injectable and biodegradable supramolecular hydrogels by inclusion complexation between...

10
Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and αCyclodextrin Zhicheng Tian, Chen Chen, and Harry R. Allcock* Department of Chemistry, Pennsylvania State University, University Park, Pennsylvania 16802, United States ABSTRACT: Biodegradable poly(organophosphazenes) containing side chains of various oligo(ethylene glycol) methyl ethers (mPEGs) and glycine ethyl ester units were synthesized and characterized. Novel supramolecular-structured hydrogel systems based on the inclusion complex between the mPEG grafted polyphosphazenes and α-cyclodextrin were prepared in aqueous media. The gelation time depended on the length of the mPEG side chains, the molar ratio between mPEG repeat units and α-cyclodextrin, and the concentration of the polymeric gel precursors. The rheological measurements of the supramolecular hydrogels indicate a fast gelation process and owable character under a large strain. The hydrogel systems demonstrate unique structure-related reversible gelsol transition properties at a certain temperature due to the reversible supramolecular assembly. The formation of a channel-type inclusion complex induced gelation mechanism was studied by DSC, TGA, 13 C CP/MAS NMR, and X-ray diraction techniques. The strong potential of the system for injectable drug delivery applications was explored with the use of bovine serum albumin as a model protein for in vitro release studies. All the supramolecular hydrogels studied showed disintegration by dethreading of the α-cyclodextrin. Polymers with longer poly(ethylene glycol) side chains had better stability and slower protein release proles. The molecular weights of the polymers were monitored by GPC to show the biodegradability of the hydrogel system. INTRODUCTION Hydrophilic hydrogels are three-dimensional cross-linked networks of water-soluble polymers. These are of special interest due to their unique properties and potential applications, such as tissue engineering, drug delivery, and cell immobilization. 15 Water-soluble polymers can be cross- linked in various ways. Chemical cross-links can be accomplished either by a reaction between two functional polymers or by the addition of small molecule cross-linking agents. 6,7 Physical noncovalent cross-linking of polymer chains can be achieved by physicochemical interactions including hydrophobic interactions, charge condensation, hydrogen bonding, or stereocomplexation. 811 Although physical inter- actions are weaker than chemical cross-linking, they allow useful properties such as processability and reversibility. Cyclodextrins (CDs), usually α-, β-, and γ-CD, which consist of 6, 7, and 8 glucose units, respectively, have been widely studied to form inclusion complexes with oligomers or polymers including poly(ethylene glycol), 12 poly(propylene glycol), 13 poly(ε-caprolactone), 14 polyisobutylene, 15 poly(ε- lysine), 16 etc. 17 Supramolecular hydrogels can be obtained in water when α-CD interacts with poly(ethylene glycol)-grafted hydrophilic polymers (heparin or dextrin) 18,19 or poly(ethylene glycol)-containing block copolymers such as poly(ε-caprolac- tone)-b-poly(ethylene glycol)-b-poly(ε-caprolactone) or poly- (ethylene glycol)- b -poly( ε -caprolactone)- b -poly[2- (dimethylamino)ethyl methacrylate]. 20,21 In these cases, the hydrogen bonding generated by inclusion complexation functions as a physical cross-linking process. Since these hydrogels show desirable properties, it is interesting to develop other alternative approaches with facile syntheses, tunable properties, and good biological behavior, especially biocompat- ibility and tunable biodegradability. Polyphosphazenes are hybrid polymers with a backbone of alternating phosphorus and nitrogen atoms with two side groups attached to each phosphorus. 22 Their synthesis is based on the reactions of poly(dichlorophosphazene) which is prepared by a thermal ring-opening polymerization of hexachlorocyclotriphosphazene at 250 °C in a sealed system. 23 This reactive polymeric intermediate can then undergo facile and tunable substitution reactions through the replacement of chlorine atoms by various nucleophiles such as alkoxides, 24 aryloxides, 25 or amines, 26 and various combinations of the above. Hydrolytically sensitive polyphosphazenes are formed when amino acid ethyl ester groups are linked to the polymer backbone via the amino terminus. 27 The nontoxic hydrolysis products are the parent amino acids, ethanol, phosphates, and ammonia, a mixture that results in a buer system of near- Received: February 27, 2013 Revised: March 14, 2013 Published: March 22, 2013 Article pubs.acs.org/Macromolecules © 2013 American Chemical Society 2715 dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 27152724

Upload: harry-r

Post on 08-Dec-2016

218 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and α-Cyclodextrin

Injectable and Biodegradable Supramolecular Hydrogels byInclusion Complexation between Poly(organophosphazenes) andα‑CyclodextrinZhicheng Tian, Chen Chen, and Harry R. Allcock*

Department of Chemistry, Pennsylvania State University, University Park, Pennsylvania 16802, United States

ABSTRACT: Biodegradable poly(organophosphazenes) containingside chains of various oligo(ethylene glycol) methyl ethers (mPEGs)and glycine ethyl ester units were synthesized and characterized.Novel supramolecular-structured hydrogel systems based on theinclusion complex between the mPEG grafted polyphosphazenesand α-cyclodextrin were prepared in aqueous media. The gelationtime depended on the length of the mPEG side chains, the molarratio between mPEG repeat units and α-cyclodextrin, and theconcentration of the polymeric gel precursors. The rheologicalmeasurements of the supramolecular hydrogels indicate a fastgelation process and flowable character under a large strain. Thehydrogel systems demonstrate unique structure-related reversiblegel−sol transition properties at a certain temperature due to thereversible supramolecular assembly. The formation of a channel-type inclusion complex induced gelation mechanism was studiedby DSC, TGA, 13C CP/MAS NMR, and X-ray diffraction techniques. The strong potential of the system for injectable drugdelivery applications was explored with the use of bovine serum albumin as a model protein for in vitro release studies. All thesupramolecular hydrogels studied showed disintegration by dethreading of the α-cyclodextrin. Polymers with longerpoly(ethylene glycol) side chains had better stability and slower protein release profiles. The molecular weights of the polymerswere monitored by GPC to show the biodegradability of the hydrogel system.

■ INTRODUCTION

Hydrophilic hydrogels are three-dimensional cross-linkednetworks of water-soluble polymers. These are of specialinterest due to their unique properties and potentialapplications, such as tissue engineering, drug delivery, andcell immobilization.1−5 Water-soluble polymers can be cross-linked in various ways. Chemical cross-links can beaccomplished either by a reaction between two functionalpolymers or by the addition of small molecule cross-linkingagents.6,7 Physical noncovalent cross-linking of polymer chainscan be achieved by physicochemical interactions includinghydrophobic interactions, charge condensation, hydrogenbonding, or stereocomplexation.8−11 Although physical inter-actions are weaker than chemical cross-linking, they allowuseful properties such as processability and reversibility.Cyclodextrins (CDs), usually α-, β-, and γ-CD, which consistof 6, 7, and 8 glucose units, respectively, have been widelystudied to form inclusion complexes with oligomers orpolymers including poly(ethylene glycol),12 poly(propyleneglycol),13 poly(ε-caprolactone),14 polyisobutylene,15 poly(ε-lysine),16 etc.17 Supramolecular hydrogels can be obtained inwater when α-CD interacts with poly(ethylene glycol)-graftedhydrophilic polymers (heparin or dextrin)18,19 or poly(ethyleneglycol)-containing block copolymers such as poly(ε-caprolac-tone)-b-poly(ethylene glycol)-b-poly(ε-caprolactone) or poly-(ethylene glycol)-b -poly(ε -caprolactone)-b -poly[2-

(dimethylamino)ethyl methacrylate].20,21 In these cases, thehydrogen bonding generated by inclusion complexationfunctions as a physical cross-linking process. Since thesehydrogels show desirable properties, it is interesting to developother alternative approaches with facile syntheses, tunableproperties, and good biological behavior, especially biocompat-ibility and tunable biodegradability.Polyphosphazenes are hybrid polymers with a backbone of

alternating phosphorus and nitrogen atoms with two sidegroups attached to each phosphorus.22 Their synthesis is basedon the reactions of poly(dichlorophosphazene) which isprepared by a thermal ring-opening polymerization ofhexachlorocyclotriphosphazene at 250 °C in a sealed system.23

This reactive polymeric intermediate can then undergo facileand tunable substitution reactions through the replacement ofchlorine atoms by various nucleophiles such as alkoxides,24

aryloxides,25 or amines,26 and various combinations of theabove. Hydrolytically sensitive polyphosphazenes are formedwhen amino acid ethyl ester groups are linked to the polymerbackbone via the amino terminus.27 The nontoxic hydrolysisproducts are the parent amino acids, ethanol, phosphates, andammonia, a mixture that results in a buffer system of near-

Received: February 27, 2013Revised: March 14, 2013Published: March 22, 2013

Article

pubs.acs.org/Macromolecules

© 2013 American Chemical Society 2715 dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 2715−2724

Page 2: Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and α-Cyclodextrin

neutral pH.28,29 Polyphosphazenes have long been proposed forhydrogel applications, and several examples have beenpublished previously. For example, hydrogels were preparedby γ-radiation cross-linking of hydrophilic methoxyethoxye-thoxy containing polyphosphazene (MEEP) or MEEPderivatives.30,31 Poly(organophosphazenes) containing isoleu-cine ethyl ester and oligo(ethylene glycol) substituents showeda reversible sol−gel transition at tunable temperatures depend-ing on the polymer compositions.32,33 Ionically cross-linkedhydrogels were achieved by serine- or threonine- containinghydrophilic polyphosphazenes when the carboxylic acid groupswere exposed to Ca2+.34 pH-responsive hydrogels were alsogenerated from hydrophilic polyphosphazenes that bear arylcarboxylate and methoxyethoxyethoxy substituents.35 Bycombining the advantages of polyphosphazenes and supra-molecular hydrogels, a novel hydrogel system can be visualizedwith desirable properties for biomedical applications.In this work, biodegradable poly(organophosphazenes) have

been synthesized by grafting poly(ethylene glycol) methyl ether(mPEG) and about 20% of glycine ethyl ester to thephosphazene backbone. After mixing aqueous solutions of thepolymers and α-CD, supramolecular hydrogels were formedrapidly by the aggregation of polyrotaxane-like side chainsresulting from the inclusion complexations (IC) betweenmPEG grafts and α-CD (Scheme 1). Rheological measure-

ments were performed to characterize the hydrogels whichshow a fast gelation process and an injectable rheologicalbehavior. It is interesting that the hydrogels obtained exhibit aunique thermoreversible gel−sol transition based on supra-molecular assembly and dissociation. The evidence of IC wasexplored by various techniques, such as TGA, DSC, 13C CP/MAS NMR, and X-ray diffraction. In vitro model protein releasewas studied using bovine serum albumin (BSA). Both thestability of the hydrogel and the biodegradability of thepolymers were studied at 37 °C by monitoring the mass lossand the molecular weight decrease, respectively.

■ EXPERIMENTAL SECTIONReagents and Equipment. Tetrahydrofuran (THF) and triethyl-

amine were purchased from EMD and dried using solvent purificationcolumns. Glycine ethyl ester hydrochloride (GlyEE, Chem Impex), α-cyclodextrin (Chem Impex), and NaH (60% in paraffin, Sigma-Aldrich) were used as received. Poly(ethylene glycol) methyl ethers(mPEGs, Sigma-Aldrich) with average Mn = 350, 550, 750, 2000, and5000 g/mol were dried under vacuum extensively before use. Bovineserum albumin (BSA) was purchased from EMD Millipore Biosciences(MW ∼ 66 000 g/mol). Poly(dichlorophosphazene) was prepared bythe thermal ring-opening polymerization of recrystallized and sublimedhexachlorocyclotriphosphazene (Fushimi Chemical, Japan) in evac-uated Pyrex tubes at 250 °C. All synthesis reactions were carried outusing standard Schlenk line techniques and a dry argon atmosphere.The glassware was dried overnight in an oven at 120 °C before use. 1Hand 31P spectra were recorded on a Bruker WM-360 NMRspectrometer operated at 360 or 145 MHz, respectively. 1H NMRspectra were referenced to solvent signals, while 31P NMR chemicalshifts are relative to 85% phosphoric acid as an external reference, withpositive shift values downfield from the reference. Molecular weightswere estimated using a Hewlett-Packard HP 1090 gel permeationchromatograph (GPC) equipped with an HP-1047A refractive indexdetector, American Polymer Standards AM gel 10 mm and AM gel 10mm 104 Å columns, and calibrated versus polystyrene standards(Polysciences). The samples were eluted at 40 °C with a 0.1 wt %solution of tetra-n-butylammonium nitrate (Sigma-Aldrich) in THF.The ultraviolet absorptions of BSA release were measured with aCARY50 Bio UV−vis spectrophotometer (Varian). The IC rates weredetermined by a LKB Biochrom Ultrospec 4050 spectrophotometerwith a wavelength of 700 nm at room temperature. The visual imageswere captured by a Canon EOS Rebel XS digital camera.

Synthesis of Cyclic Trimer Models. The synthesis of mPEG350hexasubstituted cyclic phosphazene trimer (N3P3(mPEG350)6) isrepresentative of the procedures used to synthesize all the smallmolecule trimer models. To a THF solution (50 mL) ofhexachlorocyclotriphosphazene (1 g, 4.04 mmol) was added thesodium salt of mPEG350 prepared by stirring mPEG350 (11.3 g, 32.3mmol) and NaH (1.3 g, 32.3 mmol) in THF (100 mL) overnight atroom temperature. The reaction mixture was stirred for 1 day at roomtemperature and was then dialyzed versus methanol for 4 days andmethanol/acetone/hexanes (40/20/40) for 1 day (Spectra/Pordialysis membrane, MWCO: 1000). Solvent was removed, and theproduct was dried under vacuum overnight (yield: 45−74%). For thesyntheses of N3P3(mPEG2000)6 and N3P3(mPEG5000)6, themPEG2000 and mPEG5000 were predissolved in hot THF beforetreatment with NaH. The substitutions of mPEG2000 and mPEG5000were completed under reflux for 2 days. 31P NMR (CDCl3): δ 19.01(s). 1H NMR (CDCl3): δ4.02 (s, 2H, −OCH2CH2−), δ 3.63 (m, 22H,−OCH2CH2−), δ 3.36 (s, 3H, −OCH3).

Synthesis of Polymers. The syntheses of all the polymersfollowed a similar pattern, and the preparation of PmPEG350 isdemonstrated in detail. Briefly, poly(dichlorophosphazene) (0.5 g,4.32 mmol) was dissolved in anhydrous THF (100 mL). Meanwhile,mPEG350 (2.42 g, 6.90 mmol) and NaH (0.28 g, 6.90 mmol)suspensions were stirred in anhydrous THF (100 mL) at roomtemperature overnight to give a brownish solution. It was then addedto the polymer solution dropwise and allowed to react at roomtemperature for 1 day to give a partially (80 mol %) substitutedpolymer. In a separate Schlenk flask, a suspension of glycine ethyl esterhydrochloride (0.60 g, 4.32 mmol) and triethylamine (5.0 mL, 25.89mmol) in dry THF (50 mL) was stirred at room temperature for 6 hto remove the hydrochloride component. It was then added to thepolymer flask by filter addition to complete the substitution reactionovernight at room temperature. The reaction medium was thenconcentrated and dialyzed versus water for 2 days, then methanol for 2days, and methanol/acetone/hexanes (40/20/40) for 1 day (Spectra/Por dialysis membrane, MWCO: 12−14 000). The product was driedunder vacuum to give PmPEG350 as a pale yellow viscous polymer(yield: 52%−80%). For the syntheses of PmPEG2000 and

Scheme 1. Schematic Illustration of the SupramolecularPoly(organophosphazene) Hydrogels

Macromolecules Article

dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 2715−27242716

Page 3: Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and α-Cyclodextrin

PmPEG5000, mPEG2000 and mPEG5000 needed to be predissolvedin hot THF before addition to an NaH suspension in THF. Thesubstitutions of mPEG2000 and mPEG5000 required 2 and 5 daysunder reflux, followed by completion of the substitution with glycineethyl ester for 1 and 2 days under reflux. 31P NMR (CDCl3): δ −1.11(br), δ −6.63 (br). 1H NMR (CDCl3): δ 4.07 (s, 2H, GlyEE,−OCH2CH3), δ 3.98 (br, 2H, mPEG, −OCH2CH2−), δ 3.69 (s, 2H,GlyEE, −CH2COO−), δ 3.61−3.49 (br, 22H, mPEG, −OCH2CH2−),δ 3.34 (s, 3H, mPEG, −OCH3), δ 1.20 (t, 3H, GlyEE, −CH2CH3).Supramolecular Hydrogelation. The gelation of mPEG2000

grafted polymer (PmPEG2000) with α-CD is given as an example.PmPEG2000 (80 mg) was dissolved in water (8 wt %) as a hydrogelprecursor and then mixed with aqueous α-CD (175 mg) solution (10wt %) at room temperature. The resultant mixture was shakenmechanically during the gelation. The gelation times of all thehydrogels were estimated by a vial-tilting method. The timing wasstarted immediately after mixing two components until no flow wasobserved for at least 30 s when a vial containing the hydrogel wasinverted.21,36 All the hydrogels were prepared in a similar manner, anddepending on the molar ratio between ethylene glycol repeating unitsand α-CD, the resultant mixtures could be in the form of precipitates,hydrogels, viscous liquids, or solutions. To investigate the gelationkinetics, time sweep tests were performed at a constant oscillatoryfrequency of 1 rad/s under a strain of 1% by an Advanced RheometricExpansion System (Rheometric Scientific). The samples were loadedon a parallel plate (25 mm diameter, 1 mm gap) at 25 °C immediatelyafter mixing. The measurement started at 60 s after mixing, and theresults were recorded every 10 s. The gel−sol transition temperature(Ttrans) and the reverse sol−gel transition (Ttrans′) were determined bya vial inversion method with a monotonic temperature increase ordecrease of 1 °C per step.16 Thus, the vials of hydrogel samples wereimmersed in a water bath at each temperature for 10 min, the Ttrans

and Ttrans′ were monitored visually according to whether the hydrogelsflowed when inverting the vials for 30 s. To further characterize the ICsamples, they were freeze-dried and stored under vacuum.Characterizations of the Inclusion Complexations. The

evidence for IC formation between mPEG grafts and α-CD to formpolyrotaxane-like side chain linked polyphosphazenes was collected byvarious means, including the following:Thermal Property Analysis. Thermal characteristics of samples

were measured with a TA Instruments Q10 differential scanningcalorimeter (DSC) and a PerkinElmer thermogravimetric analyzer(TGA). About 10 mg of freeze-dried sample was used for each test. Aheating rate of 10 °C/min with a temperature range from −100 to 200°C was used for DSC, while a heating rate of 20 °C/min from 25 to800 °C was applied for TGA. Both instruments used dry nitrogen asthe purge gas.

13C CP/MAS NMR. Solid-state 13C CP/MAS (cross-polarization/magic angle spinning) NMR spectra were measured at 75.5 MHz on aBruker AV-300 NMR spectrometer at room temperature. The amountof sample used was 100 mg, and the accumulated scans were 18K.Powder X-ray Diffraction. X-ray diffraction was recorded on

powdered samples of freeze-dried hydrogels using wide-angle setup ofa Bruker D8 Advance diffractometer. The radiation source used wasNi-filtered, Cu Kα radiation with a wavelength of 1.54 Å. The voltagewas set to 40 kV, and the current was set to 40 mA. Samples were

mounted on a circular sample holder, and the data were collected at arate of 2θ = 5° min−1 over the range 2θ = 5°−40°.

In Vitro BSA Release and Hydrogel Dissociation. For in situencapsulation of BSA, 8 mg of BSA was dissolved in a predissolvedPmPEG2000 solution (80 mg, 8 wt %) in a phosphate buffered saline(PBS, 0.0027 M, pH = 7.4). At the same time, an appropriate amountof α-CD (175 mg) was dissolved in PBS in a separate vial (10 wt %).The α-CD solution was then added to the polymer solution to inducesupramolecular gelation at room temperature. To study the in vitrorelease behavior, the BSA-loaded hydrogel was immersed in PBS (10mL) and secured in a shaker bath maintained at 37 °C. Atpredetermined time points, 3 mL aliquots of the solution were takenout, filtered through a 0.45 μm syringe filter before UV−vismeasurements, and then added back to maintain the same solutionvolume. The cumulative percent release of BSA was determined byUV−vis spectroscopy at 279 nm in a 1 × 1 cm quartz cuvette at roomtemperature. The molar extinction coefficient of BSA in PBS solutionwas 43 800 L mol−1 cm−1 calculated from a linear Beer−Lambert plotof absorbance at 279 nm versus concentrations. The dissociation of thehydrogels was studied by the weight loss of the supramolecularhydrogels in PBS solution. Briefly, a series of supramolecular hydrogelswith known weights were prepared as in the method above. Then, theywere immersed in 10 mL of PBS and placed in a shaker bath at 37 °C.Samples were taken out at predetermined times. The upper aqueousmedium was decanted, and the remaining hydrogels were freeze-dried.The percent mass loss of the supramolecular hydrogel was recordedbased on the mass of freeze-dried hydrogel left and the total mass ofpolymer and α-CD used initially. All the studies above were carried outin triplicate with the error bars representing the standard deviations.The degradability of the polymers was determined by monitoring themolecular weight decrease in PBS solution. The dried polymers (∼10mg each) were dissolved in PBS solution (5 mL) in 10 mL glass vialsmaintained at 37 °C. Each time, a sample was dried and the remainingsolid was redissolved in THF (2 mL) for GPC analysis.

■ RESULTS AND DISCUSSION

Polymer Synthesis and Characterization. The synthesesof mPEG-containing phosphazene small molecule cyclic trimerswere attempted initially as a prelude to polymer syntheses dueto the greater synthetic challenges of the macromolecularsubstitution reactions. As monitored by 31P NMR, fullsubstitutions of mPEG350, mPEG550, and mPEG750 on thetrimers were obtained within 1 day in THF at roomtemperature, while the full substitutions with mPEG2000 andmPEG5000 required 2 days under reflux. Even thoughmPEG2000 and mPEG5000 have longer chain length, the fullsubstitution on phosphazene cyclic trimers was still achieved,probably due to the high flexibility of mPEGs. However, thelower reactivity was expected due to the increased sterichindrance and the shielding of reaction sites by the longermPEGs.The success of the cyclic trimer reactions demonstrated the

feasibility of linking mPEGs to the linear high polymers, andthe substitution conditions for the trimers were used to guide

Scheme 2. Synthesis of mPEG Grafted Poly(organophosphazenes)

Macromolecules Article

dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 2715−27242717

Page 4: Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and α-Cyclodextrin

the polymer syntheses. All the polymers in Scheme 2 wereprepared by a simple two-step synthetic route in which 80% ofmPEG was attached by macromolecular substitution in the firststep, followed by the reaction of an excess of GlyEE in thesecond step to ensure complete replacement of the chlorineatoms. The objective was to link about 20% of glycine ethylester (GlyEE) to the phosphazene polymer backbone to createbiodegradability. The rationale for this addition sequence istwofold. First, the replacement of chlorine by mPEGs in thefirst step (P−O linkage) protects the polymer backbone fromdegradation when the amino acid is added in the second step(P−N linkage). Even though the hydrochloric acid formed inthe replacement reaction of the amino acid normally complexeswith added triethylamine, it can also attack the nitrogen atomsin the polyphosphazene backbone causing P−N bond cleavage,especially during long reaction times. Second, the addition ofsmaller side groups in the last step can ease the replacement ofthe remaining chlorine atoms due to their better accessibility tothe polymer backbone. In this case, the steric bulk of GlyEE issubstantially less than mPEGs, which can ensure completehalogen replacement if added last. During the macromolecularsubstitutions, similar reactivities were found to those of thecyclic trimer model reactions. The introduction of 80% ofmPEG350, mPEG550, and mPEG750 could be completed atroom temperature within 1 day as shown by the peak at ∼−7ppm (O−P−O) and ∼−12 ppm (O−P−Cl) in 31P NMR.However, 2 and 5 days in boiling THF were required formPEG2000 and mPEG5000, respectively. Another importantsynthetic consideration is that during the substitution ofmPEG5000 the THF solution of mPEG5000 sodium salt wasadded slowly to the vigorously stirred polymers solution at 50°C over 1 h to prevent the possible gelation caused by theentanglement between poly(dichlorophospahzene) andmPEG5000. The substitution reaction became considerablyslower after the attachment of 50% of mPEG5000 to thepolymer backbone. The reaction of GlyEE in the second stepwas completed within 1 day for PmPEG350, PmPEG550, andPmPEG750 at room temperature, while it took 1 and 2 days inboiling THF for PmPEG2000 and PmPEG5000, respectively.The completion of the reaction was monitored by theappearance of a new peak at ∼−1 ppm representing thecosubstitution pattern of GlyEE and mPEGs (O−P−N) andthe disappearance of the peak at ∼−12 ppm in the 31P NMRspectrum. With the increasing of the length of mPEG segments,the resultant polymers changed from adhesive elastomers tosolid powders. In the 1H NMR spectra, the integration at 3.59ppm (−OCH2CH2−) increases as the number of repeat unitsof mPEG increases. The percentage of GlyEE on the polymerwas calculated from the peaks of 1.20 ppm (−CH3, GlyEE) and3.34 ppm (−OCH3, mPEG). The polymers show a decreasingsolubility in THF with increasing length of the mPEGsegments, but all the polymers are soluble in chloroform andwater. PmPEG350, PmPEG550, and PmPEG750 showed Mn =3.2 × 105, 4.2 × 105, and 9.1 × 105 g/mol from GPC, while nomolecular weight information could be obtained forPmPEG2000 and PmPEG5000 due to their insolubility inTHF.Supramolecular Hydrogelation. It is well-known that low

molecular weight poly(ethylene glycol)s form white precipitatesof inclusion complexes (IC) with α-CD with a stoichiometry of2:1 (ethylene glycol unit:α-CD) in aqueous solution.37 Thisinclusion effect depends on the molecular weight of thepoly(ethylene glycol)s.38 In this work, precipitations were

observed for PmPEG550, PmPEG750, PmPEG2000, andPmPEG5000 when mixed with α-CD solution at astoichiometry of 2:1 in water. The IC rates were monitoredby the optical density at 700 nm as shown in Figure 1. With

increasing length of the mPEG side chain up to mPEG2000,the IC rate increased, suggesting that stable α-CD complexescould form when the pendent mPEG was sufficiently long togenerate more side chain conformation, while after furtherlength increase, the efficiency of IC decreased.Supramolecular hydrogels were obtained by decreasing the

amount of α-CD used (to increase the ratio to 3:1 or higher),leaving an appropriate amount of mPEG repeating unitsunthreaded. As a consequence, the complexed mPEG unitsworked as physical cross-linking units, while the uncomplexedunits maintain their hydrophilicity (Scheme 1). As the ratioincreased to 4:1−6:1 for PmPEG5000, 4:1−5:1 forPmPEG2000, or 3:1 for PmPEG750, hydrogels were obtainedinstead of precipitates (Figure 2). After mixing, the polymer

solution first became white, then the turbid mixture graduallybecame more viscous, and finally formed a white hydrogel. Thegelation time for different polymers, as shown in Table 1,corresponded well with the data in Figure 1: the α-CD firstthreaded onto the mPEG side chain (turned white), followedby supramolecular gelation as the threaded side chainsaggregated (gel). However, further increases in the molarratio only produced viscous mixtures due to weak physicalcross-links, a consequence of insufficient aggregation. More-over, the gelation required a longer time for the polymers atlower concentrations or lower molar concentrations of α-CD

Figure 1. Turbidity as a function of time after mixing α-CD solutionand the polymers.

Figure 2. Photographs of supramolecular gelation of G1 and G7(bottom up).

Macromolecules Article

dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 2715−27242718

Page 5: Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and α-Cyclodextrin

used. However, for PmPEG550, the polymer solution slowlyturned white with only a slight increase in viscosity. In this case,the side chain aggregations were not enough for effectivephysical junction formation due to the shorter length of themPEG grafts. For PmPEG350, no change in viscosity wasdetected within 30 min.Time sweep measurements were conducted for the

viscoelastic behavior of the supramolecular gelation system.Storage modulus (G′) and loss modulus (G″) were monitoredas a function of time after the mixing of the polymer and α-CDsolutions. Data for selected hydrogels are shown in Figure 3 asexamples. The measurement started from 60s after mixing thetwo components due to the time needed for sample loadingand instrument setup. As the gelation proceeded, both moduliincreased and the buildup rate of G′ was higher than G″ due tothe formation of physical cross-linked network by the IC.Generally speaking, a crossover point (G′ = G″) of G′ and G″curves was detected for a gelation process (from G′ < G″ to G′

> G″).39 In this work, no crossover point was found for thegelation processes in the time range investigated (from 60 safter mixing) since G′ was higher than G″ from the beginningof the measurement, implying a fast gelation, and the crossoverpoint was less than 60 s.To confirm the recovering property of the supramolecular

hydrogel system which is critical for injectable applications, adynamic step strain amplitude test was performed on theselected gels (Figure 4). In this measurement, the initial G′after gelation was found to be about 1.5 × 104, 1.8 × 104, 9.7 ×103, and 5.1 × 103 Pa respectively for G1, G2, G7, and G8 at asmall strain (1%). When a large strain (100%) was applied tothe hydrogel, G′ immediately decreased dramatically to 6, 5, 10,and 14 Pa. Meanwhile, G′ decreased to less than G″, suggestingthat the gelled systems were disrupted. When the strain wasreturned to the small value (1%) again, G′ of all the systemsrecovered rapidly to above G″ as the gel network reformed, andthe values finally recovered to about their original moduli

Table 1. Preparation of Supramolecular Hydrogels

gel compositiona results

gel precursor polymer (wt %) molar ratio (ru: α-CD)b gels gelation timec (min) Ttrans (Ttrans′) (°C)

PmPEG 5000 8 4:1 G1 5 60 (42)4 4:1 G2 10 58 (40)8 5:1 G3 5 56 (40)4 5:1 G4 12 55 (39)8 6:1 G5 6 54 (38)4 6:1 G6 20 52 (38)

PmPEG 2000 8 4:1 G7 3 46 (35)4 4:1 G8 10 46 (33)8 5:1 G9 5 45 (33)4 5:1 G10 20 42 (32)

PmPEG 750 8 3:1 G11 6 40 (29)4 3:1 G12 15 38 (28)

aThe concentration of α-CD solution was kept at 10 wt % for all the studies. bStoichiometry of ethylene glycol repeating unit: α-CD. cStudieslimited to 60 min.

Figure 3. Rheological measurements of the selected supramolecular gelations (frequency, 1.0 rad/s; stain, 1.0%).

Macromolecules Article

dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 2715−27242719

Page 6: Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and α-Cyclodextrin

within 300 s. This result is explained by the nature ofsupramolecular hydrogel systems, since the cross-links of suchsystems are transformable and recoverable. To visually confirmthe rheological tests, the supramolecular hydrogel (G1) wasallowed to form in a 5 mL syringe fitted with a d = 0.9 mmneedle. The hydrogel could be easily squeezed out from theneedle as a white viscous liquid, which regelled immediately.Such rheological behavior of the supramolecular hydrogel isdesirable as an injectable hydrogel matrix for drug or proteindelivery.The hydrogels exhibited a reversible gel−sol transition with

increasing or decreasing temperatures (Table 1). They becamemobile above a certain temperature (Ttrans) and returned to anopaque gel after cooling to a certain temperature (Ttrans′). Forhydrogels G1−G10, the immobile opaque gel turned into anopaque sol when T > Ttrans. The interchain physical cross-linking was interrupted by the dissociation of the polymericguests from α-CD hosts at higher temperatures. However, acomplete dethreading of α-CDs was not achieved for themPEG5000 and mPEG2000 side chains as a semiopaquemixture remained even after further increases in the temper-ature. For hydrogels G11−G12, the opaque gels also turnedinto opaque sols when T > Ttrans, and the opaque sols becameclear solutions with further increases in the temperature toabove 70 °C since most α-CDs dethreaded from the mPEG750chains and dissolved in water. This observation was consistentwith other studies as the dethreading of α-CD from a polymerchain occurs when the polymer chain is more mobile at highertemperatures. The dethreading process becomes more difficultfor mPEGs with longer chain length due to the presence ofmore chain distortions and entanglements.40,41 All thehydrogels showed reversible gel−sol transitions. When the T< Ttrans′, the opaque hydrogels were formed again, and theinconsistent temperature of Ttrans and Ttrans′ can be explained bythe hysteresis effect.18

It is interesting that the addition of an excess molar amountof sodium benzoate aqueous solution brings about rapid

dissolution of the gel to yield a semiclear solution. Smallmolecule guests such as sodium benzoate can compete withmPEG to form IC with α-CD, thus causing the dissociation ofα-CD from mPEG strands.42,43 This result also supports themechanism that the supramolecular gelation was induced bythe supramolecular interaction of α-CD and mPEG side chains.Meanwhile, the addition of β-CD solution at any molar ratio tothe polymer solutions showed no gelation or precipitation. Onthe basis of the previous works by Harada and Kamachi,13 eventhough β-CD can complex with some linear short chainpolymers such as poly(propylene glycol), the tendency forthem to complex with poly(ethylene glycol) is nearlynonexistent due to the substantially smaller cross-sectionalarea of poly(ethylene glycol) compared with the cavity of β-CD.Hydrogels could also be prepared using mPEG hexasub-

stituted starlike phosphazene cyclic trimers with α-CD.However, the resulting gels showed less desirable propertiessuch as long gelation time, low gel strength, and lowdissociation temperature even with the substitution ofmPEG5000. Thus, certain chain entanglements serving asadditional physical cross-links are essential for maintaining astrong gel network. Thus, hydrogels with mPEG-linked star-structured phosphazene trimers are not discussed in detail inthis work.

Characterization of the Inclusion Complexations.PmPEG2000 is given as an example for the evidence of IC.The thermal properties of PmPEG2000 and the freeze-dried α-CD threaded precipitate (PmPEG2000 IC) were studied bydifferential scanning calorimetry (DSC) and thermogravimetricanalysis (TGA). As shown in Figure 5a, a strong melting peakwas detected for the graft copolymer PmPEG2000 at around 60°C, resulting from the melting of the mPEG2000 chains.However, for PmPEG2000 IC, there was no observed meltingpeak corresponding to the melting of mPEG because, as long asthe mPEG chains were included, they were isolated in the α-CD inner core with the disappearance of any melting behavior.

Figure 4. Dynamic step strain tests of selected supramolecular hydrogels.

Macromolecules Article

dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 2715−27242720

Page 7: Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and α-Cyclodextrin

Meanwhile, α-CD showed no peak in the temperature range.TGA revealed that α-CD had a higher decompositiontemperature (Td = 290 °C) after threading on mPEG2000side chains than pure α-CD (Td = 275 °C), implying that thepolymer chains included inside the α-CD molecules improvethe thermal stability of the α-CD (Figure 5b). The two thermalcharacterizations corresponded well with previous studies on ICof α-CD and polymers.44

Pure α-CD showed a less symmetrical conformation whenno mPEG guests were threaded in the cavity. In this case, the13C CP/MAS NMR spectrum showed multiple resolved C1and C4 resonances of the glucose units (1,4-linked, conforma-

tionally strained C1 and C4) at ∼102 and 83 ppm in Figure 6.When polymer guests were introduced into the internal cavityof α-CD to form a channel structure (PmPEG2000 IC), C1and C4 showed as broad single resonances, and the shoulder ofC1 in pure α-CD with multiple peaks (95−98 ppm)disappeared. The result indicates that α-CD was in asymmetrical conformation and each glucose unit of α-CDwas in a similar environment.12,16,45

Results from X-ray diffraction are more persuasive in provingthe inclusion complex between α-CD and the poly-(organophosphazene). As shown in Figure 7, similar to otherstudies on CD inclusion complex,46−50 the PmPEG2000 ICshowed two characteristic diffraction peaks at 2θ = 20.2° and22.7° which were not detected in the diffraction patterns ofeither PmPEG2000 or α-CD. These two diffractions are typicalpeaks observed for the IC of α-CD with a channel-typestructure (necklace-like) such as the IC between α-CD andvaleric acid.51,52

In Vitro BSA Release and Hydrogel Dissociation. Thenew prospective injectable drug delivery system based onbiodegradable poly(organophosphazene) supramolecular hy-drogel has potential applications for the encapsulation,localization, and sustained release of bioactive compounds.BSA was used as a model protein for this study. The releasebehavior of the hydrogel was studied in physiological-mimicconditions in PBS solution at 37 °C. The in vitro cumulativepercentage of BSA released was calculated based on Beer−Lambert’s law:

ε= × = ×

mm

AMVm

released % 100% 100%released

initial initial (1)

where A is the absorption at 279 nm from UV−vis spectra ofthe PBS media, M is the molecular weight of the BSA (66 000g/mol), V is the volume of PBS solution in the vials (10 mL), εis the molar extinction coefficient (43 800 L mol−1 cm−1), andminitial of BSA is 8 mg.According to Figure 8, G11 and G7 showed burst releases in

2−5 days, while G2 and G1 gave longer release times over 10−

Figure 5. Thermal characterizations of pure PmPEG2000, freeze-driedPmPEG2000 IC, and α-CD: (a) DSC study; (b) TGA study.

Figure 6. 13C CP/MAS NMR spectra of PmPEG2000 IC and pure α-CD.

Macromolecules Article

dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 2715−27242721

Page 8: Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and α-Cyclodextrin

12 days. The release patterns are explained by combining thedissociation profiles of the hydrogels (Figure 9). G11 and G7

showed significant structural dissociations suggesting that theBSA release was not diffusion-controlled, and the morpho-logical change of the matrix was abrupt. The drastic erosion ofthe supramolecular hydrogels was induced by a significantdethreading of the α-CD from mPEG750 and mPEG2000 sidechains when the hydrogels were exposed to a large amount ofwater at 37 °C, and the physical interactions, which areessential for maintaining the hydrogel structure, fell apart.50 G2and G1 prepared by the polymers containing mPEG5000

showed a better stability and a longer release profile since thelonger length of mPEG prevents the dethreading of α-CD bymore entanglement. The hydrogels (G1) with higher polymerconcentration showed better stability and a slower releaseprofile than G2. But dissociations of hydrogels G2 and G1 werestill observed (Figure 9) at a slow and steady rate, suggestingthat the release mechanism was a combination of the structuralrelaxation of the hydrogels and the diffusion of BSA. Acalculation based on eq 2 was applied to verify the releasemechanism of G2 and G1:19,53

= ≤∞ ∞M M Kt M M/ (for / 0.6)tn

t (2)

where Mt and M∞ are the cumulative amount of the BSAreleased at a time t and equilibrium, respectively. K is the rateconstant relating to the properties of the hydrogel matrix andthe drug, and n is the release exponent characterizing thetransport mechanism. By transforming eq 2 to a logarithmicequation

= +∞M M n t Klog( / ) log( ) logt (3)

the value of n can be obtained by plotting log(Mt/M∞) versuslog(t). The n values for G2 and G1 were calculated to be 0.83 ±0.05 and 0.78 ± 0.05, respectively, suggesting an anomaloustransport process (0.5 < n < 1, case III) as the structure changeand the diffusion were comparable.54,55

All the polymers synthesized in this work were sensitive tohydrolysis. For example, the molecular weight of PmPEG750(soluble in THF) declined rapidly to almost 1/5 (from 9.1 ×105 to 1.9 × 105 g/mol) after 1 week in PBS solution at 37 °C.The glycine ethyl ester units on the polyphosphazene backboneworked as water sensitive points to cleave the polymerbackbone into shorter sections. After that, the molecularweight decreased gradually at a stable rate as the polymerbackbone continued to be broken down into oligophosphazenederivatives (molecular weight <4.0 × 104 g/mol) and smallmolecules (glycine, ethanol, ammonium, and phosphates).28 Atime−status relationship of this new supramolecular hydrogelsystem in a physiological environment can be expected. Whilereleasing any encapsulated biomolecules, the hydrogelsgradually dissociate and dissolve in an aqueous environment.Then, the water-soluble poly(organophosphazenes) graduallydegrade into shorter oligmer chains and small molecules whichcan be excreted through the renal filtration system.56,57

Figure 7. X-ray diffraction pattern of pure α-CD, pure PmPEG2000, and freeze-dried PmPEG2000 IC.

Figure 8. In vitro cumulative release of encapsulated BSA fromsupramolecular hydrogels.

Figure 9. Supramolecular hydrogels dissociation profiles.

Macromolecules Article

dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 2715−27242722

Page 9: Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and α-Cyclodextrin

■ CONCLUSIONSNovel injectable supramolecular hydrogels were obtained by ICformation between mPEG modified biodegradable polyphos-phazenes and α-CD. In this system, the aggregation of α-CDcomplexed mPEG side chains serve as physical cross-links,while the uncomplexed mPEG side chians serve as hydrophilicsegments. The gelation time varied from several minutes toover 20 min depending on the polymer structures andconcentrations. The rheological measurements showed a fastgelation process. The thermoreversible gel−sol transitions wereobserved and studied for all the hydrogels prepared. The ICinduced gelation mechanism was studied by DSC, TGA, 13CCP/MAS NMR, and X-ray diffraction techniques. The in vitrorelease of BSA and the dissociation of hydrogels showed higherstabilities and longer release profiles for polymers with longermPEG side chains. The biodegradability of the polymers wascharacterized based on the molecular weight decrease. Such abiodegradable and biocompatible hydrogel system is apromising candidate for many biomedical applications,especially in the area of injectable drug delivery.

■ AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected] authors declare no competing financial interest.

■ REFERENCES(1) Wichterle, O.; Lim, D. Nature 1960, 185, 177−178.(2) Hoffman, A. S. Adv. Drug Delivery Rev. 2002, 43, 3−12.(3) Hoare, T. R.; Kohane, D. S. Polymer 2008, 49, 1993−2007.(4) Lim, F.; Sun, M. Science 1980, 210, 908−910.(5) Yannas, I. V.; Lee, E.; Orgill, D. P. Proc. Natl. Acad. Aci. U. S. A.1989, 86, 933−937.(6) Jin, R.; Hiemstra, C.; Zhong, Z. Y.; Jan, F. J. Biomaterials 2007,28, 2791−2800.(7) Xiong, X. Y.; Tam, K. C.; Gan, L. H. J. Nanosci. Nanotechnol.2006, 6, 2638−2650.(8) Lim, D. W.; Nettles, D. L.; Setton, L. A.; Chilkoti, A.Biomacromolecules 2007, 8, 1463−1470.(9) Ricciardi, R.; Gaillet, C.; Ducouret, G.; Lafuma, F.; Laupretre, F.Polymer 2003, 44, 3375−3380.(10) Jong, S. J.; De Smedt, S. C.; Wahls, M. W. C.; Demeester, J.;Kettenes-van den Bosch, J. J.; Hennink, W. E. Macromolecules 2000,33, 3680−3686.(11) Harada, A.; Li, J.; Kamachi, M. Nature 1994, 370, 126−128.(12) Harada, A.; Li, J.; Kamachi, M. Macromolecules 1993, 26, 5698−5703.(13) Harada, A.; Okada, M.; Li, J.; Kamachi, M. Macromolecules 1995,28, 8406−8411.(14) Huang, L.; Allen, E.; Tonelli, A. E. Polymer 1998, 39, 4857−4865.(15) Harada, A.; Suzuki, S.; Okada, M.; Kamachi, M. Macromolecules1996, 29, 5611−5614.(16) Huh, K. M.; Ooya, T.; Sasaki, S.; Yui, N. Macromolecules 2001,34, 2402−2404.(17) Harada, A.; Hashidzume, A.; Yamaguchi, H.; Takashima, Y.Chem. Rev. 2009, 109, 5974−6023.(18) Huh, K. M.; Ooya, T.; Lee, W. K.; Sasaki, S.; Kwon, I. C.; Jeong,S. Y.; Yui, N. Macromolecules 2001, 34, 8657−8662.(19) Ma, D.; Tu, K.; Zhang, L. M. Biomacromolecules 2010, 11,2204−2212.(20) Zhao, S. P.; Zhang, L. M.; Ma, D. J. Phys. Chem. B 2006, 110,12225−12229.(21) Li, Z.; Yin, H.; Zhang, Z.; Liu, K. L.; Li, J. Biomacromolecules2012, 13, 3162−3172.

(22) Allcock, H. R. Chemistry and Applications of Polyphosphazene;John Wiley and Sons: Hoboken, NJ, 2003.(23) Allcock, H. R.; Kugel, R. L.; Valan, K. J. Inorg. Chem. 1966, 5,1709−1715.(24) Allcock, H. R.; Kugel, R. L. J. Am. Chem. Soc. 1965, 87, 4216−4217.(25) Allcock, H. R.; Connolly, M. S.; Sisko, J. T.; Al-Shali, S.Macromolecules 1988, 21, 323−334.(26) Singh, A.; Krogman, N. R.; Sethuraman, S.; Nair, L. S.;Sturgeon, J. L.; Brown, P. W.; Laurencin, C. T.; Allcock, H. R.Biomacromolecules 2006, 7, 914−918.(27) Allcock, H. R.; Pucher, S. R. Macromolecules 1994, 27, 1071−1075.(28) Deng, M.; Kumbar, S. G.; Wan, Y.; Toti, U. S.; Allcock, H. R.;Laurencin, C. T. Soft Matter 2010, 6, 3119−3132.(29) Nukavarapu, S. P.; Kumbar, S. G.; Allcock, H. R.; Laurencin, C.T. Biodegradable Polyphosphazene Scaffolds for Tissue Engineering.In Polyphosphazenes for Biomedical Applications; Andrianov, A. K., Ed.;John Wiley and Sons: Hoboken, NJ, 2009; p 119.(30) Allcock, H. R.; Dudley, G. K. Macromolecules 1996, 29, 1313−1319.(31) Allcock, H. R.; Pucher, S. R.; Turner, M. L.; Fitzpatrick, R. J.Macromolecules 1992, 25, 5573−5577.(32) Lee, B. H.; Lee, Y. M.; Sohn, Y. S.; Song, S. C. Macromolecules2002, 35, 3876−3879.(33) Potta, T.; Chun, C.; Song, S. C. Biomacromolecules 2010, 11,1741−1753.(34) Krogman, N. R.; Weikel, A. L.; Nguyen, N. Q.; Nair, L. S.;Laurencin, C. T.; Allcock, H. R. Macromolecules 2008, 41, 7824−7828.(35) Allcock, H. R.; Ambrosio, A. M. A. Biomaterials 1996, 17, 2295−2302.(36) Liu, K. L.; Choo, E. S. G.; Wong, S. Y.; Liu, X.; Bin He, C.;Wang, J.; Li, J. J. Phys. Chem. B 2010, 114, 7489−7498.(37) Harada, A.; Li, J.; Kamachi, M. Macromolecules 1994, 27, 4538−4543.(38) Rusa, C. C.; Tonelli, A. E.Macromolecules 2000, 33, 1813−1818.(39) Winter, H. H.; Chambon, F. J. Rheol. 1986, 30, 367−378.(40) Rajewski, R. A.; Stella, V. J. J. Pharm. Sci. 1996, 85, 1142−1169.(41) Yan, Q.; Xin, Y.; Zhou, R.; Yin, Y.; Yuan, J. Chem. Commun.2011, 47, 9594−9596.(42) Okumura, Y.; Ito, K.; Hayakawa, R. Polym. Adv. Technol. 2000,11, 815−819.(43) Yu, Z. J.; Liu, Y.; Fan, M. M.; Meng, X. W.; Li, B. J.; Zhang, S. J.Polym. Sci., Part B: Polym. Phys. 2010, 48, 951−957.(44) Lu, J.; Shin, I. D.; Nojima, S.; Tonelli, A. E. Polymer 2000, 41,5871−5883.(45) Huh, K. M.; Cho, Y. W.; Chung, H.; Kwon, I. C.; Jeong, S. Y.;Ooya, T.; Lee, W. K.; Sasaki, S.; Yui, N. Macromol. Biosci. 2004, 4, 92−99.(46) Zhao, S. P.; Zhang, L. M.; Ma, D.; Yang, C.; Yan, L. J. Phys.Chem. B 2006, 110, 16503−16507.(47) Higashi, T.; Tajima, A.; Motoyama, K.; Arima, H. J. Pham. Sci.2012, 101, 2891−2899.(48) Harada, A.; Nishiyama, T.; Kawaguchi, Y.; Okada, M.; Kamachi,M. Macromolecules 1997, 30, 7115−7118.(49) Huang, L.; Allen, E.; Tonelli, A. E. Polymer 1999, 40, 3211−3221.(50) McMullan, R. K.; Saenger, W.; Fayos, J.; Mootz, D. Carbohydr.Res. 1973, 31, 37−46.(51) Takeo, K.; Kuge, T. Agric. Biol. Chem. 1970, 34, 1787−1794.(52) Li, J.; Li, X.; Ni, X.; Wang, X.; Li, H.; Leong, K. W. Biomaterials2006, 27, 4132−4140.(53) Franson, N. M.; Peppas, N. A. J. Appl. Polym. Sci. 1983, 28,1299−1310.(54) Ritger, P. L.; Peppas, N. A. J. Controlled Release 1987, 5, 37−42.(55) Kim, S. W.; Bae, Y. H.; Okano, T. Pharm. Res. 1992, 9, 283−290.(56) Seymour, L. W.; Miyamoto, Y.; Maeda, H.; Brereton, M.;Strohalm, J. Eur. J. Cancer 1995, 31A, 766−770.

Macromolecules Article

dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 2715−27242723

Page 10: Injectable and Biodegradable Supramolecular Hydrogels by Inclusion Complexation between Poly(organophosphazenes) and α-Cyclodextrin

(57) Yokoyama, M. J. Artif. Organs 2005, 8, 77−84.

Macromolecules Article

dx.doi.org/10.1021/ma4004314 | Macromolecules 2013, 46, 2715−27242724