rheological studies of plla–peo–plla triblock copolymer hydrogels

7
Biomaterials 25 (2004) 1087–1093 Rheological studies of PLLA–PEO–PLLA triblock copolymer hydrogels Khaled A. Aamer a , Heidi Sardinha b , Surita R. Bhatia b, *, Gregory N. Tew a, * a Department of Polymer Science and Engineering, University of Massachusetts, Amherst, MA 01003, USA b Department of Chemical Engineering, University of Massachusetts, Amherst, MA 01003, USA Received 2 June 2003; accepted 25 July 2003 Abstract We report detailed rheological data on aqueous gels formed from triblock copolymers of l-lactide and ethylene oxide including the dependence of the viscoelastic moduli on frequency and applied stress of these systems for the first time. We are able to create strong gels with elastic moduli greater than 10,000 Pa, which is an order of magnitude higher than previously achieved with related biocompatible physically associated gels of similar chemistry. Moreover, the value of the elastic modulus strongly depends on PLLA block length, offering a mechanism to control the mechanical properties as desired for particular applications. At the gel point, we observe scaling that is characteristic of a percolated network, G 0 B G 00 Bo D , but with an exponent that is lower than predicted by percolation, D=0.36. Our results have implications for the design of new materials for soft tissue engineering, where native tissues have moduli in the kPa range. r 2003 Elsevier Ltd. All rights reserved. Keywords: Polylactic acid; Polyethylene oxide; Triblock; Elasticity; Hydrogel; Mechanical properties 1. Introduction Amphiphilic polymers have received considerable attention as potential biomaterials in the last decade [1–5]. In particular, materials capable of forming hydrogels remain an active area of research due to applications in drug delivery, tissue engineering and other biomedical devices [3,4]. Hydrogels are attractive for a variety of reasons including less invasive medical procedures and the fact that these materials most closely mimic natural tissues in water content, interfacial tension, and mechanical properties (soft and rubbery) [5]. Triblock copolymers of poly (l-lactide) (PLLA) and poly (ethylene oxide) (PEO) have been studied exten- sively since the initial report by Cohn [6] because of the biodegradability of the polyester and the biocompat- ibility of PEO [7–21]. In addition, these materials are known to form hydrogels of varying weight %, depending on the exact architecture, with transitions near body temperature. Most studies have focused on the sol–gel transition temperatures by vial inversion and thus do not provide detailed mechanical properties [16,22–26]. In the study presented here, we synthesized PLLA–PEO–PLLA polymers containing large PEO and varying PLLA blocks which were characterized in the gel phase by mechanical rheology. Specifically, our desire is to find strong physically cross-linked hydrogels from polyester ether materials for potential application in biomedical devices. Amphiphilic copolymers composed of PLA and PEO have been prepared with PEO end- or mid-blocks and characterized as hydrogels [16,22–26] by vial inversion methods which require elastic modulus greater than 65 Pascals (Pa) [27]. For the triblock architecture of interest here, PLA–PEO–PLA, previous reports have formed gels with polymers composed of lactic and glycolic acid, PLGA x –PEO 1000 –PLGA x , in which x, the degree of polymerization (DP), was varied both in block length and lactide to glycolide ratio [22]. These materials exhibited sol–gel transitions between 10 C and 23 C and gel–sol transitions from 27 C to 48 C depending on concentration and x [22]. The same group showed ARTICLE IN PRESS *Corresponding authors. Tel.: +1-413-545-0096; fax: +1-413-545- 1647 (S.R. Bhatia); Tel.: +1-413-577-1612; fax: +1-413-545-0082 (G.N. Tew). E-mail addresses: [email protected] (S.R. Bhatia), [email protected] (G.N. Tew). 0142-9612/$ - see front matter r 2003 Elsevier Ltd. All rights reserved. doi:10.1016/S0142-9612(03)00632-X

Upload: khaled-a-aamer

Post on 02-Jul-2016

216 views

Category:

Documents


4 download

TRANSCRIPT

Page 1: Rheological studies of PLLA–PEO–PLLA triblock copolymer hydrogels

Biomaterials 25 (2004) 1087–1093

ARTICLE IN PRESS

*Correspondin

1647 (S.R. Bha

(G.N. Tew).

E-mail addres

[email protected]

0142-9612/$ - see

doi:10.1016/S014

Rheological studies of PLLA–PEO–PLLA triblock copolymerhydrogels

Khaled A. Aamera, Heidi Sardinhab, Surita R. Bhatiab,*, Gregory N. Tewa,*aDepartment of Polymer Science and Engineering, University of Massachusetts, Amherst, MA 01003, USA

bDepartment of Chemical Engineering, University of Massachusetts, Amherst, MA 01003, USA

Received 2 June 2003; accepted 25 July 2003

Abstract

We report detailed rheological data on aqueous gels formed from triblock copolymers of l-lactide and ethylene oxide including

the dependence of the viscoelastic moduli on frequency and applied stress of these systems for the first time. We are able to create

strong gels with elastic moduli greater than 10,000 Pa, which is an order of magnitude higher than previously achieved with related

biocompatible physically associated gels of similar chemistry. Moreover, the value of the elastic modulus strongly depends on PLLA

block length, offering a mechanism to control the mechanical properties as desired for particular applications. At the gel point, we

observe scaling that is characteristic of a percolated network, G0B G0 0BoD, but with an exponent that is lower than predicted by

percolation, D=0.36. Our results have implications for the design of new materials for soft tissue engineering, where native tissueshave moduli in the kPa range.

r 2003 Elsevier Ltd. All rights reserved.

Keywords: Polylactic acid; Polyethylene oxide; Triblock; Elasticity; Hydrogel; Mechanical properties

1. Introduction

Amphiphilic polymers have received considerableattention as potential biomaterials in the last decade[1–5]. In particular, materials capable of forminghydrogels remain an active area of research due toapplications in drug delivery, tissue engineering andother biomedical devices [3,4]. Hydrogels are attractivefor a variety of reasons including less invasive medicalprocedures and the fact that these materials most closelymimic natural tissues in water content, interfacialtension, and mechanical properties (soft and rubbery)[5]. Triblock copolymers of poly (l-lactide) (PLLA) andpoly (ethylene oxide) (PEO) have been studied exten-sively since the initial report by Cohn [6] because of thebiodegradability of the polyester and the biocompat-ibility of PEO [7–21]. In addition, these materials areknown to form hydrogels of varying weight %,

g authors. Tel.: +1-413-545-0096; fax: +1-413-545-

tia); Tel.: +1-413-577-1612; fax: +1-413-545-0082

ses: [email protected] (S.R. Bhatia),

ass.edu (G.N. Tew).

front matter r 2003 Elsevier Ltd. All rights reserved.

2-9612(03)00632-X

depending on the exact architecture, with transitionsnear body temperature. Most studies have focused onthe sol–gel transition temperatures by vial inversion andthus do not provide detailed mechanical properties[16,22–26]. In the study presented here, we synthesizedPLLA–PEO–PLLA polymers containing large PEO andvarying PLLA blocks which were characterized in thegel phase by mechanical rheology. Specifically, ourdesire is to find strong physically cross-linked hydrogelsfrom polyester ether materials for potential applicationin biomedical devices.Amphiphilic copolymers composed of PLA and PEO

have been prepared with PEO end- or mid-blocks andcharacterized as hydrogels [16,22–26] by vial inversionmethods which require elastic modulus greater than65 Pascals (Pa) [27]. For the triblock architecture ofinterest here, PLA–PEO–PLA, previous reports haveformed gels with polymers composed of lactic andglycolic acid, PLGAx–PEO1000–PLGAx, in which x, thedegree of polymerization (DP), was varied both in blocklength and lactide to glycolide ratio [22]. These materialsexhibited sol–gel transitions between 10�C and 23�Cand gel–sol transitions from 27�C to 48�C depending onconcentration and x [22]. The same group showed

Page 2: Rheological studies of PLLA–PEO–PLLA triblock copolymer hydrogels

ARTICLE IN PRESSK.A. Aamer et al. / Biomaterials 25 (2004) 1087–10931088

that polymers composed of d, l lactide formed gelswith a sharper gel–sol transition compared to PLGApolymers [23]. More recently, mechanical rheology onPMG1400–PEO1450–PMG1400 (PMG-poly(d,l-3-methyl-glycolide)) showed the elastic modulus of a 27wt%sample is less than 500Pa [28]. Further, Kimura and co-workers reported hydrogel formation from stereocom-plexed PLLA1300–PEO4600–PLLA1300 and PDLA1090–PEO4600–PDLA1090 in which 10wt% solutions had anelastic modulus up to 1000 Pa at 37�C [29]. Solutions ofeither polymer independently at 10wt% did not form ahydrogel. Thus, a relatively complicated system wasrequired to obtain favorable mechanical properties. Inaddition, all rheological data reported in these studieswere taken at a single frequency and strain, so thedetailed dependence of modulus on frequency andapplied strain is completely unknown.The polymers described herein are composed solely of

l-lactide and contain PEO blocks that are larger thanthose utilized in previous studies (MW=8900Da). Wereport values for the elastic modulus (G0) and lossmodulus (G00) over a range of frequency and appliedstress, allowing for a more complete characterization ofthe rheological properties. Moreover, we focus onsystems with hydrophobic blocks that have a lowermolecular weight than the PEO block, which representsa significant difference between the majority of PLA–PEO–PLA systems reported in the literature. Thisenables us to create physically associated gels that areanalogous to reversible network gels formed fromtelechelic hydrophobically modified polymers. As dis-cussed further below, this leads to gels with elasticmoduli that are significantly higher than previouslyreported.

2. Experimental section

2.1. Materials

l-lactide (Aldrich) was purified by recrystallization indry ethylacetate and by sublimation prior topolymerization. The a, o dihydroxy polyethyleneglycol macroinitiator with molecular weight 8900(PEG 8K, Aldrich) was dried at room temperatureunder vacuum prior to polymerization. Stannous (II) 2-ethyl hexanoate (Alfa Aesar) was used without furtherpurification.

2.2. Synthesis of PLLA–PEO–PLLA triblock copolymer

PLLA–PEO–PLLA triblock copolymers were synthe-sized by bulk polymerization. PEO, was introduced intoa dried polymerization tube. The tube was purged withnitrogen, and placed in an oil bath at 150�C. Stannous(II) 2-ethyl hexanoate was introduced under nitrogen to

the molten PEO and stirred for 10min followed byaddition of l-lactide to the macroinitiator/catalyst melt.The polymerization was carried out at 150�C for 24 hwith stirring, after which it was quenched by methanol.The product was dissolved in tetrahydrofuran andprecipitated in n-hexane. The process of dissolution/reprecipitation was carried out three more times. Thecopolymer was dried under vacuum at room tempera-ture for 2 days.

2.3. Sample preparation and instrumentation

The copolymers polydispersity are measured versuspolyethylene oxide standards using GPC (HP 1050series, a HP 1047A differential refractometer, and threePLgel columns (5 mm 50 (A, two 5 mm MIXED-D) indimethylformamide as eluting solvent at a rate of 0.5ml/min rate at room temperature. The copolymer composi-tions are determined by 1H NMR (Bruker, DPX300,300MHz spectrometer, d-chloroform).Gels were prepared by slow addition of dried polymer

sample to a fixed volume of DI water (15ml) followedby stirring and heating. Gels were then transferred to aBohlin CVO rheometer for oscillatory measurements. Acone-and-plate geometry with a 4� cone, 40mmdiameter plate, and 150mm gap was used for allexperiments on hydrogels. For liquid samples with alow viscosity, a couette geometry was used. Stressamplitude sweeps were performed to ensure thatsubsequent data was collected in the linear viscoelasticregime. Frequency sweeps were performed at a constantstress (0.1–2.0 Pa, depending on the sample) in thefrequency range 0.01–100Hz. At high frequencies, aresonant frequency of the rheometer motor wasobserved; thus, data are reported up to a frequency ofapproximately 10Hz, again depending on the particularsample.

3. Results and discussion

3.1. Block copolymer synthesis

The copolymers shown in Table 1 were prepared byring-opening polymerization of l-lactide at 150�C in thebulk using stannous (II) 2-ethylhexanoate as catalyst.This method is known to limit racemization of thestereocenter and produce polymers of significant mole-cular weight and narrow polydispersity [7]. The macro-initiator, PEO, has molecular weight (MW) of 8900Daand four different polymers were prepared with increas-ing PLLA block lengths. These lengths varied from atotal DP of 26 to 74 so that the total lactide compositionis always smaller than PEO. 1H-NMR integration wasused to establish theMw for PLLA blocks as opposed toGPC standards [10,11]. In all cases, the polymerization

Page 3: Rheological studies of PLLA–PEO–PLLA triblock copolymer hydrogels

ARTICLE IN PRESS

Table 1

Molecular weight characteristics of PLLA–PEO–PLLA triblock copolymers

OHO

H202

+O

O

O

O Poly(LLA-EG-LLA)

OO

O

O

HO

Oy y150oC/Bulk polymerization

Sn(Oct) 2 H202

Sample MnPEGa MnPLLA

b MTotal MWDc Total DPPLLAb Wt%PEO Wt%PLLA

1 8900 1872 10,772 1.31 26 82.6 17.4

2 8900 3744 12,644 1.24 52 70.4 29.4

3 8900 4320 13,220 1.18 60 67.3 32.7

4 8900 5328 14,228 1.20 74 62.6 37.4

aDetermined by MALDI TOF and GPC.bDetermined by 1H NMR.cDetermined by GPC.

100

1000

10000

100000

0.01 0.1 1 10

Frequency (Hz)

G',

G"

(Pa)

G'G"

Fig. 1. Elastic modulus (filled diamonds) and viscous modulus (open

squares) as a function of frequency for a gel of sample 4 at 16wt%

polymer and T=25�C.

0.1

1

10

100

1000

10000

100000

0.01 0.1 1 10 100

Frequency (Hz)

G',

G*

(Pa)

G*, 1, 25 CG', 2, 25 CG', 3, 25 CG', 4, 25 CG', 2, 37 CG', 3, 37 CG', 4, 37 C

Fig. 2. Elastic moduli versus frequency for gels formed from a series

with increasing PLLA block length, samples 2, 3, and 4 at

concentrations of 20, 16, and 16wt%, respectively; at 25�C (filled

symbols) and 37�C (open symbols). For comparison, the complex

modulus of a 20wt% solution of sample 1 is also shown.

K.A. Aamer et al. / Biomaterials 25 (2004) 1087–1093 1089

was not run to high conversion since this broadens themolecular weight distribution.

3.2. Rheology

All samples were characterized by dynamic mechan-ical rheology. Solutions of the triblock with the smallesthydrophobes (1) did not gel up to concentrations of20wt% polymer, while the remaining three triblocksformed gels at moderate concentrations (16–20wt%).Fig. 1 shows G0 and G00 vs. frequency for 16wt% 4 at25�C. The gel displays a high degree of elasticity, with G0

only weakly dependent on frequency and greater thanG00 over the entire frequency range. Note thatG0X10,000 Pa for this system, which is an order of

magnitude higher than the stereocomplexed systemsstudied by Kimura and co-workers [29]. Biomaterials

with moduli in the kPa range are of widespread interestsince many native tissues have moduli in this range,although most have nonlinear response to strain. Forexample, human nasal cartilage (234727 kPa) [30,31],bovine articular cartilage (990750 kPa) [30,31], pigthoracic aorta (43.2715 kPa) [32], pig adventitial layer(4.7271.7 kPa) [32], right lobe of human liver(270710 kPa) [33], canine kidney cortex and medulla(B10 kPa) [34], and nucleus pulposus and eye lens(B103 Pa) [34] have moduli in this range. Moreover, forscaffolding applications, it is often desirable to ‘‘match’’mechanical properties of the polymer matrix to those ofthe surrounding tissue [35].Fig. 2 compares the elastic modulus of systems

containing 2, 3, and 4 at concentrations of 20, 16, and16wt%, respectively, at 25�C (filled symbols) and 37�C(open symbols). These represent a series in the length of

Page 4: Rheological studies of PLLA–PEO–PLLA triblock copolymer hydrogels

ARTICLE IN PRESS

1

10

100

1000

0.01 0.1 1 10

Frequency (Hz)

Ela

stic

Mo

dulu

s (P

a)

25 C

37 C

40 C

50 C

Fig. 3. Elastic modulus versus frequency for a gel with 20wt% of

sample 2 as a function of temperature. Note the non-monotonic

dependence of G0 on temperature for this sample.

1000

10000

100000

0.01 0.1 1 10

Frequency (Hz)

Ela

stic

Mo

dulu

s (P

a)

25 C 31 C37 C 49 C61 C 70 C

Fig. 4. Elastic modulus versus frequency for a gel with 16wt% of

sample 4 as a function of temperature.

1

10

100

1000

0.01 0.1 1 10

Frequency (Hz)

G',

G"

(Pa)

G'G"

1

10

100

1000

0.01 1

T = 50 C

T = 37 C

Fig. 5. Elastic modulus (filled diamonds) and viscous modulus (open

squares) versus frequency for a 20wt% gel of sample 2 at T=37�C and

(inset) T=50�C.

K.A. Aamer et al. / Biomaterials 25 (2004) 1087–10931090

the hydrophobe, with 2 having the shortest PLLAblocks. Although the gel of 2 was prepared in 20wt%, asopposed to 16wt% for the other samples, it still forms aweaker gel. It is clear from these results that thehydrophobe length or DP has a pronounced effect onelastic modulus. The only difference between thesepolymers is the total DP of the PLLA segments, whichare 52, 60, and 74, respectively. As the length of thehydrophobic block is increased from 26 (half of 52 sincethere are two PLLA ends) to >30, gel strength improvesby more than two orders of magnitude fromB100 Pa to10,000 Pa even though the percentage of polymer in thegel is decreased. For comparison, the complex modulusG� of a 20wt% solution of 1 at 25�C is also shown inFig. 2. This sample, with hydrophobic block lengthsof 13, forms a viscoelastic liquid with G�o1 Pa.Qualitatively similar trends have been observed for thehigh-frequency limit of G0 for PEO containing alkylhydrophobe end-caps [36]. However, the dependence ofG0 on hydrophobe length is much weaker for these alkyl-capped systems than in our PLLA–PEO–PLLA gels,while the high-frequency elastic moduli of gels offluoroalkyl-capped PEO were found to be insensitiveto the hydrophobe length [37,38]. Thus, the strongdependence of G0 on the PLLA block length should offera straightforward way to tune the rheological responseof our gels for specific bioapplications.The elastic modulus at 37�C is also shown on Fig. 2

for these gels. All three gels display a decrease in G0 withtemperature; however, with a much larger decrease for 2than either 3 or 4. This decrease in G0 with increasingtemperature is common for gels composed of PLA–PEO–PLA molecules [23]. It is important to note that,despite this decrease, the value of G0 for 4 remains highat physiological temperatures (roughly 10,000 Pa), andthat both 3 and 4 are still elastic gels with G0 nearlyindependent of frequency.Figs. 3 and 4 show the influence of temperature on G0

for samples 2 and 4. Interestingly, G0 for sample 2

decreases with temperature as expected; however, near40�C the sample undergoes a transition toward increas-ing G0. At 50�C, the elastic modulus is even more linearwith temperature and significantly higher at lowfrequency than found at 37�C. The physical andstructural properties of the gel responsible for thesechanges are not understood at this time. However, it ispossible that at 25�C, the initial measurement, thehydrogel is composed of slightly hydrated micelles, butas the temperature is increased near 40�C, these micellesbegin to melt and destabilize (Tg of PLLA is approxi-mately 52�C and should be lowered by the presence ofwater as a weak plasticizer). Then, as the temperature isincreased further, the PLA domains may dehydrate in amanner similar to that observed for the PPO segmentsof Pluronics [39]. The observed behavior is consistentwith this desolvation transition but more work must be

performed to prove or disprove this assumption.Alternatively, 4 behaves in a more expected manner(Fig. 4) with a steady decrease in G0 as the temperature is

Page 5: Rheological studies of PLLA–PEO–PLLA triblock copolymer hydrogels

ARTICLE IN PRESS

1

10

100

1000

10000

100000

0.00001 0.0001 0.001 0.01 0.1

Strain

Ela

stic

Mod

ulus

(Pa)

2

3

41

10

100

1000

10000

100000

0.01 0.1 1 10 100

Stress (Pa)

Ela

stic

Mod

ulus

(Pa)

2

3

4

(a) (b)

Fig. 6. Elastic modulus at a fixed frequency of 1.0Hz as a function of (a) strain and (b) stress for samples 2, 3, and 4 at concentrations of 20, 16, and

16wt%, respectively, and 25�C.

K.A. Aamer et al. / Biomaterials 25 (2004) 1087–1093 1091

increased and this would be expected because water isless likely to hydrate larger PLLA blocks. However, at70�C, the data shows a slight increase in modulusconsistent with PEO dehydration.Sol–gel transitions can be defined as G0>G00, and

Fig. 5 shows that for 2 at 37�C G0EG00 for the entirefrequency range. Interestingly, the observed increasein G0 at 50�C for sample 2 is confirmed in the inset ofFig. 5, where G0 is much larger than G00. The frequency-dependent G0 observed for sample 2 at 25�C and 37�C,as opposed to the frequency-independent behaviorobserved at 50�C, may indicate the presence of a slowrelaxation mechanism at lower temperatures.The data on 2 at the gel point (37�C) can be fit to a

power law scaling of the form G0B G00BoD [40]. Here, ois frequency and D is an exponent whose value at the gelpoint can be predicted to lie in the range 0.67–1.0 fornetworked gels through classical mean-field considera-tions [41], electrical network analogies [42], and percola-tion theory [43,44]. Modified percolation models thataccount for elasticity of the network backbone [45,46]and excluded volume effects [47] have yielded predic-tions in the range 0–1.0. Fitting the data shown in Fig. 5to a power law gives D=0.36. Although it is impossibleto draw any definite conclusions about the structure ofthe gel from this value, similar values have been reportedfor nonstoichiometric chemically crosslinked polymericgels with an excess of crosslinker present, and forchemically crosslinked gels with chains above theentanglement molecular weight [48]. By contrast,experiments on physically associated biopolymer gels[49,50] and transient networks based on triblockcopolymers [51] have yielded higher values of D in therange 0.5–0.7. Interestingly, gels with semicrystallinedomains have lower exponents, including bacteriallyderived poly(b-hydroxyoctanoate) [52] and crystallizingpolypropylene [53]. Thus, the low value of D may be aconsequence of either entanglements of the PEO chainsor rigidity of the hydrophobic PLLA domains.

Fig. 6 shows the elastic modulus at fixed frequency asa function of strain for 2, 3, and 4 at 25�C. Thesepolymers exhibit a linear viscoelastic response over awide range of strain and applied stress, and no evidenceof strain-hardening was observed. Thus, we expect thesegels will have a uniform and predictable rheologicalresponse in vivo regardless of the mechanical environ-ment to which they are subjected.

4. Conclusions

We have reported, for the first time, a completerheological characterization of physically associatedhydrogels formed from PLLA–PEO–PLLA triblockcopolymers. The elastic modulus of these gels is stronglydependent on PLLA block length, offering a means ofcontrolling the mechanical properties for particular softtissue engineering applications. The triblock copolymersthat formed gels displayed elastic moduli in the range100–18,300 Pa at ambient and physiological tempera-tures, extending the upper limit of G0 that can beachieved with physical gels of PLA/PEO copolymers byan order of magnitude. The elastic moduli of our gelsare in the same range as several soft tissues, makingthese materials excellent candidates for a variety oftissue engineering applications.

Acknowledgements

This work was partially funded by and utilized centralfacilities of the NSF-sponsored UMass MRSEC onPolymeric Materials (DMR-0213695). G.N.T. thanksONR for a Young Investigator award, and S.R.B.thanks Dupont for a Dupont Young Professor Award.G.N.T. and S.R.B. both thank 3M for NontenuredFaculty awards.

Page 6: Rheological studies of PLLA–PEO–PLLA triblock copolymer hydrogels

ARTICLE IN PRESSK.A. Aamer et al. / Biomaterials 25 (2004) 1087–10931092

References

[1] Peppas NA, Langer R. New challenges in biomaterials. Science

1994;263:1715–20.

[2] Langer R. Biomaterials in drug delivery and tissue engineering:

One laboratory’s experience. Accounts of Chem Res 2000;33:

94–101.

[3] Hoffman AS. Hydrogels for biomedical applications. Adv Drug

Deliv Rev 2002;43:3–12.

[4] Lee KY, Mooney DJ. Hydrogels for tissue engineering. Chem

Rev 2001;101:1869–79.

[5] Ratner BD, Hoffman AS. In: Series AS, editor. Hydrogels for

Medical and Related Applications, vol. 31. Washington, DC:

American Chemical Society; 1976, p. 1–36.

[6] Younes H, Cohn D. Morphological-Study of Biodegradable Peo/

Pla Block Copolymers. J Biomed Mater Res 1987;21:1301–16.

[7] Kricheldorf HR, Meierhaack J. Polylactones 0.22. Aba Triblock

Copolymers of l-Lactide and Poly(Ethylene Glycol). Makromol

Chemie-Macromol Chem Phys 1993;194:715–25.

[8] Kubies D, Rypacek F, Kovarova J, Lednicky F. Microdomain

structure in polylactide-block-poly(ethylene oxide) copolymer

films. Biomaterials 2000;21:529–36.

[9] Kissel T, Li YX, Unger F. ABA-triblock copolymers from

biodegradable polyester a-blocks and hydrophilic poly(ethylene

oxide) b-blocks as a candidate for in situ forming hydrogel

delivery systems for proteins. Adv Drug Deliv Rev 2002;54:

99–134.

[10] Li SM, Rashkov I, Espartero JL, Manolova N, Vert M. Synthesis,

characterization, and hydrolytic degradation of PLA/PEO/PLA

triblock copolymers with long poly(l-lactic acid) blocks. Macro-

molecules 1996;29:57–62.

[11] Rashkov I, Manolova N, Li SM, Espartero JL, Vert M. Synthesis,

characterization, and hydrolytic degradation of PLA/PEO/PLA

triblock copolymers with short poly(l-lactic acid) chains. Macro-

molecules 1996;29:50–6.

[12] Li YX, Kissel T. Synthesis and properties of biodegradable aba

triblock copolymers consisting of poly(l-lactic acid) or poly

(l-lactic-co-glycolic acid) a-blocks attached to central poly

(oxyethylene) b-blocks. J Control Release 1993;27:247–57.

[13] Li YX, Volland C, Kissel T. In-vitro degradation and bovine

serum-albumin release of the aba triblock copolymers consisting

of poly(l(+) lactic acid), or poly(l(+)lactic acid-co-glycolic acid)

a-blocks attached to central polyoxyethylene b-blocks. J Control

Release 1994;32:121–8.

[14] Saito N, Okada T, Horiuchi H, Murakami N, Takahashi J,

Nawata M, Ota H, Nozaki K, Takaoka K. A biodegradable

polymer as a cytokine delivery system for inducing bone

formation. Nat Biotechnol 2001;19:332–5.

[15] Molina I, Li SM, Martinez MB, Vert M. Protein release from

physically crosslinked hydrogels of the PLA/PEO/PLA triblock

copolymer-type. Biomaterials 2001;22:363–9.

[16] Jeong B, Bae YH, Lee DS, Kim SW. Biodegradable block

copolymers as injectable drug-delivery systems. Nature

1997;388:860–2.

[17] Kwon KW, Park MJ, Bae YH, Kim HD, Char K. Gelation

behavior of PEO-PLGA-PEO triblock copolymers in water.

Polymer 2002;43:3353–8.

[18] Chen XH, McCarthy SP, Gross RA. Synthesis and characteriza-

tion of l-lactide-ethylene oxide multiblock copolymers. Macro-

molecules 1997;30:4295–301.

[19] Liu L, Li CX, Liu XH, He BL. Micellar formation in aqueous

milieu from biodegradable triblock copolymer polylactide/

poly(ethylene glycol)/polylactide. Polym J 1999;31:845–50.

[20] Metters AT, Anseth KS, Bowman CN. Fundamental studies of a

novel, biodegradable PEG-b-PLA hydrogel. Polymer 2000;41:

3993–4004.

[21] Jeong B, Kibbey MR, Birnbaum JC, Won YY, Gutowska A.

Thermogelling biodegradable polymers with hydrophilic back-

bones: PEG-g-PLGA. Macromolecules 2000;33:8317–22.

[22] Lee DS, Shim MS, Kim SW, Lee H, Park I, Chang TY. Novel

thermoreversible gelation of biodegradable PLGA-block-PEO-

block-PLGA triblock copolymers in aqueous solution. Macromol

Rapid Commun 2001;22:587–92.

[23] Lee HT, Lee DS. Thermoresponsive phase transitions of PLA-

block-PEO-block-PLA triblock stereo-copolymers in aqueous

solution. Macromol Res 2002;10:359–64.

[24] Shim MS, Lee HT, Shim WS, Park I, Lee H, Chang T, Kim SW,

Lee DS. Poly(d,l-lactic acid-co-glycolic acid)-b-poly(ethylene

glycol)-b-poly (d,l-lactic acid-co-glycolic acid) triblock copoly-

mer and thermoreversible phase transition in water. J Biomed

Mater Res 2002;61:188–96.

[25] Jeong B, Bae YH, Kim SW. Thermoreversible gelation of PEG–

PLGA–PEG triblock copolymer aqueous solutions. Macromole-

cules 1999;32:7064–9.

[26] Jeong B, Kim SW, Bae YH. Thermosensitive sol–gel reversible

hydrogels. Adv Drug Deliv Rev 2002;54:37–51.

[27] Tanodekaew S, Godward J, Heatley F, Booth C. Gelation of

aqueous solutions of diblock copolymers of ethylene oxide and

d,l-lactide. Macromol Chem Phys 1997;198:3385–95.

[28] Zhong ZY, Dijkstra PJ, Jan FJ, Kwon YM, Bae YH, Kim SW.

Synthesis and aqueous phase behavior of thermoresponsive

biodegradable poly(d,l-3-methylglycolide)-block-poly(ethylene

glycol)-block-poly(d,l-3-methylglycolide) triblock copolymers.

Macromol Chem Phys 2002;203:1797–803.

[29] Fujiwara T, Mukose T, Yamaoka T, Yamane H, Sakurai S,

Kimura Y. Novel thermo-responsive formation of a hydrogel by

stereo-complexation between PLLA-PEG-PLLA and PDLA-

PEG-PDLA block copolymers. Macromol Biosci 2001;1:204–8.

[30] Stockwell R, Meachim G. Adult articular cartilage. London:

Medical P; 1979.

[31] Frank EH, Grodzinsky AJ. Cartilage electromechanics-II. A

continuum model of cartilage electrokinetics and correlation with

experiments. J Biomech Eng 1987;20:629–39.

[32] Yu QL, Zhou JB, Fung YC. Neutral axis location in bending and

Young’s modulus of different layers of arterial wall. Am J Physio

1993;265:H52–60.

[33] Carter FJ, Frank TG, Davies PJ, McLean D, Cuschieri A.

Measurements and modelling of the compliance of human and

porcine organs. Med Image Anal 2001;5:231–6.

[34] Erkamp RQ, Wiggins P, Skovoroda AR, Emelianov SY,

O’Donnell M. Measuring the elastic modulus of small tissue

samples. Ultrasonic Imag 1998;20:17–28.

[35] Hutmacher DW. Scaffold design and fabrication technologies for

engineering tissues: State of the art and future perspectives.

J Biomater Sci Polym 2001;12:107–24.

[36] Pham QT, Russel WB, Thibeault JC, Lau W. Micellar solutions

of associative triblock copolymers: The relationship between

structure and rheology. Macromolecules 1999;32:5139–46.

[37] Tae G, Kornfield JA, Hubbell JA, Lal J. Ordering transitions of

fluoroalkyl-ended poly(ethylene glycol): Rheology and SANS.

Macromolecules 2002;35:4448–57.

[38] Tae G, Kornfield JA, Hubbell JA, Johannsmann D, Hogen-Esch

TE. Hydrogels with controlled, surface erosion characteristics

from self-assembly of fluoroalkyl-ended poly(ethylene glycol).

Macromolecules 2001;34:6409–19.

[39] Wanka G, Hoffmann H, Ulbricht W. Phase diagrams and

aggregation behavior of poly(oxyethylene)-poly(oxypropylene)-

poly(oxyethylene) triblock copolymers in aqueous solutions.

Macromolecules 1994;27:4145–59.

[40] Winter HH, Chambon F. Analysis of the linear viscoelasticity

of a cross-linking polymer at the gel point. J Rheol 1986;30:

367–82.

Page 7: Rheological studies of PLLA–PEO–PLLA triblock copolymer hydrogels

ARTICLE IN PRESSK.A. Aamer et al. / Biomaterials 25 (2004) 1087–1093 1093

[41] Flory PJ. Molecular size distribution in three dimensional

polymers. I. Gelation. J Am Chem Soc 1941;63:3083–90.

[42] de Gennes PG. Relation between percolation theory and elasticity

of gels. J Phys Lett 1976;37:L1–2.

[43] Rubinstein M, Colby RH, Gilmor JR. Dynamic scaling for

polymer gelation. Polym Preprints 1989;30:81–2.

[44] Martin JE, Adolf D. The sol–gel transition in chemical gels. Ann

Rev Phys Chem 1991;42:311–39.

[45] Kantor Y, Webman I. Elastic properties of random percolating

systems. Phys Rev Lett 1984;52:1891–4.

[46] Daoud M. Viscoelasticity near the sol–gel transition. Macro-

molecules 2000;33:3019–22.

[47] Muthukumar M. Screening effect on viscoelasticity near the gel

point. Macromolecules 1989;22:4656–8.

[48] Scanlan JA, Winter HH. Composition dependence of the

viscoelasticity of end-linked poly(dimethylsiloxane) at the gel

point. Macromolecules 1991;24:47–54.

[49] Axelos MAV, Kolb M. Crosslinked biopolymers: experimental

evidence for scalar percolation theory. Phys Rev Lett

1990;64:1457–60.

[50] Matsumoto T, Kawai M, Masuda T. Viscoelastic and SAXS

investigation of fractal structure near the gel point in alginate

aqueous systems. Macromolecules 1992;25:5430–3.

[51] Yu JM, Dubois P, Teyssie P, Jerome R, Blacher S, Brouers F,

L’Homme G. Triblock copolymer based thermoreversible gels. 2.

Analysis of the sol–gel transition. Macromolecules 1996;29:

5384–91.

[52] Richtering HW, Gagnon KD, Lenz RW, Fuller RC, Winter HH.

Physical gelation of a bacterial thermoplastic elastometer.

Macromolecules 1992;25:2429–33.

[53] Lin YG, Mallin DT, Chien JCW, Winter HH. Dynamic

mechanical measurement of crystallization-induced gelation in

thermoplastic elastomeric poly(propylene). Macromolecules 1991;

24:850–4.