (co)polymers were generated using the controlled radical poly-
sumption of 100% recovery was used for the estimation of the dn/dc merization mechanism (RAFT technique) enabling the reproducible and
values needed for the calculation of the molecular weights from light efficient preparation of highly defined materials with predetermined
scattering data. molecular weights and high yield of terminal functional groups. The
hydrophilic polymer precursor (PHPMA) was synthesized by poly-
2.7. Dynamic light scattering (DLS) merizing HPMA in the presence of dithiobenzoate (DTB)-derived chain
transfer agent (CTA) yielding polymer with Mn of 7660 g/mol, narrow
The hydrodynamic diameter (DH) of the thermoresponsive copoly- molecular weights distribution (Mw/Mn = 1.01) and high functionality
mers was measured by the DLS technique at a scattering angle θ = 173° of end groups (fDTB ≈ fTT ≈ 0.9). The hydrophilic polymer chain (block
using a Nano-ZS instrument (Model ZEN3600, Malvern Instruments, A) was further extended with the thermoresponsive PDEGMA block
UK) equipped with a 632.8 nm laser. The temperature measurement (block B). Three different ratios of DEGMA to PHPMA (44:1, 88:1 and
was performed to investigate the transition of the thermoresponsive 132:1) were used to synthesize the amphiphilic diblock copolymers
copolymer chains from random coils to polymer micelles in the tem- (PHPMA)-b-(PDEGMA) with the lengths of thermoresponsive block
perature interval 293–323 K (in 1 K increments) in PBS (1.0 mg/mL, pH ranging from approximately 4000–14,000 g/mol (see Table 1). The
7.4) solutions. For the evaluation of the DLS data, the DTS(Nano) gentle broadening in the molecular weight distributions of the diblock
program was used. The mean of at least three independent measure- copolymers can be ascribed to the presence of a low quantity of the
ments was calculated. The transition temperature (Ttr), characterizing dead chains (not terminated with ~DTB groups) in the block A that
the polymer chain conformation changes, was evaluated from the could not react further with DEGMA units through the RAFT me-
temperature dependence of the hydrodynamic diameter (DH) and was chanism. Both ~DTB and ~TT polymer end groups were blocked by a
determined from the intersection point of two lines formed by the linear homolytic reaction with a high molar excess of AIBN or by an amino-
regression of a lower horizontal asymptote and a vertical section of the lysis with a high molar excess of 1-aminopropan-2-ol, respectively, to

3
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

Fig. 1. Scheme of synthesis of the hydrophilic polymer precursor (PHPMA) and the thermoresponsive diblock copolymers (PHPMA)-b-(DEGMA).

prevent their potential hydrolysis or other non-specific reactions during sized polymer micelles (above Ttr) (see Table 1 and Fig. 3). For ex-
next physico-chemical or biological experiments (for the reaction ample, the DB1 copolymer with the MW ratio of the hydrophilic
scheme see Fig. 1). (PHPMA) to the thermoresponsive (PDEGMA) blocks ~2/1 has 8 K
The thermoresponsive statistical copolymer (P higher Ttr and forms 5 nm larger-sized micelles than the DB3 copolymer
[(DEGMA)–co–(PGMA)]) was prepared by copolymerizing DEGMA having the MW of the hydrophobic block ~2 times higher than MW of
with PGMA in the presence of dithiobenzoate (DTB)-derived chain the hydrophilic block. It is a quite interesting finding showing that the
transfer agent (CTA) yielding polymer with Mn of 37,320 g/mol, narrow individual blocks of the diblock copolymers do not behave in the mi-
molecular weights distribution (Mw/Mn = 1.09) and high functionality celles as independent structural units but they may influence each
of ~DTB end group (fDTB ≈ 1.0). The content of PGMA units in the other.
copolymer, determined using the 1H NMR, was 7.3 mol% (see Fig. S1 in In the case of the statistical copolymer (SP), we observed that above
Supplementary material). The reactive PGMA co-monomer units were Ttr the copolymer chains assembled into hydrophobic globules with the
incorporated to the copolymer due to the possibility of their future hydrodynamic size ~1500 nm. The reason for the formation of such
modification with any azide group-containing compound through the large-size and polydisperse objects (aggregates) instead of precisely
CuI-catalysed cycloaddition reaction (so called click chemistry). As in defined polymer micelles observed in the case of diblock copolymers is
the case of the diblock copolymers, terminal ~DTB group of the P absence of the hydrophilic blocks which condense and stabilize the
[(DEGMA)–co–(PGMA)] copolymer was removed by a homolytic re- hydrophobic PDEGMA chains with a hydrophilic PHPMA. In an
action with a high molar excess of the azoinitiator (for the reaction agreement with the theory, statistical incorporation of the hydrophobic
scheme see Fig. 2). PGMA units to the structure of the SP induced shift of Ttr to the lower
value (294 K) than it was measured for the unmodified PDEGMA
3.2. Solution behavior of copolymers studied by DLS homopolymer (299 K) with the similar molecular weight.

The temperature-dependent changes in solution behavior of the 3.3. Copolymers behavior on molecular level: 1H NMR spectra and fraction
thermoresponsive diblock copolymers (DB1 – DB3) and thermo- p of proton groups (units) with significantly reduced mobility
responsive statistical copolymer (SP) were studied by DLS in the solu-
tions mimicking physiological conditions. Specifically, we evaluated Fig. 4 shows high-resolution 1H NMR spectra of a D2O solution
how the thermoresponsive block (PDEGMA) length influences the (c = 0.67 wt%) of the SP statistical copolymer measured under the
transition temperature (Ttr) of the diblock copolymers or how the same instrumental conditions at three temperatures. The assignment of
presence of the hydrophobic co-monomer (PGMA) influences the self- resonances to various proton types is shown directly in the spectrum
assembly of the PDEGMA chains. We demonstrated that an increase in measured at 290 K and chemical structure of statistical copolymer is
molecular weight of the PDEGMA block (at a constant MW of the hy- shown at the Figure. The broad signals “a” (δ ≈ 2 ppm) and “b”
drophilic block) caused not only a decrease in transition temperatures (δ ≈ 1 ppm) are respectively related to methylene CH2 and methyl CH3
(Ttr) of the diblock copolymers but it also resulted in the formation of protons from the main chain of SP. DEGMA side chain C(O)OCH2 group
larger-sized polymer coils (below Ttr) and to a certain extent in larger- was assigned as “c” (δ ≈ 4.2 ppm) while peaks marked as “d, e, f”

4
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

Fig. 2. Scheme of synthesis of the thermoresponsive P[(DEGMA)–co-(PGMA)] statistical copolymer.

Fig. 3. Temperature dependence of the hydrodynamic diameter (DH) of the thermoresponsive diblock copolymers (DB1 – DB3) (a) and statistical copolymer (SP) (b)
measured by DLS in PBS buffer (0.15 mM, pH 7.4) at a concentration c = 0.1 wt%.

(δ ≈ 3.9–3.6 ppm) correspond to OCH2 protons. Moreover strong signal LCST (310 K) of SP. In this section we shall concentrate on changes in
of OCH3 “g” is observed at δ = 3.45 ppm. Due to the hydrophobic integrated intensities of polymer signals. The most significant effect
character and low content (7 mol%) signals related to PGMA monomer observed in the spectra is a visible reduction in integral intensities and
repeating units are not visible in the spectrum (for chemical shifts see disappearance of all signals of DEGMA units. This result is evidently
spectrum in DMSO; Fig. S1 in Supplementary material). 1H NMR related to the fact that with increasing temperature, the mobility of the
spectra presented in Fig. 4 were measured at temperatures below the part of polymer segments which form globular-like structures (me-
LCST (290 K), in the middle of the transition (297 K) and above the soglobules) decreases to such an extent that they escape detection in

5
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

Fig. 4. 600.2 MHz 1H NMR spectra of SP statistical copolymer in D2O solution (c = 0.67 wt%) measured at 290, 297 and 310 K under the same instrumental
conditions.

high-resolution NMR spectra. It has been shown for other thermo- in Fig. 5. Similarly as in Fig. 4 peaks assignments of the various proton
responsive polymers studied by us and other authors previously that types are shown in the spectrum measured at 295 K. Signals of
proton linewidths of segments in globular structures are approx. 100 PDEGMA units (a, b, c, d, e, f, g) are at the same positions as in the
times larger and spin-spin relaxation times T2 are approx. 100 times spectra of statistical copolymer (SP cf. Fig. 4). Additional two peaks of
shorter in comparison with polymers segments retaining a high mobi- HPMA monomer repeating units side chain are detected: “k” (CH,
lity [32,33]. This shows that mobility of polymer segments in globular δ ≈ 3.9 ppm) and “j” (CH2, δ ≈ 3.3 ppm). Signals from remaining
structures is affected in direction to solid-like systems. At the same time HPMA proton groups (marked as a, b, l) are overlapped by DEGMA
linewidths and T2 values are for globules probably associated with backbone peaks. In the spectra recorded at 307 K and 320 K DEGMA
isotropic Brownian tumbling of globules as a whole as well as with signals show similar effect as observed for SP, i.e., a decreasing integral
internal segmental mobility. Similar behavior as depicted in Fig. 4 was intensity of “a, b, d, e, f, g” signals and disapperance of “c” peak with
previously observed also for PDEGMA homopolymer [44] and other increasing temperature. Nevertheless, in comparison with SP copo-
thermoresponsive polymer systems [13,28,32,33,38]. lymer (cf. Fig. 4) this effect is much weaker and intensity changes de-
High-resolution 1H NMR spectra of the D2O solution (c = 0.67 wt%) pend on distance of proton group from the main chain of the block;
of the block copolymer DB2 recorded at three temperatures 295 K, increasing distance (i.e., increasing number of bonds from the back-
307 K and 320 K under the same instrumental conditions are presented bone) - smaller change in the intensity: “c” fully disappears, “g” shows a

Fig. 5. 600.2 MHz 1H NMR spectra of DB2 block copolymer in D2O solution (c = 0.67 wt%) measured at 295, 307 and 320 K under the same instrumental conditions.

6
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

certain broadening, which might be directly related to the hydro- we shall deal only with such signals which are not created by protons
phobicity of the respective proton groups [28,36]. On the other hand, from the both blocks (i.e., signals c, d + e + f, g, j, k). In comparison
intensity of peaks assigned to HPMA protons remains virtually un- with SP (Fig. 6a) a different temperature behavior of PDEGMA proton
changed. This effect in combination with DLS measurements imply groups is observed: only methylene protons “c” closest to block back-
nanoparticle formation at temperatures above the LCST. bone show a sharp transition with rather high pmax (≈0.80) value and
From temperature dependent integrated intensities of NMR signals LCST shifted to ≈305 K. Next OCH2 (d + e + f) side chain proton
it is possible to quantitatively characterize changes occurring during groups reach a smaller pmax (≈0.36) value at 310 K which with in-
the heating and cooling processes. For this purpose the values of the creasing temperature slightly drops. Moreover, methoxy end-group of
fraction p of proton groups of the given type with significantly reduced the side chain (g) exhibits a weak phase transition at higher tempera-
mobility were obtained using the relation [32,33,38,45,46]: ture (starting at 315 K) with pmax (≈0.20). On the other hand, inter-
esting behavior of the PHPMA block is observed: signals “j” and “k”.
I (T )
p=1− T0
The starting values of p-fraction of both PHPMA proton groups
I (T0) × T (1) (j ≈ 0.20 and k ≈ 0.30) decrease to reach minimum value at 306 K and
304 K, respectively, and then after a shallow minimum start to increase
where I(T) is the integrated intensity of given polymer signal in the
again at 315 K. The described behavior is in contrast to various ther-
spectrum at given absolute temperature T and I(T0) is the integrated
moresponsive homopolymers [PNIPAm, poly(N-iso-
intensity of this signal when no phase transition or other reason for the
propylmethacrylamide) (PNIPMAm), poly(N,N-diethylacrylamide),
reduced mobility of polymer segments occurs. For T0 we chose the
poly(vinyl methylether) (PVME), poly(N-vinyl caprolactam) (PVCL),
temperature where the integrated intensity of the given signal was the
poly(2-ethyl-2-oxazoline) (PEOx) [32,38,47], random or gradient co-
highest and therefore p(T0) = 0. Additionally, in denominator of the
polymers containing NIPAm or EOx [38,39] units as well to block co-
Eq. (1) we took into account the fact that the integrated intensities
polymers with PNIPAm or PEOx blocks [13,46] where temperature
should decrease with temperature as 1/T (Curie law holds also for
dependences of the p-fraction are virtually the same for all proton
nuclear magnetization). In Fig. 6a, temperature dependences of the p-
groups of the thermoresponsive component. In all these systems side
fraction of various proton types of SP in D2O solution (c = 0.67 wt%)
chains in thermoresponsive units are relatively short. On the other hand
are shown. For all proton groups (signal assignment in Fig. 4) p-fraction
a similar situation to described above by us for P[(HPMA)-b-(DEGMA)]
first slightly decreases and has a minimum at 293 K. At temperatures
diblock copolymers has been reported by Hiller et al. [48] who studied
above 293 K the values of the p-fraction are drastically increasing and
diblock copolymers composed of hydrophilic poly(2-methyl-2-oxazo-
phase transition occurs in accord with DLS results, (see Fig. 3b). Tem-
line) and hydrophobic poly(2-alkyl-2-oxazoline) where alkyl is pentyl,
perature dependences of the p-fraction determined from integrated in-
heptyl or nonyl, i.e. again systems where side chains are fairly long.
tensities of various DEGMA signals are for SP copolymer virtually the
Values of the p-fraction decreasing with temperature as observed for
same so confirming that p-fraction relates to PDEGMA units as a whole.
side chain proton groups of PHPMA units (signals j and k in Fig. 6b)
This means that all SP groups are similarly restricted in their mobility
resemble rather an UCST behavior like in κ-carrageenan [45] than
and the SP copolymer forms aggregates. Maximum values of the p-
normal temperature behavior of polymers in solution where mobility of
fraction (pmax ≈1) give quantitative information on the fraction of
polymer chains increases with temperature without any changes in p-
polymer chains which participate in the phase transition and have been
fraction values. Explanation of abovementioned effects is related to
achieved around 310 K. The LCST (defined as the temperature at pmax/
different changes in mobility of each proton groups. Mobility of
2) was for SP estimated as 298 K in agreement with DLS results when
PDEGMA block main chain decreasing with temperature caused by
we take into account the temperature in the middle of the transition
vanishing of hydrophilic interactions and their substitution by hydro-
interval (cf. Fig. 3b).
phobic interactions, affects also close methylene group (c). Side chain,
Temperature behavior of block copolymers was similarly char-
due to oxygen atoms, in some fraction (65% for d + e + f groups and
acterized by measuring the temperature dependences of their aqueous
80% for g methoxy group) remains hydrophilic. Temperature behavior
solutions. As an example temperature dependences of p-fraction of
of PHPMA block is connected with the PDEGMA block: mobility of
various proton types of DB3/D2O solution (c = 0.67 wt%) are pre-
PHPMA (j, k) is increasing to the phase transition temperature of
sented in Fig. 6b. Temperature dependences shown in the Fig. 6b are for
PDEGMA. In this temperature region copolymer chains are reorganizing
all measured signals (signal assignment in Fig. 5), but in the discussion

Fig. 6. Temperature dependences of the fraction p as determined for signals of various proton types in D2O solutions (c = 0.67 wt%) of SP (a) and DB3 (b) during
gradual heating.

7
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

Fig. 7. Temperature dependences of the fraction p for the PDEGMA OCH2 protons (signal “c”) in D2O solutions (c = 0.67 wt%) of SP and DB1-DB3 copolymers (a)
and DB3 at two concentrations (c = 0.67 and 3 wt%) (b) during gradual heating.

and form nanoparticles. Similar behavior was observed also for other composition.
two block copolymers (DB1 and DB2, cf. Fig. S2 in the Supplementary Characterization of the reversibility of the phase transition after
material) and it is in good agreement with the DLS measurements (see heating process was obtained directly by similar measurements during
Fig. 3a). subsequent gradual cooling. During cooling process behavior of all
For further considerations and comparisons as the most suitable proton groups was similar to heating. Temperature dependences of the
signal “c” of PDEGMA C(O)OCH2 protons was chosen. Temperature fraction p of the PDEGMA OCH2 protons with significantly reduced
dependences of the fraction p of D2O solutions (c = 0.67 wt%) of all mobility (signal “c”) in D2O solutions (c = 0.67 wt%) of SP and DB3
investigated polymers are shown in Fig. 7a. From Fig. 7a it follows that during gradual heating and subsequent gradual cooling are shown in
in comparison with the SP (LCST = 298 K), in all block copolymers Fig. 8a and b, respectively. In contrast to the cloud point measurements
values of the LCST are shifted to higher temperatures and depend on the presented in the literature [21,22,49], where hysteresis between
PDEGMA content in the copolymer with dependence: lower PDEGMA heating and cooling process was observed, in NMR measurements this
content – higher LCST (DB1 = 310 K, DB2 = 308 K and DB3 = 305 K). effect is not detected both in case of the SP and in diblock copolymer
This result is in agreement with DLS results (Fig. 3) and with literature solutions (for comparisons of the DB1, DB2 c = 0.67 wt% and DB3
[20,22,49,50]. Moreover, there is a visible change in the pmax values c = 3 wt% see Figs. S6 and S7 in the Supplementary material).
which decrease with reduction of the PDEGMA content. These effects Therefore the same way of the phase transition during heating and
are caused by changes in overall hydrophilicity of copolymer chains: cooling process indicates that on molecular level the phase transition of
higher PHPMA content – higher hydrophilicity of the copolymer and all investigated samples is fully reversible without any hysteresis.
coil- to- globule transition becomes weaker due to smaller amount of
hydrophobic interactions above the LCST (shorter PDEGMA block) 3.4. Behavior of water (HDO) and copolymer molecules as shown by spin-
[20]. Additionally, smaller pmax values which are directly connected to spin relaxation times T2
mobility of the proton groups of the given type (here proton group “c”)
show that some percentage of them are still mobile (i.e. 40% for DB1), Information on behavior of water and polymer-solvent interactions
but from DLS results (see Fig. 3a) it follows that nanoparticles are (hydration) during phase-transition in aqueous solutions can be pro-
formed. Therefore one can assume some hydration of polymer chains in vided by NMR relaxation measurements of solvent, or in some cases
nanoparticles [51,52]. When nanoparticle is more hydrated, polymer polymer molecules [53–55]. It was shown that especially spin-spin re-
chains are more mobile and achieve smaller values of the pmax. Similar laxation times T2 of water (HDO protons or D2O deuterons) are useful in
behavior was observed for other PDEGMA proton groups and compar- this respect [13,14,32,33,38]. Fig. 9a shows temperature dependence of
isons are shown in the Supplementary material, Figs. S3 and S4. From 1
H spin-spin relaxation time T2 of HDO in D2O solution (c = 0.67 wt%)
studies based on cloud point or transmittance measurements it follows of the SP. Measurements were done at temperatures based on the
that LCST of PDEGMA- based copolymers in aqueous solutions slightly temperature dependence of the p-fraction (see Fig. 6a). At all tem-
depends on polymer concentration [22,49]. For this reason we chose peratures there was a single line of HDO in 1H NMR spectrum and this
DB3 copolymer as an example to study its solutions at two polymer holds for all investigated samples. Low starting values of T2 (≈1 s) are
concentrations, c = 0.67 wt% and 3 wt%. Temperature dependences of related to the relatively low mobility of water molecules and imply that
the p-fraction show similar temperature behavior for both polymer water molecules interact with polymer chains of the SP by hydrogen
concentrations for all copolymer proton types (see Fig. 6b and Fig. S5 in bonding. Additionally, values of T2 slightly increase with temperature
the Supplementary material). Fig. 7b shows temperature dependences which means that during the phase transition polymer - water hydrogen
of the fraction p in D2O solutions of the DB3 copolymer with polymer bonding becomes weaker and HDO mobility somewhat increases. This
concentrations c = 0.67 and 3 wt% recorded during gradual heating. effect is also connected to formation of polymer aggregates when
From Fig. 7b a slight shift of the LCST to smaller value (from 305 K to polymer - polymer interactions outweigh: some water molecules are
304 K) with increasing polymer concentration is clearly visible. This slowly releasing and the other are still present in the globular structure.
dependence is in agreement with abovementioned literature, as well as This result is in contrast to results obtained for other thermoresponsive
with other thermoresponsive systems based on PNIPAM [12,13], PVCL polymer systems like PNIPAm, PNIPMAm or PVME where T2 values
[47] or polyoxazolines [38]. Additionally, Fig. 7b shows that there is no were reduced at elevated temperatures [13,32,33]. In Fig. 9b time de-
change in pmax values which were noticeably affected by copolymer pendence of 1H spin-spin relaxation time T2 of HDO in D2O solution

8
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

Fig. 8. Temperature dependences of the fraction p of the PDEGMA OCH2 protons (signal “c”) with significantly reduced mobility in D2O solutions (c = 0.67 wt%) of
SP (a) and DB3 (b) during gradual heating and subsequent gradual cooling.

(c = 0.67 wt%) of the SP measured at 320 K (temperature above the which are 3 orders of magnitude shorter (T2 ≈ 8 ms at 330 K), Similar
phase transition) is presented. A continuous increase of T2 values with behavior was previously observed for PEOx homopolymers and PEOx-
time is observed showing that portion of water molecules released from based gradient or block copolymers [38,46] and in analogy with these
aggregates increases with time a little. Moreover, no sedimentation of systems we assigned “free water” to HDO molecules in solution and
SP aggregates was visually observed after the measurement. Almost the “bound water” to HDO molecules bound (confined) inside the nano-
same effect as shown in Fig. 9b was observed in PVCL [47] and PEO-b- particle. Intensities of the two T2 components show that the content of
PNIPAM [13] aqueous solutions where releasing of originally bound the “bound” water is roughly 30%. The existence of two separate T2
water with time starts directly without any induction period. components implies that exchange between “bound” and “free” water
Fig. 10a shows temperature dependence of 1H spin-spin relaxation molecules must be slow regarding T2 values. Fig. 10b shows time de-
time T2 of HDO in D2O solution (c = 0.67 wt%) of the DB3 block co- pendences of two T2 components measured in the D2O solution of the
polymer. Similarly to the SP sample measurements were done at tem- DB3 at 330 K. From Fig. 10b it follows that T2 values are virtually
peratures based on the temperature dependence of the p-fraction (see constant for both types of water molecules showing that arrangement
Fig. 7a). In comparison with SP aqueous solution a different behavior is with “free” and “bound” water is stable at least for 12 h. Comparable
observed. Firstly, much higher T2 value at temperature below the LCST behavior as shown in Fig. 10 was observed also with DB2 and DB1
(T2 ≈ 8.5 s at 295 K) was measured which means that polymer - water copolymers as well as DB3 copolymer with concentration c = 3 wt%;
interactions are much weaker in comparison with SP solution. Sec- results are depicted in Figs. S8–S10 in the Supplementary material,
ondly, T2 values significantly decrease with increasing temperature where Fig. S8 also illustrates marked differences in the fitting of ex-
which can be explained that some water molecules are hidden in na- perimental spin-spin relaxation curves at 295 K and 330 K.
noparticle structures during their formation. Thirdly, at temperatures In measurements of NMR relaxation times of polymer protons one
above phase transition, the relaxation curves were bi-exponential and has to take into account the fact that in the LCST transition region and
two T2 components were necessary to fit well experimental relaxation above the transition these measurements provide information only on
curves. This result shows the existence of two types of water at tem- the fraction of polymer protons (segments) (fraction 1 - p) which remain
perature T = 330 K. First type is “free water” with longer relaxation mobile and therefore are directly detected in high-resolution NMR
times (T2 ≈ 3 s) and second type is “bound” water, with T2 values spectra. Due to the fact that the p-fraction values of OCH3 group (signal

Fig. 9. Temperature dependence (a) and time dependence at 320 K (b) of 1H spin–spin relaxation times T2 of HDO in D2O solution (c = 0.67 wt%) of the SP.

9
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

Fig. 10. Temperature dependence (a) and time dependence at 330 K (b) of 1H spin-spin relaxation times T2 of HDO in D2O solution (c = 0.67 wt%) of the DB3.

Fig. 11. Temperature dependence (a) and time dependence at 330 K (b) of 1H spin-spin relaxation times T2 of OCH3 group “g” from PDEGMA block in D2O solution
(c = 0.67 wt%) of the DB3.

“g”) of the PDEGMA block are only slightly changing with temperature from the side chain of the PHPMA block it is possible to follow changes
(see Fig. S3b in the Supplementary material), 1H spin-spin relaxation in mobility of hydrophilic part of the copolymer during the phase
times T2 for this group were also measured. Obtained results for tem- transition. Temperature and time dependence of 1H spin-spin relaxation
perature and time dependences for the DB3 block copolymer solution time T2 of the “k” CH group in D2O solution (c = 0.67 wt%) of the DB3
(c = 0.67 wt%) are shown in Fig. 11. From Fig. 11a it follows that p- block copolymer are shown in Fig. 12. Similarly to the PDEGMA OCH3
fraction values for the OCH3 group “g” are directly connected with its “g” group, p-fraction values of the PHPMA CH “k” group (Fig. 6b) are
mobility. T2 are highest (and p-fraction is lowest) at the 300 K – tem- also directly connected with its T2 values (mobility). Fig. 12a shows
perature where the phase transition starts. This behavior is directly significantly increasing T2 values with increasing temperature. They are
connected with the changes in polymer-water interactions during the caused by temperature as well as by changes in interactions in solution,
heating processes: at room temperature hydrogen bonds between water PDEGMA block exhibits the phase transition and gives more space and
and DEGMA units result in slightly reduced mobility of OCH3 group “g”, less spherical restrictions to the PHPMA block. On the other hand,
next in pretransition region hydrogen bonding DEGMA-water becomes during first 6 h above the phase transition T2 values of the “k” CH group
weaker and mobility of the methoxy group increases and finally mo- decrease and imply some changes in the behavior of the PHPMA block,
bility of the OCH3 group decreases in nanoparticle structure (but it indicating probable changes in interactions in nanoparticles. Similar
remains still relatively mobile that its detection in high-resolution in 1H results as shown in Fig. 12 were obtained also with DB2 and DB1 co-
NMR spectra is possible). From 1H spin-spin relaxation time T2 time polymers and DB3 copolymer with concentration c = 3 wt% (cf. Figs.
dependence at 330 K (Fig. 11b), we see that T2 values remain virtually S14–S16 in the Supplementary material).
unchanged for at least 12 h showing stability of this group in formed
nanoparticles. Similarly to water, analogous behavior as shown in
Fig. 11 was observed also with DB2 and DB1 copolymers (c = 0.67 wt 3.5. Conformational changes of block copolymer: 2D 1H–1H NOESY NMR
%) and DB3 copolymer with polymer concentration c = 3 wt% (see spectra
Figs. S11–S13 in the Supplementary material).
By measuring 1H spin-spin relaxation time T2 of CH group (“k”) To understand the changes occurring during the phase separation
and obtain information on spatial proximity between proton groups of

10
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

Fig. 12. Temperature dependence (a) and time dependence at 330 K (b) of 1H spin-spin relaxation time T2 of the PHPMA group “k” in D2O solution (c = 0.67 wt%) of
the DB3.

Fig. 13. 2D NOESY spectrum of DB3 block copolymer in D2O solution (c = 3 wt%) measured at 295 K with mixing time 600 ms. On the right up there is expanded
part of the spectrum close to PHPMA “k” group cross-peaks, on the right down there is 1D slice spectrum extracted from the PHPMA “k” signal of the NOESY
spectrum.

PDEGMA and PHPMA units, 2D nuclear Overhauser effect spectroscopy The presence of all these cross-peaks implies that distances between
(NOESY) was employed. From NOESY NMR it is possible to obtain in- respective protons are smaller than 0.5 nm. PDEGMA and PHPMA units
formation on the spatial interactions of different nuclear spins in dis- which are in close proximity can be both from the same chain of the
tances to maximum 0.5 nm [56–58]. Experimental parameters (espe- copolymer, assuming a random-coil conformation of copolymer chains
cially the mixing time) used in the NOESY NMR measurements were at room temperature, and from different copolymer chains. 2D NOESY
carefully chosen based on our previous studies of other thermo- spectra with 1D slices for other temperatures are shown in the
responsive polymer systems [13,35,38] as well as studies of other au- Supplementary material (Figs. S17–S19). Fig. 14 then shows tempera-
thors [56,58]. DB3 (c = 3 wt%) D2O solution was chosen for that ture dependences of the integrated intensities of signals in 1D slices,
purpose and 2D 1H–1H NOESY NMR spectra were measured at four extracted from the signal of CH protons (“k”) of PHPMA units of the 2D
temperatures: at 295 K (starting temperature, below the transition), NOESY spectra measured with the mixing time 600 ms for D2O solution
300 K (temperature directly below the transition), 304 K (in the middle of the DB3 diblock copolymer. Dependences of the respective in-
of the transition) and 315 K (above the transition). We assumed to use tegrated intensities in 1D slices extracted from NOESY spectra mea-
the fact that even at 315 K the value of the p-fraction of the PDEGMA sured with mixing time 100 ms were similar, only the signal/noise ratio
OCH3 group “g” is low (pmax = 0.09, cf. Fig. 6b) and therefore a major in these slices was somewhat lower. This is consistent with build-up
part of these side chain end-groups is directly detected in high-resolu- curves (measured in 259 K) which show that in the range 100–600 ms
tion NMR spectra at this temperature. In NOESY spectrum measured at intensities of the crosspeaks increased with the mixing time as illu-
295 K (Fig. 13) we detected not only cross-peaks between various strated in Fig. S20 in the Supplementary material. Due to the fact that
proton groups within PDEGMA or PHPMA units, but also weaker cross- distance between CH (“k”) and CH2 (“j”) PHPMA protons cannot
peaks between side chain CH (“k”) protons of PHPMA units (signal at change, cross-peak between these groups was used as internal standard
3.93 ppm) and PDEGMA side chain protons (Fig. 13 on the right down). with integral intensity 1. From the Fig. 14 it follows that unexpectedly

11
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

Fig. 14. Chemical structure and inter-


molecular correlations in DB3 block copo-
lymer (left). Temperature dependences of
integrated intensities of various signals in
1D slices extracted from the signal of CH
protons (“k”) of PHPMA units (at 3.93 ppm)
of the NOESY NMR spectra of D2O solution
(c = 3 wt%) of the DB3 copolymer (right).

integral intensity of cross-peaks between CH (“k”) protons of PHPMA significantly different behavior of the water molecules in SP and DB
and PDEGMA side chain protons (signals “d + e + f” as well as “g”) solutions was observed. While in case of SP solution the single T2 values
increases with temperature. This result excludes typical core- to- shell increase with temperature, in case of DB solutions two types of water
formation of nanoparticles. Moreover, these findings, together with molecules –“free” and “bound” - with long and very short T2 values,
values of the p-fraction and T2 values mentioned above, suggest that respectively, were detected at temperatures above the transition. The T2
above the phase transition temperature block copolymer chains create values of the two types of water remain stable for at least 12 h at 330 K
multi-chain aggregates with partially hydrated (PHPMA, PDEGMA side and no release of the “bound” water was observed during this time. The
chains) and dehydrated (PDEGMA main chain) volumes (parts) in their existence of two T2 components indicates that the exchange between
nanostructures. Similar result was obtained also from the 2D NOESY “bound” and “free” water must be slow relative to T2 values. In case of
spectra measured with the mixing time 100 ms. the DB copolymers, T2 values were also measured for polymer side
chain groups of both PDEGMA and PHPMA blocks. In case of PDEGMA,
4. Conclusions OCH3 group was chosen due to the small values of the p-fraction. Its T2
values decrease with temperature, indicating that mobility of this group
In the present study we report on temperature behavior of ther- is reduced in the form of nanoparticles. For PHPMA side chain CH
moresponsive PDEGMA-based statistical and diblock copolymers in group, a different temperature effect is observed: T2 values increase
D2O solutions as investigated by 1H NMR methods which were com- with temperature, which is normal behavior of hydrophilic polymers in
pared with DLS measurements. Measurements of 1H NMR spectra, 1H water, but at temperature above the phase transition they decrease with
spin-spin relaxation time T2 and 2D 1H–1H NOESY spectra were used for time because mobility of the hydrophilic block is reduced due to
the characterization of structural changes on molecular level and be- polymer-polymer interactions in nanoparticles.
havior of water and copolymer molecules during the temperature-in- 2D 1H–1H NOESY NMR spectra show that in case of DB3 block
duced phase transition. In comparison with other thermoresponsive copolymer in D2O solution (c = 3 wt%) some HPMA and DEGMA units
polymers (PNIPAm, PNIPMAm, PVME, PVCL, PEOx) some important are in close contact already at room temperature. As temperature in-
differences were revealed for PDEGMA copolymers. creases in some NOESY cross-peaks between the PHPMA and PDEGMA
P[(DEGMA)–co-(PGMA)] statistical copolymer (SP) and three side chains their integral intensity increases. Based on these facts, we
(PHPMA)-b-(PDEGMA) (DB1-3) block copolymers (with constant assume that polymer chains are closer to each other in nanoparticles
length of PHPMA block: 53 monomer units and different lengths of and therefore core-to-shell morphology is improbable in these systems.
PDEGMA blocks: 26; 53 and 78 monomer units) were studied. In the It is expected that the results reported herein will assist in designing of
transition region, we observed at similar temperatures an increase in “smart” thermoresponsive polymer drug delivery systems.
the hydrodynamic sizes of the respective nanoparticles, as determined
by lower diffusion of the particles in DLS experiments and fraction p of
proton groups (units in the case of the SP copolymer) with significantly CRediT authorship contribution statement
reduced mobility in NMR experiments. Temperature dependences of
the p-fraction show that for D2O solution of the SP the phase transition Rafał Konefał: Conceptualization, Formal analysis, Investigation,
is relatively sharp with a slight difference between backbone and side Writing - original draft. Jiří Spěváček: Writing - review & editing,
chain. In contrast to the SP, the phase separation of DB in D2O solutions Supervision, Funding acquisition. Gabriela Mužíková: Investigation.
occurs in PDEGMA block backbone and in its nearest methylene group. Richard Laga: Formal analysis, Investigation, Funding acquisition.
Values of the p-fraction of PDEGMA side chain proton groups decrease
with increasing distance from the backbone. Moreover, transition
temperatures substantially depend on polymer composition in accord Declaration of Competing Interest
with DLS results and slightly on the concentration of the solution. The
temperature dependences of the p-fraction measured during gradual The authors declare that they have no known competing financial
cooling after preceding gradual heating show that for all samples the interests or personal relationships that could have appeared to influ-
phase transition is fully reversible on the molecular level. ence the work reported in this paper.
From measurements of 1H spin-spin relaxation times T2 of HDO, a

12
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

Acknowledgement behavior, Polimery/Polymers 62 (2017) 298–310 https://doi.org/10.14314/


polimery.2017.298.
[21] S. Han, M. Hagiwara, T. Ishizone, Synthesis of thermally sensitive water-soluble
Support by the Czech Science Foundation (Project 15-13853S) and polymethacrylates by living anionic polymerizations of oligo(ethylene glycol) me-
by the Ministry of Education, Youth and Sports of the Czech Republic thyl ether methacrylates, Macromolecules 36 (2003) 8312–8319, https://doi.org/
(National Sustainability Program II – project BIOCEV-FAR LQ1604 and 10.1021/ma0347971.
[22] J.F. Lutz, A. Hoth, Preparation of ideal PEG analogues with a tunable thermo-
INTER-EXCELLENCE/INTER-ACTION program – project LTAUSA18173 sensitivity by controlled radical copolymerization of 2-(2-methoxyethoxy)ethyl
is gratefully acknowledged. methacrylate and oligo(ethylene glycol) methacrylate, Macromolecules 39 (2006)
893–896, https://doi.org/10.1021/ma0517042.
[23] J.F. Lutz, Thermo-switchable materials prepared using the OEGMA-platform, Adv.
Data availability statement Mater. 23 (2011) 2237–2243, https://doi.org/10.1002/adma.201100597.
[24] M.A.C. Stuart, W.T.S. Huck, J. Genzer, M. Müller, C. Ober, M. Stamm,
The raw/processed data required to reproduce these findings cannot G.B. Sukhorukov, I. Szleifer, V.V. Tsukruk, M. Urban, F. Winnik, S. Zauscher,
I. Luzinov, S. Minko, Emerging applications of stimuli-responsive polymer mate-
be shared at this time as the data also form part of an ongoing study.
rials, Nat. Mater. 9 (2010) 101–113, https://doi.org/10.1038/nmat2614.
[25] C. Lu, M.W. Urban, Stimuli-responsive polymer nano-science: shape anisotropy,
Appendix A. Supplementary material responsiveness, applications, Prog. Polym. Sci. 78 (2018) 24–46, https://doi.org/
10.1016/j.progpolymsci.2017.07.005.
[26] Z.L. Yao, K.C. Tam, Temperature induced micellization and aggregation of bio-
Supplementary data to this article can be found online at https:// compatible poly (oligo(ethylene glycol)methyl ether methacrylate) block copo-
doi.org/10.1016/j.eurpolymj.2020.109488. lymer analogs in aqueous solutions, Polymer (Guildf). 53 (2012) 3446–3453,
https://doi.org/10.1016/j.polymer.2012.06.002.
[27] K. Gawlitza, A. Radulescu, R. Von Klitzing, S. Wellert, On the structure of bio-
References compatible, thermoresponsive poly(ethylene glycol) microgels, Polymer (Guildf).
55 (2014) 6717–6724, https://doi.org/10.1016/j.polymer.2014.10.069.
[1] M.W. Urban, Stratification, stimuli-responsiveness, self-healing, and signaling in [28] F. Zuppardi, F.R. Chiacchio, R. Sammarco, M. Malinconico, G. Gomez d’Ayala,
polymer networks, Prog. Polym. Sci. 34 (2009) 679–687, https://doi.org/10.1016/ P. Cerruti, Fluorinated oligo(ethylene glycol) methacrylate-based copolymers:
j.progpolymsci.2009.03.004. tuning of self assembly properties and relationship with rheological behavior,
[2] F. Liu, M.W. Urban, Recent advances and challenges in designing stimuli-responsive Polymer (Guildf) 112 (2017) 169–179, https://doi.org/10.1016/j.polymer.2017.
polymers, Prog. Polym. Sci. 35 (2010) 3–23, https://doi.org/10.1016/j. 01.080.
progpolymsci.2009.10.002. [29] W. Yuan, J. Wang, Oligo(ethylene glycol) and quaternary ammonium-based block
[3] A.K. Bajpai, J. Bajpai, R. Saini, R. Gupta, Responsive polymers in biology and copolymer micelles: from tunable thermoresponse to dual salt response, RSC Adv. 4
technology, Polym. Rev. 51 (2011) 53–97, https://doi.org/10.1080/15583724. (2014) 38855–38858, https://doi.org/10.1039/c4ra05096g.
2010.537798. [30] Y. Guo, X. Dong, W. Ruan, Y. Shang, H. Liu, A thermo-sensitive OEGMA-based
[4] A. Bordat, T. Boissenot, J. Nicolas, N. Tsapis, Thermoresponsive polymer nano- polymer: synthesis, characterization and interactions with surfactants in aqueous
carriers for biomedical applications, Adv. Drug Deliv. Rev. 138 (2019) 167–192, solutions with and without salt, Colloid Polym. Sci. 295 (2017) 327–340, https://
https://doi.org/10.1016/j.addr.2018.10.005. doi.org/10.1007/s00396-016-4006-4.
[5] P. Zarrintaj, M. Jouyandeh, M.R. Ganjali, B.S. Hadavand, M. Mozafari, S.S. Sheiko, [31] Y. Maeda, T. Nakamura, I. Ikeda, Hydration and phase behavior of poly (N-vi-
M. Vatankhah-Varnoosfaderani, T.J. Gutiérrez, M.R. Saeb, Thermo-sensitive poly- nylcaprolactam) and poly (N-vinylpyrrolidone) in water, Macromolecules 32
mers in medicine: a review, Eur. Polym. J. 117 (2019) 402–423, https://doi.org/10. (2002) 217–222, https://doi.org/10.1021/ma011034+.
1016/j.eurpolymj.2019.05.024. [32] J. Spěváček, NMR investigations of phase transition in aqueous polymer solutions
[6] M. Wei, Y. Gao, X. Li, M.J. Serpe, Stimuli-responsive polymers and their applica- and gels, Curr. Opin. Colloid Interface Sci. 14 (2009) 184–191, https://doi.org/10.
tions, Polym. Chem. 8 (2017) 127–143, https://doi.org/10.1039/C6PY01585A. 1016/j.cocis.2008.10.003.
[7] D. Schmaljohann, Thermo- and pH-responsive polymers in drug delivery, Adv. Drug [33] J. Spěváček, Application of NMR spectroscopy to study thermoresponsive polymers,
Deliv. Rev. 58 (2006) 1655–1670, https://doi.org/10.1016/j.addr.2006.09.020. in: T.K. Khutoryanskiy, V.V. Georgiou (Eds.), Temp. Polym. Chem. Prop. Appl. first
[8] M.A. Ward, T.K. Georgiou, Thermoresponsive polymers for biomedical applications, ed., John Wiley & Sons Ltd., 2018, pp. 225–247.
Polymers (Basel) 3 (2011) 1215–1242, https://doi.org/10.3390/polym3031215. [34] C.H. Hofmann, M. Schönhoff, Dynamics and distribution of aromatic model drugs in
[9] V.R. De La Rosa, Poly(2-oxazoline)s as materials for biomedical applications, J. the phase transition of thermoreversible poly(N-isopropylacrylamide) in solution,
Mater. Sci. Mater. Med. 25 (2014) 1211–1225, https://doi.org/10.1007/s10856- Colloid Polym. Sci. 290 (2012) 689–698, https://doi.org/10.1007/s00396-011-
013-5034-y. 2577-7.
[10] P.-F. Cao, J.D. Mangadlao, R.C. Advincula, Stimuli-responsive polymers and their [35] J. Spěváček, R. Konefał, E. Čadova, NMR study of thermoresponsive block copo-
potential applications in oil-gas industry, Polym. Rev. 55 (2015) 706–733, https:// lymer in aqueous solution, Macromol. Chem. Phys. 217 (2016) 1370–1375, https://
doi.org/10.1080/15583724.2015.1040553. doi.org/10.1002/macp.201600025.
[11] V. Aseyev, H. Tenhu, F.M. Winnik, Non-ionic thermoresponsive polymers in water, [36] C. Zhang, H. Peng, A.K. Whittaker, NMR investigation of effect of dissolved salts on
Adv. Polym. Sci. 242 (2010) 29–89, https://doi.org/10.1007/12_2010_57. the thermoresponsive behavior of oligo(ethylene glycol)-methacrylate-based poly-
[12] A. Halperin, M. Kroeger, F.M. Winnik, Poly(N-isopropylacrylamide) phase dia- mers, J. Polym. Sci. Part A Polym. Chem. 52 (2014) 2375–2385, https://doi.org/10.
grams: fifty years of research, Angew. Chemie - Int. Ed. 54 (2015) 15342–15367, 1002/pola.27252.
https://doi.org/10.1002/anie.201506663. [37] J. Spěváček, L. Hanyková, J. Labuta, Behavior of water during temperature-induced
[13] J. Spěváček, R. Konefał, J. Dybal, E. Čadová, J. Kovářová, Thermoresponsive be- phase separation in poly(vinyl methyl ether) aqueous solutions. NMR and optical
havior of block copolymers of PEO and PNIPAm with different architecture in microscopy study, Macromolecules 44 (2011) 2149–2153, https://doi.org/10.
aqueous solutions: a study by NMR, FTIR, DSC and quantum-chemical calculations, 1021/ma200010h.
Eur. Polym. J. 94 (2017) 471–483, https://doi.org/10.1016/j.eurpolymj.2017.07. [38] R. Konefał, J. Spěváček, P. Černoch, Thermoresponsive poly(2-oxazoline) homo-
034. polymers and copolymers in aqueous solutions studied by NMR spectroscopy and
[14] M. Rusu, S. Wohlrab, D. Kuckling, H. Möhwald, M. Schönhoff, Coil-to-globule dynamic light scattering, Eur. Polym. J. 100 (2018) 241–252, https://doi.org/10.
transition of PNIPAM graft copolymers with charged side chains: a 1H and 2H NMR 1016/j.eurpolymj.2018.01.019.
and spin relaxation study, Macromolecules 39 (2006) 7358–7363, https://doi.org/ [39] H. Kouřilová, J. Šťastná, L. Hanyková, Z. Sedláková, J. Spěváček, 1H NMR study of
10.1021/ma060831a. temperature-induced phase separation in solutions of poly(N-iso-
[15] H.G. Schild, Poly(N-isopropylacrylamide): experiment, theory and application, propylmethacrylamide-co-acrylamide) copolymers, Eur. Polym. J. 46 (2010)
Prog. Polym. Sci. 17 (1992) 163–249, https://doi.org/10.1016/0079-6700(92) 1299–1306, https://doi.org/10.1016/j.eurpolymj.2010.03.006.
90023-R. [40] K. Ulbrich, V. Šubr, J. Strohalm, D. Plocová, M. Jelínková, B. Říhová, Polymeric
[16] D. Roy, J.N. Cambre, B.S. Sumerlin, Future perspectives and recent advances in drugs based on conjugates of synthetic and natural macromolecules. I. Synthesis
stimuli-responsive materials, Prog. Polym. Sci. 35 (2010) 278–301, https://doi.org/ and physico-chemical characterisation, J. Control. Release 64 (2000) 63–79,
10.1016/j.progpolymsci.2009.10.008. https://doi.org/10.1016/S0168-3659(99)00141-8.
[17] Debashish Roy, William L.A. Brooks, Brent S. Sumerlin, New directions in ther- [41] V. Šubr, L. Kostka, J. Strohalm, T. Etrych, K. Ulbrich, Synthesis of well-defined
moresponsive polymers, Chem. Soc. Rev. 42 (17) (2013) 7214, https://doi.org/10. semitelechelic poly[N-(2-hydroxypropyl) methacrylamide] polymers with func-
1039/c3cs35499g. tional group at the α-end of the polymer chain by RAFT polymerization,
[18] R. Hoogenboom, H. Schlaad, Thermoresponsive poly(2-oxazoline)s, polypeptoids, Macromolecules 46 (2013) 2100–2108, https://doi.org/10.1021/ma400042u.
and polypeptides, Polym. Chem. 8 (2017) 24–40, https://doi.org/10.1039/ [42] R. Laga, O. Janoušková, K. Ulbrich, R. Pola, J. Blažková, S.K. Filippov, T. Etrych,
C6PY01320A. M. Pechar, Thermoresponsive polymer micelles as potential nanosized cancero-
[19] O. Sedláček, P. Černoch, J. Kučka, R. Konefal, P. Štěpánek, M. Vetrík, T.P. Lodge, statics, Biomacromolecules 16 (2015) 2493–2505, https://doi.org/10.1021/acs.
M. Hrubý, Thermoresponsive polymers for nuclear medicine: which polymer is the biomac.5b00764.
best? Langmuir 32 (2016) 6115–6122, https://doi.org/10.1021/acs.langmuir. [43] E.D. Farrar, T.C. Becker, Pulse and Fourier Transform NMR, Academic Press, New
6b01527. York, 1971.
[20] D. Szweda, R. Szweda, A. Dworak, B. Trzebicka, Thermoresponsive poly[oligo [44] Q. Zhang, C. Weber, U.S. Schubert, R. Hoogenboom, Thermoresponsive polymers
(ethylene glycol) methacrylate]s and their bioconjugates - Synthesis and solution with lower critical solution temperature: from fundamental aspects and measuring

13
R. Konefał, et al. European Polymer Journal 124 (2020) 109488

techniques to recommended turbidimetry conditions, Mater. Horiz. 4 (2017) [52] V.O. Aseyev, H. Tenhu, F.M. Winnik, Temperature dependence of the colloidal
109–116, https://doi.org/10.1039/C7MH00016B. stability of neutral amphiphilic polymers in water, Adv. Polym. Sci. 196 (2006)
[45] L. Loukotová, A. Bogomolova, R. Konefal, M. Špírková, P. Štěpánek, M. Hrubý, 1–85, https://doi.org/10.1007/12_052.
Hybrid κ-carrageenan-based polymers showing “schizophrenic” lower and upper [53] C. Fu, S. Herbst, C. Zhang, A.K. Whittaker, Polymeric 19F MRI agents responsive to
critical solution temperatures and potassium responsiveness, Carbohydr. Polym. reactive oxygen species, Polym. Chem. 8 (2017) 4585–4595, https://doi.org/10.
210 (2019) 26–37, https://doi.org/10.1016/j.carbpol.2019.01.050. 1039/c7py00986k.
[46] R. Konefał, J. Spěváček, E. Jäger, S. Petrova, Thermoresponsive behaviour of ter- [54] S.Y. Park, S.H. Cho, S.H. Yuk, M.S. Jhon, Characterization of temperature-induced
polymers containing poly(ethylene oxide), poly(2-ethyl-2-oxazoline) and poly(ε- phase transition of polymer complex composed of poly(N, N-dimethylamino)ethyl
caprolactone) blocks in aqueous solutions: an NMR study, Colloid Polym. Sci. 294 methacrylate and poly(ethylacrylamide) by -NMR relaxation time measurement,
(2016) 1717–1726, https://doi.org/10.1007/s00396-016-3930-7. Eur. Polym. J. 37 (2001) 1785–1790, https://doi.org/10.1016/S0014-3057(01)
[47] J. Spěváček, J. Dybal, L. Starovoytova, A. Zhigunov, Z. Sedláková, Temperature- 00049-0.
induced phase separation and hydration in poly(N-vinylcaprolactam) aqueous so- [55] P.Y. Ghi, D.J.T. Hill, A.K. Whittaker, 1H NMR study of the states of water in
lutions: a study by NMR and IR spectroscopy, SAXS, and quantum-chemical cal- equilibrium poly(HEMA-co-THFMA) hydrogels, Biomacromolecules 3 (2002)
culations, Soft Matter 8 (2012) 6110–6119, https://doi.org/10.1039/c2sm25432h. 991–997, https://doi.org/10.1021/bm0200332.
[48] W. Hiller, N. Engelhardt, A.L. Kampmann, P. Degen, R. Weberskirch, Micellization [56] C. Zhang, H. Peng, S. Puttick, J. Reid, S. Bernardi, D.J. Searles, A.K. Whittaker,
and mobility of Amphiphilic Poly(2-oxazoline) based block copolymers character- Conformation of hydrophobically modified thermoresponsive poly(OEGMA-co-
ized by 1H NMR spectroscopy, Macromolecules 48 (2015) 4032–4045, https://doi. TFEA) across the LCST revealed by NMR and molecular dynamics studies,
org/10.1021/acs.macromol.5b00149. Macromolecules 48 (2015) 3310–3317, https://doi.org/10.1021/acs.macromol.
[49] S.-I. Yamamoto, J. Pietrasik, K. Matyjaszewski, The effect of structure on the 5b00641.
thermoresponsive nature of well-defined poly(oligo(ethylene oxide) methacrylates) [57] C. Zhang, R.J.P. Sanchez, C. Fu, R. Clayden-Zabik, H. Peng, K. Kempe,
synthesized by ATRP, J. Polym. Sci. Part A Polym. Chem. 46 (2008) 194–202, A.K. Whittaker, Importance of thermally induced aggregation on 19 F magnetic
https://doi.org/10.1002/pola.22371. resonance imaging of perfluoropolyether-based comb-shaped poly(2-oxazoline)s,
[50] X. Han, X. Zhang, H. Zhu, Q. Yin, H. Liu, Y. Hu, Effect of composition of PDMAEMA- Biomacromolecules. 20 (2019) 365–374, https://doi.org/10.1021/acs.biomac.
b-PAA block copolymers on their pH- and temperature-responsive behaviors, 8b01549.
Langmuir 29 (2013) 1024–1034, https://doi.org/10.1021/la3036874. [58] C. Zhang, H. Peng, W. Li, L. Liu, S. Puttick, J. Reid, S. Bernardi, D.J. Searles,
[51] G. Pasparakis, C. Alexander, Sweet talking double hydrophilic block copolymer A. Zhang, A.K. Whittaker, Conformation transitions of thermoresponsive den-
vesicles, Angew. Chemie - Int. Ed. 47 (2008) 4847–4850, https://doi.org/10.1002/ dronized polymers across the lower critical solution temperature, Macromolecules
anie.200801098. 49 (2016) 900–908, https://doi.org/10.1021/acs.macromol.5b02414.

14

You might also like