S.
Petrova
*,
C. G.
Venturini
*,
A.
Jäger
,
E.
Jäger
,
M.
Hrubý
,
E.
Pavlova
and
P.
Štěpánek
Institute of Macromolecular Chemistry, Heyrovsky Sq. 2, 162 06 Prague 6, Czech Republic. E-mail: petrova@imc.cas.cz; cgventurini@gmail.com; Tel: +420 296 809 322
First published on 15th July 2015
Nonlinear amphiphilic block copolymer architectures with precisely controlled structures bring new challenges to biomedical materials research. The paper describes the straightforward synthesis of new “snake tongue“ Y-shaped terpolymers containing poly(ethylene oxide) (PEO), poly(2-ethyl-2-oxazoline) (PEtOx) and poly(ε-caprolactone) (PCL) blocks into structure [AB(C)2] (herein referred to as [MPEO44-b-PEtOx252-b-(PCL)2×44], [MPEO44-b-PEtOx252-b-(PCL)2×87], [MPEO44-b-PEtOx252-b-(PCL)2×131]). A series of well-defined Y-shaped terpolymers were successfully synthesised by a combination of living cationic and anionic ring-opening polymerization (ROP). The selected Y-shaped [MPEO44-b-PEtOx252-b-(PCL)2×44] terpolymer self-assembly was characterised in detail by static and dynamic light scattering, nanoparticle tracking analysis and cryo-transmission electron microscopy. The physico-chemical properties as well as the molecular architecture effect on the self-assembled structures and on the LCST were compared with the Y-shaped [MPEO44-b-PEtOx252-b-(PCL)2×87] and the [MPEO44-b-PEtOx252-b-(PCL)2×131] terpolymers. The results indicated a temperature-induced aggregation with an LCST between 60–63 °C for the [MPEO44-b-PEtOx252-b-(PCL)2×44], at 60 °C for the [MPEO44-b-PEtOx252-b-(PCL)2×87] and between 45–50 °C for the [MPEO44-b-PEtOx252-b-(PCL)2×131] with significant differences in the supramolecular self-assembly behaviour compared with the analogous linear structure, clearly indicating the crucial effect of the molecular architecture. Furthermore, the increase of the molecular weight fraction of the hydrophobic block on the Y-shaped triblock terpolymers likely induced a decrease of the LCST.
Star-shaped block copolymers consisting of at least three linear polymeric arms with a radial arrangement around a central molecular fragment (core)8,9 are usually prepared by the “arm-first” or “core-first” methods. The “arm-first” approach involves the construction of polymer arms on a macroinitiator that contains a precise number of reactive sites.10,11 The “core-first” approach utilises multifunctional low-molecular-weight initiators, allowing for the synthesis of block copolymer chains.12 A method based on difunctional monomers is mentioned in the literature as a third approach for the synthesis of star-shaped copolymers.6 However, this method does not allow for strict control of the number of arms.
It is well known that amphiphilic star-shaped copolymers can easily self-assemble in aqueous media to form nanosized unimolecular micelles containing hydrophobic cores surrounded by hydrophilic shells.13,14 These micellar systems are of great interest for medical uses such as the construction of micellar drug delivery systems.9,15–19 Significant differences in the physicochemical properties of star-shaped copolymers compared with their linear analogues can be observed, such as smaller hydrodynamic volume and, radius of gyration and low melt and solution viscosities, which are beneficial to drug loading and delivery.20–24 It has been shown that linear amphiphilic copolymers have limited applications in drug delivery because they suffer from an initial burst release effect. Especially in systems with non-covalently incorporated drugs, the micellar stability and drug release are difficult to control.25,26 Therefore, various star-shaped copolymers with varying arm numbers and chemical compositions have received considerable attention because of the unique properties and advantages that they possess.27–29 Y-shaped copolymers (typically referred to star copolymers) are another interesting vehicle for drug delivery because they exhibit distinct a micellization behaviour (special stability) compared with the amphiphilic copolymers with a linear architecture.30–32 Furthermore, star copolymers bearing distinct polymeric arms have shown dynamic morphological changes (e.g., micelle-to-unimer transition under certain conditions) because the constituting polymeric components could be designed to be individually responsive to external stimuli such as pH, temperature and solvent.7,33–36 Moreover, a particular subject of even greater interest is the study of biocompatible thermo-responsive self-assembled polymer micelles of amphiphilic, double-hydrophilic and hydrophobic species of star copolymers. It should be noted that the number of such studies is quite limited, and more work is greatly needed to understand the particular characteristics involved in the micellar behaviour of such polymer systems.
In this paper, we describe the synthesis and the study of the self-assembly properties of new “snake tongue” Y-shaped terpolymers based on poly(ethylene oxide) (PEO), poly(2-ethyl-2-oxazoline) (PEtOx) and poly(ε-caprolactone) (PCL) with the general architecture [PEO-b-PEtOx-b-(PCL)2] (Scheme 1). The newly synthesised Y-shaped terpolymers combine environmentally friendly blocks with possible applications in biomedicine. PEtOx was chosen because it exhibits similar chemical and biological properties to PEO;37–39 both polymers are water-soluble and non-toxic.40–43 PEO and PEtOx can be eliminated from the human body if they possess a low enough molar mass. Furthermore, PEtOx in an aqueous solution exhibits a lower critical solution temperature (LCST).44–47 The LCST of PEtOx is ∼61–66.5 °C and strongly depends on the polymer molecular weight (20–500 kDa) and polymer concentration.48,49 PCL is a hydrophobic, nontoxic, biocompatible and fully biodegradable aliphatic polyester.50 To the best of our knowledge, this is the first time that these three blocks were combined in a nonlinear architecture using the “arm-first” method and their supramolecular self-assembly behaviour was compared with analogous blocks of linear architecture and identical weight ratios and molecular weights.
1H NMR, δ (TMS, ppm): 2.45 (s, 3H, CH3–), 3.38 (s, 3H, –OCH3), 3.65 (m, 4H, –OCH2CH2–), 4.16 (t, 2H,–CH2O(SO2)), 7.36–7.33 (d, 2H, ArH), 7.82–7.8 (d, 2H, ArH), Mn(NMR) = 2200 g mol−1:
Mn(SEC) = 2355 g mol−1, Mw/Mn(SEC) = 1.15. |
Here, a [MPEO-b-PEtOx(OH)2] double-hydrophilic block copolymer was synthesised by CROP of EtOx using ω-tosyl-MPEO as a macroinitiator and subsequent in situ end-capping by diethanolamine of the living oxazolinium species converted into ω,ω′-dihydroxyl groups, (compound 3, in Scheme 1). The double-hydrophilic block copolymer was obtained with a MPEO molecular weight of approximately 2200 g mol−1 and a PEtOx block molecular weight of approximately 23550 g mol−1. After purification, the structure of the double-hydrophilic block copolymer was confirmed by 1H NMR and FT-IR spectroscopy. The Mn of [MPEO-b-PEtOx(OH)2] was determined by 1H NMR spectroscopy and SEC. The 1H NMR spectrum of the diblock copolymer (Fig. 1) showed the characteristic signals of the protons belonging to the ethylene oxide (EO) and EtOx repeat units. The methylene protons of the EO repeat units marked with b were observed at δ = 3.64 ppm. The spectrum also showed a broad singlet signal observed at δ = 3.46 ppm and labelled as d that corresponded to the chemical shifts of the protons in –N–CH2–CH2 from the EtOx repeat units. Furthermore, the spectrum detected signals corresponding to the pendant group of the main polymer chain of PEtOx at δ = 2.40 ppm (e) that was attributed to the methylene protons of N–C(O)–CH2–CH3 and another labelled as f at δ = 1.12 ppm that was assigned to the methyl group of N–C(O)–CH2–CH3. The 1H NMR spectrum also showed a signal identified as c at δ = 3.70 from the last monomer unit of PEO that was connected to the PEtOx polymer chain. A triplet signal at δ = 2.84 ppm, which corresponded to the four protons from the N–CH2–CH2–HO end-capped group (g and g′) that formed after termination by diethanolamine, was observed. Nevertheless, the signals of the four –CH2–OH (h and h′) hydrogen nuclei were not detectable because they were hidden by the signal of the CH2O units of the PEO repeat unit at δ = 3.64 ppm. Furthermore, the singlet signal at δ = 3.39 ppm attributed to CH3–O– was not detected in the spectrum because it was hidden by the peak of the –N–CH2–CH2– units of the PEtOx repeat unit at δ = 3.46 ppm.
![]() | ||
Fig. 1 1H NMR spectrum of the [MPEO-b-PEtOx(OH)2] diblock copolymer in CDCl3 (1, Table 1). |
Based on the molecular weight of the initiating ω-tosyl-MPEO macroinitiator, the number-average molecular weight Mn(NMR) of the diblock copolymer was calculated by eqn (1).
Mn(NMR)[MPEO-b-PEtOx(OH)2] = [(Id/4)/(Ib/4)] × DPMPEO × 99 + Mn(NMR) (macroinitiator) | (1) |
No | Sample | M n a, (theor.) | M n b, (NMR) | M n c, (SEC) | M w/Mnd, (SEC) | Weight fraction MPEO | Weight fraction PEtOx | Weight fraction PCL |
---|---|---|---|---|---|---|---|---|
a M n = [M]o/[I]o × 99 + Mn(NMR) (macroinitiator) and Mn = [M]o/[I]o × 114 + Mn (double-hydrophilic block copolymer). b M n was calculated by 1H NMR spectroscopy according to eqn (1) and (2). c M n values relative to linear PS standards. d M w/Mn values relative to linear PS standards. | ||||||||
1 | [MPEO44-b-PEtOx252-(OH)2] | 27![]() |
25![]() |
30![]() |
1.24 | |||
2 | [MPEO44-b-PEtOx252-b-(PCL)2×44] | 37![]() |
39![]() |
44![]() |
1.38 | 0.052 | 0.68 | 0.27 |
3 | [MPEO44-b-PEtOx252-b-(PCL)2×87] | 47![]() |
48![]() |
50![]() |
1.39 | 0.04 | 0.53 | 0.42 |
4 | [MPEO44-b-PEtOx252-b-(PCL)2×131] | 57![]() |
59![]() |
63![]() |
1.33 | 0.034 | 0.44 | 0.53 |
The structure of the obtained diblock copolymer was also confirmed by FT-IR spectroscopy (Fig. 2). The spectrum showed the absorption peaks characteristic of both components (MPEO and PEtOx): at 1625 cm−1 corresponding to the -amide bond (CO stretching) of the 2-ethyl-2-oxazoline repeat units of the PEtOx block; at 1421 cm−1 (δ CH from the CH3 of POx); at 1197 cm−1 attributed to the -ether bond (C–O–C stretching) of the EO repeat units of the PEO backbone; at 1051 cm−1 (C–N stretching from POx); and at 2976 and 2885 cm−1 corresponding to the –C–H vibrations typical for both monomer units. The presence of hydroxyl groups at the ω and ω′ positions in the double-hydrophilic block copolymer structure was proven by the appearance of a broad and intense absorption band with a maximum at 3442 cm−1. Furthermore, a low-intensity absorption band at 3739 cm−1 indicating the existence of intramolecular hydrogen bonding between the hydroxyl groups, was observed.
The SEC analysis of the [MPEO-b-PEtOx(OH)2] diblock copolymer (bold line, Fig. 3) (1, Table 1) showed a monomodal and narrow molecular weight distribution. The main molecular characteristics of the double-hydrophilic block copolymer are listed in Table 1. The obtained data reported in Table 1 confirmed that the polymerisation was controlled with a low polydispersity index and an experimental molecular weight dictated by the monomer-to-initiator ratio and monomer conversion.
![]() | ||
Fig. 3 SEC chromatograms in DMF of the [MPEO-b-PEtOx(OH)2] diblock copolymer (bold line) (1, Table 1) and the Y-shaped [MPEO-b-PEtOx-b-(PCL)2] terpolymers (dotted line) (2, Table 1), (dashed line) (3, Table 1), (dash dotted line) (4, Table 1). |
Mn(NMR)[MPEO-b-PEtOx-b-(PCL)2] = [(Il/2)/(Ib/4)] × DPMPEO × 114 + Mn(NMR)[MPEO-b-PEtOx(OH)2] | (2) |
![]() | ||
Fig. 4 1H NMR spectrum of the Y-shaped [MPEO-b-PEtOx-b-(PCL)2] terpolymer in CDCl3 (2, Table 1). |
The molecular weights and polydispersity indices of the synthesised Y-shaped [MPEO-b-PEtOx-b-(PCL)2] terpolymers were determined by SEC. As evidenced the SEC curves were overlapped (see the dotted, dashed and dash dotted lines, Fig. 3). The SEC traces revealed relatively narrow molecular weight distributions commonly observed for “living” controlled polymerisation techniques. The SEC profiles of the Y-shaped terpolymers were monomodal, and the elution times were shifted towards lower values corresponding to higher molecular weights compared with those of the diblock copolymer macroinitiator (bold line, Fig. 3). A slight asymmetry was observed at longer elution time in the SEC curve (dashed line, Fig. 3), which strongly suggested that some unreacted double-hydrophilic block copolymer was present. The main characteristic molecular features of the Y-shaped [AB(C)2] terpolymers are listed in Table 1.
The DLS data showed a single population of nanoparticles for each of the four solution concentrations. The size distributions of the particles were narrow, as indicated by NTA that displayed a span value of 0.7 (eqn (4) see ESI†). The RH values (Table 2) were calculated using eqn (1) (see ESI†) through the diffusion coefficient values obtained from the slope of the linear fits of the relaxation rate dependence on q2 (Fig. S1†). The RG parameter was calculated from the linear fit of the angular dependence of Kc/R(θ) based on the data in Fig. 6.
![]() | ||
Fig. 6 Dependence of Kc/R(θ) on q2 for [MPEO44-b-PEtOx252-b-(PCL)2×44] NPs at solution concentrations of (![]() ![]() ![]() ![]() |
The diffusion coefficient (D) values and the RG/RH ratios were similar (Table 2) for the different concentrations (0.5 to 2.0 mg mL−1). At 25 °C, the D values were constant regardless of the angle of observation, which was in agreement with the scattering contribution of spherical structures.64 The RG/RH ratio, a sensitive structural parameter of the particles in solution,65,66 agreed with a concept of spherical micellar shape in the [MPEO44-b-PEtOx252-b-(PCL)2×44] NP solutions at room temperature. In our previous results,51 a linear triblock terpolymer (MPEO44-b-PEtOx263-b-PCL87) presented similar RH and RG values. Therefore, the resulting assemblies in aqueous media are expected to be related to diffuse core-shell-like or soft ball structures since the hydrophilic (MPEO-b-PEtOx diblock copolymer) corona is much longer then the hydrophobic PCL core. In comparison to the size of the particle shell a small and diffuse core might be expected. Such particles are characterised by high amounts of water entrapped inside the assemblies, lower densities and RG/RH values similar to those obtained in our experiments (RG/RH ∼ 1.0).51,67–71
T (°C) | R H (nm) | R G (nm)a | R G/RH | M w(NP) (107 g mol−1)a | ρ (g mL−1)b | N agg c |
---|---|---|---|---|---|---|
a All the SLS data were obtained from the data in Fig. S6 and 7. b NP density (ρ) was calculated by eqn (3) see ESI. c The aggregation number (Nagg) was calculated by Nagg = Mw(NP)/Mw(unimer). | ||||||
5 | 73 | 71 | 0.97 | 1.55 | 0.016 | 386 |
15 | 73 | 71 | 0.97 | 1.56 | 0.016 | 389 |
25 | 70 | 68 | 0.97 | 1.53 | 0.018 | 382 |
40 | 67 | 62 | 0.92 | 1.40 | 0.019 | 349 |
45 | 67 | 62 | 0.92 | 1.64 | 0.022 | 410 |
50 | 69 | 67 | 0.97 | 1.85 | 0.022 | 462 |
55 | 72 | 81 | 1.12 | 2.11 | 0.022 | 527 |
60 | 72 | 80 | 1.11 | 4.72 | 0.050 | 1177 |
62 | 82 | 134 | 1.63 | 8.73 | 0.063 | 2179 |
For the [MPEO44-b-PEtOx252-b-(PCL)2×44] the DLS data (Fig. S3 and 4†) showed only one population in all of the concentrations up to 55 °C. The zeta potential values for the nanoparticle solutions (2 mg mL−1) varied from −2 to −6 mV (5–70 °C) showing that the values did not change as a function of the temperature. The RG/RH ratio values (0.92–0.97) indicated the presence of structures corresponding to spherical nanoparticles at temperatures up to 50 °C. In addition, nearly constant molecular weight values (1.53–1.85 × 107 g mol−1) and aggregation numbers (349–462) suggested that no particle aggregation was observed in this temperature range (5 to 50 °C). For the [MPEO44-b-PEtOx252-b-(PCL)2×87] (Table 4) the DLS data showed only one scattering population for all temperatures up to 60 °C (data not shown). The zeta potential values for the nanoparticle solutions varied from −4 to −8 mV (5–60 °C) and as for the [MPEO44-b-PEtOx252-b-(PCL)2×44] did not change as a function of the temperature. According to the RG/RH ratio values (0.86–0.97) the presence of structures corresponding to spherical nanoparticles at temperatures up to 50 °C are expected. In addition, nearly constant molecular weight values (0.60–0.74 × 107 g mol−1) and aggregation numbers (118–132) suggested that no particle aggregation was observed in this temperature range (5 to 50 °C). For the last synthesised [MPEO44-b-PEtOx252-b-(PCL)2×131] terpolymer (Table 5) the DLS data showed only one population in all of the concentrations up to 40 °C. The zeta potential values for the nanoparticle solutions varied from −2 to −10 mV (5–50 °C). For these terpolymers the RG/RH ratio values (0.97–1.06) indicated the presence of structures corresponding to spherical nanoparticles at temperatures up to 40 °C. The molecular weight values (0.13–0.21 × 107 g mol−1) and aggregation numbers (20–33) suggested no particle aggregation in the temperature range of 5 to 40 °C.
T (°C) | R H (nm) | R G (nm)a | R G/RH | M w(NP) (107 g mol−1)a | ρ (g mL−1)b | N agg c |
---|---|---|---|---|---|---|
a All the SLS data were obtained from the data in Fig. S6 and 7. b NP density (ρ) was calculated by eqn (3) see ESI. c The aggregation number (Nagg) was calculated by Nagg = Mw(NP)/Mw(unimer). d Sample precipitation. | ||||||
5 | 33 | 32 | 0.97 | 0.67 | 0.073 | 132 |
15 | 37 | 32 | 0.86 | 0.72 | 0.061 | 142 |
25 | 37 | 34 | 0.92 | 0.61 | 0.048 | 120 |
40 | 37 | 32 | 0.86 | 0.60 | 0.047 | 118 |
45 | 37 | 36 | 0.97 | 0.63 | 0.050 | 124 |
50 | 38 | 35 | 0.92 | 0.65 | 0.047 | 128 |
55 | 38 | 41 | 1.08 | 0.78 | 0.052 | 154 |
60 | 45 | 48 | 1.07 | 1.15 | 0.051 | 227 |
62d | — | — | — | — | — | — |
T (°C) | R H (nm) | R G (nm)a | R G/RH | M w(NP) (107 g mol−1)a | ρ (g mL−1)b | N agg c |
---|---|---|---|---|---|---|
a All the SLS data were obtained from the data in Fig. S6 and 7. b NP density (ρ) was calculated by eqn (3) see ESI. c The aggregation number (Nagg) was calculated by Nagg = Mw(NP)/Mw(unimer). d Sample precipitation. | ||||||
5 | 63 | 67 | 1.06 | 0.13 | 0.0021 | 20 |
15 | 69 | 67 | 0.97 | 0.14 | 0.0017 | 22 |
25 | 69 | 73 | 1.06 | 0.16 | 0.0020 | 25 |
40 | 77 | 81 | 1.05 | 0.21 | 0.0018 | 33 |
45 | 87 | 102 | 1.17 | 0.29 | 0.0017 | 45 |
50 | 115 | 132 | 1.15 | 0.32 | 0.0008 | 50 |
55d | — | — | — | — | — | — |
60d | — | — | — | — | — | — |
62d | — | — | — | — | — | — |
At temperatures below the LCST, hydrogen bonds between the polymer carbonyl group and the water hydrogens were abundant,72 and the NPs were swollen by water. The density values confirmed this swollen state, which was also previously verified for linear MPEO44-b-PEtOx263-b-PCL87 NPs.51
When the temperature approached 55 °C for [MPEO44-b-PEtOx252-b-(PCL)2×44] and [MPEO44-b-PEtOx252-b-(PCL)2×87] as well as 40 °C for [MPEO44-b-PEtOx252-b-(PCL)2×131], a slight increase in the values of the RG/RH ratio, molecular weight and aggregation number were observed for the nanoparticles (Tables 3–5). The temperature dependence behaviour observed for the 3 terpolymer NPs was related to the thermodynamic effects of the LCST on the PEtOx block.66 This result was observed because with the increase in temperature, the hydrogen bonds between the carbonyl (CO) group of the PEtOx block and the water hydrogens were weakened.73 The weakening of the hydrogen bonds favoured the expulsion of the water entrapped inside of the particles. Therefore, with an increase in the temperature, an increase in the particle interaction was expected due to the increase in its hydrophobicity. According the results, the onset of NP aggregation driven by PEtOx dehydration started to be observed at ∼55 °C by DLS/SLS measurements for the [MPEO44-b-PEtOx252-b-(PCL)2×44] at ∼50–55 for the [MPEO44-b-PEtOx252-b-(PCL)2×87] and at ∼40 °C for the [MPEO44-b-PEtOx252-b-(PCL)2×131]. Although the increases observed for the NP molecular weight and aggregation number starting from ∼55 °C, [MPEO44-b-PEtOx252-b-(PCL)2×44], ∼50–55, [MPEO44-b-PEtOx252-b-(PCL)2×87], and at 40 °C, [MPEO44-b-PEtOx252-b-(PCL)2×131], were strong indicators of NP aggregation, the RG/RH ratio values also provided valuable information related to this process.64 Commonly, in the aggregated state, the increase in the RG of the NPs is more pronounced in comparison with RH; thus, an increase in the RG/RH ratio is observed. According to previous observations, this behaviour is related to a shift in the position of the scattering centres between the non-aggregated and aggregated NPs.51 In the non-aggregated NPs, the polymer density decreases from the centre to the shell, whereas in the aggregated NPs, the density is also relatively high in the periphery. The scattering centres of the aggregated particles are composed of a collection of hydrophobic collapsed particles, whereas they are water swollen in the non-aggregated particles.
Therefore, higher RG/RH ratios would be expected as the aggregation proceeds with the temperature increase. At 60 °C ([MPEO44-b-PEtOx252-b-(PCL)2×44]) and for nanoparticle concentrations of 0.5 and 1.0 mg mL−1, two peaks were observed in the distribution of the hydrodynamic radii (Fig. S4 c†). The scattering intensity related to the largest peak was approximately 450 nm for both concentrations, indicating a collapsed system at 60 °C. At 61 °C, for the nanoparticle concentration of 1.5 mg mL−1, the presence of a scattering intensity peak corresponding to a size larger than 100 nm (∼429 nm; aggregates) (Fig. S4 d†) was observed. For the nanoparticle concentration of 2 mg mL−1, the collapse was observed at 63 °C (Fig. S5†). Another evidence of the onset of aggregation and nanoparticle collapse was the increase in the diffusion coefficient with the increase on the temperature up to 60 °C (1.95–7.48 × 10−8 cm2 s−1; Fig. S2†). However, at 62 °C a decrease (6.67 × 10−8 cm2 s−1) in the diffusion coefficient indicated the increase in the RH and nanoparticles aggregation. The scattering patterns for the [MPEO44-b-PEtOx252-b-(PCL)2×87] and the [MPEO44-b-PEtOx252-b-(PCL)2×131] follow similar trends with the largest peak (or sample precipitation) indicating a collapsed system at 60–62 °C for [MPEO44-b-PEtOx252-b-(PCL)2×87] and at 45–50 °C for [MPEO44-b-PEtOx252-b-(PCL)2×131] (Fig. S9†). Taking into account the aforementioned results, we may infer that the LCST for the Y-shaped [MPEO44-b-PEtOx252-b-(PCL)2×44] nanoparticle solutions was between 60-63 °C, for the [MPEO44-b-PEtOx252-b-(PCL)2×87] was around 60 °C and for the [MPEO44-b-PEtOx252-b-(PCL)2×131] was between 45–50 °C. Our previous results for linear [MPEO44-b-PEtOx263-b-PCL87] nanoparticle solutions showed a LCST slightly lower, at 56–60 °C51 when compared to [MPEO44-b-PEtOx252-b-(PCL)2×44 and to [MPEO44-b-PEtOx252-b-(PCL)2×87]. The LCST for poly(2-ethyl-2-oxazoline) was observed in the range of 61–66.5 °C and it was dependent of the concentration and molecular weight of the polymer.48 Poly(2-ethyl-2-oxazoline) with concentration between 0.5% and 1% (wt %) and molecular weight of 20000 g mol−1 showed a LCST at 66.5 and 64.5 °C, respectively. In the present work the synthesised Y-shaped [MPEO44-b-PEtOx252-b-(PCL)2×44] and [MPEO44-b-PEtOx252-b-(PCL)2×87] with similar poly(2-ethyl-2-oxazoline) molecular weight presented LCST values of ∼60 °C at the same concentration range. Therefore, the presence of a hydrophobic polymer on the triblock terpolymers induced a decrease of the LCST as previously observed.51 The lowest LCST values (between 45–50 °C, Table 5, Fig. S9†) observed for the largest hydrophobic terpolymer synthesised [MPEO44-b-PEtOx252-b-(PCL)2×131] supports the statement. From these results, a schematic representation of the NPs [MPEO44-b-PEtOx252-b-(PCL)2×44] could be designed (Fig. 7).
![]() | ||
Fig. 7 Schematic representation for Y-shaped [MPEO44-b-PEtOx252-b-(PCL)2×44] nanoparticles solutions at different temperatures: (a) in a swollen state, (b) in an aggregation state. |
Cryo-TEM microscopy (Fig. 8) was also performed on the NPs to verify the possible morphological changes induced by the temperature. For samples measured at 25 °C, spherical NPs with a mean particle size of 70 nm were observed for [MPEO44-b-PEtOx252-b-(PCL)2×44] (Fig. 8a, top) around 50–60 nm for [MPEO44-b-PEtOx252-b-(PCL)2×87] (Fig. 8c, middle) and 100–120 nm for [MPEO44-b-PEtOx252-b-(PCL)2×131] (Fig. 8e, bottom). NPs measured at higher temperatures showed the onset of aggregation (Fig. 8b, top) for [MPEO44-b-PEtOx252-b-(PCL)2×44] (60 °C) with a mean diameter of 95 nm, around 200 nm (aggregates) for [MPEO44-b-PEtOx252-b-(PCL)2×87] (62 °C) and 300 nm (aggregates) for [MPEO44-b-PEtOx252-b-(PCL)2×131] (55 °C) (Fig. 8e, bottom).
The experimental data suggest that the temperature increase caused size and morphological changes in the NPs. The mean diameter of the NPs determined from the cryo-TEM images was in agreement with that determined by DLS (Tables 3–5). It was possible to observe some morphologically ill-defined structures, which were related to inorganic salts (NaCl and KCl) presented in the saline buffer solution.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra08298f |
This journal is © The Royal Society of Chemistry 2015 |