Next Article in Journal
Statistical and Dynamical Properties of Topological Polymers with Graphs and Ring Polymers with Knots
Next Article in Special Issue
Cure Behavior and Thermomechanical Properties of Phthalonitrile–Polyhedral Oligomeric Silsesquioxane Copolymers
Previous Article in Journal
Particle Distribution of Solid Flame Retardants in Infusion Moulded Composites
Previous Article in Special Issue
A Rapid One-Pot Synthesis of Novel High-Purity Methacrylic Phosphonic Acid (PA)-Based Polyhedral Oligomeric Silsesquioxane (POSS) Frameworks via Thiol-Ene Click Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis and Self-Assembly of Amphiphilic Polyether-Octafunctionalized Polyhedral Oligomeric Silsesquioxane via Thiol-Ene Click Reaction

1
Key Laboratory of Advanced Packaging Materials and Technology of Hunan Province, Hunan University of Technology, Zhuzhou 412007, China
2
College of Chemistry and Chemical Engineering, Southeast University, Nanjing 211189, China
*
Author to whom correspondence should be addressed.
Polymers 2017, 9(7), 251; https://doi.org/10.3390/polym9070251
Submission received: 17 May 2017 / Revised: 12 June 2017 / Accepted: 26 June 2017 / Published: 28 June 2017
(This article belongs to the Collection Silicon-Containing Polymeric Materials)

Abstract

:
We demonstrated here a facile and efficient synthesis of polyhedral oligomeric silsesquioxane-based amphiphilic polymer by thiol-ene click chemistry. The properties of polyhedral oligomeric silsesquioxane (POSS)–PEG amphiphilic polymers were studied in detail by a combination of 1H NMR, 13C NMR, 29Si NMR FT-IR, GPC, and TG analysis. The newly-designed thiol-ene protocol obtains only anti-Markovnikov addition POSS-based amphiphilic polymers when compared with platinum-catalysed hydrosilylation method. The critical micelle concentration (CMC) of the resulting polymers are in the range of 0.011 to 0.050 mg/mL, and dynamic light scattering (DLS) results revealed that the obtained amphiphilic polymers can self-assemble into nanoparticles in aqueous solutions with a bimodal (two peaks) distribution. Furthermore, the specific polymer showed obvious thermo-sensitive behaviour at 45.5 °C.

Graphical Abstract

1. Introduction

In recent years, organic/inorganic hybrid nanomaterials have aroused widespread interest. Since they are combined with the traditional organic polymers’ easy workability, toughness, and thermal stability and oxidation resistance properties of inorganic compounds, this kind of materials play a key role in the development of high performance and high functional materials [1,2,3,4]. At present, there are many different types of nanocomposites, such as three-dimensional sol-gel materials [5,6,7,8,9], ceramic polymer [10,11,12], organic/inorganic polymer blend [13,14,15], and polyhedral oligomeric silsesquioxane (POSS) nanocomposites [16,17,18,19,20,21]. POSS-based polymer is a kind of nanoscale silicate functional material with cage-like segments directly bound to the polymer chains. Such newly-developed nanocomposite materials show synergistic properties of organic/inorganic materials, and present some new characteristics [22,23].
POSS has a nanoscale silicon core and eight organic functional groups on the surface, with sizes from 0.7 to 1.5 nm. The three-dimensional nanostructures of POSS can be used to build all types of hybrid materials with specific performance and controllable nanostructures [24]. Attracted by the fascinating vistas, POSS has potential applications in high-performance materials and biomedical fields [25,26,27,28], but its low water-solubility is the main impediment for real-world applications. Grafting hydrophilic groups on POSS is one of the effective methods to achieve water-soluble materials [29,30]. As is known to all, polyethylene glycol (PEO) is a water-soluble polymer with excellent properties, such as non-toxicity, lubricity, and biocompatibility. PEO is widely applied in biotechnology, in addition, it is a type of very important solid polymer electrolyte in lithium ion batteries because of its ability of dissolve the lithium ion [31]. Kim et al. [32] reported the synthesis and characterization of amphiphilic telechelics incorporating polyhedral oligosilsesquioxane (POSS–PEG). They obtained amphiphilic telechelic polymers with different POSS content by changing the PEG molecular weight, and studied the effect of POSS content on the crystallization behaviour of PEG. The results showed that different morphology and thermodynamic properties of the amphiphilic telechelic polymers could be obtained through controlling the ratio of hydrophilic PEG homopolymer and hydrophobic POSS macromolecular monomer. Subsequently, Kim et al. [33,34] systematically studied the polymer morphology, microstructure and rheological properties of POSS–PEG, and the impacts of structure and solvent polarity on bonding behavior in dilute solution using dilute solution viscosity method. Maitra et al. [35] synthesized a series of ethylene oxide oligomers functional silsesquioxane, reveals how the POSS silicon surface affect thermal behavior of ethylene oxide oligomers, Tg increases with the increase of the length of ethylene oxide chain, and inhibiting crystallization.
Although there have been many reports about the synthesis and application of polyethylene oxide functionalized silsesquioxane [36,37,38,39,40,41,42,43], but the preparation methods generally involve the Pt-catalyzed hydrosilylation between the terminal Si–H groups on the POSS cube with an unsaturated carbon double bond. The traditional method possesses some disadvantages: this method requires noble metal platinum, which is very expensive. It is difficult to get rid of the residues in purification process, which not only affects the product performance, but also limits its application in the field of biological medicine. This reaction generally requires strict water-free and oxygen-free conditions. The hydrosilylation reaction usually results in a mixture of Markovnikov and anti-Markovnikov addition products [44], which might complicate the structure-activity relationship of the amphiphilic copolymer. In addition, it is very difficult to obtain the high hydrophilic eight substitution product due to the limit of reaction efficiency and the influence of steric hindrance. Li et al. [45] successfully synthesize the eight functional POSS–PEG polymer using the copper-catalyzed alkyne-azide cycloaddition, which has faster reaction rate and higher yield. Nevertheless, the residues of heavy metal copper catalyst still have a certain influence on its application in the biomedical field. Therefore, a series of eight polyethylene oxide functionalized POSS amphiphilic polymers are synthesized through the metal-free thiol-ene click reaction, theirs self-assembly behaviour in aqueous solution, thermal stability, and thermo-responsive are systematically studied.

2. Experimental Section

2.1. Materials

Allyl-terminated polyether A (1000 g/mol, EO/PO = 70/30), B (1200 g/mol, EO/PO = 85/15), and C (2000 g/mol, EO/PO = 40/60) were provided by Jiangsu Maysta Chemical Co. Ltd. (Nanjing, China) 2,2-dimethoxy-2-phenylacetophenone (DMPA) was purchased from Aladdin (Shanghai, China) and used as received. Mercaptopropyltrimethoxysilane (MPT) and tetramethylammonium hydroxide ((CH3)4NOH) were purchased from J and K Chemical (Shanghai, China) and used without any further purification. Tetrahydrofuran (THF), diethyl ether, acetone, and concentrated hydrochloric acid were of analytical grade. In this work, the thiol-ene reaction was carried out under ambient conditions, the other reactions were carried out under a dry nitrogen atmosphere. The photo-induced reactions was carried out by a UV lamp (20 mW·cm−2, λ = 365 nm; LP-40A; LUYOR Corporation, Shanghai, China).

2.2. Synthesis of Octamercaptopropyl-POSS (POSS–SH)

POSS–SH is prepared according to our previous work [46]. The resulting viscous solution was dissolved in CH2Cl2 and then washed three times with H2O. The CH2Cl2 phase was dried with anhydrous Na2SO4 and concentrated to obtain POSS–SH in 76% yield. IR (KBr; ν, cm−1): 2929 (m), 2555 (w), 2849 (m), 1401 (m), 1122 (vs), 1032 (s), 562 (m), 472 (m). 1H NMR (300 MHz, CDCl3; δ, ppm): 0.76 (double, Si–CH2–), 1.71 (s,–CH2–), 2.56 (s, –CH2–S), 1.37 (br, –SH). 13C NMR (300 MHz, CDCl3; δ, ppm): 13.46 (Si–CH2–), 29.87 (Si–CH2–CH2–), 30.16 (–CH2–SH). 29Si NMR (400 MHz, CDCl3; δ, ppm): −67.20. GPC-result, Mn: 1005 g/mol, PDI: 1.02.

2.3. Synthesis of Polyether-Octafunctionalized POSS (POSS–PEG)

For the purpose of discussions, the POSS–PEG amphiphilic polymers containing different allyl-terminated polyethers are defined as P1, P2, and P3, respectively. The synthetic route of POSS–PEG amphiphilic polymer is shown in Scheme 1. The P1 was synthesized by thiol-ene click chemistry as follows: 0.5 g octamercaptopropyl-POSS (POSS–SH), 4.4 g allyl-terminated polyether A and 0.08 g DMPA were dissolved by 8 mL dry THF solvent. The system was stirred gently for 30 min under UV light irradiation, then placed in a dialysis bag (cutoff Mn: 3.5 kDa) and dialyzed against water for three days to get rid of residual polyether, replacing dialysate every 4 h and, finally, obtaining the target product by drying the inside mixture. Yield: 98%, 1H NMR (300 MHz, CDCl3, ppm): δ = 0.71 (m, SiCH2–), 1.08–1.13 (d, –CH3), 1.64 (t, –SiCH2CH2–), 1.81 (m, –SCH2CH2–), 2.51 (t, –CH2SCH2–), 3.27–3.85 (m, –OCH2CH2O–, –OCH2CH(CH3)O–). GPC-result, Mn: 13510 g/mol, PDI: 1.15.
The synthetic methods of P2 (polyether: Mn = 1200 g/mol, EO/PO = 85/15) and P3 (polyether: Mn = 2000 g/mol, EO/PO = 40/60) are identical to P1.

2.4. Characterizations

FT-IR was carried out using a Bruker TENSOR27 (Ettlingen, Germany) with the KBr pellet technique within the 4000 to 400 cm−1 region. 1H NMR was carried out on a Bruker Avance (Billerica, MA, USA) 300 and 400 MHz at 25 °C. The solvent was deuterated chloroform (CDCl3) without the interior label. The measurements of GPC were recorded on PL-GPC220 (Agilent, Palo alto, CA, USA) with THF as the eluent (1.0 mL/min) at 40 °C. TG analysis was performed with a STARe System (Mettler Toledo, Columbus, OH, USA). The heating rate was 20 °C/min under N2. TEM was recorded on a JEOL 2100 high-resolution TEM (JEOL Ltd., Tokyo, Japan) and Philips (CM 200, 80 kV, Amsterdam, The Netherlands).

2.5. Critical Micelle Concentration (CMC) Measurements

The CMC of the copolymers in ultrapure water was determined by fluorescence spectroscopy using pyrene as a hydrophobic fluorescent probe [47,48]. Briefly, a pyrene solution (6.2 × 10−6 M in acetone, 1 mL) was placed in different vials, and the solvent was evaporated. Then, different concentrations of copolymer (10 mL) were added to those vials, respectively. Thus, the final concentration of pyrene was 6.2 × 10−7 M and the concentrations of polymer micelles were from 1.0 × 10−4 to 1.0 g/L. Fluorescence measurements were carried out on a FluoroLog 3-TCSPC (Horiba Jobin Yvon Inc., Edison, NJ, USA) with a xenon light source (UXL-150S, Ushio, Japan). The slit widths of emission and excitation were 2 nm. The excitation spectra were recorded from 290 to 360 nm with the emission 373 nm. The emission spectra values I338 and I335, at 338 and 335 nm, respectively, were used for subsequent calculations. The CMC was determined by plotting the I338/I335 ratio against the polymer concentration. The CMC was taken as the intersection of regression lines calculated from the linear portions of the plot.

2.6. Dynamic Light Scattering (DLS)

The DLS measurements were performed by a Brookhaven BI-200SM instrument (Holtsville, NY, USA) equipped with a 4 mW He-Ne laser (λ = 532 nm) at an angle of 90°. The micelle copolymer solutions, with concentration 1 g/L, were prepared by direct dissolution of copolymers in ultrapure water. These samples, which were standing for 24 h before measurement, were filtered through 0.45 μm PTFE microfilters. The results of the test were recorded using the intensity-weighted distribution of particle sizes.

3. Results and Discussion

3.1. Characterization of Octamercaptopropyl-POSS (POSS–SH)

The absorption bands of Si–O–Si asymmetric stretching vibration are situated at 1109 and 1037 cm−1, showing in Figure S1. The large peak area may indicate that the obtained POSS–SH is composed primarily of Si–O–Si structure. The bands at 562 and 469 cm−1 are ascribed to the deformation vibration of POSS skeletal, and the stretching vibration of the bridge between caged silicon core and organic ligand Si–C is located at 693 cm−1. The bands at 2930, 2819, and 1407 cm−1 are assigned to the stretching and bending vibrations of –CH, while those at 1261 and 802 cm−1 are attributed to C–C stretching vibration. It is even more exciting to clearly see the characteristic absorption of –SH stretching vibration, which is located at 2555 cm−1.
1H NMR spectrum of POSS–SH (Figure S2) illustrates four strong resonance signals corresponding to different H atoms in its structure. The chemical shifts of these protons are assigned at 0.76 ppm (double, SiCH2–), 1.71 ppm (s, –CH2–), 2.56 ppm (s, –CH2S), and 1.37 ppm (br, –SH), which agree well with a, b, c and d groups on the side chains of the POSS–SH framework. 13C NMR spectrum (Figure S3) shows that the chemical shift at 13.46 ppm is assigned to the –CH2 group (carbon a) directly bound to the Si atom, while the others at 29.87 and 30.16 ppm are attributed to carbon b and c of the mercaptopropyl group, respectively. On the other hand, the 29Si NMR spectrum (Figure S3) reveals a sharp peak with a chemical shift at 67.20 ppm, ascribing to the Si atom of the POSS–SH. The results suggest that the cage-like Si–O–Si structure is well formed by the hydrolytic condensation of MTP and mercaptopropyl groups are symmetrically distributed on the cage-like POSS skeletal.

3.2. Characterization of Amphiphilic Polymers

Amphiphilic POSS–PEG polymers were synthesized by thiol-ene click chemistry with polyhedral oligomeric silsesquioxane and allylic polyether as starting materials.
In 1H NMR (Figure 1) of P1, all of the proton signal belonging to the polyether fragment and the POSS–SH fragment can be observed. The disappearance of S–H signal at 1.37 ppm and vinyl proton signal at 5.11 to 5.88 ppm confirmed the formation of P1. As previously reported, the thiol-ene reaction (Si–Vi and –SH) may give either of two possible products (Markovnikov and anti-Markovnikov) [49,50,51]. The anti-Markovnikov addition product is dominant. However, the result of 1H NMR indicates that there seem to be only anti-Markovnikov addition product in the current system. The number of epoxy propane (PO) unit in each polyether molecule was calculated by 1H NMR of allyl polyether, then the number of polyether chains grafted on POSS could be determined by the integral area ratio of methyl at 1.08–1.13 ppm and methylene at 0.71 ppm. The calculated results indicate that nearly eight polyether chains were grafted on each POSS, achieving entirely replace. However, due to the steric effect of high molecular weight polyether, amphiphilic POSS–PEG polymers synthesized by the hydrosilation method were difficult to be fully substituted. For example, Mya et al. [52] prepared POSS–PEG polymers by the hydrosilation, in which each polymer had only six polyether chains grafted onto silsesquioxane. The 1H NMR of P2 and P3 were similar to the result of P1.
The IR spectrum of P1 (Figure 2) clearly showed the symmetrical stretching vibration of Si–O–Si at 1111 cm−1, corresponding to the cage structure of silsesquioxane. The stretching vibration and bending vibration of Si–C were located at 1252 and 954 cm−1, respectively. Absorption bands at 954 and 843 cm−1 were characteristic peaks for Si–C groups on its outside surface and the crystallization phases of polyether. Characteristic peaks of thiol at 2555 cm−1, and those peaks of olefin C–H and C=C disappeared, which suggested the completion of the click reaction.
Further verification of the successful synthesis was obtained from GPC analyses. Figure 3 shows representative GPC profiles of the Allyl-terminated polyether A, B, and C and the subsequent polymers (P1, P2 and P3). A significant shift in the GPC traces (to a higher molecular weight) of the PEG and monomodal molecular weight distributions of the polymers indicates a successful click chemistry producing well-defined polymers. Furthermore, all of the three polymers have narrow molecular weight distribution, which are 1.15, 1.23, and 1.23, respectively. The molecular weight of P1, P2, and P3, which increases with the increasing of the molecular weight of polyether, was 13,510, 15,200, and 20,800 g/mol, respectively. Obviously, the number-average molecular weight of polymers is greater than those of the sum of POSS molecule and eight polyether segments. However, the difference between the number-average molecular weight of the three polymers (1690 and 5600 g/mol) is very consistent with the result of the octa-substituted polymer.

3.3. Self-Assembly Behaviour of Polymers

The micelle formation of the amphiphilic polymers was confirmed by fluorescence technique using pyrene as a probe. Figure 4 shows the fluorescence-excitation spectra of pyrene in POSS–PEG micelles at different concentrations. A red shift of the absorption peak from 335 to 338 nm was observed when the concentration of polymer was increased from 1.0 × 10−4 to 2.0 mg/mL. This red shift is attributed to the transfer of pyrene molecules from the hydrophilic environment to the hydrophobic micellar core. The results supply information on the location of the pyrene molecule in the system, that is, micelles are formed [53].
CMC is one of the key parameters for studying the formation of micelles and evaluating the stability of the resulting micelles. The intensity ratio of I338/I335 versus logC of the polymers in aqueous solution is shown in Figure 5. From these plots, the CMC of P1 was determined to be approximately 1.1 × 10−2 mg/mL through the intersection of two straight lines. The CMC of P2 and P3 were also obtained by the same method and shown in Figure S4. The CMC values of the three polymers increased as the proportion of polyether segment increased, which is consistent with the expected order (but it is different from polysiloxane system [54]). For polysiloxane system, because of the formation of vesicles and lamellae, the CMC values decreased as the proportion of polyether segment increased. Generally, amphiphilic polymers with higher content of the hydrophobic segments will cause stronger interactions between hydrophobic chains, leading to a more stable structure and, therefore, to a lower CMC value.
To study the self-assembly nanoparticle formation, the particle size was studied using dynamic light scattering, and the results are shown in Figure 6.
The DLS results showed that a bimodal (two peaks) distribution was observed at all concentrations. The small peak represents the unassociated unimolecular micelle with Rh = ~7 nm and the dominant peaks attributed to the scattering from the large particles attributable to the aggregation of POSS–PEG. The unassociated unimolecular micelle and aggregate micelle are in a state of dynamic balance. The aggregation peak is observed, even at a concentration of 0.5 mg/mL, suggesting that the CMC of POSS–PEG is lower than 0.5 mg/mL, which is in agreement with the fluorescence measurements (CMC = 0.011 mg/mL). Figure 6 shows that both of the micelle sizes in aqueous solution do not depend on the polymer concentration, indicating that the measurements were carried out in the dilute concentration regime and that the micelle formation followed a closed association mechanism [55]. This is also consistent with the previous literature [56]. The particle size obtained from TEM micrographs are smaller than those obtained by DLS (Figure 7). TEM micrographs show the micelle with the corona shrinkage after the evaporation of water, while DLS gives the size of the swollen nanoparticles in aqueous solutions. From the TEM images, there are two kinds of size of micelles in the three polymers; the small micelles are several nanometers and the large micelles are dozens of nanometers. At the same time, there is a slight difference of the morphological micelles between the three polymers. These results are consistent with the measurements of DLS.

3.4. Thermal Response Aggregation Behavior

POSS–PEG polymer has the temperature sensitivity because of the hydrogen bonding interaction with the water molecules, and the hydrogen bond is destroyed gradually with the increase of temperature. In particular, the polyether backbone containing structural units of propylene glycol becomes hydrophobic with the increment of temperature. Micelles formed by the POSS–PEG polymer are comprised of hydrophobic POSS as the core and a hydrophilic polyether chain as the outer shell, which optimizes the surface contacting with water and prevent the micelles to aggregate further. If this hydrophilic outer shell becomes less hydrophilic with the increase of temperature, the micelles will aggregate, then resulting in turbidity of solution. Thus, the polymer has a lower critical solution temperature (LCST). The thermo-sensitive behaviors of POSS–PEG nanoparticles were investigated by measuring the cloud point (CP). In order to ensure the formation of polymer nanoparticles, the concentrations of polymer aqueous solution were of 5 mg/mL.
The polymers P1 and P2 did not show a significant temperature response behaviour, they were colourless and transparent, while the P3 changed from a colourless solution to turbid white (Figure 8). This is mainly because of the different composition of the polyether. In P1 and P2, the EO/PO of the polyether chain were 70/30 and 85/15, in which the PO content dominated thermo-sensitive behaviours [57,58]. While in the polymer P3, the EO/PO was 40/60, thus P3 had obvious temperature sensitivity. The transmittance curve of the polymer P3 was shown in Figure 9. When the temperature was above CP, the transparent P3 aqueous solution suddenly became turbid, caused by the aggregation of P3 nanoparticles. P3 aqueous solution was sensitive to temperature, and the range of the transition temperature was less than 2 °C. It is worth noting that the thermo-sensitive behaviour of P3 nanoparticles was completely reversible. The transmittance of the P3 nanoparticles was about 100% at room temperature, and it was about 0 when heated to 60 °C, and when cooled to room temperature, it was nearly 100% again.

3.5. Thermal Stability Analysis

The thermodynamic behaviour of allyl-polyether and octa-substituted POSS in the nitrogen atmosphere was studied by TGA, and the results were shown in Figures S5 and S6, respectively. It could be found that the allyl-polyether was one step degradation, the initial decomposition temperature was from 270 to 310 °C, and the decomposition temperature increased with the increase of the molecular weight of polyether. These results were consistent with that reported results about oligomeric allyl-polyether (molecular weight ranges from 146 to 322 g/mol) [37]. The results showed that the high molecular weight allyl-polyether had good thermodynamic stability.
The thermal decomposition temperature of the polymers were not much different from that of the corresponding polyether, which indicated the thioether bond formed by click reaction had no effect on the thermal stability of the polyether. The thermal analysis results of the small molecular weight polyether substituted POSS polymer showed that the thermal decomposition process was divided into two steps [37]. The first step was the polyether chain segment at 164–268 °C, and the second step was the POSS cage to SiO2 in the range of 359–394 °C. However, the current POSS–PEG polymers were different. Their curves of TGA had only one major degradation event similar to high molecular weight polyether, which might attribute to approximate decomposition temperature of high molecular weight polyether and the first decomposition temperature of POSS–SH (Figure 10). Although the second decomposition temperature of POSS–SH (decomposition temperature of POSS cage) presented about 400 °C (Figure S6), its decomposition process was not displayed in the TGA curves because the contents of POSS in polymers were very low. When the temperature was up to 600 °C, the residual amounts of three allyl-polyethers were close to almost zero, while those of polymer P1, P2, and P3 were 5.5%, 4.0%, and 3.2%, respectively. The results showed that the contents of POSS in polymers were P1 > P2 > P3, which was consistent with the theoretical calculation results.

4. Conclusions

Three octa-substituted POSS amphiphilic polymers (P1, P2, and P3) were successfully prepared via a newly-designed thiol-ene click chemistry protocol. The results of 1H NMR, 13C NMR, 29Si NMR FT-IR, and gel permeation chromatography indicated the well-defined structures of the polymers. The experimental results showed that the amphiphilic polymers could form stable and bimodal nanoparticles directly by self-assembly in aqueous solution, and the formation of micelle followed closed association mechanism, in which the particle size was independent of the polymer concentration. Furthermore, the polymer P3 showed obvious temperature responsive behaviour. When the temperature rose to 45.5 °C, the solution quickly transformed into a turbid white solution and the range of the transition temperature was less than 2 °C. Thermal analysis showed that the polymer had excellent thermal stability and the presence of the thioether bond did not affect its thermal stability. Due to the excellent properties of POSS, facile synthesis, amphipathy, direct self-assembly to form polymer nanoparticles, and thermo-sensitive behaviour, this kind of polymer would shine in many fields, especially in biomedical engineering.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4360/9/7/251/s1, Figure S1: The infrared spectrum of POSS–SH, Figure S2: 1H NMR spectra of POSS–SH, Figure S3: 13C NMR spectra (a) and 29Si NMR spectra (b) of POSS–SH, Figure S4: Variation of the intensity ratio (I338/I335) as a function of P2 and P3 concentrations, the dotted line shows the CMC value, Figure S5: TGA curves of PEG-A, PEG-B, and PEG-C (in N2), Figure S6: TGA curve of POSS–SH (in N2).

Acknowledgments

The author wishes to acknowledge the financial supports from the National Natural Science Foundation of China (No. 11372108) and the Natural Science Foundation of Hunan Province (No. 14JJ5021).

Author Contributions

Yong Xia conceived and designed the experiments, and wrote articles, Sha Ding performed the experiments and analyzed the data, Yuejun Liu collected and analysed the data, and Zhengjian Qi supervised the project.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sanchez, C.; Belleville, P.; Popall, M.; Nicole, L. Applications of advanced hybrid organic–inorganic nanomaterials: From laboratory to market. Chem. Soc. Rev. 2011, 40, 696–753. [Google Scholar] [CrossRef] [PubMed]
  2. Descalzo, A.B.; Martinez-Manez, R.; Sancenon, R.; Hoffmann, K.; Rurack, K. The supramolecular chemistry of organic–inorganic hybrid materials. Angew. Chem. Int. Ed. 2006, 45, 5924–5948. [Google Scholar] [CrossRef] [PubMed]
  3. Kaushik, A.; Kumar, R.; Arya, S.K.; Nair, M.; Malhotra, B.D.; Bhansali, S. Organic–inorganic hybrid nanocomposite-based gas sensors for environmental monitoring. Chem. Rev. 2015, 115, 4571–4606. [Google Scholar] [CrossRef] [PubMed]
  4. Sharma, R.K.; Sharma, S.; Dutta, S.; Zboril, R.; Gawande, M.B. Silica-nanosphere-based organic–inorganic hybrid nanomaterials: Synthesis, functionalization and applications in catalysis. Green Chem. 2015, 17, 3207–3230. [Google Scholar] [CrossRef]
  5. Li, Z.; Chen, X.; Lin, C.; Zhao, N. The in vitro bioactive of sol–gel bioactive glass powders with three-dimensional lamellar structure. Adv. Powder Technol. 2012, 23, 13–15. [Google Scholar] [CrossRef]
  6. Costa, D.O.; Dixon, S.J.; Rizkalla, A.S. One- and three-dimensional growth of hydroxyapatite nanowires during sol–gel-hydrothermal synthesis. ACS Appl. Mater. Interfaces 2012, 4, 1490–1499. [Google Scholar] [CrossRef] [PubMed]
  7. Tiwari, J.N.; Tiwari, R.N.; Kim, K.S. Zero-dimensional, one-dimensional, two-dimensional and three-dimensional nanostructured materials for advanced electrochemical energy devices. Prog. Mater. Sci. 2012, 57, 724–803. [Google Scholar] [CrossRef]
  8. Pan, A.; Wu, H.B.; Yu, L.; Zhu, T.; Lou, X.W. Synthesis of hierarchical three-dimensional vanadium oxide microstructures as high-capacity cathode materials for lithium-ion batteries. ACS Appl. Mater. Interfaces 2012, 4, 3874–3879. [Google Scholar] [CrossRef] [PubMed]
  9. Kim, S.I.; Lee, J.S.; Ahn, H.J.; Song, H.K.; Jang, J.H. Facile route to an efficient nio supercapacitor with a three-dimensional nanonetwork morphology. ACS Appl. Mater. Interfaces 2013, 5, 1596–1603. [Google Scholar] [CrossRef] [PubMed]
  10. Ionescu, E.; Kleebe, H.J.; Riedel, R. Silicon-containing polymer-derived ceramic nanocomposites (PDC-NCs): Preparative approaches and properties. Chem. Soc. Rev. 2012, 41, 5032–5052. [Google Scholar] [CrossRef] [PubMed]
  11. Mera, G.; Gallei, M.; Bernard, S.; Ionescu, E. Ceramic nanocomposites from tailor-made preceramic polymers. Nanomaterials 2015, 5, 468–540. [Google Scholar] [CrossRef] [PubMed]
  12. Bhandavat, R.; Kuhn, W.; Mansfield, E.; Lehman, J.; Singh, G. Synthesis of polymer-derived ceramic Si(B) CN-carbon nanotube composite by microwave-induced interfacial polarization. ACS Appl. Mater. Interfaces 2011, 4, 11–16. [Google Scholar] [CrossRef] [PubMed]
  13. Ren, S.; Chang, L.Y.; Lim, S.K.; Zhao, J.; Smith, M.; Zhao, N.; Bulovic, V.; Bawendi, M.; Gradecak, S. Inorganic–organic hybrid solar cell: Bridging quantum dots to conjugated polymer nanowires. Nano Lett. 2011, 11, 3998–4002. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, Y.Y.; Ford, W.T.; Materer, N.F.; Teeters, D. Facile conversion of colloidal crystals to ordered porous polymer nets. J. Am. Chem. Soc. 2000, 122, 10472–10473. [Google Scholar] [CrossRef]
  15. Kim, J.; Kim, S.S.; Kim, K.H.; Jin, Y.H.; Hong, S.M.; San Hwang, S.; Cho, B.G.; Shin, D.Y.; Im, S.S. Applications of telechelic polymers as compatibilizers and stabilizers in polymer blends and inorganic/organic nanohybrids. Polymer 2004, 45, 3527–3533. [Google Scholar] [CrossRef]
  16. Wang, F.K.; Lu, X.H.; He, C.B. Some recent developments of polyhedral oligomeric silsesquioxane (POSS)-based polymeric materials. J. Mater. Chem. 2011, 21, 2775–2782. [Google Scholar] [CrossRef]
  17. Ghanbari, H.; Cousins, B.G.; Seifalian, A.M. A nanocage for nanomedicine: Polyhedral oligomeric silsesquioxane (POSS). Macromol. Rapid Commun. 2011, 32, 1032–1046. [Google Scholar] [CrossRef] [PubMed]
  18. Ayandele, E.; Sarkar, B.; Alexandridis, P. Polyhedral oligomeric silsesquioxane (POSS)-containing polymer nanocomposites. Nanomaterials 2012, 2, 445–475. [Google Scholar] [CrossRef] [PubMed]
  19. Lin, Y.; Jin, J.; Song, M.; Shaw, S.; Stone, C. Curing dynamics and network formation of cyanate ester resin/polyhedral oligomeric silsesquioxane nanocomposites. Polymer 2011, 52, 1716–1724. [Google Scholar] [CrossRef]
  20. Geng, Z.; Huo, M.; Mu, J.; Zhang, S.; Lu, Y.; Luan, J.; Huo, P.; Du, Y.; Wang, G. Ultra low dielectric constant soluble polyhedral oligomeric silsesquioxane (POSS)–poly (aryl ether ketone) nanocomposites with excellent thermal and mechanical properties. J. Mater. Chem. C 2014, 2, 1094–1103. [Google Scholar] [CrossRef]
  21. Li, Z.; Wu, D.; Liang, Y.; Fu, R.; Matyjaszewski, K. Synthesis of well-defined microporous carbons by molecular-scale templating with polyhedral oligomeric silsesquioxane moieties. J. Am. Chem. Soc. 2014, 136, 4805–4808. [Google Scholar] [CrossRef] [PubMed]
  22. Fatieiev, Y.; Croissant, J.; Julfakyan, K.; Deng, L.; Anjum, D.H.; Gurinov, A.; Khashab, N. Enzymatically degradable hybrid organic–inorganic bridged silsesquioxane nanoparticles for in vitro imaging. Nanoscale 2015, 7, 15046–15050. [Google Scholar] [CrossRef] [PubMed]
  23. Raftopoulos, K.N.; Pielichowski, K. Segmental dynamics in hybrid polymer/POSS nanomaterials. Prog. Polym. Sci. 2015, 7, 15046–15050. [Google Scholar] [CrossRef]
  24. Wu, J.; Mather, P.T. Poss polymers: Physical properties and biomaterials applications. Polym. Rev. 2009, 49, 25–63. [Google Scholar] [CrossRef]
  25. Croissant, J.G.; Cattoen, X.; Durand, J.O.; Man, M.W.C.; Khashab, N.M. Organosilica hybrid nanomaterials with a high organic content: Syntheses and applications of silsesquioxanes. Nanoscale 2016, 8, 19945–19972. [Google Scholar] [CrossRef] [PubMed]
  26. Au-Yeung, H.L.; Tam, A.Y.Y.; Leung, S.Y.L.; Yam, V.W.W. Supramolecular assembly of platinum-containing polyhedral oligomeric silsesquioxanes: An interplay of intermolecular interactions and a correlation between structural modifications and morphological transformations. Chem. Sci. 2017, 8, 2267–2276. [Google Scholar] [CrossRef] [PubMed]
  27. Lee, J.H.; Lee, A.S.; Lee, J.C.; Hong, S.M.; Hwang, S.S.; Koo, C.M. Multifunctional mesoporous ionic gels and scaffolds derived from polyhedral oligomeric silsesquioxanes. ACS Appl. Mater. Interfaces 2017, 9, 3616–3623. [Google Scholar] [CrossRef] [PubMed]
  28. Schafer, S.; Kickelbick, G. Simple and high yield access to octafunctional azido, amine and urea group bearing cubic spherosilicates. Dalton Trans. 2017, 46, 221–226. [Google Scholar] [CrossRef] [PubMed]
  29. Gupta, J. Nanotechnology applications in medicine and dentistry. J. Investig. Clin. Dent. 2011, 2, 81–88. [Google Scholar] [CrossRef] [PubMed]
  30. Chaloupka, K.; Malam, Y.; Seifalian, A.M. Nanosilver as a new generation of nanoproduct in biomedical applications. Trends Biotechnol. 2010, 28, 580–588. [Google Scholar] [CrossRef] [PubMed]
  31. Ismail, I.; Noda, A.; Nishimoto, A.; Watanabe, M. Xps study of lithium surface after contact with lithium-salt doped polymer electrolytes. Electrochim. Acta 2001, 46, 1595–1603. [Google Scholar] [CrossRef]
  32. Kim, B.S.; Mather, P.T. Amphiphilic telechelics incorporating polyhedral oligosilsesquioxane: 1. Synthesis and characterization. Macromolecules 2002, 35, 8378–8384. [Google Scholar] [CrossRef]
  33. Kim, B.S.; Mather, P.T. Morphology, microstructure, and rheology of amphiphilic telechelics incorporating polyhedral oligosilsesquioxane. Macromolecules 2006, 39, 9253–9260. [Google Scholar] [CrossRef]
  34. Kim, B.S.; Mather, P.T. Amphiphilic telechelics with polyhedral oligosilsesquioxane (POSS) end-groups: Dilute solution viscometry. Polymer 2006, 47, 6202–6207. [Google Scholar] [CrossRef]
  35. Maitra, P.; Wunder, S.L. Oligomeric poly(ethylene oxide)-functionalized silsesquioxanes: Interfacial effects on Tg, Tm, and ΔHm. Chem. Mater. 2002, 14, 4494–4497. [Google Scholar] [CrossRef]
  36. Lee, W.; Ni, S.L.; Deng, J.J.; Kim, B.S.; Satija, S.K.; Mather, P.T.; Esker, A.R. Telechelic poly(ethylene glycol)–POSS amphiphiles at the air/water interface. Macromolecules 2007, 40, 682–688. [Google Scholar] [CrossRef]
  37. Markovic, E.; Ginic-Markovic, M.; Clarke, S.; Matisons, J.; Hussain, M.; Simon, G.P. Poly(ethylene glycol)-octafunctionalized polyhedral oligomeric silsesquioxane: Synthesis and thermal analysis. Macromolecules 2007, 40, 2694–2701. [Google Scholar] [CrossRef]
  38. Markovic, E.; Matisons, J.; Hussain, M.; Simon, G.P. Poly(ethylene glycol) octafunctionalized polyhedral oligomeric silsesquioxane: Waxd and rheological studies. Macromolecules 2007, 40, 4530–4534. [Google Scholar] [CrossRef]
  39. Miao, J.J.; Cui, L.; Lau, H.P.; Mather, P.T.; Zhu, L. Self-assembly and chain-folding in hybrid coil-coil-cube triblock oligomers of polyethylene–b–poly(ethylene oxide)–b–polyhedral oligomeric silsesquioxane. Macromolecules 2007, 40, 5460–5470. [Google Scholar] [CrossRef]
  40. Mu, J.F.; Liu, Y.H.; Zheng, S.X. Inorganic-organic interpenetrating polymer networks involving polyhedral oligomeric silsesquioxane and poly(ethylene oxide). Polymer 2007, 48, 1176–1184. [Google Scholar] [CrossRef]
  41. Li, Z.B.; Tan, B.H.; Jin, G.R.; Li, K.; He, C.B. Design of polyhedral oligomeric silsesquioxane (POSS) based thermo-responsive amphiphilic hybrid copolymers for thermally denatured protein protection applications. Polym. Chem. 2014, 5, 6740–6753. [Google Scholar] [CrossRef]
  42. Wang, D.K.; Varanasi, S.; Strounina, E.; Hill, D.J.T.; Symons, A.L.; Whittaker, A.K.; Rasoul, F. Synthesis and characterization of a POSS–PEG macromonomer and POSS–PEG–PLA hydrogels for periodontal applications. Biomacromolecules 2014, 15, 666–679. [Google Scholar] [CrossRef] [PubMed]
  43. Mohamed, M.G.; Jheng, Y.R.; Yeh, S.L.; Chen, T.; Kuo, S.W. Unusual emission of polystyrene-based alternating copolymers incorporating aminobutyl maleimide fluorophore-containing polyhedral oligomeric silsesquioxane nanoparticles. Polymers 2017, 9, 103. [Google Scholar] [CrossRef]
  44. Mukbaniani, O.; Titvinidze, G.; Tatrishvili, T.; Mukbaniani, N.; Brostow, W.; Pietkiewicz, D. Formation of polymethylsiloxanes with alkyl side groups. J. Appl. Polym. Sci. 2007, 104, 1176–1183. [Google Scholar] [CrossRef]
  45. Li, Y.J.; Wan, L.Q.; Huang, F.R.; Du, L. Core–shell octa(azidopropyl) POSS–PEO micelle via “click” chemistry. J. Mol. Liq. 2014, 196, 238–243. [Google Scholar] [CrossRef]
  46. Xia, Y.; Yao, H.; Cui, M.; Ma, Y.; Kong, Z.; Wu, B.; Qi, Z.; Sun, Y. Theoretical and experimental investigations on mono-substituted and multi-substituted functional polyhedral oligomeric silsesquioxanes. RSC Adv. 2015, 5, 80339–80345. [Google Scholar] [CrossRef]
  47. Bakkour, Y.; Darcos, V.; Li, S.M.; Coudane, J. Diffusion ordered spectroscopy (DOSY) as a powerful tool for amphiphilic block copolymer characterization and for critical micelle concentration (CMC) determination. Polym. Chem. 2012, 3, 2006–2010. [Google Scholar] [CrossRef]
  48. Pineiro, L.; Novo, M.; Al-Soufi, W. Fluorescence emission of pyrene in surfactant solutions. Adv. Colloid Interface Sci. 2015, 215, 1–12. [Google Scholar] [CrossRef] [PubMed]
  49. Xue, L.; Wang, D.X.; Yang, Z.Z.; Liang, Y.; Zhang, J.; Feng, S.Y. Facile, versatile and efficient synthesis of functional polysiloxanes via thiol-ene chemistry. Eur. Polym. J. 2013, 49, 1050–1056. [Google Scholar] [CrossRef]
  50. Shipp, D.A.; McQuinn, C.W.; Rutherglen, B.G.; McBath, R.A. Elastomeric and degradable polyanhydride network polymers by step-growth thiol-ene photopolymerization. Chem. Commun. 2009, 6415–6417. [Google Scholar] [CrossRef] [PubMed]
  51. Rissing, C.; Son, D.Y. Thiol-ene reaction for the synthesis of multifunctional branched organosilanes. Organometallics 2008, 27, 5394–5397. [Google Scholar] [CrossRef]
  52. Mya, K.Y.; Li, X.; Chen, L.; Ni, X.P.; Li, J.; He, C.B. Core-corona structure of cubic silsesquioxane-poly(ethylene oxide) in aqueous solution: Fluorescence, light scattering, and tem studies. J. Phys. Chem. B 2005, 109, 9455–9462. [Google Scholar] [CrossRef] [PubMed]
  53. Kabanov, A.V.; Nazarova, I.R.; Astafieva, I.V.; Batrakova, E.V.; Alakhov, V.Y.; Yaroslavov, A.A.; Kabanov, V.A. Micelle formation and solubilization of fluorescent-probes in poly(oxyethylene-b-oxypropylene-b-oxyethylene) solutions. Macromolecules 1995, 28, 2303–2314. [Google Scholar] [CrossRef]
  54. Xia, Y.; Yao, H.T.; Miao, Z.H.; Ma, Y.; Cui, M.F.; Yan, L.Q.; Ling, H.H.; Qi, Z.J. Facile synthesis and self-assembly of amphiphilic polydimethylsiloxane with poly(ethylene glycol) moieties via thiol-ene click reaction. RSC Adv. 2015, 5, 50955–50961. [Google Scholar] [CrossRef]
  55. Chu, B. Laser Light Scattering: Basic Principles and Practice; Academic Press: New York, NY, USA, 1991. [Google Scholar]
  56. Hussain, H.; Tan, B.H.; Seah, G.L.; Liu, Y.; He, C.B.; Davis, T.P. Micelle formation and gelation of (PEG–P(MA-POSS)) amphiphilic block copolymers via associative hydrophobic effects. Langmuir 2010, 26, 11763–11773. [Google Scholar] [CrossRef] [PubMed]
  57. Louguet, S.; Rousseau, B.; Epherre, R.; Guidolin, N.; Goglio, G.; Mornet, S.; Duguet, E.; Lecommandoux, S.; Schatz, C. Thermoresponsive polymer brush-functionalized magnetic manganite nanoparticles for remotely triggered drug release. Polym. Chem. 2012, 3, 1408–1417. [Google Scholar] [CrossRef]
  58. Persson, J.; Kaul, A.; Tjerneld, F. Polymer recycling in aqueous two-phase extractions using thermoseparating ethylene oxide-propylene oxide copolymers. J. Chromatogr. B 2000, 743, 115–126. [Google Scholar] [CrossRef]
Scheme 1. The synthesis of POSS–PEG polymer.
Scheme 1. The synthesis of POSS–PEG polymer.
Polymers 09 00251 sch001
Figure 1. 1H NMR spectra of P1.
Figure 1. 1H NMR spectra of P1.
Polymers 09 00251 g001
Figure 2. The FT-IR spectra of P1.
Figure 2. The FT-IR spectra of P1.
Polymers 09 00251 g002
Figure 3. The GPC profiles of allyl-polyether and octa-substituted POSS polymers.
Figure 3. The GPC profiles of allyl-polyether and octa-substituted POSS polymers.
Polymers 09 00251 g003
Figure 4. Excitation spectra of pyrene in aqueous solution of P1 at various concentrations.
Figure 4. Excitation spectra of pyrene in aqueous solution of P1 at various concentrations.
Polymers 09 00251 g004
Figure 5. Variation of the intensity ratio (I338/I335) as a function of P1 concentrations, the dotted line shows the CMC value.
Figure 5. Variation of the intensity ratio (I338/I335) as a function of P1 concentrations, the dotted line shows the CMC value.
Polymers 09 00251 g005
Figure 6. Size distribution of nanoparticles formed by different concentration of P1 determined by DLS at 25 °C in aqueous solutions.
Figure 6. Size distribution of nanoparticles formed by different concentration of P1 determined by DLS at 25 °C in aqueous solutions.
Polymers 09 00251 g006
Figure 7. Typical TEM images of copolymers micelles (1 mg/mL), from left to right are P1, P2, and P3, respectively.
Figure 7. Typical TEM images of copolymers micelles (1 mg/mL), from left to right are P1, P2, and P3, respectively.
Polymers 09 00251 g007
Figure 8. Digital photographs of transmittance variation of POSS–PEG aqueous solution (5 mg/mL).
Figure 8. Digital photographs of transmittance variation of POSS–PEG aqueous solution (5 mg/mL).
Polymers 09 00251 g008
Figure 9. Temperature dependence of optical transmittance and reversible change of transmittance at 550 nm for P3 aqueous solution (5 mg/mL).
Figure 9. Temperature dependence of optical transmittance and reversible change of transmittance at 550 nm for P3 aqueous solution (5 mg/mL).
Polymers 09 00251 g009
Figure 10. TGA curves of allyl-polyether and octa-substituted POSS (in N2).
Figure 10. TGA curves of allyl-polyether and octa-substituted POSS (in N2).
Polymers 09 00251 g010

Share and Cite

MDPI and ACS Style

Xia, Y.; Ding, S.; Liu, Y.; Qi, Z. Facile Synthesis and Self-Assembly of Amphiphilic Polyether-Octafunctionalized Polyhedral Oligomeric Silsesquioxane via Thiol-Ene Click Reaction. Polymers 2017, 9, 251. https://doi.org/10.3390/polym9070251

AMA Style

Xia Y, Ding S, Liu Y, Qi Z. Facile Synthesis and Self-Assembly of Amphiphilic Polyether-Octafunctionalized Polyhedral Oligomeric Silsesquioxane via Thiol-Ene Click Reaction. Polymers. 2017; 9(7):251. https://doi.org/10.3390/polym9070251

Chicago/Turabian Style

Xia, Yong, Sha Ding, Yuejun Liu, and Zhengjian Qi. 2017. "Facile Synthesis and Self-Assembly of Amphiphilic Polyether-Octafunctionalized Polyhedral Oligomeric Silsesquioxane via Thiol-Ene Click Reaction" Polymers 9, no. 7: 251. https://doi.org/10.3390/polym9070251

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop