Patrícia V. Mendonçaa,
Maria João Morenob,
Arménio C. Serra
a,
Sérgio Simõesc and
Jorge F. J. Coelho
*a
aCEMUC, Department of Chemical Engineering, University of Coimbra, 3030-790 Coimbra, Portugal. E-mail: jcoelho@eq.uc.pt
bCQC, Department of Chemistry, University of Coimbra, Largo D. Dinis, Rua Larga, 3004-535 Coimbra, Portugal
cBluepharma, Indústria Farmacêutica, SA, São Martinho do Bispo, 3045-016 Coimbra, Portugal
First published on 23rd May 2016
This work reports the synthesis of tailor-made polymeric bile acid sequestrants (BAS) by supplemental activator and reducing agent atom transfer radical polymerization (SARA ATRP) using ecofriendly conditions. The new materials were based on amphiphilic poly(methyl acrylate)-b-poly((3-acrylamidopropyl)trimethylammonium chloride) (PMA-b-PAMPTMA) star block copolymers and cationic hydrogels (PAMPTMA). The in vitro sequestration ability of the polymers was investigated in simulated intestinal fluid using sodium cholate (NaCA) as the bile salt model molecule. Both polymeric structures investigated showed higher affinity towards NaCA micelles than unimers. The cationic hydrogels proved to be attractive BAS candidates, with binding parameters similar to those of the most effective commercial BAS: Colesevelam hydrochloride. Several polymer features were investigated for the star block copolymers in order to understand the structure/performance relationship. It was found that the binding parameters can be tuned by targeting different compositions of the block copolymers and, typically, longer cationic arms led to enhanced binding capacity.
There are several reports in the literature dedicated to the preparation of BAS.10–16 Cationic hydrogels based on (meth)acrylates, vinyl polymers, allyl polymers, poly(meth)acrylamides and polyethers are the most common used polymeric structures. Also, polymers based on polystyrene backbones (as a hydrophobic segment), containing amine/ammonium pendant groups (as a cationic hydrophilic segment) have been employed.2,3,17 It is known that the binding process is mainly ruled by both electrostatic and hydrophobic interactions.18 For this reason, an appropriate balance between the charged groups and the hydrophobic segments of the polymer is required in order to achieve high maximum binding capacity. In addition, the degree of crosslinking of the hydrogel (swelling characteristics) should be adjusted to allow the diffusion of bile acids through the polymer network. Despite the findings on the importance of controlling the BAS structure, these materials have been prepared mainly by conventional polymerization techniques, which hamper any fine control over the polymeric structures. To the best of our knowledge, there is only one work11,19 which describes the controlled synthesis of polystyrene-b-poly(N,N,N-trimethylammoniumethylene acrylamide chloride) block copolymers by anionic polymerization, as new BAS candidates in the form of micelles. However, the reaction conditions employed were very strict (e.g., T = −78 °C) and it was necessary to perform post-polymerization reactions to obtain the cationic segment. Reversible deactivation radical polymerization (RDRP) techniques could be a very useful tool for the design of tailor-made BAS. By this means, it could be also possible to deepen the understanding of the polymer structure–performance relationship.
The aim of this work was to prepare alternative structures to the commercial hydrogels, namely amphiphilic star block copolymers, using an ecofriendly supplemental activator and reducing agent atom transfer radical polymerization (SARA ATRP) system to afford a fine control over the polymers structure and molecular weight. To the best of our knowledge, this is the first time that such polymer structures are studied as potential BAS. Controlled poly((3-acrylamidopropyl)trimethylammonium chloride) (PAMPTMA)-based cationic hydrogels were also synthesized by SARA ATRP to assess the influence of the architecture on the performance of the materials. The binding capacity of the synthesized polymers was evaluated by in vitro equilibrium binding experiments with sodium cholate (NaCA).
Colesevelam hydrochloride was kindly supplied by Bluepharma and it was used as received.
Dipentaerythritol hexakis(2-bromoisobutyrate) (6f-BiB) (yield = 11%),20 pentaerythritol tetrakis(2-bromoisobutyrate) (4f-BiB)20 (yield = 20%) and tris[2-(dimethylamino)ethyl]amine (Me6TREN)21 (yield = 90%) were synthesized according the procedures described in the literature. Both 1H and 13C NMR spectra of 6f-BiB, 4f-BiB, and Me6TREN are provided in the ESI (Fig. S1–S3,† respectively).
Metallic copper (Cu(0), d = 1 mm, Sigma Aldrich) was washed with HCl in methanol and subsequently rinsed with methanol and dried under a stream of nitrogen following the literature procedure.22
Methyl acrylate (MA) (Acros, 99% stabilized), was passed through a sand/alumina column before use in order to remove the radical inhibitor.
Purified water (Milli-Q®, Millipore, resistivity > 18 MΩ cm) was obtained by reverse osmosis.
Simulated gastric fluid (SGF) was prepared by dissolving sodium chloride (100 mg) and pepsin (160 mg) in hydrochloric acid (350 μL). The volume was adjusted to 50 mL with water to give a solution with pH = 1.2.
Simulated intestinal fluid (SIF) was prepared by dissolving dipotassium hydrogen phosphate (203 mg), potassium dihydrogen phosphate (182 mg) and pancreatin (500 mg) in water (40 mL). The pH was adjusted with 0.2 N NaOH or 0.2 N HCl to 6.8 and the volume was completed to 50 mL with water.
PMA samples were analyzed by a high performance size exclusion chromatography HPSEC; Viscotek (Viscotek TDAmax) with a differential viscometer (DV), right-angle laser-light scattering (RALLS, Viscotek), low-angle laser-light scattering (LALLS, Viscotek) and RI detectors. The column set consisted of a PL 10 mm guard column (50 × 7.5 mm2) followed by one Viscotek T200 column (6 μm), one MIXED-E PLgel column (3 μm) and one MIXED-C PLgel column (5 μm). HPLC dual piston pump was set with a flow rate of 1 mL min−1. The eluent (THF) was previously filtered through a 0.2 μm filter. The system was also equipped with an on-line degasser. The tests were done at 30 °C using an Elder CH-150 heater. Before the injection (100 μL), the samples were filtered through a PTFE membrane with 0.2 μm pore. The system was calibrated with narrow polystyrene standards. The dn/dc was determined as 0.063 for PMA. Molecular weight (MSECn) and Đ of the synthesized polymers were determined by Multidetectors calibration using OmniSEC software version: 4.6.1.354.
400 MHz 1H NMR spectra of reaction mixture samples were recorded on a Bruker Avance III 400 MHz spectrometer, with a 5 mm TIX triple resonance detection probe, in D2O or CDCl3. Conversion of monomers was determined by integration of monomer and polymer NMR signals using MestRenova software version: 6.0.2-5475.
Fourier transform infrared attenuated total reflection (FTIR-ATR) spectroscopy was performed using a Jasco, model 4000 UK spectrometer. The samples were analyzed with 64 scans and 4 cm−1 resolution, between 500 and 3500 cm−1.
The critical micellar concentration (CMC) of NaCA in SIF at both pH = 6.8 and pH = 7.6 at 37 °C was determined using a method described in the literature,23 based on the increase in fluorescence quantum yield and blue shift of the fluorescence emission spectra of NPN upon association with micelles (Fig. S4 and S5†). The total concentration of NPN was 1 μM, and the concentrations of NaCA were investigated between 0 and 40 mM. The fluorescent emission of NPN was followed using an excitation wavelength of 356 nm on a Cary Eclipse fluorescence spectrophotometer from Varian (Victoria, Australia) equipped with a thermostated multi-cell holder accessory. The CMC was determined to be 6.7 and 6.0 for pH = 6.8 and pH = 7.6, respectively. Based on the asymptotic statistical theory,24 for a 95% confidence level, the confidence intervals obtained for the determination of the CMC values were [6.2–7.1] (CMC = 6.7) and [5.8–6.3] (CMC = 6.0).
The concentration of free NaCA remaining after the equilibrium binding experiments was determined by HPLC. The analyses were conducted at 25 °C using an Agilent Zorbax ODS, 5 μm, 4.6 × 250 mm column, under isocratic elution. The mobile phase used was 20 mM tetrabutylammonium hydroxide 30-hydrate (pH adjusted to 7.5 with phosphoric acid)/acetonitrile = 55/45 (v/v), with a flow rate of 0.7 mL min−1. The NaCA was detected by ultraviolet light at 210 nm. A NaCA standard calibration curve (R2 > 0.99) was constructed for the quantification of the NaCA present in the injected sample. For the quantification of low concentrations (0.1–3 mM), a 200 μL injection loop was used. For concentrations above 3 mM, a 20 μL injection loop was used. To check the system stability, one NaCA standard solution was injected after every 10 sample injections.
Isothermal titration calorimetry (ITC) analyses were performed on a VP-ITC instrument from MicroCal (Northampton, MA) with a reaction cell volume of 1410.9 μL, containing an amphiphilic star block copolymer solution, at 37 °C. A 5 mM solution of NaCA was injected at a speed of 0.5 μL s−1, stirring speed was 459 rpm, and the reference power was 10 μcal s−1. As recommended by the manufacturer, a first injection of 4 μL was performed before the experiment was considered to start in order to account for diffusion from/into the syringe during the equilibration period. The titration proceeded with additions of 10 μL per injection. All solutions were previously degassed for 15 min under reduced pressure.
The content of nitrogen in the amphiphilic star block copolymers was determined by elemental analysis using a Fisons – EA-1108 CHNS-O Element Analyzer.
![]() | (1) |
In the eqn (1), ws is the weight of the swollen sample after immersion and wd is the weight of the dry sample before immersion. The samples were analyzed in duplicate.
000 (Pall Corporation). The first 5 mL of filtrate were discarded and then an aliquot of 2.5 mL was collected for further analysis. The filtrates were analyzed by HPLC in order to determine the free concentration of NaCA. The amount of NaCA bonded to the polymers was calculated from the difference between the initial amount of NaCA introduced in the binding assays and the amount of free NaCA in the filtrates. The Hill equation (eqn (2)) was used to fit the experimental data, using the SigmaPlot10 software, in order to determine the binding parameters for each polymer.
![]() | (2) |
In the eqn (2), qe is the amount of NaCA bonded to the polymer in mg g−1 polymer, qmax is the apparent maximum amount of NaCA bonded to the polymer (binding capacity) in mg g−1, K is the intrinsic binding constant (relative to the strength of binding) in L mg−1, Ce is the free NaCA concentration in mg L−1 and n is the cooperative parameter (measure of the cooperativity of binding). For n = 1, the eqn (2) corresponds to the Langmuir isotherm (noncooperative binding).
![]() | ||
| Fig. 1 Generic structure of the amphiphilic star block copolymers (PMA-b-PAMPMTA) prepared by SARA ATRP. | ||
Star-shaped polymeric architectures with controlled structure can only be prepared by advanced polymerization techniques, such as ATRP. Previous publications by our research group26,27 reported the development of ecofriendly SARA ATRP systems for the synthesis of both PMA and PAMPTMA in ethanol/water mixtures near room temperature. In this work, the SARA ATRP method developed was extended to the preparation of the new star-shaped PMA-b-PAMPTMA block copolymers by “one-pot” polymerization in ethanol/water = 80/20 (v/v). It should be stressed that the direct copolymerization of MA and AMPTMA is a notable achievement, considering the very different nature of these two monomers. Due to solubility issues, it was not possible to find an appropriate solvent for the SEC analysis, in order to determine both block copolymers molecular weight and dispersity. Alternatively, the molecular weight of the block copolymers was estimated taking into account the molecular weight of the star-shaped PMA (first stage of the polymerization), which was determined by SEC, and the content of nitrogen from the PAMPTMA block on the final block copolymer. Additionally, the success of the copolymerization was confirmed by the appearance of the characteristic adsorption bands of both PMA and PAMPMTA blocks in the FTIR spectrum of the copolymers (Fig. 2): quaternary ammonium group at 1480 cm−1 and amide C
O stretching, amide II band and amide I band at 1550 cm−1 and 1650 cm−1, respectively, from the PAMPTMA and ester C
O stretching at 1729 cm−1 from the PMA.28
![]() | ||
| Fig. 2 FTIR spectra of PMA, PAMPTMA and amphiphilic star block copolymer (PMA-b-PAMPTMA) samples prepared by SARA ATRP. | ||
Block copolymers with different targeted properties, such as the number of arms, were prepared in order to evaluate the influence of the polymer structure/composition on the binding of bile acids. Table 1 summarizes the main features of the amphiphilic star block copolymers produced by SARA ATRP.
| Sample code | 1st block | 2nd block | Block copolymer | ||
|---|---|---|---|---|---|
| MA DP/arma | Đb | AMPTMA DP/arma | Number of arms | Mn × 10−3c | |
| a [Monomer]0/[initiator]0/number of arms.b Determined by SEC in THF at 30 °C.c The molecular weight of the second block (PAMPTMA) was estimated by elemental analysis. | |||||
| 4(PMA95-b-PAMPTMA50) | 95 | 1.09 | 50 | 4 | 74.8 |
| 4(PMA94-b-PAMPTMA24) | 94 | 1.13 | 24 | 4 | 52.9 |
| 4(PMA94-b-PAMPTMA16) | 94 | 1.07 | 16 | 4 | 46.3 |
| 4(PMA199-b-PAMPTMA52) | 199 | 1.08 | 52 | 4 | 112.2 |
| 6(PMA98-b-PAMPTMA52) | 98 | 1.09 | 52 | 6 | 116.2 |
| 6(PMA106-b-PAMPTMA38) | 106 | 1.09 | 38 | 6 | 103.0 |
| 6(PMA104-b-PAMPTMA18) | 104 | 1.10 | 18 | 6 | 77.2 |
The SARA ATRP method was also employed for the preparation of PAMPTMA-based hydrogels (Fig. 3), a common polymeric architecture used as BAS. It is expected that hydrogels prepared by ATRP present a more homogeneous structure that the ones obtained by conventional free radical polymerization techniques, due to the fast initiation and reversible deactivation reactions.29–31
ECP was chosen to initiate the polymerization since it was previously demonstrated that this is an effective initiator for the SARA ATRP of AMPTMA in ethanol/water mixtures.27 The monomer concentration was maintained in all the polymerizations ([AMPTMA]0 = 2 M) and two different diacrylate crosslinkers with distinct hydrophilicities were tested (BDDA and TEGDA). Different targeted DP values of both monomer and crosslinker were investigated and the main results are summarized in Table 2.
| Sample code | Crosslinker | AMPTMA DPb | Crosslinker DPb | AMPTMA conv. (%) | Crosslinker conv. (%) |
|---|---|---|---|---|---|
| a No macroscopic gelation was observed.b [AMPTMA or crosslinker]0/[ECP]0.c Data not available. | |||||
| AT 100/10 | TEGDA | 100 | 10 | 50 | 49 |
| AT 50/10 | TEGDA | 50 | 10 | c | c |
| AT 50/3a | TEGDA | 50 | 3 | c | c |
| AB 100/10 | BDDA | 100 | 10 | 35 | 35 |
| AB 50/10 | BDDA | 50 | 10 | 61 | 42 |
| AB 50/5 | BDDA | 50 | 5 | 61 | 63 |
| AB 50/3a | BDDA | 50 | 3 | c | c |
It was reported that the concentration of monomer influences the gelation point and ultimately it could prevent the formation of a crosslinked polymeric structure, due to the occurrence of intramolecular cyclization under dilute conditions.32 In this work, the AMPTMA concentration proved to be sufficient to allow macroscopic gelation (change from viscous liquid to elastic gel)29 of the polymer, for [crosslinker]0/[ECP]0 = 5 or 10 (Table 2). For a targeted ratio [crosslinker]0/[ECP]0 = 3, the polymer remained soluble after the predetermined time of reaction (24 h) (AT 50/3 and AB 50/3 in Table 2). In this case, the system did not reach the gel point, probably due to a low conversion of the crosslinker achieved after 24 h (data not available). Macroscopic gelation should occur when, on an average basis, each primary chain contains at least one crosslink point, meaning that the ratio [reacted crosslinker]/[initiator] should be one, considering a high initiation efficiency (each molecule of initiator originates one primary polymer chain).33
Regarding the AMPTMA conversion, the value achieved was not high as it was expected from the results obtained in the reported AMPTMA homopolymerization by SARA ATRP (conv. > 90% in less than 2 hours).27 Similarly, both crosslinkers did not reach very high conversion. This observation could mean that the reaction should be kept longer to allow full monomer conversion, considering the diffusion limitations experienced by monomers and catalytic complex in the polymer network after the gelation point.31 A control SARA ATRP reaction was allowed to proceed during seven days and led to similar results (compared with a 24 h reaction) in terms of both monomer and crosslinker conversions. In order to prepare PAMPMTA-based hydrogels with targeted properties, such as crosslink density, further investigation of the gelation process is required. This includes kinetic studies and the determination of both the gelation points and the gel fractions for different hydrogels. Nevertheless, the results are reproducible and this work clearly demonstrates the potential of the SARA ATRP as versatile tool for the preparation of cationic hydrogels.
| Sample code | R2 | K (mM−1) | n | qmaxa (mg g−1) × 10−3 |
|---|---|---|---|---|
| a mg NaCA per g polymer. | ||||
| Colesevelam | 0.9771 | 0.260 ± 0.090 | 1.0 | 2.23 ± 0.24 |
| AT 100/10 | 0.9949 | 0.17 ± 0.005 | 5.4 ± 0.9 | 1.63 ± 0.04 |
| AT 50/10 | 0.9868 | 0.22 ± 0.011 | 7.3 ± 4.3 | 1.45 ± 0.10 |
| AB 100/10 | 0.9988 | 0.22 ± 0.004 | 6.5 ± 0.8 | 1.73 ± 0.03 |
| AB 50/10 | 0.9983 | 0.22 ± 0.004 | 5.9 ± 0.6 | 1.49 ± 0.02 |
| AB 50/5 | 0.9962 | 0.17 ± 0.003 | 10.7 ± 2.3 | 1.82 ± 0.06 |
In this work, Colesevelam was used as a reference material since it is the most effective commercial BAS present in the market. The equilibrium binding of NaCA by Colesevelam (black symbols in Fig. 4) followed a typical Langmuir isotherm (Hill equation with n = 1, meaning no cooperativity), as reported by other authors for the binding of conjugated bile salts.36 The strength of binding of NaCA by the cationic hydrogels prepared was comparable to that of Colesevelam, as judged by the similar K values (Table 3). In addition, the cooperative parameter was higher than 1 for all the hydrogels, which indicates positive cooperativity in the binding. This parameter remained relatively constant, regardless the type of crosslinker used.
In order to assess the overall efficacy of the prepared polymers, it is important to evaluate not only the binding capacity (qmax), but also the dissociation constant (Kd = 1/K). This will provide information about the ability of the polymers to bind NaCA molecules, under the physiological conditions. Kd, which represents the concentration of bile salt at which the polymer is half-saturated, should be much lower than the concentration of bile salts in the intestine in order to provide an effective binding. In average, the total concentration of bile salts in the human intestine is above 15 mM.37 Therefore, the PAMPTMA-based hydrogels revealed to be promising BAS candidates (Kd = [4.5–5.9] mM), with an affinity towards NaCA similar to the one of Colesevelam. In addition, the binding capacity of the hydrogels (Table 3) is at least two times higher than the one reported in the literature for other hydrogel-based BAS candidates, as well as for the commercial BAS Cholestyramine, under similar conditions.13,15,35
As previously mentioned, due to the limited monomer conversions achieved, some of the hydrogel samples prepared by SARA ATRP presented similar composition (different from the one targeted) which did not allow the investigation of the polymer composition/binding parameters relationship. Besides the hydrogels composition, their swelling capacity can also influence the binding parameters. It was reported that a more elastic polymer network could promote the cooperativity of the binding.14 The swelling capacities of the hydrogels prepared by SARA ATRP were determined in duplicate and the results (Table 4) showed that the swelling capacity of the hydrogels prepared by SARA ATRP is at least two times higher than the one of the commercial BAS Colesevelam, regardless the crosslinker used. Considering the final application, one could expect that these new hydrogels could form a softer gel in the intestine environment, which can decisively contribute to better patient compliance (less constipation).
| Sample code | Crosslinker | AMPTMA DPa | Crosslinker DPa | Swelling capacity × 10−3 (%) |
|---|---|---|---|---|
| a Determined by 1H NMR spectroscopy.b Data not available. | ||||
| Colesevelam | — | — | — | 0.64 ± 0.07 |
| AT 100/10 | TEGDA | 50 | 4.9 | 2.11 ± 0.47 |
| AT 50/10 | TEGDA | b | b | 1.77 ± 0.09 |
| AB 100/10 | BDDA | 35 | 3.5 | 1.96 ± 0.08 |
| AB 50/10 | BDDA | 30 | 6.1 | 2.00 ± 0.29 |
| AB 50/5 | BDDA | 30 | 3.0 | 2.42 ± 0.33 |
| Sample code | R2 | K (mM−1) | n | qmaxa (mg g−1) × 10−3 |
|---|---|---|---|---|
| a mg NaCA per g polymer. | ||||
| Colesevelam | 0.9771 | 0.260 ± 0.090 | 1.0 | 2.23 ± 0.24 |
| 4(PMA95-b-PAMPTMA50) | 0.9968 | 0.172 ± 0.003 | 7.0 ± 0.9 | 1.37 ± 0.03 |
| 4(PMA94-b-PAMPTMA24) | 0.9997 | 0.129 ± 0.004 | 2.0 ± 0.1 | 0.95 ± 0.02 |
| 4(PMA94-b-PAMPTMA16) | 0.9791 | 0.215 ± 0.025 | 2.6 ± 0.6 | 0.90 ± 0.07 |
| 4(PMA199-b-PAMPTMA52) | 0.9971 | 0.301 ± 0.012 | 4.2 ± 0.5 | 0.80 ± 0.02 |
| 6(PMA98-b-PAMPTMA52) | 0.9905 | 0.215 ± 0.013 | 3.4 ± 0.6 | 1.43 ± 0.08 |
| 6(PMA106-b-PAMPTMA38) | 0.9896 | 0.172 ± 0.009 | 4.8 ± 1.1 | 1.36 ± 0.05 |
| 6(PMA104-b-PAMPTMA18) | 0.9987 | 0.172 ± 0.013 | 1.6 ± 0.1 | 0.64 ± 0.02 |
In addition, the cooperative parameter was lower than the one observed for the hydrogels (compare Tables 3 and 5), suggesting that either larger polymer chains in solution or higher density of cationic charges (or a combination of both factors) could enhance the cooperativity of binding. In fact, it is interesting to note that even within the same set of star block copolymers, the cooperativity increases with the increase of both the size of the copolymer and the number of positive charges. Regarding the size of the polymers in solution, these results suggest that polymer networks could be more beneficial for the binding process, in comparison to some nanoassembled polymer structures, probably due to their high charge density and due to the fact that the swollen hydrogels could facilitate the diffusion of the bile salt molecules and their access to the polymer binding sites.
Nevertheless, the results also highlight the importance of fine-tuning the polymers composition, since it was possible to significantly increase the maximum binding capacity by changing, for instance, the size of the PMA hydrophobic core of the star (e.g., red and pink symbols in Fig. 5). Such control over the polymers features is only possible by using RDRP techniques, as the one investigated in this work (SARA ATRP).
As previously mentioned, the library of block copolymers synthesized in this work was designed to allow the investigation of the polymers structure/sequestration performance relationship. Fig. 6 shows the variation of the maximum binding capacity (qmax) with the increase of the hydrophilic cationic arm length (AMPTMA), for both 4-arm and 6-arm star block copolymers with a fixed core size (MA DP/arm = 100). Also, the concentration of free bile salt in solution at 95% of polymer saturation (Ce0.95) was calculated from the binding parameters and it is shown in Fig. 6. This parameter, which depends on both K and n, provides a good estimate of the overall efficacy of the polymers under physiological conditions, for comparison purposes. Generally, it seems that there is a tendency of the star polymers to be more effective BAS for longer cationic segments (AMPTMA), regardless the number of arms of the star. This conclusion is supported by the increase of the cooperativity (considering similar K values; see details on Table 5), as well as by the decrease of the Ce0.95 (black symbols in Fig. 6) to values below the physiological concentration of bile salts (Ce0.95 < 15 mM). It is also worth to mention that the significant increase of the efficacy of the star polymers is observed for different lengths of the cationic segment, depending on the number of arms of the star. In addition, the increase of the hydrophilic cationic arm length led to an increase of the maximum binding capacity (red and green symbols in Fig. 6), regardless the number of arms of the stars. These results confirm that the electrostatic interactions play an important role in the binding process. The same trend was also observed by other authors11 for BAS candidates based on cationic micelles.
The influence of the balance between the cationic charges (PAMPMTA) and the hydrophobic regions (PMA) of the star block copolymers on the binding capacity, was also investigated (Fig. 7). The increase of the ionic density of the polymers led to an increase of the maximum NaCA binding capacity, regardless the number of arms of the stars. These results suggest that electrostatic interactions are the major driving force for the association between the bile salt and the polymer at high bile salt concentrations. Nevertheless, the interpretation of the results should be done with caution, since the binding parameters could not have a linear variation with the polymers features.12–14 However, the results presented in this work suggest that the cationic monomer (AMPTMA) binds NaCA with high affinity under conditions mimicking the intestine environment.
This work represents the first approach towards the design of BAS based on new polymeric structures (amphiphilic star block copolymers) and shows that the advanced polymerization techniques could be valuable tools for the understanding of the binding process.
In this work, ITC was used for the preliminary investigation of the binding thermodynamics between the NaCA molecules and the amphiphilic block copolymers. This technique has been employed for the study of the binding of bile salts by several receptors.38–40 Contrary to the determination of the binding isotherms by HPLC, ITC can provide valuable information about the binding process, such as the binding enthalpy change.41 The preliminary ITC experiments were conducted using an initial concentration of NaCA below the CMC to avoid the formation of micelles, which could difficult the analysis of the titration results. However, in those conditions, all the star block copolymers showed to have low affinity towards the NaCA molecules (Fig. S6†). Therefore, the titration curves did not allow the determination of the binding parameters. Nevertheless, it was possible to observe that the enthalpy of binding was positive (endothermic process), suggesting that the binding process was mainly governed by hydrophobic interactions, for the low NaCA concentration investigated (below 1 mM). Additionally, the heat involved in the binding was strongly dependent on the buffer ionization enthalpy indicating changes on the ionization of the NaCA upon binding, corresponding to an increase in the NaCA neutral form (Fig. S7†). These results indicate the prevalence of hydrophobic interactions under the conditions investigated, contrary to what is proposed by the binding model reported in the literature.13,14
It is worth to mention that the interpretation of the ITC results should be done with caution, since the NaCA concentration investigated corresponds to the low affinity region of the binding isotherm. Further analyses will be needed to clarify the binding mechanism. One of the parameters that could influence a cooperative binding process is the starting concentration of both polymer and NaCA, which ultimately will affect their conformation in solution (e.g., formation of micelles) and their affinity towards each other. To test this hypothesis, different initial concentrations of a selected hydrogel sample were used for the equilibrium binding assays (Fig. 8).
The cooperative binding effect showed to be dependent on the concentration of polymer used in the binding assay (Fig. 8(a)). However, it was possible to observe that the abrupt increase of the binding capacity (qe) of the hydrogel occurred for the same concentration of free NaCA in the solution (Ce) (Fig. 8(b)), irrespectively of the polymer concentration. The region of high affinity was observed for starting concentrations of NaCA above 7 mM (Fig. 8(a)), which is above its CMC (NaCACMC (pH = 6.8) = 6.7 mM).42 Therefore, the results obtained suggest that instead of the existence of a cooperative binding process, the behavior observed could be in fact the result of a stronger interaction between the PAMPTMA-based materials and NaCA micelles (electrostatic interaction), than with NaCA unimers (both hydrophobic and electrostatic interactions). This aspect could be advantageous considering BAS applications, since the bile salts are usually present in the intestine at concentrations above their CMC.37 Based on these results, it seems that the cooperative mechanism proposed in the literature13,14 is not the most suitable one to describe the binding of NaCA by the AMPTMA-based materials prepared by SARA ATRP. In this case, one might consider the existence of two different binding mechanisms corresponding to NaCA concentrations below and above the bile salt CMC.
As previously mentioned, the star block copolymers prepared by SARA ATRP present a PMA-based hydrophobic core and hydrophilic cationic arms, with high charge density (one charge per each AMPTMA repeating unit). The NaCA is an amphiphilic molecule that possesses both nonpolar (steroid skeleton) and polar (negatively charged) regions. In this case, one might consider that the NaCA unimers (concentration below the CMC) can interact either with the hydrophobic part of the polymer, which is less accessible, or/and with the cationic arms of the polymer by electrostatic interactions. On this matter, the ITC results suggested that the binding of NaCA unimers by the star block copolymers is relatively weak (low affinity constant K1) and it is mainly governed by hydrophobic interactions with the nonpolar region of the polymer. This hypothesis was also confirmed by the low binding capacity obtained in the equilibrium binding experiments using [NaCA]0 < NaCACMC (e.g., Fig. 8(a)). In addition, data reported in the literature12 indicates that the binding of NaCA unimers can be suppressed in the presence of small electrolytes (e.g., NaCl). This fact might have also contributed for the low affinity showed by the PAMPTMA-based materials towards NaCA molecules below the CMC, due to the intrinsic high concentration of Cl− anions from the PAMPTMA segment. On the other hand, for initial concentrations of NaCA above its CMC, the NaCA micelles/aggregates formed will be able to compete with small Cl− anions and establish electrostatic interactions with the polymer, since the nonpolar segment of the NaCA will be in the micelles core. In this case the binding process will be characterized by an affinity constant K2, higher than K1, and could be described by the Langmuir isotherm. This hypothesis is currently under investigation.
During the course of this work, we have noticed that the NaCA precipitated from the SIF solution at pH = 6.8, after several hours, for [NaCA] ≥ 15 mM. In fact, it is known that the pKa of bile salts increases when they are in polar media43 or aggregated in the form of micelles.44 Therefore, most probably, there was a small fraction of NaCA molecules in their neutral form at pH = 6.8. To confirm the validity of the results obtained at pH = 6.8 and to understand the influence of the ionization state of the NaCA on the binding mechanism, new binding experiments were conducted at pH = 7.6. At this pH value, there is a larger fraction of NaCA in the ionized form both as monomers and when associated in micelles, and the solutions are stable for several days in the range of concentrations investigated (0.1–30 mM). Colesevelam hydrochloride and the hydrogel showing the best performance (AB 50/5) were chosen as model polymers. The results (Fig. S8†) showed that the shape of the binding isotherms was the same as the one obtained at pH = 6.8, which confirms the validity of the results obtained. In addition, it is interesting to note that a slightly lower binding capacity was obtained at pH = 7.6. This supports the interpretation that the driving force for the association is not simply the electrostatic interaction between the bile salt and the polymer. This observation is in agreement with information found in the literature.18
| Sample | Medium | Degradation time (h) | MSECn × 10−3 | Đ |
|---|---|---|---|---|
| PAMPTMA | SGF (pH = 1.2) | 0 | 19.6 | 1.09 |
| 2 | 19.6 ± 0.1 | 1.09 ± 0.01 | ||
| PMA | SGF (pH = 1.2) | 0 | 51.7 | 1.07 |
| 2 | 51.3 ± 0.7 | 1.07 ± 0.01 | ||
| PAMPTMA | SIF (pH = 6.8) | 0 | 19.6 | 1.09 |
| 3 | 19.4 ± 0.1 | 1.08 ± 0.01 | ||
| PMA | SIF (pH = 6.8) | 0 | 51.7 | 1.07 |
| 3 | 51.7 ± 0.8 | 1.08 ± 0.01 |
The results obtained by SEC suggest that there was no degradation of the polymers backbones, since there was no variation on the molecular weight values. Nevertheless, one must be careful in the interpretation of the SEC results, since degradation of some polymer side chains (amide and ester linkages) could have minimal influence in the polymers molecular weight value. Therefore, additional 1H NMR spectroscopy analyses were performed. The results showed that there was no degradation of the side chains of both PMA and PAMPTMA, since the ratio between the integral of the protons of the backbone and the protons of the side chains was constant for the samples studied (Fig. S9 and S10†).
Footnote |
| † Electronic supplementary information (ESI) available: ITC titration curves, binding parameters of the star block copolymers and 1H NMR spectra. See DOI: 10.1039/c6ra06087k |
| This journal is © The Royal Society of Chemistry 2016 |