Next Article in Journal
Self-Sensitization and Photo-Polymerization of Diacetylene Molecules Self-Assembled on a Hexagonal-Boron Nitride Nanosheet
Next Article in Special Issue
Erratum: Synergistic Effect of Binary Mixed-Pluronic Systems on Temperature Dependent Self-Assembly Process and Drug Solubility. Polymers 2018, 10, 105.
Previous Article in Journal
Neodymium Recovery by Chitosan/Iron(III) Hydroxide [ChiFer(III)] Sorbent Material: Batch and Column Systems
Previous Article in Special Issue
Elucidation of Spatial Distribution of Hydrophobic Aromatic Compounds Encapsulated in Polymer Micelles by Anomalous Small-Angle X-ray Scattering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Water-soluble Polyion Complex (PIC) Micelles Covered with Amphoteric Random Copolymer Shells with Pendant Sulfonate and Quaternary Amino Groups

Department of Applied Chemistry, University of Hyogo, 2167 Shosha, Himeji, Hyogo 671-2280, Japan
*
Author to whom correspondence should be addressed.
Polymers 2018, 10(2), 205; https://doi.org/10.3390/polym10020205
Submission received: 28 January 2018 / Revised: 15 February 2018 / Accepted: 17 February 2018 / Published: 19 February 2018
(This article belongs to the Special Issue Polymer Micelles)

Abstract

:
An amphoteric random copolymer (P(SA)91) composed of anionic sodium 2-acrylamido-2-methylpropanesulfonate (AMPS, S) and cationic 3-acrylamidopropyl trimethylammonium chloride (APTAC, A) was prepared via reversible addition-fragmentation chain transfer (RAFT) radical polymerization. The subscripts in the abbreviations indicate the degree of polymerization (DP). Furthermore, AMPS and APTAC were polymerized using a P(SA)91 macro-chain transfer agent to prepare an anionic diblock copolymer (P(SA)91S67) and a cationic diblock copolymer (P(SA)91A88), respectively. The DP was estimated from quantitative 13C NMR measurements. A stoichiometrically charge neutralized mixture of the aqueous P(SA)91S67 and P(SA)91A88 formed water-soluble polyion complex (PIC) micelles comprising PIC cores and amphoteric random copolymer shells. The PIC micelles were in a dynamic equilibrium state between PIC micelles and charge neutralized small aggregates composed of a P(SA)91S67/P(SA)91A88 pair. Interactions between PIC micelles and fetal bovine serum (FBS) in phosphate buffered saline (PBS) were evaluated by changing the hydrodynamic radius (Rh) and light scattering intensity (LSI). Increases in Rh and LSI were not observed for the mixture of PIC micelles and FBS in PBS for one day. This observation suggests that there is no interaction between PIC micelles and proteins, because the PIC micelle surfaces were covered with amphoteric random copolymer shells. However, with increasing time, the diblock copolymer chains that were dissociated from PIC micelles interacted with proteins.

Graphical Abstract

1. Introduction

A mixture of oppositely charged polyelectrolytes in water forms a water-insoluble polyion complex (PIC) due to attractive electrostatic interactions between the polymer chains [1]. Many researchers study polymer aggregates formed by electrostatic interactions. When oppositely charged diblock copolymers containing nonionic water-soluble poly(ethylene glycol) (PEG) and polyelectrolyte blocks are mixed in water, the polymers spontaneously form water-soluble PIC micelles covered with hydrophilic PEG shells [2,3]. PIC micelles can encapsulate charged compounds such as metal ions, proteins, and nucleic acid in the PIC core via electrostatic interactions [4,5,6,7,8,9,10,11]. Recently, water-soluble PIC micelles were prepared with hydrophilic poly(2-methacryloyloxyethyl phosphorylcholine) (PMPC) coronas without net charge [12,13,14]. PMPC has pendant phosphorylcholine groups with the same chemical structure as the hydrophilic part of phospholipids comprising cell membranes. The phosphorylcholine group comprises an anionic phosphonium anion and cationic quaternary amino groups. The charges in PMPC are neutralized within a single polymer chain. PMPC is a betaine polymer with good biocompatibility and antithrombogenicity [15,16,17].
Protein fouling on the surfaces of medical devices due to hydrophobic, electrostatic, and hydrogen bonding interactions causes deterioration of the functionality. Therefore, much attention has been given to the surface modification of medical devices using polymer coatings with antifouling properties. In general, protein antifouling polymers can bind water molecules strongly and are electrically neutral. When an antifouling polymer is coated on a substrate, proteins have minimal contact with the polymer on the substrate due to the presence of water molecules between the proteins and the polymers [18]. In particular, zwitterionic polymers can suppress protein adsorption, because they contain many bound water molecules [19]. Among these, betaine polymers effectively suppress protein adsorption [20,21,22]. Nanoparticles that are surface-modified with betaine polymers have an increased circulation time in the body compared to bare nanoparticles [23]. Shih et al. reported that amphoteric random copolymers with pendant anionic sulfonate and cationic quaternary amino groups inhibit protein adsorption for stoichiometrically charge neutralized compositions [24]. It is expected that amphoteric random copolymers can be applied to the surface modification of medical devices.
In the present study, we prepared an amphoteric random copolymer (P(SA)91) macro chain transfer agent (CTA) composed of the same amounts of anionic sodium 2-acrylamido-2-methylpropanesulfonate (AMPS, S) and cationic 3-acrylamidopropyl trimethylammonium chloride (APTAC, A) via reversible addition-fragmentation chain transfer (RAFT) radical polymerization [25]. AMPS and APTAC were polymerized using P(SA)91 macro-CTA to prepare an anionic diblock copolymer (P(SA)91S67) and a cationic diblock copolymer (P(SA)91A88), respectively (Figure 1). The subscripts in the abbreviations indicate the degree of polymerization (DP). Water-soluble PIC micelles were prepared by stoichiometrically charge neutralized mixing of the oppositely charged diblock copolymers, P(SA)91S67 and P(SA)91A88, in water. To the best of our knowledge, this is the first report on such micelles. The PIC micelles inhibit interactions with proteins, because their surface is covered with amphoteric random P(SA)91 copolymer shells.

2. Materials and Methods

2.1. Materials

2-Acrylamido-2-methylpropanesulfonic acid (AMPS, 98%) from Tokyo Chemical Industry (Tokyo, Japan) and 4,4’-azobis(4-cyanopentanoic acid) (V-501, 98%) from Wako Pure Chemical (Osaka, Japan) were used as received without further purification, and 3-acrylamidopropyl trimethylammonium chloride (APTAC, 75 wt % in water) from Tokyo Chemical Industry was passed through an Aldrich (St Louis, MO, USA) disposable inhibitor removal column. Methanol was dried with molecular sieves 3A and purified by distillation. Phosphate buffered saline (PBS) was prepared by dissolving one PBS tablet (Aldrich) in predetermined amounts of water, while 4-cyanopentanoic acid dithiobenzoate (CPD) was synthesized according to a reported method [26]. Bovine serum albumin (BSA, pH 5.0–5.6 solution) from Wako Pure Chemical and fetal bovine serum (FBS) from GE Healthcare Life Sciences HyClone were used without further purification. Water was purified using an ion-exchange column. Other reagents were used as received.

2.2. Preparation of P(SA)91

AMPS (10.0 g, 48.3 mmol) was neutralized with 6 M aqueous NaOH (10.9 mL) to adjust the pH to 6.0. APTAC (9.98 g, 48.3 mmol), V-501 (135 mg, 0.482 mmol), and CPD (270 mg, 0.965 mmol) were dissolved in a mixed solvent of MeOH (5 mL) and water (32 mL), which was added to the aqueous AMPS. The solution was degassed by purging with Ar for 30 min. Polymerization was carried out at 70 °C for 24 h. After the reaction, the total conversion of AMPS and APTAC estimated from 1H NMR was 90.9%. The reaction mixture was dialyzed against 1.5 M aqueous NaCl for one day, and then pure water for one day using a dialysis membrane with a molecular weight cutoff of 14 kDa (EIDIA Co. Ltd, Tokyo, Japan). The polymer (P(SA)91) was recovered by a freeze-drying technique (15.3 g, 71.8%). The theoretical degree of polymerization (DP(theory)) and number-average molecular weight (Mn(theory)) were 91 and 1.91 × 104 g/mol, respectively, estimated from the conversion and molar ratio of monomer to CPD (Formulars 1 and 2). The number-average molecular weight (Mn(GPC)) and molecular weight distribution (Mw/Mn) were 1.54 × 104 g/mol and 1.27, respectively, estimated from gel-permeation chromatography (GPC). The APTAC content was 50 mol%, estimated from quantitative 13C NMR measurements. The synthesis route of P(SA)91 is shown in Figure S1.

2.3. Preparation of P(SA)91S67

AMPS (771 mg, 3.72 mmol) was neutralized with 6 M aqueous NaOH (6.2 mL) to adjust the pH to 6.0. P(SA)91 (703 mg, 3.68 × 10−2 mmol, Mn(theory) = 1.91 × 104 g/mol, Mw/Mn = 1.27) and V−501 (4.15 mg, 1.48 × 10−2 mmol) were added to the aqueous solution. The solution was degassed by purging with Ar for 30 min. Polymerization was carried out at 70 °C for 24 h (conversion = 90.9%). The reaction mixture was dialyzed against 1.5 M aqueous NaCl for one day, and then pure water for one day. The polymer (P(SA)91S67) was recovered by a freeze-drying technique (1.23 g, 67.6%). The degree of polymerization (DP(NMR)) of the PAMPS block was 67, estimated from quantitative 13C NMR measurements. Mn(GPC) and Mw/Mn were 2.38 × 104 g/mol and 1.04, respectively, estimated from GPC. The synthesis route of P(SA)91S67 is shown in Figure S1.

2.4. Preparation of P(SA)91A88

APTAC (765 mg, 3.69 mmol), P(SA)91 (703 mg, 3.68 × 10−2 mmol, Mn(theory) = 1.91 × 104 g/mol, Mw/Mn = 1.27), and V-501 (4.14 mg, 1.48 × 10−2 mmol) were dissolved in water (7.1 mL). The solution was degassed by purging with Ar for 30 min. Polymerization was carried out at 70 °C for 24 h (conversion = 87.7%). The reaction mixture was dialyzed against 1.5 M aqueous NaCl for one day, and then pure water for one day. The polymer (P(SA)91A88) was recovered by a freeze-drying technique (1.10 g, 74.9%). The DP(NMR) of the PAPTAC block was 88, estimated from quantitative 13C NMR measurements. Mn(GPC) and Mw/Mn were 1.87 × 104 g/mol and 1.14, respectively, estimated from GPC. The synthesis route of P(SA)91A88 is shown in Figure S1.

2.5. Preparation of PIC micelles

P(SA)91S67 and P(SA)91A88 were dissolved separately in 0.1 M aqueous NaCl. The aqueous P(SA)91A88 was added to the P(SA)91S67 solution over a period of 5 min with stirring to prepare the PIC micelles. The mixing ratio of the polymers was represented by the molar fraction of cationic charge (f+ = [APTAC]/([AMPS] + [APTAC])). The PIC micelles were prepared at f+ = 0.5 unless otherwise noted, which represents complete charge neutralization.

2.6. Measurements

The 1H NMR and inverse-gated decoupling 13C NMR spectra were obtained in D2O using a Bruker (Yokohma, Japan) DRX-500 spectrometer. GPC measurements were performed using a Jasco (Tokyo, Japan)UV-2075 detector equipped with a Shodex (Tokyo, Japan) OHpak SB-804 HQ column working at 40 °C with a flow rate of 0.6 mL/min. An acetic acid (0.5 M) solution containing sodium sulfate (0.3 M) was used as the eluent. Sample solutions were filtered with a 0.2 µm pore size membrane filter. Mn and Mw/Mn for the polymers were calibrated using standard poly(2-vinylpyridine) (P2VP) samples. Dynamic light scattering (DLS) measurements were obtained using a Malvern (Worcestershire, UK) Zetasizer Nano ZS with a He-Ne laser (4 mW at 633 nm) at 25 °C. The hydrodynamic radius (Rh) was calculated using the Stokes-Einstein equation, Rh = kBT/(6πηD), where kB is the Boltzmann constant, T is the absolute temperature, and η is the solvent viscosity. The DLS data was analyzed using Malvern Zetasizer software version 7.11. The angular dependence of DLS and static light scattering (SLS) was measured using an Otsuka Electronics Photal (Osaka, Japan) DLS-7000HL light scattering spectrometer equipped with a multi-τ digital time correlator (ALV-5000E), at 25 °C. A He-Ne laser (10 mW at 633 nm) was used as the light source [27,28,29]. The weight-average molecular weight (Mw), z-average radius of gyration (Rg), and second virial coefficient (A2) were estimated from SLS measurements [30]. Sample solutions for light scattering measurements were filtered using a membrane filter with 0.2 µm pores. The known Rayleigh ratio of toluene was used to calibrate the instrument. Plots of the refractive index increments against polymer concentration (dn/dCp) at 633 nm were obtained with an Otsuka Electronics Photal DRM-3000 differential refractometer at 25 °C. The zeta potential was measured using a Malvern Zetasizer Nano-ZS equipped with a He-Ne laser light source (4 mW, at 632.8 nm) at 25 °C. The zeta potential (ζ) was calculated from the electrophoretic mobility (μ) using the Smoluchowski relationship, ζ = ημ/ε (κa >> 1), where η is the viscosity of the solvent, ε is the dielectric constant of the solvent, and κ and a are the Debye-Hückel parameter and particle radius, respectively [31]. TEM was performed with a JEOL (Tokyo, Japan) JEM-2100 microscope at an accelerating voltage of 200 kV. A sample for TEM was prepared by placing one drop of the aqueous solution on a copper grid coated with a thin film of Formvar. Excess water was blotted using filter paper. The sample was stained with sodium phosphotungstate and dried under vacuum for one day.

3. Results and Discussion

3.1. Preparation of P(SA)91S67 and P(SA)91A88

The monomer conversions (p) of AMPS and APTAC could not be determined individually from 1H NMR measurements, because the peaks of vinyl groups in the monomers completely overlapped around 5.5–6.5 ppm. The sum of p for the AMPS and APTAC monomers was estimated from the integral intensity ratio of the vinyl groups compared to the sum of AMPS and APTAC pendant methylene protons at 3.0–3.5 ppm after polymerization. The monomer reactivity ratios for AMPS and APTAC are assumed to be one, because the polymerizable functional groups had the same chemical structure, i.e., acrylamide-type. When random copolymerization was performed to prepare P(SA)91 with equal concentrations of AMPS ([MAMPS]0) and APTAC ([MAPTAC]0), DP(theory) and Mn(theory) were calculated from the following formulas:
DP ( theory ) = 2 [ M AMPS ] 0 [ CTA ] 0 × p 100
M n ( theory ) = DP ( theory ) × MW MAV + MW CTA
where [CTA]0 is the initial concentration of CTA, MWMAV is the average molecular weight of AMPS and APTAC, and MWCTA is the molecular weight of CTA. When [MAMPS]0 = [MAPTAC]0, the total monomer concentration becomes 2[MAMPS]0. The values of DP(theory) and Mn(theory) are listed in Table 1.
The composition of P(SA)91 could not be determined from 1H NMR, because the peaks of the AMPS and APTAC units overlapped (Figure S2). To determine the compositions of P(SA)91, quantitative inverse gated decoupling 13C NMR measurements were performed in D2O (Figure 2a). The contents of AMPS and APTAC in the amphoteric random copolymer were determined from the integral intensity ratio of the peaks at 57.6 and 64.2 ppm, attributed to the pendant methylene carbons in AMPS and APTAC, respectively. The compositions of AMPS and APTAC in P(SA)91 were both 50 mol %. The DP(NMR) values for the PAMPS and PAPTAC blocks in P(SA)91S67 and P(SA)91A88 were found to be 67 and 88, respectively, using quantitative 13C NMR measurements (Figure 2b,c). The Mn(NMR) values for P(SA)91S67 and P(SA)91A88 were close to the Mn(theory) values. However, the Mn(GPC) values for P(SA)91S67 and P(SA)91A88 deviated from the theoretical values, which indicates that there are interactions between the sample polymers and that the GPC column or the GPC standard sample (P2VP) was unsuitable for determining Mn(GPC) (Figure S3). The Mw/Mn values for P(SA)91S67 and P(SA)91A88 were relatively small (1.04–1.15), which suggests that the obtained polymers have well-controlled structures.

3.2. Preparation and Characterization of PIC Micelles

The Rh values for P(SA)91S67 and P(SA)91A88 in 0.1 M aqueous NaCl were 5.7 and 6.0 nm, respectively (Figure 3). These had unimodal distributions. The polydispersity indices (PDI) for P(SA)91S67 and P(SA)91A88 were 0.147–0.169. The Rh for the stoichiometrically charge neutralized mixture of P(SA)91S67 and P(SA)91A88 increased to 29.0 nm, which suggests the formation of PIC micelles. Assuming that the polymer main chain has a planar zig-zag structure, the expanded chain lengths of P(SA)91S67 and P(SA)91A88 are 39.5 and 44.8 nm, respectively. These are longer than the Rh of the PIC micelles (= 29.0 nm). This suggests that a PIC micelle has a simple spherical core-shell structure composed of a core formed from the anionic PAMPS and cationic PAPTAC blocks and an amphoteric P(SA)91 shell. The PDI for PIC micelles was narrower than for P(SA)91S67 and P(SA)91A88, which indicates that the PIC micelles have a uniform size.
The relationship between the relaxation rate (Γ) and light scattering angle (θ) was measured for PIC micelles in 0.1 M aqueous NaCl (Figure S4). The scattering vector (q) was calculated using q = (4πn/λ)sin(θ/2), where n is the refractive index of the solvent and λ is the wavelength of the light source (= 632.8 nm). The Γ-q2 plot was a straight line passing through the origin. Therefore, Rh was determined at θ = 90°, because the diffusion coefficient (D) was independent of θ.
The Rh, LSI, and zeta potential for the mixture of P(SA)91S67 and P(SA)91A88 at various f+ were measured in 0.1 M aqueous NaCl (Figure 4). The total polymer concentration in the solution was maintained at 1.0 g/L. We performed the LSI and zeta-potential experiments at f+ = 0 and 1 were performed at Cp = 10.0 g/L, because LSIs were too low to measure Rh and zeta-potential. The compensated LSIs at f+ = 0 and 1 are plotted in Figure 4a. At f+ = 0, the values indicated for aqueous P(SA)91S67, and the zeta potential was –11.4 mV due to the negative charge of the PAMPS block. At f+ = 1, the zeta potential for P(SA)91A88 was 8.40 mV due to the positive charge of the PAPTAC block. The zeta potential of the stoichiometrically charge neutralized mixture of P(SA)91S67 and P(SA)91A88 at f+ = 0.5 was close to 0 mV. At f+ = 0.5, the Rh and LSI had the highest values for PIC micelles. We will discuss the structure of PIC micelles in more detail, together with the results of the DLS, SLS, and TEM measurements, later in this paper.
The Rh and LSI of PC micelles were plotted against Cp (Figure S5). Rh remained almost constant at about 30 nm and was independent of Cp in the range 0.02 ≤ Cp ≤ 1.0 g/L. LSI increased linearly with Cp. These observations suggest that the shape and aggregation number (Nagg), which is the number of polymer chains that form one PIC micelle, may be constant and independent of Cp in this region (0.02–1.0 g/L). At Cp ≤ 0.02 g/L, the LSI for PIC micelles was too low to obtain Rh. We studied the time dependence on Rh and LSI for PIC micelles. The Rh and LSI values remained constant until at least 41 h (Figure S6).
SLS measurements for P(SA)91S67, P(SA)91A88, and PIC micelles were performed in 0.1 M aqueous NaCl (Table 2). Mw was estimated by extrapolating Cp and θ to zero in Zimm plots (Figure S7). Rg was estimated from the slope of θ at Cp → 0. A2 was estimated from the slope of Cp at θ → 0. The Nagg of the PIC micelles was 218, as calculated from the Mw ratio of the PIC micelles and the unimer states of P(SA)91S67 and P(SA)91A88. The Mw values of the unimer states of the block copolymers estimated from SLS were close to the weight-average molecular weight calculated from Mn(NMR) and Mw/Mn (Table 1). Rg/Rh is useful for determining the shape of molecular aggregates. The theoretical value of Rg/Rh for a homogeneous hard sphere is 0.778, while that of a random coil is about 1, and this increases substantially for a structure with a lower density and polydispersity, e.g., Rg/Rh = 1.5–1.7 for flexible linear chains, and Rg/Rh ≥ 2 for a rod [32]. The PIC micelles may be spherical, because Rg/Rh = 0.91, which is close to 1. The Rg/Rh ratios for P(SA)91S67 and P(SA)91A88 were above 1, which suggests that the block copolymer chains were relatively expanded due to electrostatic repulsions in the polyelectrolyte blocks in 0.1 M aqueous NaCl. In general, A2 relates to the interactions between a polymer chain and solvent. A large value of A2 indicates a good solvent; however, a small or negative value of A2 indicates a poor solvent [33,34]. For the PIC micelles, A2 was 1.27 × 10−6 cm3·mol/g, which was less than the corresponding values for P(SA)91S67 and P(SA)91A88 (3.73 × 10−4 and 3.77 × 10−4 cm3·mol/g). This indicates that the solubility of PIC micelles in 0.1 M NaCl decreased compared with those of the unimer states of the block copolymers because of the insoluble PIC core in the PIC micelles. The density (d) for P(SA)91S67, P(SA)91A88, and PIC micelles was calculated using the following formula:
d = M w N A × V
where V is the volume of the block copolymers or PIC micelles calculated from 4/3πRh3. The d values for P(SA)91S67 and P(SA)91A88 were 0.116 and 0.0914 g/cm3, respectively, while d was 0.122 g/cm3 for the PIC micelles, which was slightly larger than for the diblock copolymers in the unimer states. This suggests that the polymer chains in the PIC micelles were more densely packed than those in the unimer states, because the PIC micelle core was formed by strong electrostatic interactions.
TEM was performed on PIC micelles in 0.1 M aqueous NaCl (Figure 5). The average radius of a PIC micelle was 20.3 nm, estimated from TEM, which was smaller than Rh (= 29.0 nm) estimated from DLS, because the TEM sample was in the dry state. Recently, we have reported the solution properties of amphoteric random copolymers in aqueous solutions [25]. We found out that there are no interpolymer interactions of amphoteric random copolymers in 0.1 M NaCl aqueous solutions. Also, we confirmed that there are on interactions between P(SA)91 and anionic and cationic homopolymers using DLS measurements. From these findings, there are no interactions between amphoteric random copolymer shells in PIC micelles. Moreover, there are no interactions between amphoteric random copolymer shells and the PIC core. The PIC micelle system has the same chemical structure in the core and corona, as both domains have the same composition [35].
The critical micelle concentration (cmc) for PIC micelles in 0.1 M aqueous NaCl was determined from the relationship between the LSI ratio (I/I0) and Cp (Figure 6). I and I0 are the LSIs of the solution and solvent, respectively. Although Rh for the PIC micelles cannot be measured for Cp ≤ 0.02 g/L, I/I0 can be measured below 0.02 g/L. The crossing point of the linear portions in the low and high Cp regions was 0.002 g/L, as the cmc. Below the cmc, PIC micelles may dissociate into charge neutralized small aggregates composed of a P(SA)91S67/P(SA)91A88 pair [36]. PIC micelles thus have a cmc, suggesting that they are in a dynamic equilibrium state.
PIC micelles formed due to electrostatic interactions of the anionic and cationic blocks. When a low molecular weight salt such as NaCl is added to aqueous PIC micelles, their shape may change because of a screening effect [37,38]. Rh and LSI for the aqueous PIC micelles were plotted as a function of the NaCl concentration ([NaCl]) (Figure 7). At [NaCl] ≤ 0.7 M, Rh gradually decreased with increasing [NaCl], and Rh was 25.8 nm at [NaCl] = 0.7 M. Rh decreased rapidly at 0.7 M < [NaCl] ≤ 0.8 M, and Rh was about 12 nm at [NaCl] = 0.8 M. The LSI decreased almost monotonously with increasing [NaCl] until [NaCl] = 0.8 M. At [NaCl] > 0.8 M, the LSI was almost constant and independent of [NaCl]. This observation indicates that the PIC micelles partially dissociated due to the screening effect of NaCl. Rh was almost constant at 9–12 nm at [NaCl] > 0.8 M. The Rh values for the unimer states of P(SA)91S67 and P(SA)91A88 were 5.7–6.0 nm (Figure 3). The Rh values for PIC micelles at [NaCl] > 0.8 M were larger than those for the unimer states. These observations suggest that PIC micelles cannot dissociate into their unimer states at [NaCl] > 0.8 M, presumably because of salting out effects.

3.3. Interactions between PIC Micelles and Proteins

We evaluated the interactions between PIC micelles (0.1 g/L) and BSA (5.0 g/L) in PBS using DLS (Figure 8). Before mixing, the Rh distributions were unimodal, and the Rh values for PIC micelles and BSA were 29.0 and 4.9 nm, respectively. A mixed solution of PIC micelles and BSA was prepared in PBS, and then DLS was performed on the solution within 1 h. After mixing, the distribution became bimodal, with Rh = 5.1 and 31.5 nm. The Rh value of the large size distribution was close to the Rh (= 29.0 nm) of the PIC micelles. Before mixing, the LSI values for the PIC micelles and BSA were 1.9 and 2.0 Mcps, respectively. The LSI for the mixture of PIC micelles and BSA was 2.1 Mcps, which was close to the values before mixing. This indicates that the PIC micelles do not interact considerably with BSA.
The time dependences of Rh and LSI for the mixture of PIC micelles and BSA were determined (Figure S8). The Rh distributions and LSI remained almost constant for one day. After two days, the distribution became trimodal. A small third distribution peak with an Rh of about 200 nm was observed after two days. The large aggregates with Rh ≥ 200 nm may be formed from BSA and unit PIC of a pair of P(SA)91S67/P(SA)91A88 dissociated from the PIC micelles. After four days, the area of the third distribution peak increased, and Rh increased to 281 nm. The average Rh and LSI for the mixture were almost constant for three days; however, these increased after four days. These observations suggest that the interaction between PIC micelles and proteins increased with time. The block copolymer chains can dissociate from the PIC micelles because these are in a dynamic equilibrium state. The unit PICs of P(SA)91S67/P(SA)91A88 detached from the PIC micelles may interact with proteins due to electrostatic attractions. The total charges in the unit PIC are compensated. However, dangling charge loops may be formed in the complex of cationic PAMPTAC and anionic PAMSP blocks in the unit PIC (Figure S9). In the case of PIC micelles, the surface of the PIC core was completely covered with amphoteric random copolymer shells which have protein antifouling properties. On the other hand, in the case of the unit PIC, the dangling charge loops may be exposed. The dangling charge groups in the unit PIC strongly interacted with proteins to form large aggregates.
The interactions between PIC micelles (0.1 g/L) and FBS (40 g/L) in PBS were evaluated by DLS (Figure 9). While BSA is a negatively charged protein, FBS is a mixture of negatively and positively charged proteins. The Rh distribution for FBS was bimodal with Rh = 5.3 and 31.8 nm. Mixed PBS solutions of PIC micelles and FBS were prepared, and then DLS was performed within 1 h. The Rh values for the mixture of PIC micelles and FBS were similar to those for FBS. The LSI of the mixed solution was 2.1 Mcps, which was similar to the values for the PIC micelles (1.9 Mcps) and FBS (2.1 Mcps) before mixing. This indicates that there is no interaction between PIC micelles and FBS. The time dependences of Rh distributions for the mixture of PIC micelles and FBS were measured (Figure S10). After two days, the distribution became trimodal. A small third distribution peak with an Rh of about 200 nm was observed after two days, which suggests that the large aggregates were formed from FBS and unit PIC formed.

4. Conclusions

Oppositely charged diblock copolymers of anionic P(SA)91S67 and cationic P(SA)91A88 were prepared via RAFT using P(SA)91 macro-CTA. Water-soluble PIC micelles were formed from a stoichiometrically charge neutralized mixture of 0.1 M aqueous NaCl solutions of P(SA)91S67 and P(SA)91A88. The maximum values of Rh and LSI were observed when f+ was 0.5, and the zeta potential was close to 0 mV. The PIC micelles were in a dynamic equilibrium state with PIC micelles and charge neutralized small aggregates composed of a P(SA)91S67/P(SA)91A88 pair. The particle sizes in the mixture of PIC micelles and proteins in PBS remained almost constant for at least one day, which suggests that there is no interaction between PIC micelles and proteins. The PIC micelles were covered with amphoteric P(SA)91 shells that suppressed their interaction with proteins. However, the interactions of the diblock copolymer chains dissociated from PIC micelles and proteins increased with time. If the lengths of the PAMPS block in P(SA)91S67 and the PAPTSC block in P(SA)91A88 increased, the dynamic equilibrium may shift to form PIC micelles. Therefore, it is expected that the interaction of PIC micelles with long polyelectrolyte chains and proteins can be suppressed for a long time.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4360/10/2/205/s1; Figure S1: Synthesis routes of P(SA)91S67 and P(SA)91A88; Figure S2: 1H NMR spectra for (a) P(SA)91, (b) P(SA)91S67, and (c) P(SA)91A88 in D2O; Figure S3: GPC elution curves for (a) P(SA)91S67 and (b) P(SA)91A88 using an acetic acid (0.5 M) solution containing sodium sulfate (0.3 M) as an eluent; Figure S4: Relationship between the relaxation rate (Γ) and the square of the magnitude of the scattering intensity vector (q2) for PIC micelles at Cp = 1 g/L in 0.1 M aqueous NaCl at 25 °C; Figure S5: Hydrodynamic radius (Rh, ○) and light scattering intensity (LSI, △) of PIC micelles as a function of polymer concentration (Cp) in 0.1 M aqueous NaCl. Figure S6: Relationship between Rh (○) and light scattering intensity (LSI, △) as a function of time for PIC micelle with f+ = 0.5 at Cp = 1.0 g/L in 0.1 M NaCl aqueous solution; Figure S7: A typical Zimm plot for PIC micelles in 0.1 M aqueous NaCl at 25 °C; Figure S8: (a) Relationship between Rh (○) and light scattering intensity (LSI, △) as a function of time, and (b) Rh distributions for a mixture of PIC micelles/BSA at Cp = 0.1 g/L and [BSA] = 5.0 g/L in PBS at 25 °C; Figure S9: Conceptual illustration of dangling charge groups in the unit PIC of P(SA)91S67/P(SA)91A88; Figure S10: Rh distributions for mixture of PIC micelle/FBS at Cp = 0.1 g/L and [FBS] = 40 g/L in PBS at 25 °C.

Acknowledgments

This work was financially supported by a Grant-in-Aid for Scientific Research (25288101 and 16K14008) from the Japan Society for the Promotion of Science (JSPS), JSPS Bilateral Open Partnership Joint Research Projects, and the Research Program of “Dynamic Alliance for Open Innovation Bridging Human, Environment and Materials” in “Network Joint Research Center for Materials and Devices.”

Author Contributions

Rina Nakahata was a research student who performed the experimental work. Shin-ichi Yusa was responsible for analyzing the experimental data and writing the paper. All authors approved the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Michaels, A.S.; Miekka, R.G. Polycation-polyanion complexes: preparation and properties of poly(vinylbenzyltrimethylammonium) poly(styrensulfonate). J. Phys. Chem. 1961, 65, 1765–1773. [Google Scholar] [CrossRef]
  2. Wibowo, A.; Osada, K.; Matsuda, H.; Anraku, Y.; Hirose, H.; Kishimura, A.; Kataoka, K. Morphology control in water of polyion complex nanoarchitectures of double-hydrophilic charged block copolymers through composition tuning and thermal treatment. Macromolecules 2014, 47, 3086–3092. [Google Scholar] [CrossRef]
  3. Yusa, S.; Yokoyama, Y.; Morishima, Y. Synthesis of oppositely charged block copolymers of poly(ethylene glycol) via reversible addition–fragmentation chain transfer (RAFT) radical polymerization and characterization of their polyion complex (PIC) micelles in water. Macromolecules 2009, 42, 376–383. [Google Scholar] [CrossRef]
  4. Oparin, A.I. Origin and evolution of metabolism. Comp. Biochem. Physiol. 1962, 4, 371–377. [Google Scholar] [CrossRef]
  5. Oishi, M.; Nagasaki, Y.; Itaka, K.; Nishiyama, N.; Kataoka, K. Lactosylated poly(ethylene glycol)-siRNA conjugate through acid-labile β-thiopropionate linkage to construct pH-sensitive polyion complex micelles achieving enhanced gene silencing in hepatoma cells. J. Am. Chem. Soc. 2005, 127, 1624–1625. [Google Scholar] [CrossRef] [PubMed]
  6. Van der Gucht, J.; Spruijt, E.; Lemmers, M.; Cohen Stuart, M.A. Polyelectrolyte complexes: Bulk phases and colloidal systems. J. Colloid Interface Sci. 2011, 361, 407–422. [Google Scholar] [CrossRef] [PubMed]
  7. Müller, M. Sizing, shaping and pharmaceutical applications of polyelectrolyte complex nanoparticles. Adv. Polym. Sci. 2014, 256, 197–260. [Google Scholar]
  8. Pergushov, D.V.; Mueller, A.H.; Schacher, F.H. Micellar interpolyelectrolyte complexes. Chem. Soc. Rev. 2012, 41, 6888–6901. [Google Scholar] [CrossRef] [PubMed]
  9. Voets, I.K.; de Keizer, A.; Stuart, M.A.C. Complex coacervate core micelles. Adv. Colloid Interface Sci. 2009, 147, 300–318. [Google Scholar] [CrossRef] [PubMed]
  10. Steinschulte, A.A.; Gelissen, A.P.H.; Jung, A.; Brugnoni, M.; Caumanns, T.; Lotze, G.; Mayer, J.; Pergushov, D.V.; Plamper, F.A. Facile screening of various micellar morphologies by blending miktoarm stars and diblock copolymers. ACS Macro Lett. 2017, 6, 711–715. [Google Scholar] [CrossRef]
  11. Dähling, C.; Lotze, G.; Mori, H.; Pergushov, D.V.; Plamper, F.A. Thermoresponsive segments retard the formation of equilibrium micellar interpolyelectrolyte complexes by detouring to various intermediate structures. J. Phys. Chem. B 2017, 121, 6739–6748. [Google Scholar] [CrossRef] [PubMed]
  12. Nakai, K.; Nishiuchi, M.; Inoue, M.; Ishihara, K.; Sanada, Y.; Sakurai, K.; Yusa, S. Preparation and characterization of polyion complex micelles with phoshobetaine shells. Langmuir 2013, 29, 9651–9661. [Google Scholar] [CrossRef] [PubMed]
  13. Nakai, K.; Ishihara, K.; Yusa, S. Preparation of giant polyion complex vesicles (G-PICsomes) with polyphosphobetaine shells composed of oppositely charged diblock copolymers. Chem. Lett. 2017, 46, 824–827. [Google Scholar] [CrossRef]
  14. Nakai, K.; Ishihara, K.; Kappl, M.; Fujii, S.; Nakamura, Y.; Yusa, S. Polyion complex vesicles with solvated phosphobetaine shells formed from oppositely charged diblock copolymers. Polymers 2017, 9, 49. [Google Scholar] [CrossRef]
  15. Iwasaki, Y.; Ishihara, K. Cell membrane-inspired phospholipid polymers for developing medical devices with excellent biointerfaces. Sci. Technol. Adv. Mater. 2012, 13, 064101. [Google Scholar] [CrossRef] [PubMed]
  16. Iwasaki, Y.; Ijuin, M.; Mikami, A.; Nakabayashi, N.; Ishihara, K. Behavior of blood cells in contact with water-soluble phospholipid polymer. J. Biomed. Mater. Res. 1999, 6, 360–367. [Google Scholar] [CrossRef]
  17. Ishihara, K.; Ueda, T.; Nakabayasi, N. Preparation of phospholipid polymers and their properties as hydrogel sheet. Polym. J. 1990, 22, 355–360. [Google Scholar] [CrossRef]
  18. Ostuni, E.; Chapman, R.G.; Holmlin, R.E.; Takayama, S.; Whitesides, G.M. A survey of structure-Property relationships of surfaces that resist the adsorption of protein. Langmuir 2001, 17, 5605–5620. [Google Scholar] [CrossRef]
  19. Holmlin, R.E.; Chen, X.; Chapman, R.G.; Takayama, S.; Whitesides, G.M. Zwitterionic SAMs that resist nonspecific adsorption of protein from aqueous buffer. Langmuir 2001, 17, 2841–2850. [Google Scholar] [CrossRef]
  20. Zhao, T.; Chen, K.; Gu, H. Investigations on the interactions of proteins with polyampholyte-coated magnetite nanoparticles. J. Phys. Chem. B 2013, 117, 14129–14135. [Google Scholar] [CrossRef] [PubMed]
  21. Chang, Y.; Chen, S.; Zhang, Z.; Jiang, S. Highly protein-resistant coatings from well-defined diblock copolymers containing sulfobetaines. Langmuir 2006, 22, 2222–2226. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, Z.; Chen, S.; Jiang, S. Dual-functional biomimetic materials: Nonfouling poly(carboxybetaine) with active functional groups for protein immobilization. Biomacromolecules 2006, 7, 3311–3315. [Google Scholar] [CrossRef] [PubMed]
  23. Muro, E.; Pons, T.; Lequeux, N.; Fragola, A.; Sanson, N.; Lenkei, Z.; Dubertret, B. Small and stable sulfobetaine zwitterionic quantum dots for functional live-cell imaging. J. Am. Chem. Soc. 2010, 132, 4556–4557. [Google Scholar] [CrossRef] [PubMed]
  24. Shih, Y.; Chang, Y.; Quemener, D.; Yang, H.; Jhong, J.; Ho, F.; Higuchi, A.; Chang, Y. Hemocompatibility of polyampholyte copolymers with well-defined charge bias in human blood. Langmuir 2014, 30, 6489–6496. [Google Scholar] [CrossRef] [PubMed]
  25. Nakahata, R.; Yusa, S. Solution properties of amphoteric random copolymers bearing pendant sulfonate and quaternary ammonium groups with controlled structures. Langmuir 2018. [Google Scholar] [CrossRef] [PubMed]
  26. Mitsukami, Y.; Donovan, S.M.; Lowe, B.A.; McCormick, L.C. Water-soluble polymers. 81. Direct synthesis of hydrophilic styrenic-based homopolymers and block copolymers in aqueous solution via RAFT. Macromolecules 2001, 34, 2248–2256. [Google Scholar] [CrossRef]
  27. Koppel, E.D. Analysis of macromolecular polydispersity in intensity correlation spectroscopy: The method of cumulants. J. Chem. Phys. 1972, 57, 4814–4820. [Google Scholar] [CrossRef]
  28. Huber, K.; Bantle, S.; Lutz, P.; Burchard, W. Hydrodynamic and thermodynamic behavior of short-chain polystyrene in toluene and cyclohexane at 34.5 °C. Macromolecules 1985, 18, 1461–1467. [Google Scholar] [CrossRef]
  29. Konishi, T.; Yoshizaki, T.; Yamakawa, H. On the “Universal Constants” ρ and Φ. of flexible polymers. Macromolecules 1991, 24, 5614–5622. [Google Scholar] [CrossRef]
  30. Zimm, H.B. Apparatus and methods for measurement and interpretation of the angular variation of light scattering; Preliminary results on polystyrene solutions. J. Chem. Phys. 1948, 16, 1099–1116. [Google Scholar] [CrossRef]
  31. Ali, I.S.; Heuts, A.P.J.; van Herk, A.M. Controlled synthesis of polymeric nanocapsules by RAFT-based vesicle templating. Langmuir 2010, 26, 7848–7858. [Google Scholar] [CrossRef] [PubMed]
  32. Akcasu, A.Z.; Han, C.C. Molecular weight and temperature dependence of polymer dimensions in solution. Macromolecules 1979, 12, 276–280. [Google Scholar] [CrossRef]
  33. Matsuda, Y.; Kobayashi, M.; Annaka, M.; Ishihara, K.; Takahara, A. Dimensions of a free linear polymer immobilized on silica nanoparticles of a zwitterionic polymer in aqueous solutions with various ionic strength. Langmuir 2008, 28, 8772–8778. [Google Scholar] [CrossRef] [PubMed]
  34. Cheng, L.; Hou, G.; Miao, J.; Chen, D.; Jiang, M.; Zhu, L. Efficient synthesis of unimolecular polymeric janus nanoparticles and their unique self-assembly behavior in a common solvent. Macromolecules 2008, 41, 8159–8166. [Google Scholar] [CrossRef]
  35. Savoji, M.T.; Strandman, S.; Zhu, X.X. Switchable vesicles formed by diblock random copolymers with tunable pH-and thermo-responsiveness. Langmuir 2013, 29, 6823–6832. [Google Scholar] [CrossRef] [PubMed]
  36. Anraku, Y.; Kishimura, A.; Oba, M.; Yamasaki, Y.; Kataoka, K. Spontaneous formation of nanosized unilamellar polyion complex vesicles with tunable size and properties. J. Am. Chem. Soc. 2010, 132, 1631–1636. [Google Scholar] [CrossRef] [PubMed]
  37. Santis, D.S.; Ladogana, D.R.; Diociaiuti, M.; Masci, G. Pegylated and thermosensitive polyion comples micelles by self-assembly of two oppositely and permanently charged diblock copolymers. Macromolecules 2010, 43, 1992–2001. [Google Scholar] [CrossRef]
  38. Maggi, F.; Ciccarelli, S.; Diociaiuti, M.; Casciardi, S.; Masci, G. Chitosan nanogels by template chemical cross-linking in polyion complex micelle nanpreactors. Biomacromolecules 2011, 12, 3499–3507. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Chemical structures of P(SA)91S67 and P(SA)91A88. (b) Schematic representation of water-soluble polyion complex (PIC) micelle formed from a mixture of P(SA)91S67 and P(SA)91A88; the PIC micelle shows protein antifouling properties because of its amphoteric random copolymer shell.
Figure 1. (a) Chemical structures of P(SA)91S67 and P(SA)91A88. (b) Schematic representation of water-soluble polyion complex (PIC) micelle formed from a mixture of P(SA)91S67 and P(SA)91A88; the PIC micelle shows protein antifouling properties because of its amphoteric random copolymer shell.
Polymers 10 00205 g001
Figure 2. Inverse gated decoupling 13C NMR spectra of (a) P(SA)91, (b) P(SA)91S67, and (c) P(SA)91A88 in D2O.
Figure 2. Inverse gated decoupling 13C NMR spectra of (a) P(SA)91, (b) P(SA)91S67, and (c) P(SA)91A88 in D2O.
Polymers 10 00205 g002
Figure 3. Hydrodynamic radius (Rh) distributions with polydispersity indices (PDI) for (a) P(SA)91S67; (b) P(SA)91A88; and (c) PIC micelles in 0.1 M aqueous NaCl.
Figure 3. Hydrodynamic radius (Rh) distributions with polydispersity indices (PDI) for (a) P(SA)91S67; (b) P(SA)91A88; and (c) PIC micelles in 0.1 M aqueous NaCl.
Polymers 10 00205 g003
Figure 4. (a) Hydrodynamic radius (Rh, ○) and light scattering intensity (LSI, △) of PIC micelles as a function of f+ (= [APTAC]/([APTAC] + [AMPS])) in 0.1 M aqueous NaCl; (b) Zeta potential of PIC micelles as a function of f+ in 0.1 M aqueous NaCl.
Figure 4. (a) Hydrodynamic radius (Rh, ○) and light scattering intensity (LSI, △) of PIC micelles as a function of f+ (= [APTAC]/([APTAC] + [AMPS])) in 0.1 M aqueous NaCl; (b) Zeta potential of PIC micelles as a function of f+ in 0.1 M aqueous NaCl.
Polymers 10 00205 g004
Figure 5. TEM image of PIC micelles with f+ = 0.5 at Cp = 1.0 g/L in 0.1 M aqueous NaCl.
Figure 5. TEM image of PIC micelles with f+ = 0.5 at Cp = 1.0 g/L in 0.1 M aqueous NaCl.
Polymers 10 00205 g005
Figure 6. Light scattering intensity ratio (I/I0) for the aqueous PIC micelles containing 0.1 M NaCl as a function of polymer concentration (Cp). I and I0 are the light scattering intensities of the solution and solvent, respectively.
Figure 6. Light scattering intensity ratio (I/I0) for the aqueous PIC micelles containing 0.1 M NaCl as a function of polymer concentration (Cp). I and I0 are the light scattering intensities of the solution and solvent, respectively.
Polymers 10 00205 g006
Figure 7. Hydrodynamic radius (Rh, ○) and light scattering intensity (LSI, △) of PIC micelles with f+ = 0.5 at Cp = 1.0 g/L as a function of NaCl concentration ([NaCl]).
Figure 7. Hydrodynamic radius (Rh, ○) and light scattering intensity (LSI, △) of PIC micelles with f+ = 0.5 at Cp = 1.0 g/L as a function of NaCl concentration ([NaCl]).
Polymers 10 00205 g007
Figure 8. Rh distributions of (a) PIC micelles; (b) BSA; and (c) a mixture of PIC micelles with BSA in PBS at 25 °C.
Figure 8. Rh distributions of (a) PIC micelles; (b) BSA; and (c) a mixture of PIC micelles with BSA in PBS at 25 °C.
Polymers 10 00205 g008
Figure 9. (a) Rh distributions for (a) PIC micelles, (b) FBS, and (c) a mixture of PIC micelles with FBS in PBS buffer at 25 °C.
Figure 9. (a) Rh distributions for (a) PIC micelles, (b) FBS, and (c) a mixture of PIC micelles with FBS in PBS buffer at 25 °C.
Polymers 10 00205 g009
Table 1. Degree of polymerization (DP), number-average molecular weight (Mn), and molecular weight distribution (Mn/Mn).
Table 1. Degree of polymerization (DP), number-average molecular weight (Mn), and molecular weight distribution (Mn/Mn).
SampleDP
(theory) a
Mn(theory) b × 104
(g/mol)
DP
(NMR) c
Mn(NMR) × 104
(g/mol)
Mn(GPC) × 104
(g/mol)
Mw/Mn
P(SA)91911.91- d- d1.541.27
P(SA)91S67613.30673.272.381.04
P(SA)91A88903.73883.701.871.15
a Theoretical degree of polymerization estimated from Formula 1. b Theoretical number-average molecular weight estimated from Formula 2. c Estimated from quantitative inverse-gated decoupling 13C NMR spectra in D2O. d The values could not be determined from 1H NMR, because the terminal phenyl protons overlapped with the pendant amino protons.
Table 2. Dynamic light scattering (DLS) and static light scattering (SLS) data for P(SA)91S67, P(SA)91A88, and PIC micelles in 0.1 M NaCl.
Table 2. Dynamic light scattering (DLS) and static light scattering (SLS) data for P(SA)91S67, P(SA)91A88, and PIC micelles in 0.1 M NaCl.
SamplesMw × 104
(g/mol)
NaggRg
(nm)
Rh
(nm)
Rg/RhA2 × 10−4
(cm3⋅mol/g2)
d a
(g/cm3)
dn/dCp
(mL/g)
P(SA)91S675.4117.35.71.283.730.1160.118
P(SA)91A884.9816.76.01.123.770.09140.138
PIC micelles75221826.529.00.910.01270.1220.128
a Estimated from the values of Mw and Rh using Formula 3.

Share and Cite

MDPI and ACS Style

Nakahata, R.; Yusa, S.-i. Preparation of Water-soluble Polyion Complex (PIC) Micelles Covered with Amphoteric Random Copolymer Shells with Pendant Sulfonate and Quaternary Amino Groups. Polymers 2018, 10, 205. https://doi.org/10.3390/polym10020205

AMA Style

Nakahata R, Yusa S-i. Preparation of Water-soluble Polyion Complex (PIC) Micelles Covered with Amphoteric Random Copolymer Shells with Pendant Sulfonate and Quaternary Amino Groups. Polymers. 2018; 10(2):205. https://doi.org/10.3390/polym10020205

Chicago/Turabian Style

Nakahata, Rina, and Shin-ichi Yusa. 2018. "Preparation of Water-soluble Polyion Complex (PIC) Micelles Covered with Amphoteric Random Copolymer Shells with Pendant Sulfonate and Quaternary Amino Groups" Polymers 10, no. 2: 205. https://doi.org/10.3390/polym10020205

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