Sarah L.
Canning
ab,
Joseph M. F.
Ferner
a,
Natalie M.
Mangham
a,
Trevor J.
Wear
c,
Stuart W.
Reynolds
c,
Jonathan
Morgan
c,
J. Patrick A.
Fairclough
d,
Stephen M.
King
e,
Tom
Swift
f,
Mark
Geoghegan
*b and
Stephen
Rimmer
*af
aDepartment of Chemistry, University of Sheffield, S3 7HF, UK. E-mail: s.rimmer@bradford.ac.uk
bDepartment of Physics and Astronomy, University of Sheffield, S3 7RH, UK. E-mail: mark.geoghegan@sheffield.ac.uk
cDomino UK Ltd, Bar Hill, Cambridge, CB23 8TU, UK
dDepartment of Mechanical Engineering, University of Sheffield, S3 7HQ, UK
eISIS Pulsed Neutron & Muon Source, STFC Rutherford Appleton Laboratory, Didcot, OX11 0QX, UK
fDepartment of Chemistry and Biosciences, University of Bradford, Bradford BD7 1DP, UK
First published on 12th November 2018
Uniform onion micelles formed from up to ten nano-structured polymer layers were produced by the aqueous self-assembly of highly-branched copolymers. Highly-branched poly(alkyl methacrylate)s were chain extended with poly(acrylic acid) in a two-step reversible addition–fragmentation chain transfer-self-condensing vinyl polymerization (RAFT-SCVP) in solution. The resulting polymers were dispersed into water from oxolane (THF) using a self-organized precipitation-like method and the self-assembled particles were studied by phase-analysis light scattering, small-angle neutron scattering, and electron microscopy techniques. The relative hydrophobicity of the blocks was varied by changing the alkyl methacrylate (methyl, butyl, or lauryl) and this was found to affect the morphology of the particles. Only the poly(butyl methacrylate)-containing macromolecule formed an onion micelle structure. The formation of this morphology was observed to depend on: the evaporation of the good solvent (THF) during the self-assembly process causing kinetic trapping of structures; the pH of the aqueous phase; and also on the ratio of hydrophobic to hydrophilic segments within the copolymer. The lamellar structure could be removed by annealing the dispersion above the glass transition temperature of the poly(butyl methacrylate). To exemplify how these onion micelles can be used to encapsulate and release an active compound, a dye, rhodamine B (Rh B), was encapsulated and released. The release behaviour was dependent on the morphology of the particles. Particles formed containing the poly(methyl methacrylate) or poly(lauryl methacrylate) core did not form onions and although these materials absorbed Rh B, it was continuously released at room temperature. On the other hand, the lamellar structure formed from branch-poly(butyl methacrylate)-[poly(butyl methacrylate)-block-poly(acrylic acid)] allowed for encapsulation of approximately 45% of the dye, without release, until heating disrupted the lamellar structure.
An important means of controlling the formation of particles is that of SORP,5 in which two miscible solvents (one of which is a good solvent and the other is poor) are used. The faster drying component is a good solvent, which, when evaporated, allows spherical particles of controlled structure to be formed as a dispersion in the poor solvent. Block copolymers with immiscible components are ideal candidate materials5 for SORP because order is inherent in their structure. Nevertheless, a two-solvent process has been shown to create particles of a hyperbranched polymer relatively uniform in size.16 In the case of block copolymers, most examples of the formation of particles in a poor solvent are of concentric layered (onion) micelles,5–8 but hollow spheres have been observed,17 and other structures can be formed, for example by blending.5
Branched polymers, encompassing dendritic, multi-branched and highly-branched architectures, exhibit unique properties in terms of solution behaviour and rheology in comparison to linear analogues.18,19 Additionally, the large number of chain ends per molecule offers the possibility of adding further chemical functionality to the polymer.20–23 These branched materials can be produced by chain growth polymerization via the use of a branching monomer which acts as either monomer and transfer agent24 or monomer and initiator,25 in a process known as self-condensing vinyl polymerization (SCVP). A similar approach uses monomers that both polymerize and undergo addition–fragmentation with the propagating chain end.26 An extension of SCVP uses reversible addition–fragmentation chain transfer (RAFT) polymerization,27 a type of controlled radical polymerization which has been widely used to produce polymers with well-defined architectures, such as block copolymers, graft copolymers, and star polymers.28–33
Highly-branched polymers do not entangle in solution,34 which means that such materials are liable to form films quickly on the evaporation of their solvent,35 and may require careful control if there are specific requirements for molten structures. Perhaps surprisingly given the heterogeneous nature of highly-branched polymers, it has been possible to create ordered macroscopic objects from them,36,37 and micellar37–39 or vesicular37,40,41 structures of various sizes have also been possible.
Since SORP works readily with linear block copolymers, it is pertinent to question the kind of structures that can be formed with highly-branched block copolymers. Here, the conditions for the SORP-induced formation of multi-lamellar onion micelles from copolymers of poly(acrylic acid) (PAA) and different highly-branched poly(alkyl methacrylates)s are presented. Three different polymers were prepared, with hydrophobic highly-branched poly(methyl methacrylate) (PMMA), poly(butyl methacrylate) (PBMA), or poly(lauryl methacrylate) (PLMA). A SORP-like procedure was used to disperse these molecules into water, a good solvent for the hydrophilic poly(acrylic acid) (PAA) blocks but a poor solvent for the hydrophobic poly(alkyl methacrylate) (PnMA). Gradual addition of a selective solvent to a solution of copolymer in a good solvent for both blocks, followed by slow evaporation of the good solvent, led to the formation of self-assembled structures. This self-assembly was studied by a combination of transmission electron microscopy (TEM), particle-sizing and zeta-potential measurements using phase-analysis light scattering (PALS), small-angle neutron scattering (SANS), and scanning electron microscopy (SEM). It is generally accepted that a deterioration of solvent quality leads to the formation of aggregates.42–44 Indeed onion micelles are formed under specific conditions and only for branch-poly(butyl methacrylate)-[poly(butyl methacrylate)-block-poly(acrylic acid)], which is abbreviated here as HB-PBMA-PAA. (The nomenclature used here is extended to the other PnMA highly-branched cores: HB-PMMA-PAA and HB-PLMA-PAA.) The requirement for the poor solvent evaporation step and the effect of varying the ratio of hydrophilic to hydrophobic blocks were investigated. A mechanism for the formation of these self-assembled structures is postulated. Finally, these onion micelles are investigated for application delivery by studying the encapsulation and release of a model compound, rhodamine B (Rh B).
![]() | ||
Scheme 1 Synthesis of HB-PnMA-PAA block copolymers with pyrrole chain ends in a two-step RAFT-SCVP polymerization. |
Polymer | Conversiona /% | Molar feed ratio nMA![]() ![]() |
Polymer molar ratio PnMA![]() ![]() |
DBa |
M
n![]() |
M
w![]() |
Đ |
---|---|---|---|---|---|---|---|
a Monomer conversion and degree of branching determined by 1H NMR. b Molar masses and dispersity (Đ) determined by GPC (THF, PMMA standards). | |||||||
HB-PMMA | 55 | — | — | 0.059 | 13.5 | 27.3 | 2.0 |
HB-PBMA | 88 | — | — | 0.030 | 18.0 | 35.8 | 2.0 |
HB-PLMA | 77 | — | — | 0.062 | 30.8 | 64.7 | 2.1 |
HB-PMMA-PAA | 93 | 0.7![]() ![]() |
1.1![]() ![]() |
0.029 | 29.5 | 118.3 | 4.0 |
HB-PBMA1.4-PAA1.0 | 97 | 0.5![]() ![]() |
1.4![]() ![]() |
0.013 | 40.0 | 139.9 | 3.5 |
HB-PLMA-PAA | 99 | 0.3![]() ![]() |
0.7![]() ![]() |
Unknown | 34.8 | 328.0 | 9.4 |
HB-PBMA0.9-PAA1.0 | 92 | 0.5![]() ![]() |
0.9![]() ![]() |
0.014 | 2.7 | 7.6 | 2.8 |
HB-PBMA1.3-PAA1.0 | 91 | 0.75![]() ![]() |
1.3![]() ![]() |
0.011 | 4.3 | 16.7 | 3.9 |
HB-PBMA1.55-PAA1.0 | 94 | 1.0![]() ![]() |
1.55![]() ![]() |
0.018 | 2.6 | 8.0 | 3.1 |
HB-PBMA1.67-PAA1.0 | 96 | 1.5![]() ![]() |
1.67![]() ![]() |
0.029 | 7.5 | 30.1 | 4.0 |
HB-PBMA1.71-PAA1.0 | 97 | 2.0![]() ![]() |
1.71![]() ![]() |
0.033 | 14.4 | 126.8 | 8.8 |
The HB-PMMA macro-CTA (0.4 g) was dissolved in 7.2 g dioxane, then AA (0.4 g, 5.55 mmol) and ACVA (0.0022 g, 80 μmol, CTA/ACVA molar ratio = 5.0) were added and allowed to mix until the solid initiator had dissolved. The resulting solution was transferred into a glass ampoule, freeze–pump–thawed on a high vacuum line, flame-sealed, and heated in a water bath as described above to undergo polymerization. Products were then precipitated into rapidly stirring ice-cold petroleum ether 40–60 °C. The petroleum ether was removed by decanting and the polymer was dried under vacuum at room temperature for 24 h. The procedure was repeated once more to remove any traces of residual monomer, giving polymer products as a pale-yellow powder. HB-PBMA-PAA and HB-PLMA-PAA were synthesized following the same method.
Table 1 displays the results of the syntheses of both the homopolymer macro-CTAs, and also the copolymers. 1H NMR spectra peaks of the copolymers at δ = 1.63 and 2.22 ppm were due to PAA. The ratio of methacrylate to acrylic acid within the copolymers could also be calculated from the 1H NMR spectra. Equal mass fractions of PnMA and PAA were targeted for each copolymer, in addition to equal molar masses for all three copolymers, with the intention that any observed differences in self-assembly could be attributed to the change in hydrophobicity. The amount of PAA in all three copolymers was less than the feed ratio. However, all three copolymerizations of AA proceeded to high conversion, indicating the formation of PAA homopolymer, as expected for RAFT and RAFT-SCVP polymerizations.21 Homopolymer of the monomer used to form the second block is generally formed during the RAFT polymerization reaction. This lower molar mass peak in the chromatogram is due to the presence of a small amount of PAA homopolymer. DB was calculated for all of the macro-CTAs and the block copolymers, except for HB-PLMA-PAA where the pyrrole and styryl groups could not be clearly seen in the 1H NMR spectrum. DB values for the PnMA cores were close to the 0.06 targeted, then values decreased following PAA addition as the branch lengths increased.
THF GPC analysis of the copolymers following methylation with trimethylsilyldiazomethane demonstrated an increase in molar mass in each case compared to that of the macro-CTA. A large increase in dispersity was also observed for all three copolymers compared to their macro-CTA, which indicates a considerable degree of branching within the copolymer structure (Fig. 1). The large dispersities obtained here are not unusual for copolymerizations of this nature.25,50
The three HB-PnMA-PAA copolymers self-assembled into very different structures when dispersed into water, as shown by the TEM images in Fig. 2. Spherical micelles with a rough appearance to the particle surface were formed from HB-PMMA-PAA. These are consistent with the granular appearance of large particles formed by the aggregation of unimicellar aggregates.38 The BMA analogue, however, formed lamellar onion micelles. The HB-PLMA-PAA copolymer formed much smaller particles which were more elliptical in shape with a “dimple” in the centre. These results clearly demonstrate the strong effect that changing the alkyl group on the methacrylate monomer has on the self-assembly of these polymers. The architecture, degree of branching, and molar masses were similar and thus should not account for the differences in the self-assembled structures observed.
A combination of TEM, particle-sizing, and zeta-potential measurements (made by PALS) and SANS (by fitting a simple sphere model to scattering data) was used to characterize the different dispersions. The results are summarized in Table 2. Diameters measured from TEM images are expected to be smaller than those determined by PALS and SANS due to the effect of drying, whereas PALS and SANS measure particles in their hydrated state. This is true for the HB-PLMA-PAA sample. However, in the case of HB-PMMA-PAA and HB-PBMA1.4-PAA1.0 there is no clear trend in size measured by the different techniques. It must be noted that the diameter obtained using TEM represents the particles measured in a limited number of images analysed, whereas the SANS and PALS techniques give an ensemble average and therefore are more representative of the whole sample. Particle-size measurements of the HB-PBMA1.4-PAA1.0 dispersion indicated the presence of a bimodal population of larger and smaller onion structures which correlated with the TEM images (see Fig. 2). The HB-PLMA-PAA dispersion was notably uniform. All three dispersions were shown to be stable by the zeta-potential results, which indicate a negative charge on the particle surfaces due to ionization of the carboxylic acid groups of the PAA segments. At the native pH of the deionized water (pH ∼ 5) the PAA is partially ionized.
Sample | Mean diameter (PALS)/nm | PDI (PALS)a | Mean zeta potential/mV | Mean diameter (TEM)/nm | Mean diameter (SANS)b/nm | DI (SANS)c |
---|---|---|---|---|---|---|
a Dispersity of particles measured by PALS. b Using a simple sphere model. c Dispersity of particles determined by SANS. | ||||||
HB-PMMA-PAA | 159 ± 1 | 0.13 ± 0.01 | −55 ± 1 | 169 ± 5 | 137 ± 6 | 0.07 ± 0.02 |
HB-PBMA1.4-PAA1.0 | 82 ± 1 | 0.07 ± 0.02 | −50 ± 1 | 164 ± 6 | 171 ± 2 | 0.06 ± 0.01 |
191 ± 1 | ||||||
HB-PLMA-PAA | 60 ± 1 | 0.02 ± 0.02 | −41 ± 10 | 44 ± 1 | 63 ± 1 | 0.18 ± 0.01 |
Scanning electron microscopy (SEM) was used to study the surface of the sample and TEM to visualize the internal structure. SEM was used to confirm that the onion micelles were spherical. A TEM grid to which a stained HB-PBMA1.4-PAA1.0 dispersion was adsorbed was sputter-coated with gold and used for SEM imaging to allow direct comparison between the TEM and SEM images. The SEM images in Fig. 3 show that the onion micelles were indeed spherical, and the lamellar structure (Fig. 2B) from TEM was a series of internal concentric shells.
SANS was used to study dispersions of the copolymers in D2O. Diverse scattering profiles were obtained for the three samples, as shown in Fig. 4. A clear Bragg peak was observed for the HB-PBMA1.4-PAA1.0 sample which is characteristic of a multilayered structure. The layer spacing can be obtained from Q = 2π/d. Using the Q value associated with the Bragg peak, an average spacing of 12 nm is predicted for HB-PBMA1.4-PAA1.0. Despite the apparent differences in the profiles, the scattering of all three polymers was successfully fitted to a lamellar paracrystal model,52 and indicated a structure composed of layers with the hydrophobic PnMA on the inside to avoid the D2O and PAA segments on the outer surfaces. This model (which is more sophisticared than that used to calculate the particle sizes) is used to calculate the scattering from a stack of ordered lamellae, which is considered as a paracrystal to account for the repeat spacing. The model has been used previously to model the scattering from large multilamellar vesicles.52,53
The values of the fitting parameters obtained from the model are summarized in Table 3. The model fit gives a layer spacing of 11.5 nm for HB-PBMA1.4-PAA1.0, which is consistent with the spacing obtained from the Bragg peak, and shows some associated dispersity which can be observed in the TEM images. The model suggests an average of ten layers for this sample in agreement with the TEM results. The scattering length density (SLD) of the layers is approximately 2.2 × 10−6 Å−2, which indicates that there is water within the layered structure, as the SLD for the copolymer alone is 1.4 × 10−6 Å−2. A water content of 16.1% by volume can therefore be calculated using the calculated SLDs of the copolymer and the D2O, and the SLD of the layers determined from the model fit. Indeed, contrast in the experiment is due to trapped D2O within the onion micelle. This is likely to be in contact with the PAA.
Model parameter | HB-PMMA-PAA | HB-PBMA1.4-PAA1.0 | PLMA-PAA |
---|---|---|---|
Number of layers | 2.4 ± 0.4 | 10.3 ± 0.2 | 2.6 ± 0.1 |
Thickness/nm | 19 ± 2 | 4.2 ± 0.2 | 14.5 ± 0.2 |
Spacing/nm | 18 ± 2 | 11.5 ± 0.5 | 16.9 ± 0.1 |
Dispersity of spacing | 0.26 ± 0.09 | 0.22 ± 0.04 | 0.22 ± 0.02 |
SLD layers/Å−2 (×10−6) | 2.1 ± 0.2 | 2.20 ± 0.02 | 1.9 ± 0.2 |
% D2O within layers | 14.4 | 16.1 | 10.2 |
Lamellar structures were also indicated for HB-PMMA-PAA and HB-PLMA-PAA, but with fewer layers. The results suggest a mixture of bi- and tri-layered micelles, with much thicker layers than obtained for HB-PBMA1.4-PAA1.0. Increased SLD again indicates the presence of D2O within the micelle structure. Because of the greater amount of noise in the scattering of the HB-PMMA-PAA particles, the model fit is much less reliable than those of the other copolymers and therefore the parameters in Table 3 are subject to greater uncertainty.
The stability of the onion micelles under ambient conditions was also investigated. A sample of HB-PBMA1.4-PAA1.0 dispersed in water was prepared and analysed by TEM and PALS to confirm that onion micelles were formed. This sample was then kept at room temperature for a period of 5 weeks and subsequently re-analysed. Fig. 5 shows TEM images of the sample before and after storage. It is clear that the lamellar structure remains intact during this period. A bimodal distribution was again observed, with an increase in diameter of both the smaller and larger particles, accompanied by an increase in dispersity. This suggests some growth or swelling of the onion micelles had occurred during the storage period; but significantly the lamellar structure was maintained.
Annealing was carried out on the dispersions to investigate temperature-dependent behaviour. Sample tubes containing copolymer dispersions in water were sealed and immersed in an oil bath heated to 45 °C for 12 h. Samples were removed and imaged by TEM before and after the annealing period (Fig. 6). Particle-sizing measurements were also performed. The HB-PMMA-PAA and HB-PLMA-PAA structures underwent a small increase in size and associated uncertainty, and their shapes appeared to become slightly more irregular. The HB-PBMA1.4-PAA1.0 micelles also underwent a small increase in particle size and the uncertainty increased. More significantly, the lamellar structure was no longer present, with a more vesicle-like structure observed. These results can be rationalized by considering the glass transition temperature (Tg) of the three alkyl methacrylate polymers. PMMA has a high Tg of 105 °C, similar to that of PAA (106 °C), and PLMA has a low Tg of −65 °C.54 PBMA, on the other hand, has Tg = 25 °C,55i.e. close to ambient temperature at which the onion micelles are assembled. When the temperature of the dispersion is increased to 45 °C, this has little effect on the self-assembled HB-PMMA-PAA and HB-PLMA-PAA structures since they remain well above or well below their respective Tg. However, the PBMA segment of HB-PBMA1.4-PAA1.0 is heated to above its Tg during the annealing process and hence the lamellae are able to coalesce to form an amorphous ‘vesicular’ structure.
The rate of self-assembly of the onion micelles was investigated by varying the rate of water addition from 0.05 to 1.00 ml min−1. No differences in micelle structure were observed (see ESI Fig. S4 and S5†). One possibility is that, if equilibrium structures were being formed, then the evaporation of the good solvent would not be important. On the other hand, if solvent evaporation proved to be important then this would confirm the formation of dynamic, non-ergodic structures. In order to test this, the micelle preparation procedure was carried out as usual but instead of allowing the THF to evaporate once the water was added, the solution was injected into a dialysis cassette and the THF removed by dialysis against water. After 24 h, the dispersion was removed from the cassette and imaged by TEM; Fig. 7 shows some of the images obtained. Instead of onion micelles, many small spherical micelles were formed. The mean diameter of these micelles was 18 ± 4 nm. This was comparable to the thickness of the individual layers of the onion micelles formed when the THF was evaporated. Some coalescence of small spheres to form worm-like structures was observed, and in the final image an onion micelle apparently in the process of formation was observed. However, this represents a very small population of the overall sample, the majority of which are present as small spheres. These results suggest that the evaporation step is a significant part of the process of onion micelle formation from this HB-PBMA1.4-PAA1.0 copolymer, and removal of this step impedes the formation of the kinetically-trapped onion structures.
The ratio of hydrophilic to hydrophobic segments in a block copolymer is known to play an important role in self-assembly. Consequently, the ratio of poly(alkyl methacrylate) to poly(acrylic acid) in these copolymers was expected to affect their self-assembly behaviour. To test this, HB-PBMA-PAA copolymers with varying ratios of BMA to AA monomer were synthesized. Five different polymers were prepared in all, with BMA:
AA ratios (from NMR) of 0.9
:
1.0, 1.3
:
1.0, 1.55
:
1.0, 1.67
:
1.0 and 1.71
:
1.0. Dispersions in water were prepared as previously and imaged by TEM; Fig. 8 shows the images obtained.
The results shown in Fig. 8 confirm that the ratio of hydrophobic to hydrophilic segments within these copolymers has a significant effect on their self-assembly behaviour. Large spheres are seen when less AA is present relative to the amount of BMA, whereas very small spheres are formed where copolymers contain more AA relative to BMA. A longer hydrophilic ‘stabilizer’ block disfavours micelle fusion, whilst a shorter stabilizer block is more likely to lead to micelle fusion. In the case of linear block copolymers, this leads to the formation of fibres or ‘worms’ when the stabilizer block is shorter. In the case of these highly branched polymers, when there is less hydrophilic polymer the small unimolecular micelles fuse to form more energetically favourable large multimicellar aggregates, whereas when more hydrophilic polymer is present it can adequately stabilize the smaller micelles. Onion micelles are observed at only one of these ratios, 1.32:
1.0 PBMA
:
PAA, indicating the presence of an optimum composition with an intermediate hydrophilic component where onion micelle formation is favourable.
The pH dependence of onion micelle formation was also studied. Dispersions were prepared in buffer solutions at pH 4, 7 and 10, with NaCl added as required to maintain constant ionic strength. Representative TEM images are shown in Fig. 9. These experiments showed that pH was not a strong factor in the formation of the self-assembled structures except for pH 10. The HB-PBMA1.4-PAA1.0 onion micelles were formed at all pH but were somewhat smaller at pH 10. At pH 4 the PAA is not expected to be significantly charged (PAA has a pKa of 4.5,56–59 which will shift to higher pH for dense systems such as these described here to avoid Coulombic frustration) but at pH 7 more than half the acidic groups would be expected to be charged, and most would be charged by pH 10; yet onion micelles are still formed. That pH does not play a significant role in the micelle formation indicates that the PAA may not become fully solvated when the water is added and that the THF promotes hydrogen bonding between PAA branches. Such a consideration would dictate that it is the hydrophobic PnMA that is driving the self-assembly.
The critical role of solvent evaporation indicates that the formation of the onions is due to SORP. Amphiphilic block copolymers have not been observed to form onions,17 but nevertheless there is no reason why the mechanism cannot apply to specific systems. The resultant morphology is driven by a balance between the minimization of interfacial energies and the repulsive interactions generated by swelling of the PAA segments. As indicated in earlier work,60 the attainment of the observed morphologies almost certainly involves distortion of the segments within the hydrophobic phase; and clearly the distortion of highly branched polymers has a greater relative energy cost compared to that of linear polymers. Block copolymer worm-like micelles and vesicles can be considered as the products of the fusion of spherical micelles or single polymer chains.61,62 However, fusion of the unimolecular micellar spheres reported here is limited because of the branched architecture.63,64 The limitations on fusion will impose a restriction on the system and can contribute to the formation of unusual morphologies.
THF is a good solvent for both PAA and PBMA, so the copolymers are initially present as random coils in solution. Water, however, is a good solvent for PAA and a poor solvent for PBMA (meaning it is a traditional block-selective solvent for amphiphilic block copolymer self-assembly). As water is added and THF begins to evaporate, PAA becomes more soluble and PBMA less so. The PBMA segments are driven inside the micelle to minimize the contact with water, with PAA on the outside acting to stabilize the micelles. It is proposed that the presence of some remaining THF in the mixture promotes the hydrogen bonding between PAA segments rather than the occurrence of hydrogen bonding between PAA and water. Particle coalescence leads to the formation of lamellar phases. Curvature is induced to reduce PBMA contact with water and increase PAA–PAA contacts. For the onion micelles, as polymers are added to the micelle the entropic cost of coating each layer increases because the curvature decreases with each added layer. This curvature effect could explain the limit on micelle size of ∼10 layers.
A solution of Rh B in water was added to the HB-PBMA1.4-PAA1.0 in THF solution, following the previous methodology. Once the THF had evaporated, the solution was injected into a dialysis cassette with a 3500 MWCO membrane. Firstly, the solution was dialysed against water at room temperature to remove any unbound dye. The water was changed regularly, at which point a sample of this water was analysed for Rh B. After 48 h, a negligible amount of Rh B was being released so the cassette was placed into a sealed vessel containing fresh deionized water heated to 45 °C. Rh B release was monitored by UV-visible absorption spectroscopy.
Measurement of the concentration of dye released during dialysis at room temperature allowed calculation of a ‘loading efficiency’ by comparing this to the initial amount of dye added. An average loading efficiency of 45% was obtained over the two release studies. Of this 45% encapsulated dye, 10% was released during 24 h of heating at 45 °C (Fig. 10). The same encapsulation procedure was carried out using the methyl and lauryl analogues, HB-PMMA-PAA and HB-PLMA-PAA. As expected, the encapsulation was significantly less efficient without the layered structure. The dye was steadily released from these micelles during dialysis at room temperature; after 15 days with regular water changes, 78% and 71% of the initial amounts of Rh B dye had been released for HB-PMMA-PAA and HB-PLMA-PAA respectively with release still ongoing so that it was not possible to obtain a loading efficiency.
![]() | ||
Fig. 10 Release profile of encapsulated Rh B from HB-PBMA1.4-PAA1.0 onion micelles monitored by UV-visible absorption spectroscopy (λmax = 553 nm) during heating at 45 °C. |
The binding of Rh B with the PAA makes it likely that the rhodamine is located in the PAA layers of the micelle, even if the release does not occur in stages. This has also been described for other onion micelles.13
These data show the utility of the HB-PBMA-PAA onion micelles as vectors for encapsulation and release, initiated by thermal stimulus, of active compounds. Thermal release is applicable to a number of technologies involving the release of drugs or fragrances. Nanocapsules involving alkyl methacrylates and methacrylic acid have also been shown to release fragrances with a pH trigger.65
The HB-PBMA1.4-PAA1.0 copolymer self-assembled into unusual onion micelles with lamellar structure. SEM was used to show that the HB-PBMA1.4-PAA1.0 onion particles had a spherical 3D structure, which confirmed that the lamellar structure seen in the TEM images was internal. The lamellar structure was confirmed by SANS studies and the scattering data were fitted to a lamellar paracrystal model, with the parameters obtained from the fit in agreement with observations made by TEM. The onion particles were found to be stable when stored at room temperature over a period of five weeks. Annealing the sample at 45 °C for 24 h was found to remove the lamellar structure as this heating traversed the Tg of the PBMA segment, whilst no change was observed when the same procedure was carried out on the PMMA and PLMA analogues.
Several experiments were carried out to aid with elucidating the mechanism by which the onion micelles form. The evaporation of THF was found to play an important role in onion micelle formation as when this step was omitted, small spheres were formed instead. This is consistent with previous reports of SORP. It was also found that the ratio of PBMA to PAA within the polymer does affect the self-assembly behaviour, as expected. Onion micelles were only formed at one ratio of hydrophilic to hydrophobic component, while either small or large spheres were formed at the other ratios depending on the degree of steric stabilization.
A proof-of-concept study demonstrated that Rh B dye could be encapsulated during the onion micelle assembly process, and subsequently released by heating to 45 °C. This opens the system to potential applications in delivery of active agents.
Footnote |
† Electronic supplementary information (ESI) available: Assigned 1H NMR spectra; characterization details; investigation into rate of water addition on copolymer self-assembly; UV-vis release data; details of lamellar paracrystal model used to fit SANS data; and additional TEM images. See DOI: 10.1039/c8py00800k |
This journal is © The Royal Society of Chemistry 2018 |