Yangmiao Moabcd,
Shudong Linabcd,
Yuanyuan Tuabd,
Guojun Liuace,
Jiwen Hu*abcd,
Feng Liuac and
Jun Songabcd
aGuangzhou Institute of Chemistry, Chinese Academy of Sciences, Guangzhou, P. R. China 510650. E-mail: hjw@gic.ac.cn; Fax: +86-20-85232307
bKey Laboratory of Cellulose and Lignocellulosics Chemistry, Chinese Academy of Sciences, P. R. China 510650
cThe University of the Chinese Academy of Science, Beijing, P. R. China 100039
dGuangdong Provincial Key Laboratory of Organic Polymer Materials for Electronics, 510650, P. R. China
eDepartment of Chemistry, Queen's University, 90 Bader Lane, Kingston, Ontario, Canada K7L 3N6
First published on 8th June 2016
A novel amphiphilic binary graft copolymer poly(glycidyl methacrylate)-graft-[poly(2-cinnamoyl-oxyethyl methacrylate)-random-methoxy polyethylene glycol] (PGMA-g-(PCEMA-r-MPEG)) was successfully synthesized by a combination of atom transfer radical polymerization (ATRP) and click reaction, in which alkyne-end-functionalized poly(2-cinnamoyloxyethyl methacrylate) (PCEMA–CCH) and poly(ethylene glycol) methyl ether (MPEG–C
CH) were grafted onto a poly(3-azide-2-hydroxy-propyl methacrylate) (P(GMA-N3) backbone. This polymer was used to prepare stable unimolecular micelles (UMMs), which could be produced using either high or low polymer concentrations. Since water is a good solvent only for MPEG but a poor solvent for both PGMA and PCEMA, the hydrophobic PGMA and PCEMA segments aggregated together to form a dense core that was surrounded by a corona based on the soluble MPEG segments. PCEMA was photo-crosslinkable, and thus the UMMs could be crosslinked by shining UV light on the system to yield permanent UMMs. The morphologies of the UMMs were characterized by TEM, AFM, and DLS. Both the TEM and AFM observations indicated that the crosslinked UMMs had a diameter of ∼13 nm, while the DLS measurements indicated they had a diameter of ∼34 nm. The unimolecular state of the micelles was confirmed by SEC, as well as a comparison of the theoretical mass per graft copolymer molecule with that of an individual micelle. Moreover, the morphology of the UMMs was unperturbed by the crosslinking reaction although they became more compact and had a slightly smaller diameter.
To date, UMMs have been successfully prepared from a number of amphiphilic polymers with various topologies (Fig. 1), such as linear copolymers,19–21 macrocyclic copolymers,22 star-shaped copolymers,16,23 graft copolymers,4,24–26 and hyperbranched copolymers.27,28
Graft copolymers possess unique architectures containing a backbone chain and many side-chains.29 They bear a cylindrical brush conformation if D/Rg < 2, where D is the average distance between the grafted chains and Rg is the radius of gyration of the polymer in solution.30 It is noteworthy that the morphology and the size of UMMs based on amphiphilic graft polymers can be conveniently and precisely tuned by changing the length of the backbone or the side-chains, as well as the grafting density. Moreover, their synthesis is relatively simple. Therefore, graft copolymers are promising precursors for UMMs.
However, in view of documents, while copolymers with other topologies have been widely used to prepare UMMs, research focusing on the preparation of UMMs based on graft copolymer is relatively scarce. There are a few examples on graft copolymers that have been used to investigate the formation of UMMs by the collapse of their backbone or side-chains, i.e., the flower-like unimolecular structures of poly(methyl methacrylate)-graft-polystyrene (PMMA-g-PS)25 formed in selective solvents and the core–shell nanostructures based on PGA-g-(PCL-b-PEO).24 Moreover, all of the graft copolymers that have been used as precursors for UMMs were homograft copolymers (Fig. 1d(1)) or graft copolymers bearing block copolymer side-chains (Fig. 1d(2)). That is, both of these examples possess one type of pendent polymer chain and they are thus described as unitary graft copolymers. However, unitary graft copolymers suffer from difficulty in adjusting the morphology of the UMMs and size of the hydrophobic core simultaneously.
We envisioned that this problem may be solved by using binary graft copolymers as the precursors for UMMs with the topology as shown in Fig. 1d(3). Here, different from unitary graft copolymers, the hydrophobic and hydrophilic segments of binary graft copolymers randomly distribute in the backbone chains of the graft copolymers. This unique structure of binary graft copolymers were expected to make it possible to adjust the morphology of the obtained UMMs and the size of their hydrophobic core simultaneously in one step by controlling the initial molar ratio of the side chains to the backbone chain. However, to the best of our knowledge, there have been no reports on UMMs from binary graft copolymers.
Herein, to confirm our hypothesis, we design a new kind of binary graft polymers and investigated the formation of UMMs in a selective solvent via intramolecular collapse. Specifically, a novel well-defined amphiphilic binary graft copolymer PGMA-g-(PCEMA-r-MPEG), denoted as GP, was synthesized by grafting MPEG–CCH and PCEMA–C
CH as side-chains onto a third polymer backbone P(GMA-N3) via Cu catalyzed alkyne–azide cycloaddition (CuAAC). The photo-crosslinkable PCEMA component was introduced in this work to help investigate the formation of the unimolecular micelles. The one-pot grafting reactions were quantitative at a [C
CH] to [N3] molar ratio of ≤39/260 or when x + y of Scheme 1 was ≤15%. The residual N3 groups were then deactivated by reacting them with propargyl alcohol.
In order to obtain UMMs, we initially adopted Liu's approach,31 which was an effective strategy to inhibit intermolecular association. This strategy involved adding a GP solution in DMF into deionized water under constant stirring and UV irradiation (if photo-crosslinking was desired). Since water is a good solvent only for MPEG but a poor solvent for both PGMA and PCEMA, the hydrophobic PGMA and PCEMA segments aggregated together to form a dense core while the soluble MPEG segments stretched into the water phase and served as a corona. The structure of the obtained UMMs could be locked by shining UV light onto the system in order to crosslink the PCEMA segment, thus yielding permanent crosslinked UMMs. Importantly, we also found that while UMMs could be obtained without subjecting them to crosslinking under UV irradiation, to our surprise, GP also underwent assembly at a relative high polymer concentration of 10 mg mL−1 by directly adding deionized water into a GP solution in DMF and then dialyzing this dispersion again with a large amount of deionized water.
Usually, the high dilution method (i.e., <0.004 mg mL−1) was normally employed to avoid undesired intermolecular association.32,33 However, this method is very noneconomical in terms of solvent use. This inefficient and costly method can be improved by a continuous addition stategy,34 which was employed by Liu to prepare unimolecular tadpoles.31 This strategy ensured that UMMs were prepared at final graft copolymer concentrations that were thousands of times higher than what could have been achieved with the high dilution method. Up to now, UMMs using unitary graft copolymers as the precursors have been prepared at a relative high concentration (∼0.3 mg mL−1) in some published literature.35,36 For example, Takaya Terashima and coworkers found two of their amphiphilic random copolymer underwent intramolecular folding even at 60 mg mL−1.20 This was because the proper hydrophilic/hydrophobic ratio successfully suppressed intermolecular association. In our case, new binary graft copolymer could also be assembled into UMMs at high polymer concentrations of 10 mg mL−1. This finding could be not attractive in fundamental study on preparation of UMMs based on designed binary graft copolymers, but may be useful for potential application, i.e., drug delivery.
Glycidyl methacrylate (GMA, Aladdin, 98%) was passed through a basic alumina column to remove the inhibitor. 2-Hydroxyethyl methacrylate (HEMA, Aladdin, 98%) was purified by initially washing an aqueous solution of this monomer (25 vol% HEMA, 100 mL) with hexane (4 × 200 mL). The monomer HEMA was subsequently precipitated from the aqueous phase by the addition of NaCl, and then it was dried over MgSO4 prior to distillation under reduced pressure.37 2-Methoxyethyl 2-bromoisobutyrate and 3-(trimethylsilyl)propargyl 2-bromoisobutyrate were synthesized according to a literature procedure.38
Copper(I) bromide (CuBr, Aladdin, 98%) and copper(I) chloride (CuCl, Aladdin, 98%) were rinsed with glacial acetic acid and subsequently washed with methanol before they were dried under vacuum. Propargyl alcohol was dried over anhydrous magnesium sulfate for 12 h and distilled immediately before use. Meanwhile, triethylamine (99%, Aladdin) was refluxed over sodium and distilled before use. Pyridine, diethyl ether, 1,4-dioxane and diphenyl ether were purchased from Aladdin Reagents of China and refluxed over CaH2 overnight and distilled prior to use.
P(GMA-N3) was prepared by a method described in the literature.39 PGMA260 (5.61 g, bearing 39.5 mmol of epoxy groups), NaN3 (10.3 g, 158 mmol) and AlCl3 (0.133 g, 1.00 mmol) was dissolved in 100 mL of DMF and stirred for 48 h in a temperature-controlled oil bath at 50 °C. After the reaction, insoluble impurities were removed via filtration. The filtrate was concentrated under reduced pressure and then precipitated in water three times. The product was dried under vacuum to provide 7.18 g of a white solid in a 98.2% yield.
To remove the trimethylsilyl protecting group introduced by the initiator, the product obtained above (10.0 g, bearing 1.18 mmol of trimethylsilyl groups), DMF (80 mL) and TBAF (1.23 g, 4.71 mmol) were stirred together at room temperature for 72 h. After most of the solvent was removed under reduced pressure, the polymer solution was added to an excess of deionized water (300 mL) in order to precipitate the polymer. The polymer was redissolved in ∼30 mL of methanol and precipitated into 300 mL of water again. The obtained product was dried under vacuum for 24 h, thus generating 8.0 g of PHEMA–CCH as a light yellow solid in an 81% yield.
In the second step, poly(ethylene glycol) methyl ether was reacted with 4-oxo-4-(prop-2-ynyloxy)butanoic acid to yield MPEG–CCH. In a typical procedure, MPEG (Mn = 5000 g mol−1, 20.0 g, 4.00 mmol), 4-oxo-4-(prop-2-ynyloxy)butanoic acid (1.87 g, 12.0 mmol), DMAP (1.95 g, 16.0 mmol) and EDC·HCl (4.60 g, 24.0 mmol) were dissolved in 100 mL CH2Cl2. The solution was stirred at room temperature for 72 h and then sequentially washed twice with 2 M HCl (5 mL), saturated NaHCO3 (10 mL) and distilled water (10 mL). The organic phase was collected and dried over anhydrous MgSO4 for 12 h and filtered. The filtrate was concentrated via rotary evaporation and precipitated from an excess of cold diethyl ether. The precipitate was then dried under vacuum for 12 h, thus yielding 17.5 g of MPEG–C
CH as a white powder in an 85.1% yield.
Sample | [M0]/[I0] | Yielda (%) | DPb NMR | Mn NMR (kg mol−1) | Mn SEC (kg mol−1) | Mw/Mn SEC |
---|---|---|---|---|---|---|
a Determined by gravimetric analysis.b Determined via 1H NMR integration. | ||||||
PGMA260 | 700![]() ![]() |
40.1 | 260 | 36.9 | 93.3 | 1.22 |
P(GMA-N3)260 | 98.2 | 260 | 48.1 | 1610.2 | 2.15 | |
MPEG–C![]() |
85.1 | 114 | 5.0 | 28.3 | 1.01 | |
PCEMA–C![]() |
60![]() ![]() |
78.1 | 65 | 20.8 | 41.9 | 1.44 |
P(GMA-N3) was prepared by ring-opening reaction of the epoxy groups with NaN3 in the presence of AlCl3. As shown in Fig. 3a, the signals corresponding to the CH (3.23 ppm) and CH2 (2.64 ppm, 2.84 ppm) protons of the epoxide ring had shifted to 3.87 and 3.38 ppm, respectively. The signals at 3.79 and 4.29 ppm of the methylene (–COOCH2–) moiety had also shifted to one signal at 3.87 ppm in P(GMA-N3) (d peaks for P(GMA-N3) in Fig. 3a). All of these chemical shifts confirmed the 100% conversion of the epoxy groups to azido groups.
The P(GMA-N3)260 sample was also characterized by SEC using DMF as the eluent. A relatively broad peak (PDI = 2.15) was observed, which was probably due to the enhanced interaction between the azide groups or/and the hydroxyl groups of P(GMA-N3) and the SEC columns according to previous reports.41 In order to determine whether the azide groups and/or hydroxyl groups of P(GMA-N3) influenced the SEC trace, the azide groups and hydroxyl groups of P(GMA-N3) were protected by reacting P(GMA-N3) with propargyl alcohol and acetic anhydride in sequence. The resultant polymers obtained from each step were then analyzed by SEC. Fig. 2 shows the SEC traces of P(GMA-N3) prior to protection (a), P(GMA-N3) after protection with propargyl alcohol (b) and P(GMA-N3) after protection by both propargyl alcohol and acetic anhydride (c). The peak remained broad after P(GMA-N3) had reacted with propargyl alcohol, but became narrow (PDI = 1.50) after the reaction with acetic anhydride. This result suggested that the broad distribution was likely caused by the hydroxyl groups.
PHEMA–CC–TMS was synthesized according to a published literature procedure.42 3-(Trimethylsilyl)propargyl 2-bromoisobutyrate was used as the initiator to prevent possible side reactions between the radicals and the terminal alkyne group during the polymerization.43 The TMS protecting group was readily removed by stirring the polymer in a DMF solution with TBAF. The 1H NMR spectrum of PHEMA–C
CH is shown in Fig. 3d. It was apparent that the signal corresponding to the TMS protecting group at 0.14 ppm had disappeared from the spectrum, which demonstrated that this protecting group had been successfully removed.
![]() | ||
Fig. 3 1H NMR spectra and peak assignments for (PGMA-N3)260 (a), PGMA260 (b), PCEMA–C![]() ![]() ![]() |
PCEMA–CCH was obtained by reacting PHEMAC
CH with cinnamoyl chloride in anhydrous pyridine. This cinnamation reaction proceeded quantitatively. Its quantitative occurrence can be appreciated by comparing the 1H NMR spectra of PHEMA–C
CH and PCEMA–C
CH shown in Fig. 3. The peaks appearing at 3.59 and 3.90 ppm in the spectrum of PHEMA–C
CH (signals c and d, respectively in Fig. 3d) had completely disappeared and were replaced by the signals appearing at 4.08 and 4.21 ppm after the cinnamation reaction as shown in the spectrum of PCEMA–C
CH (signals c and d, respectively in Fig. 3c). The DP of PCEMA–C
CH was determined from the integration ratio of the peak at 4.59 ppm (signal e in Fig. 3c) with respect to that of the ethyl groups at 4.08 ppm (signal c in Fig. 3c). The repeat unit number of PCEMA–C
CH was determined to be 65. SEC characterization revealed that the number-average molecular weight of PCEMA–C
CH was 4.19 × 104 g mol−1. Meanwhile, the polydispersity index was 1.44.
The 1H NMR spectra of the purified GP is shown in Fig. 4. The PGMA backbone signals were absent because the backbone chain was surrounded by side-chains and tumbled slowly, and thus these signals were too broad to be detected via NMR.38 In contrast, all of the protons of the grafted PCEMA and MPEG chains were visible in the spectrum. A quantitative comparison of the integrations measured at 7.50 and 3.64 ppm yielded a molar ratio of 1.00:
7.99 for the PCEMA and MPEG repeat units, respectively, while the theoretical value was 1.00
:
7.89. The mass ratio of MPEG–C
CH to PCEMA–C
CH was calculated to be 7
:
3.
The SEC traces of PCEMA–CCH and MPEG–C
CH and of the crude GP product obtained after the click reactions had been performed are shown in Fig. 5. DMF rather than THF was used as the eluent in this case because GP had poor solubility in THF.
![]() | ||
Fig. 5 SEC traces of GP and its three precursors: GP (a), P(GMA-N3) (b), PCEMA–C![]() ![]() |
As expected, GP had a much shorter retention time than those of its precursors. P(GMA-N3) had an abnormally short retention time, probably due to the excessive swelling of the P(GMA-N3) chains in DMF. At these low feed ratios, all of the added polymers underwent quantitative grafting under our reaction conditions, and thus the obtained grafting density matched its targeted value of 15%. The mass fraction of the hydrophilic MPEG chains was 58.6%, and that of hydrophobic PGMA and PCEMA chains combined was 41.4%. The polydispersity index of GP was 1.62 (Table 2).
Sample | Mass feed ratioa | Molar feed ratiob | Mn,SEC (kg mol−1) | Mn,theorc (kg mol−1) | Mw/Mn |
---|---|---|---|---|---|
a Feed mass ratio between P(GMA-N3), PCEMA–C![]() ![]() ![]() ![]() ![]() |
|||||
GP | 1.0![]() ![]() ![]() ![]() |
260![]() ![]() ![]() ![]() |
2190 | 294 | 1.62 |
As shown in Scheme 3, the prepared crosslinked UMM solution (c) was as transparent and colorless as pure water (a) and the solvated GP solution in DMF (b), but the pathway of a laser beam was not visible in water. The laser beam's pathway in the crosslinked UMM sample was much stronger than that in the GP solution in DMF, suggesting some new sources of scattering were introduced during the above steps. In order to determine the crosslinking degree of the obtained micelle sample, UV-vis spectroscopy was used to monitor the crosslinking process of the non-crosslinked micelle sample by measuring the decrease in the absorbance at 274 nm corresponding to the CEMA double-bond. Its crosslinking degree was determined to be ∼60.1% by comparing the reduced absorbance at 274 nm of the micelle sample with the original absorbance of the non-crosslinked micelle sample.
Size Exclusion Chromatography (SEC) was performed to confirm whether the micelles formed were UMMs or multimolecular micelles. The crosslinked UMMs described above were subjected to dialysis against deionized water to remove DMF in advance. The crosslinked UMM sample was obtained by freeze-drying treatment and redispersal in DMF prior to characterization by SEC. Fig. 6a compares the SEC traces of GP and a micellar GP sample. It was apparent that the SEC peak of the GP micelle sample had shifted to a longer retention time, which was due to a reduction of the hydrodynamic volume VP.
To calculate the hydrodynamic volume VP of the GP micelle sample at its chromatogram peak, we can correlate its peak PS-equivalent molecular weight Mp with VP using eqn (1):31
![]() | (1) |
Herein, the intrinsic viscosity [η]P of PS with a molecular weight Mw can be calculated from eqn (2):
[η]P = 1.10 × 10−2MP0.725 mL g−1 | (2) |
The use of both eqn (1) and (2) yielded a hydrodynamic volume VP reduction of 15.2% for the GP micelle sample relative to its graft copolymer precursor GP. This pronounced decrease in volume suggested that PCEMA and PGMA underwent intramolecular collapse in water. Based on the above analysis, we concluded that GP underwent intramolecular folding in water to form compact UMMs, rather than multimolecular micelles.
The morphology of the crosslinked UMMs was analyzed by TEM. For this measurement, the micelles were aero-sprayed onto nitrocellulose-coated copper grids and then negatively stained with phosphotungstic acid. As shown in Fig. 6c, the path length of the electron beam through the particles increased from the edge to the center because the stain precipitated onto the nitrocellulose-coated grid after deposition of the particles. Thus, the particles appeared white on a dark background. The average diameters of the crosslinked UMMs shown in Fig. 6c were 13 ± 2 nm. AFM measurements were also performed to quantify the structures of the crosslinked UMMs. As shown in Fig. 6e, robust spherical nanoparticles were clearly observed with a mean diameter of 15 nm, which was very close to that determined by TEM. The height of the nanoparticles was ∼6 nm, which closely matched their radius. The size distribution of the crosslinked UMMs was studied by DLS. As shown in Fig. 6b (dashed line), a unimodal peak in the range of 20–60 nm was observed via DLS with a narrow PDI (0.12). The mean hydrodynamic diameter (Dh) of the unimolecular micelles was 34 nm. The Dh value was much larger than that determined by both TEM and AFM because the samples prepared for the TEM and AFM measurements had been dried, but an aqueous solution was employed for the DLS measurements. Consequently, the hydrophilic MPEG chains assumed a stretched conformation in the latter case.
The theoretical mass of a UMM was calculated to be 4.9 × 10−19 g from mtheor = (M(P(GMA-N3)) + M(PCEMA) × n × x + M(MPEG) × n × y)/NA, where NA is Avogadro's number. Assuming that the crosslinked UMMs measured by AFM was in the shape of a hemisphere with a radius of 6.5 nm, the mass of an individual nanoparticle would be 6.8 × 10−19 g, as given by mmeasu = ρ(GP) × (1/2) × (4πr3/3). Herein, the densities of the GP, PCEMA, MPEG, and PGMA were denoted as ρ(GP), ρ(PCEMA) = 1.25 mg mL−1,47 ρ(MPEG) = 1.10 mg mL−1,48 and ρ(PGMA) = 1.08 mg mL−1.49 Assuming that the volume of the GP is the sum of the volumes of individual components, then ρ(GP) was calculated from ρ(GP) = M(GP)/(M(PGMA)/ρ(PGMA) + (M(PCEMA) × 4.4)/ρ(PCEMA) + (M(MPEG) × 34.6)/ρ(MPEG)) to be 1.18 mg mL−1. The values of mtheor and mmeasu were very similar, further proving the validity of the conclusion that was reached based on the SEC measurements.
In order to determine whether the same results were obtained without a crosslinking process, a non-crosslinked sample was prepared and observed via TEM. The sample was prepared by adding a GP solution (5 mg mL−1 in DMF) dropwise into deionized water under constant stirring. The DMF was subsequently removed by dialysis against deionized water. The other experimental conditions were the same as those used to prepare the crosslinked UMMs, except that the crosslinking process was not employed in this case. As shown in Fig. 6d, spherical solid particles with a diameter of 15 ± 1 nm were observed via TEM for the non-crosslinked sample. This non-crosslinked micelles was larger than the crosslinked UMMs because the hydrophobic non-crosslinked PCEMA domain was not as compact as the crosslinked PCEMA,50 rather than because multimolecular micelles formed in the absence of crosslinking treatment. Therefore, we concluded that stable UMMs with the PCEMA and PGMA components forming the core and the MPEG components forming the corona were prepared through the above non-crosslinking protocol. Therefore, crosslinking did not influence the morphologies of the UMMs, but did cause the crosslinked UMMs to become contracted and exhibit a slightly smaller diameter. For comparison, DLS measurements were also performed on the non-crosslinked UMMs as shown in Fig. 6b (solid line). It can be seen that there was no obvious difference encountered when the micelles were not subjected to crosslinking treatment.
The micellar solution prepared from the above protocol was added dropwise onto nitrocellulose-coated grids and negatively stained with 1.3 wt% phosphotungstic acid for 30 s prior to TEM observation. Fig. 7a shows a TEM image of the micelles prepared at a high GP concentration. It could be seen that the crosslinked micelles had a spherical morphology and an average diameter of ∼13 nm, which was very similar to that of the crosslinked UMMs obtained in ultra-dilute conditions. This morphology and size of the crosslinked micelles was also confirmed by AFM measurement (Fig. 7b). The hydrodynamic diameters and polydispersity indices of the crosslinked micelles prepared at ultra dilute conditions (dash line) and at a high concentration (solid line) were determined by DLS (Fig. 7c). The polydispersity index and the average hydrodynamic diameter (Dh) of the crosslinked micelles prepared at a high concentration was very close to that of the crosslinked micelles obtained using dilute conditions. Therefore, it was apparent that only UMMs were obtained from the assembly of the binary graft polymer PGMA-g-(PCEMA-r-MPEG), even at a high polymer concentration.
This phenomenon was unexpected because intermolecular interactions were supposed to become dominant at a high concentration of GP. This may be because the GP had a particular hydrophilic/hydrophobic ratio that caused it to favor the exclusive assembly of UMMs. Takaya Terashima and coworkers have previously investigated the single-chain folding of amphiphilic random copolymers in water.20 They found that with a 20 mol% hydrophobic segment, all of copolymers formed unimolecular micelles in water, regardless of the hydrophobic pendant structures and the degree of polymerization, while those with over 50 mol% hydrophobic segment formed multichain aggregates. More importantly, the compact structure of the copolymers with a 20 mol% hydrophobic segment in water was maintained even at a relatively high concentration of 60 mg mL−1. They concluded this was because both the steric repulsion between multiple hydrophilic chains and the intramolecular hydrophobic interactions successfully isolated the polymer backbone in water to prevent intermolecular aggregation from occurring when the hydrophilic segment reached a certain mol% value. In this study, PGMA-g-(PCEMA-r-MPEG) could serve as an amphiphilic random copolymer bearing longer hydrophilic and hydrophobic pendant chains. The molar fraction of the hydrophobic PCEMA component relative to the side-chains was 11 mol% in this work, which was much smaller than the hydrophobic content of 20 mol% employed by Terashima and coworkers. Therefore, we suspected that the GP's ability to form UMMs even at a high concentration was probably due to the high proportion of its hydrophilic segment. However, it was also worth noting that, in contrast with random copolymers, the hydrophobic main chain of this graft copolymer accounted for a significant portion of its mass. Therefore, not only the molar ratio of PCEMA relative to the total number of side-chains, but also the mass fraction of the hydrophobic segments (PGMA and PCEMA) relative to the total mass of an individual molecule determined whether the final micelles would be unimolecular or multimolecular. In addition, the graft length could also influence the self-assembly behavior. It has been reported that unimolecular spherical micelles were obtained from graft copolymers with longer side-chains, and multimolecular spherical micelles were obtained from those with shorter side-chains.27 From the analysis above, we can conclude that the structure of precursor polymer was critical to control the intra- or intermolecular association through the external factors (concentration,33 switch rate of solvent quality,36 etc.) would have an effect on it.
This journal is © The Royal Society of Chemistry 2016 |